Abstract
Strained cyclic allenes are a class of in situ-generated fleeting intermediates that, despite being discovered more than 50 years ago, has received significantly less attention from the synthetic community compared to related strained intermediates. Examples of trapping strained cyclic allenes that involve transition metal catalysis are especially rare. We report the first annulations of highly reactive cyclic allenes with in situ-generated π-allylpalladium species. By varying the ligand employed, either of two isomeric polycyclic scaffolds can be obtained with high selectivity. The products are heterocyclic and sp3-rich and bear two or three new stereocenters. This study should encourage the further development of fragment couplings that rely on transition metal catalysis and strained cyclic allenes for the rapid assembly of complex scaffolds.
In situ-generated strained intermediates that bear a functional group with a preferred linear geometry within a small ring have emerged as valuable synthetic building blocks. Arynes1–6 and cyclic alkynes7,8 (e.g., 1–3, Figure 1A) have been most well-studied with applications spanning the synthesis of heterocycles,9,10 ligands (such as XPhos),11 natural products,12–14 agrochemicals,15 and organic materials.16,17 A related class of strained intermediates discovered around the same time as arynes and cyclic alkynes is strained cyclic allenes, such as 1,2-cyclohexadiene (4).18,19 The cumulated diene confined to a small ring leads to ~30 kcal/mol of strain energy, making cyclic allenes well-suited for strain-promoted reactions.20–22 However, these intermediates have received significantly less attention compared to arynes and cyclic alkynes despite possessing many attractive attributes23 that substantiate their value as building blocks for the synthesis of complex sp3-rich scaffolds. Recent studies of strained cyclic allenes have led to advances in cyclic allene generation protocols,24–26 cycloaddition reactions,27–36 metal-catalyzed processes,24,37,38 trapping in a single-electron process,39 DNA-encoded library synthesis,40 and total synthesis.41
Figure 1.

(A) Strained cyclic intermediates 1–4, (B) approaches for metal-catalyzed reactions of strained cyclic allenes, and (C) overview of current study.
An attractive yet underdeveloped approach to leveraging strained cyclic allenes in synthesis utilizes transition metal catalysis.42,43 Only three reports of such transformations have been demonstrated in the literature.44 In these reactions, an organometallic intermediate 5 is generated catalytically and intercepts a transient cyclic allene (e.g., 6) (Figure 1B). This gives rise to annulated product 7. In the known literature reports, only σ-bound metal species have been used, namely, palladacycle 8,24,45 nickelacycle 9,37 and arylpalladium species 10.38 With the aim of expanding the types of organometallic intermediates that can be used in reactions of strained cyclic allenes, we became interested in the use of π-allylpalladium species 11. The ability of 11 to undergo C–C bond formation at either of its termini46–48 is distinct from the reactivity of organometallic intermediates used previously in cyclic allene annulations. As will be described herein, the use of π-allylpalladium complexes49,50 to trap cyclic allenes provides opportunities to control regiochemistry and stereoselectivity, while building diverse and sp3-rich scaffolds.
In this manuscript, we demonstrate the first use of π-allylmetal complexes to trap strained cyclic allenes (12 and 6, respectively, Figure 1C). By varying the ligand employed, either of two different polycyclic scaffolds, 13 or 14, can be formed preferentially, via the formation of multiple bonds and stereocenters. Studies toward an enantioselective variant are also reported. These studies demonstrate the merger of transition metal catalysis and strained cyclic allenes for the rapid assembly of complex, sp3-rich scaffolds.
To initiate our studies, we investigated the reaction of vinyl benzoxazinone 15 with silyl triflate 16 using palladium catalysis (Table 1). Of note, benzoxazinone 15, a known π-allylmetal precursor,51–55 was selected for these studies as it had been demonstrated to be compatible with fluoride-mediated generation of benzyne.56 Based on known reactivity of 15, two products could plausibly arise, tricycle 17 or tetracycle 18. To generate soluble fluoride ions, CsF and Bu4NOTf57 were used in combination. Furthermore, elevated temperatures were employed to facilitate efficient generation of reactive intermediates. Select key results from reaction optimization are depicted, and additional data are available in the Supporting Information (SI).
Table 1.
