Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2025 Jun 5.
Published in final edited form as: J Am Chem Soc. 2024 May 20;146(22):15420–15427. doi: 10.1021/jacs.4c03369

Access to Complex Scaffolds Through [2+2] Cycloadditions of Strained Cyclic Allenes

Matthew S McVeigh 1,, Jacob P Sorrentino 1,, Allison T Hands 1, Neil K Garg 1
PMCID: PMC11459239  NIHMSID: NIHMS2026452  PMID: 38768558

Abstract

We report the strain-induced [2+2] cycloadditions of strained cyclic allenes for the assembly of highly substituted cyclobutanes. By judicious choice of trapping agent, complex scaffolds bearing heteroatoms, fused rings, contiguous stereocenters, spirocycles, and quaternary centers are ultimately accessible. Moreover, we show that the resulting cycloadducts can undergo thermal isomerization. This study provides an alternative strategy to photochemical [2+2] cycloadditions for accessing highly functionalized cyclobutanes, while validating the use of underexplored strained intermediates for the assembly of complex architectures.

Graphical Abstract

graphic file with name nihms-2026452-f0001.jpg

INTRODUCTION

Stereochemically complex scaffolds have become increasingly desirable in the discovery of bioactive molecules.1,2 One such motif, the cyclobutane, is present in biologically relevant compounds and many natural products.3,4,5,6 Its rigid scaffold possesses distinct exit vectors that can be used to ultimately improve biological properties.3,6 Three representative cyclobutane-containing compounds are shown in Figure 1A: antiviral agent Lobucavir (1),7 kinase inhibitor 2,8 cannabinoid 3,9 and natural products such as welwitindolinone A isonitrile (4),10,11 arnamial (5),12 and scopariusicide A (6).13

Figure 1.

Figure 1.

(A) Cyclobutane-containing, biologically relevant molecules 16. (B) Strained cyclic intermediates 7a–d and 8. (C) Overview of the current study.

Cyclobutanes are commonly introduced synthetically using feedstock fragments or cycloaddition reactions. Methods to introduce cyclobutanes have been reviewed3,14,15 and several elegant approaches have recently been reported, including those by Wilkerson–Hill,16 Yoon,17,18 and Brown.19,20,21 With regard to cycloadditions, photochemical [2+2] cycloadditions, which leverage in situ-generated photoexcited intermediates, are most common.22,23 Notably, harnessing these intermediates for regio- and stereoselective syntheses intermolecularly can be difficult, necessitating specialized reagents or catalysts.24 As such, new methods for the synthesis of highly functionalized cyclobutanes bearing multiple stereocenters remain coveted.

An attractive yet underutilized approach to access highly functionalized cyclobutanes relies on strain-promoted reactivity. For example, arynes (e.g., 7a and 7b)25,26,27,28,29,30,31 and non-aromatic cyclic alkynes (e.g., 7c and 7d)32,33,34 are highly reactive, in situ-generated intermediates that can readily engage in [2+2] cycloadditions35 due to their significant strain energies, typically ranging from 40–50 kcal/mol (Figure 1B).36 A comparatively less well-studied class of compounds is strained cyclic allenes 8.37,38,39 Like cyclic alkynes, strained cyclic allenes 8 contain a functional group that typically prefers a linear geometry, but is bent due to ring constraint, leading to significant strain energy (~30 kcal/mol).40,41 Although historically understudied, cyclic allenes 8 are emerging building blocks in the synthetic community.42,43,44,45,46,47,48,49 Recent advances include cycloaddition chemistry by West,44,50,51,52,53,54 the synthesis of DNA-encoded libraries by Schreiber,55 and several studies from our laboratory,45,56,57,58,59,60,61,62,63,64,65,66,67 including the recent use of strained cyclic allenes in total synthesis.68 With regard to [2+2] cycloadditions of strained cyclic allenes, seminal studies by Christl and others validated reactivity.69,70,71,72,73 More recently, elegant studies by West have demonstrated [2+2] cycloadditions of acetoxy-substituted carbocyclic allenes,53 as well as intramolecular trappings,54 using silyl bromide precursors to cyclic allenes. Our laboratory has also reported scattered examples of [2+2] cycloadditions of cyclic allenes.56,57,62,66,74