Select Results from Optimization Studies
|
Reactions conditions: 15 (3.0 equiv), 16 (1.0 equiv), catalyst (5 mol %), ligand (as shown), CsF (10 equiv), Bu4NOTf (5.0 equiv), additive (as shown), DMF (0.050 M), 70 °C, 2 h.
Yield determined by an isolation experiment.
Ratio of 17:18 and dr were determined by 1H NMR analysis. In entries where 17 is the major product, dr = 11:1; in entries where 18 is the major product dr = >20:1.
[Pd] (10 mol %), 0.025 M.
The use of the dialkylbiaryl ligand DavePhos, which had previously been identified as a competent ligand in a cyclic allene annulation,38 did not lead to a productive reaction (Table 1, entry 1). However, switching to the PPh3 ligand gave tricyclic annulation product 17, albeit in a low yield of 20% (entry 2). Alternatively, the use of the bidentate ferrocene ligand dppf delivered constitutional isomer 18 as the major product (entry 3). With initial results for the selective formation of annulation product 17 or 18, we performed optimization to selectively access each of the constitutional isomers. For practical reasons, we elected to pursue the use of Pd(PPh3)4, which gave 17 in slightly improved yield and excellent selectivity (entry 4; compared with entry 2). Because a significant amount of decomposition of benzoxazinone 15 was observed under these conditions,58 we explored additives to mitigate deleterious side reactions. Empirically, the addition of water led to significant improvement, generating tricycle 17 in 65% yield after 2 h (entry 5). Extensive efforts were made to selectively generate tetracycle 18. We ultimately identified that the use of Buchwald precatalyst dppf Pd G3 and water as an additive delivered 18 in 73% yield after 2 h (entry 6; compared with entry 3). Thus, either tricycle 17 or tetracycle 18 could be accessed by the judicious choice of ligand. It should be noted that the annulation of benzyne with benzoxazinone 15 only delivers the tetracyclic counterpart to 18, without formation of the tricyclic counterpart to 17.56
Having identified suitable reaction conditions for the selective formation of either constitutional isomer 17 or 18, we examined the scope of the methodology beginning with the formation of tricyclic products 21 (Figure 2). The parent substrate, along with its N-mesyl derivative, each gave a 60% yield of tricyclic product (17 and 22, respectively). A variety of substituents on the arene (R2 = Me, F, OMe, Ph, 2-thiophene) were also tolerated, as evidenced by the formation of tricycles 23–29.59 The use of a naphthalene-derived substrate proceeded smoothly to furnish 30 in 64% yield. Additionally, we evaluated a benzoxazinone bearing an isopropenyl group, which afforded tricycle 31. Lastly, heterocyclic allene precursors (i.e., 20) could be employed under modified reaction conditions to enable further modulation of the product structure. More specifically, interception of oxa-29 and aza-derivatives28 of 20 delivered heterocycles 32 and 33, respectively. It should be noted that formation of the corresponding tetracyclic products was either negligible or not observed for the results shown in Figure 2. Additionally, in most cases, the observed diastereoselectivities are synthetically useful.
Figure 2.

Scope of the annulation reaction using PPh3 as the ligand. Reactions performed on 0.05 mmol scale with 19 (3.0 equiv) and 20 (1.0 equiv) at 0.050 M. Yields reflect an average of two isolation experiments. a Reactions were performed with CsF (5.0 equiv) in MeCN (0.050 M) at 60 °C for 6 h, without Bu4NOTf additive.
Subsequently, the scope of the methodology for preparing tetracycles utilizing the dppf Pd G3 catalyst was assessed (Figure 3). Although our conditions shown in Table 1, entry 6 were generally useful, some modifications were made to obtain optimal yields based on empirical observations. Beginning with the parent substrate combination, 18 was isolated in 70% yield and >20:1 dr. Variation of the N-substituent, as well as steric and electronic perturbations of the aryl ring, were tolerated as seen by the formation of 35–43 in good yields. Finally, trapping of an oxacyclic allene gave tetracycle 44. Notably, the annulations to afford tetracycles 34 proceed via the formation of three new bonds and three stereocenters with excellent diastereoselectivity (>20:1 dr in all cases).
Figure 3.

Scope of the annulation reaction using dppf as the ligand. Reactions performed on 0.05 mmol scale with 19 (3.0 equiv) and 20 (1.0 equiv) at 0.025 M. Yields reflect an average of two isolation experiments. For all isolations, dr = >20:1 and ratios of 34:21 range from 6.3:1 to >20:1; see SI for details. a Reaction was performed at 100 °C. b Reaction was performed in MeCN (0.0125 M) at 30 °C for 14 h, without Bu4NOTf additive.
Plausible mechanisms for the transformations we have developed are depicted in Figure 4. The Pd(0) catalyst coordinates to 45. Oxidative addition of the Pd(0) catalyst into the allylic C–O bond of 46 followed by decarboxylation affords zwitterionic π-allylpalladium intermediate 12.51 Ligand-controlled migratory insertion of cyclic allene 6 (formed from fluoride-mediated 1,2-elimination of silyl triflate 20) gives π-allylpalladium intermediates 47 and 48.60 These intermediates arise from C–C bond formation at either of the two sites of reactivity of π-allylpalladium species 12. The branched or linear selectivity of the migratory insertion is proposed to be a result of the ligands employed.61 Cyclizations of 47 and 48 afford tricycle 1362 and tetracycle 14, respectively, through the formation of one or two new bonds.
Figure 4.

Proposed catalytic cycle.
Given the notable impact of ligands on selectivity in these transformations, the feasibility of an enantioselective variant was investigated using vinyl benzoxazinone 49 and oxacyclic allene precursor 50. Notably, the use of two chiral racemic starting materials would represent a departure from previous studies that use an achiral trapping partner.37,38 Select key results are shown in Table 2, with additional information being available in the SI.63 The use of PHOX ligand L1 and Trost’s ligand L2, which are commonly employed in π-allylpalladium chemistry,50 provided no desired reactivity (entries 1 and 2, respectively). However, using phosphoramidite ligand L3 (entry 3), annulation product 51 was formed as the major isomer, thus demonstrating the feasibility of rendering the annulation to form both products enantioselective. Given the greater structural complexity of isomer (−)-52, we elected to assess other chiral diphosphine ligands in pursuit of (−)-52. The use of Phanephos L4, Segphos ligand L5, and Josiphos ligand L6 each gave the desired product (−)-52 in varying yields and enantioselectivities (entries 4, 5, and 6, respectively). The employment of Walphos ligand L7 and Mandyphos ligand L8 each resulted in increased yields and enantioselectivities (entries 7 and 8, respectively). Of these, use of L8 provided superior results and delivered (−)-52 in 54% yield and 70% ee.64 We surmise that the absolute configuration of (−)-52 is as depicted, on the basis of an X-ray crystal structure obtained for carbocyclic derivative (−)-43. Although improving the enantioselectivity was challenging in our initial efforts,65 the results shown provide the most structurally complex products accessible from the merger of strained cyclic allenes and asymmetric catalysis to date. As enantioselective transformations of fleeting strained cyclic intermediates remain exceedingly rare, we hope these results will promote further efforts in this area.
Table 2.
Evaluation of Select Chiral Ligands in the Annulation Reaction
|
Reaction conditions: 49 (3.0 equiv), 50 (1.0 equiv), Pd2(dba)3 (5 mol%), ligand (20 mol%), CsF (10 equiv), H2O (9.0 equiv), MeCN (0.0125 M), 30 °C, 14 h.
Yield determined by an isolation experiment.
See SI, section F for details and crystallographic data.
We have developed the first strategy for engaging strained cyclic allenes with π-allyl metal species. These are rare examples of cyclic allene reactions in transition metal-catalyzed processes. By the judicious choice of ligands, selectivity can be controlled, leading to the preferential formation of tri- or tetracyclic heterocycles with significant structural complexity. Of note, the transformations proceed by formation of either two or three new bonds and two or three new stereocenters. We also demonstrate the feasibility of an enantioselective variant, thus providing a crucial proof-of-principle study for further asymmetric reaction development. Overall, the union of strained cyclic allenes and π-allylpalladium chemistry expands the types of organometallic intermediates that can be used in cyclic allene reactions to access complex sp3-rich scaffolds. We hope this study promotes the further development and understanding of transition metal-catalyzed reactions of strained cyclic allenes.
Supplementary Material
ACKNOWLEDGMENTS
The authors are grateful to the NIH-NIGMS (R35 GM139593 for N.K.G. and T320GM136614 for M.S.M.), the Trueblood Family (for N.K.G.), the Foote Family (for M.S.M., S.M.A.), the Stone Family (for D.C.W.), and the California Tobacco-Related Disease Research Program (for S.M.A.). We thank Dr. Saeed Khan (UCLA) for X-ray analysis. These studies were supported by shared instrumentation grants from the NSF (CHE-1048804), the National Center for Research Resources (S10RR025631), and the NIH Office of Research Infrastructure Programs (S10OD028644).
Footnotes
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c03102.
Detailed experimental procedures, compounds characterization data, and crystallographic data for (−)-43 (PDF)
Accession Codes
CCDC 2238913 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
Complete contact information is available at: https://pubs.acs.org/10.1021/jacs.3c03102
The authors declare no competing financial interest.
Contributor Information
Dominick C. Witkowski, Department of Chemistry and Biochemistry, University of California, Los Angeles, California 90095, United States
Matthew S. McVeigh, Department of Chemistry and Biochemistry, University of California, Los Angeles, California 90095, United States
Georgia M. Scherer, Department of Chemistry and Biochemistry, University of California, Los Angeles, California 90095, United States
Sarah M. Anthony, Department of Chemistry and Biochemistry, University of California, Los Angeles, California 90095, United States
Neil K. Garg, Department of Chemistry and Biochemistry, University of California, Los Angeles, California 90095, United States
REFERENCES
- (1).Roberts JD; Simmons HE; Carlsmith LA; Vaughan CW Rearrangement in the reaction of chlorobenzene-1-C14 with potassium amide. J. Am. Chem. Soc 1953, 75, 3290–3291. [Google Scholar]
- (2).Wenk HH; Winkler M; Sander W One century of aryne chemistry. Angew. Chem., Int. Ed 2003, 42, 502–528. [DOI] [PubMed] [Google Scholar]
- (3).Dhokale RA; Mhaske SB Transition-metal-catalyzed reactions involving arynes. Synthesis 2018, 50, 1–16. [Google Scholar]
- (4).Bhunia A; Yetra SR; Biju AT Recent advances in transition-metal-free carbon–carbon and carbon–heteroatom bond-forming reaction using arynes. Chem. Soc. Rev 2012, 41, 3140–3152. [DOI] [PubMed] [Google Scholar]
- (5).Goetz AE; Garg NK Enabling the use of heterocyclic arynes in chemical synthesis. J. Org. Chem 2014, 79, 846–851. [DOI] [PubMed] [Google Scholar]
- (6).Hosoya T; Yoshida S; Nakamura Y Recent advances in synthetic hetaryne chemistry. Hetereocycles 2019, 98, 1623–1677. [Google Scholar]
- (7).Scardiglia F; Roberts JD Evidence for cyclohexyne as an intermediate in the coupling of phenyllithium with 1-chlorocyclohexene. Tetrahedron 1957, 1, 343–344. [Google Scholar]
- (8).For a review describing a wide range of in situ generated strained intermediates, see:; Shi J; Li L; Li Y o-Silylaryl triflates: A journey of Kobayashi aryne precursors. Chem. Rev 2021, 121, 3892–4044. [DOI] [PubMed] [Google Scholar]
- (9).Dubrovskiy AV; Markina NA; Larock RC Use of benzynes for the synthesis of heterocycles. Org. Biomol. Chem 2013, 11, 191–218. [DOI] [PubMed] [Google Scholar]
- (10).Goetz AE; Shah TK; Garg NK Pyridynes and indolynes as building blocks for functionalized heterocycles and natural products. Chem. Commun 2015, 51, 34–45. [DOI] [PubMed] [Google Scholar]
- (11).Mauger CC; Mignani GA An efficient and safe procedure for the large-scale Pd-catalyzed hydrazonation of aromatic chlorides using Buchwald technology. Org. Proc. Res. Dec 2004, 8, 1065–1071. [Google Scholar]
- (12).Tadross PM; Stoltz BM A comprehensive history of arynes in natural product total synthesis. Chem. Rev 2012, 112, 3550–3577. [DOI] [PubMed] [Google Scholar]
- (13).Gampe CM; Carreira EM Arynes and cyclohexyne in natural product synthesis. Angew. Chem., Int. Ed 2012, 51, 3766–3778. [DOI] [PubMed] [Google Scholar]
- (14).Takikawa H; Nishii A; Sakai T; Suzuki K Aryne-based strategy in the total synthesis of naturally occurring polycyclic compounds. Chem. Soc. Rev 2018, 47, 8030–8056. [DOI] [PubMed] [Google Scholar]
- (15).Schleth F; Vettiger T; Rommel M; Tobler H Process for the preparation of pyrazole carboxylic acid amides. International patent WO 2011131544 A1, Oct 27, 2011. [Google Scholar]
- (16).Lin JB; Shah TJ; Goetz AE; Garg NK; Houk KN Conjugated trimeric scaffolds accessible from indolyne cyclotrimerizations: Synthesis, structures, and electronic properties. J. Am. Chem. Soc 2017, 139, 10447–10455. [DOI] [PubMed] [Google Scholar]
- (17).Darzi ER; Barber JS; Garg NK Cyclic alkyne approach to heteroatom-containing polycyclic aromatic hydrocarbon scaffolds. Angew. Chem., Int. Ed 2019, 58, 9419–9424. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).Wittig G; Fritze P On the intermediate occurrence of 1,2-cyclohexadiene. Angew. Chem., Int. Ed 1966, 5, 846. [Google Scholar]
- (19).Moser WR The reactions of gem-dihalocyclopropanes with organometallic reagents Ph.D. Dissertation, Massachusetts Institute of Technology: Cambridge, MA, 1964. [Google Scholar]
- (20).Christl M Cyclic allenes up to seven-membered rings. Modern Allene Chemistry; Krause N, Kashmi SAK, Eds.; Wiley-VCH: Weinheim, 2004; pp 243–357. [Google Scholar]
- (21).Angus RO Jr.; Schmidt MW; Johnson RP Small-ring cyclic cumulenes: Theoretical studies of the structure and barrier to inversion of cyclic allenes. J. Am. Chem. Soc 1985, 107, 532–537. [Google Scholar]
- (22).Daoust KJ; Hernandez SM; Konrad KM; Mackie ID; Winstanley J; Johnson RP Strain estimates for small-ring cyclic allenes and butatrienes. J. Org. Chem 2006, 71, 5708–5714. [DOI] [PubMed] [Google Scholar]
- (23). Strained cyclic allenes have several attractive attributes that include their (a) generation under mild reaction conditions; (b) ability to undergo two or one electron processes leading to a wide scope of reactions; (c) inherent chirality, which can allow for stereoselective transformations; (d) participation in regioselective reactions based on substituent effects; and (e) relatively understudied nature, which allows numerous opportunities for discovery.
- (24).Quintana I; Peña D; Pérez D; Guitián E Generation and reactivity of 1,2-cyclohexadiene under mild reaction conditions. Eur. J. Org. Chem 2009, 2009, 5519–5524. [Google Scholar]
- (25).McVeigh MS; Kelleghan AV; Yamano MM; Knapp RR; Garg NK Silyl tosylate precursors to cyclohexyne, 1,2-cyclohexadiene, and 1,2-cycloheptadiene. Org. Lett 2020, 22, 4500–4504. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (26).Hioki Y; Mori A; Okano K Steric effects on deprotonative generation of cyclohexynes and 1,2-cyclohexadienes from cyclohexenyl triflates by magnesium amides. Tetrahedron 2020, 76, 131103. [Google Scholar]
- (27).Barber JS; Styduhar ED; Pham HV; McMahon TC; Houk KN; Garg NK Nitrone cycloadditions of 1,2-cyclohexadiene. J. Am. Chem. Soc 2016, 138, 2512–2515. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (28).Barber JS; Yamano MM; Ramirez M; Darzi ER; Knapp RR; Liu F; Houk KN; Garg NK Diels–Alder cycloadditions of strained azacyclic allenes. Nat. Chem 2018, 10, 953–960. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Yamano MM; Knapp RR; Ngamnithiporn A; Ramirez M; Houk KN; Stoltz BM; Garg NK Cycloadditions of oxacyclic allenes and a catalytic asymmetric entryway to enantioenriched cyclic allenes. Angew. Chem., Int. Ed 2019, 58, 5653–5657. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (30).Ramirez M; Svatunek D; Liu F; Garg NK; Houk KN Origins of endo selectivity in Diels–Alder reactions of cyclic allene dienophiles. Angew. Chem., Int. Ed 2021, 60, 14989–14997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Lofstrand VA; West FG Efficient trapping of 1,2-cyclohexadienes with 1,3-dipoles. Chem. - Eur. J 2016, 22, 10763–10767. [DOI] [PubMed] [Google Scholar]
- (32).Lofstrand VA; McIntosh KC; Almehmadi YA; West FG Strain-activated Diels–Alder trapping of 1,2-cyclohexadienes: Intramolecular capture by pendant furans. Org. Lett 2019, 21, 6231–6234. [DOI] [PubMed] [Google Scholar]
- (33).Almehmadi YA; West FG A mild method for the generation and interception of 1,2-cycloheptadienes with 1,3-dipoles. Org. Lett 2020, 22, 6091–6095. [DOI] [PubMed] [Google Scholar]
- (34).Wang B; Constantin M-G; Singh S; Zhou Y; Davis RL; West FG Generation and trapping of electron-deficient 1,2-cyclohexadienes. Unexpected hetero-Diels–Alder reactivity. Org. Biomol. Chem 2021, 19, 399–405. [DOI] [PubMed] [Google Scholar]
- (35).Jankovic CL; West FG 2 + 2 Trapping of acyloxy-1,2-cyclohexadienes with styrenes and electron-deficient olefins. Org. Lett 2022, 24, 9497–9501. [DOI] [PubMed] [Google Scholar]
- (36).Nendel M; Tolbert LM; Herring LE; Islam MN; Houk KN Strained allenes as dienophiles in the Diels–Alder reaction: An experimental and computational study. J. Org. Chem 1999, 64, 976–983. [DOI] [PubMed] [Google Scholar]
- (37).Yamano MM; Kelleghan AV; Shao Q; Giroud M; Simmons BJ; Li B; Chen S; Houk KN; Garg NK Interception fleeting cyclic allenes with asymmetric nickel catalysis. Nature 2020, 586, 242–247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (38).Kelleghan AV; Witkowski DC; McVeigh MS; Garg NK Palladium-catalyzed annulations of strained cyclic allenes. J. Am. Chem. Soc 2021, 143, 9338–9342. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (39).McVeigh MS; Garg NK Interception of 1,2-cyclohexadiene with TEMPO radical. Tetrahedron Lett 2021, 87, 153539–153543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (40).Westphal MV; Hudson L; Mason JW; Pradeilles JA; Zécri FJ; Briner K; Schreiber SL Water-compatible cycloadditions of oligonucleotide-conjugated strained allenes for DNA-encoded library synthesis. J. Am. Chem. Soc 2020, 142, 7776–7782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (41).Ippoliti FM; Adamson NJ; Wonilowicz LG; Nasrallah DJ; Darzi ER; Donaldson JS; Garg NK Total synthesis of lissodendoric acid A via stereospecific trapping of a strained cyclic allene. Science 2023, 379, 261–265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (42).Anthony SM; Wonilowicz LG; McVeigh MS; Garg NK Leveraging fleeting strained intermediates to access complex scaffolds. JACS Au 2021, 1, 897–912. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (43).Spence KA; Tena Meza A; Garg NK Merging metals and strained intermediates. Chem. Catalysis 2022, 2, 1870–1879. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (44). In contrast, over 50 studies of cycloadditions or nucleophilic additions of 6-membered strained cyclic allenes are available in the literature. Similarly, arynes have been widely used in metal-catalyzed processes (for reviews, see refs 3, 8, 42, and 43).
- (45).For a report of palladacycle 8 being invoked in strained intermediate cocyclotrimerizations, see:; Peña D; Pérez D; Guitián E; Castedo L Selective palladium-catalyzed cocyclotrimerization of arynes with dimethyl acetylenedicarboxylate: A versatile method for the synthesis of polycyclic aromatic hydrocarbons. J. Org. Chem 2000, 65, 6944–6950. [DOI] [PubMed] [Google Scholar]
- (46).Baker R π-Allylmetal derivatives in organic synthesis. Chem. Rev 1973, 73, 487–530. [Google Scholar]
- (47).Trost B Metal catalyzed allylic alkylations: Its development in the Trost laboratories. Tetrahedron 2015, 71, 5708–5733. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (48).Dutta S; Bhattacharya T; Werz DB; Maiti D Transition-metal-catalyzed C–H allylation reactions. Chem 2021, 7, 555–605. [Google Scholar]
- (49).Trost BM; Crawley ML Asymmetric transition-metalcatalyzed allylic alkylations: Applications in total synthesis. Chem. Rev 2003, 103, 2921–2944. [DOI] [PubMed] [Google Scholar]
- (50).Pàmies O; Margalef J; Cañellas S; James J; Judge E; Guiry PJ; Moberg C; Bäckvall J-E; Pfaltz A; Pericàs MA; Diéguez M Recent advances in enantioselective Pd-catalyzed allylic substitution: From design to applications. Chem. Rev 2021, 121, 4373–4505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (51).Wang C; Pahadi N; Tunge JA Decarboxylative cyclizations and cycloadditions of palladium-polarized aza-ortho-xylylenes. Tetrahedron 2009, 65, 5102–5109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (52).Li T-R; Wang Y-N; Xiao W-J; Lu L-Q Transition-metal-catalyzed cyclization reactions using vinyl and ethynyl benzoxazinones as dipole precursors. Tetrahedron Lett 2018, 59, 1521–1530. [Google Scholar]
- (53).Zuo L; Liu T; Chang X; Guo W An update of transition metal-catalyzed decarboxylative transformations of cyclic carbonates and carbamates. Molecules 2019, 24, 3930–3945. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (54).Tian Y; Duan M; Liu J; Fu S; Dong K; Yue H; Hou Y; Zhao Y Recent advances in metal-catalyzed decarboxylative reactions of vinyl benzoxazinones. Adv. Synth. Catal 2021, 363, 4461–4474. [Google Scholar]
- (55).You Y; Li Q; Zhang Y-P; Zhao J-Q; Wang Z-H; Yuan W-C Advances in palladium-catalyzed decarboxylative cycloadditions of cyclic carbonates, carbamates, and lactones. ChemCatChem 2022, 14, No. e202101887. [Google Scholar]
- (56).Duan S; Cheng B; Duan X; Bao B; Li Y; Zhai H Synthesis of cis-5,5a,6,10b-tetrahydroindeno[2,1-b]indoles through palladium-catalyzed decarboxylative coupling of vinyl benzoxazinanones with arynes. Org. Lett 2018, 20, 1417–1420. [DOI] [PubMed] [Google Scholar]
- (57).Feng M; Tang B; Wang N; Xu H-X; Jiang X Ligand controlled regiodivergent C1 insertion on arynes for construction of phenanthridinone and acridone alkaloids. Angew. Chem., Int. Ed 2015, 54, 14960–14964. [DOI] [PubMed] [Google Scholar]
- (58).We hypothesize that fluoride-mediated detosylation of the vinyl benzoxazinone could be responsible for decomposition; see:; Carato P; Yous S; Sellier D; Poupaert JH; Lebegue N; Berthelot P Efficient and selective deprotection method for N-protected 2(3H)benzoxazolones and 2(3H)-benzothiazolones. Tetrahedron 2004, 60, 10321–10324. [Google Scholar]
- (59). Yield of 26 represents a combined yield of tri- and tetracyclic isomers (16:1); see Supporting Information, section C for details.
- (60).Zimmer R; Dinesh CU; Nandanan E; Khan FA Palladium-catalyzed reactions of allenes. Chem. Rev 2000, 100, 3067–3125. [DOI] [PubMed] [Google Scholar]
- (61). How specific ligands influence regioselectivity in the migratory insertion step of the proposed mechanism is not presently understood. However, ligand-guided regioselectivity in trapping π-allylpalladium species with conventional nucleophiles is not uncommon; see refs 46, 47.
- (62). We hypothesize that the trans diastereomer is obtained as the major product due to a destabilizing steric interaction in the presumed transition state leading to the cis diastereomer. Specifically, a steric interaction between the vinyl group and the π-allyl fragment may disfavor formation of the cis product.
- (63). Empirically, 49 and 50 gave optimal yields and ee in comparison to other substrate combinations and, therefore, were used in subsequent optimization.
- (64). 49 was recovered (<30%) in 10% ee. This finding will be subject to further investigation.
- (65). Efforts to improve ee by varying easily modifiable parameters, such as solvent, concentration, and stoichiometry, led to no improvement. Future efforts beyond the present proof-of-principle study will be aimed toward improving enantioselectivity.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