As part of a general program aimed at exploring the use of unconventional strained intermediates for the construction of complex scaffolds, we sought to investigate strain-driven [2+2] cycloadditions of cyclic allenes using the reaction design shown in Figure 1C. Cyclic allenes 8, generated via fluoride-mediated 1,2-elimination of silyl triflates 9, would engage with highly substituted alkenes 10 to afford densely functionalized cycloadducts 11. As in other [2+2] cycloadditions, two new C–C bonds would form. Moreover, through the judicious choice of trapping partner, features indicative of structural complexity,75,76 such as heterocyclic frameworks,77 spiro centers, contiguous stereocenters, and vicinal quaternary centers could be accessible under mild conditions. The transformation is accomplished without the need for metal-, Lewis acid-, or Brønsted acid catalysts; directing or auxiliary groups are also not required. Herein, we report the success of this approach to access highly functionalized cyclobutanes, which pushes the limits of complexity accessible from strained cyclic allene intermediates.

RESULTS AND DISCUSSION

To initiate these studies, we surveyed a range of exocyclic alkenes 14 in cyclic allene trapping experiments (Figure 2). Exocyclic alkenes are historically understudied in strained cyclic allene [2+2] cycloadditions, but if generally useful, would lead to the formation of desirable spirocyclic products.6,78 Silyl triflate 12, a precursor to oxacyclic allene 13,57,79 was treated with CsF in the presence of exocyclic alkenes 14 at ambient temperatures to afford cycloadducts 15. Reaction outcomes were then determined by 1H NMR analysis. Notably, using unactivated alkene 16 afforded cycloadduct 17 in 42% yield (entry 1). Improved yields were seen when substrates containing an alkene with delocalized π-electrons were employed. For example, the electron-deficient alkene in α,β-unsaturated lactone 18 reacted to give cycloadduct 19 in high yield and moderate diastereoselectivity (entry 2). Similarly, electron-rich alkenes 20, 22, and 24 bearing heteroatoms could be employed and afforded cycloadducts 21, 23, and 25, respectively (entries 3–5).80 Lastly, leveraging styrene 26 in the transformation provided cycloadduct 27 in 94% yield and 2.8:1 diastereomeric ratio (dr) (entry 6). Based on the synthetic efficiency observed, we elected to further pursue trappings of substrates related to dienamine 24 and styrene 26 and further push the limits of complexity in the products.81

Figure 2.

Figure 2.

Survey of olefin partners 14 in [2+2] cycloaddition of oxacyclic allene 13. Reactions conditions are as follows: 12 (1.0 equiv, 0.050 mmol), trapping partners 14 (5.0 equiv), CsF (5.0 equiv), MeCN (0.1 M), 23 °C, 3–4 h. Yields and dr determined by 1H NMR analysis of the crude reaction mixture using mesitylene as an external standard.

Several features of the reactions shown in Figure 2 should be noted: a) the reactions occur at ambient temperatures, b) yields and diastereomeric ratios are generally synthetically useful, although some variation in the latter was observed,80 c) excellent regioselectivity is observed in each transformation,82 d) the general transformation allows for the formation of two new C–C bonds and typically two contiguous stereocenters, with access to further complexity being foreseeable based on the potential use of modified substrates, e) despite that strained cyclic allenes have historically been avoided, they can be used to readily access products containing three or more rings, including the desired spirocyclobutane and various heterocycles.

Having selected two promising classes of cycloaddition partners (i.e., 24 and 26)81 for further investigation, we assessed the scope of the methodology with respect to dienamines 28 (Figure 3). Silyl triflate 12 and dienamines 28 were combined in the presence of CsF to afford cyclobutanes 29. During early optimization efforts, it was found that substoichiometric amounts of tetrabutylammonium triflate (Bu4NOTf) in dichloromethane as solvent was preferred when compared to acetonitrile.83 Tetrabutylammonium salts are known to improve the solubility of fluoride ion in organic solvents84,85 and have been beneficial in prior studies of cyclic allene trapping reactions.60,62,64,68 Tosyl (Ts) and tert-butyloxycarbonyl (Boc) N-substituted substrates 24 and 30 could be employed, providing cycloadducts 25 and 31 in high yields and diastereoselectivities (entry 1). The use of tetrahydroazepine derivative 32 afforded 33 in 88% yield and 10:1 dr (entry 2). Increasingly substituted dienamines were tested as well. When 34 and 36 were utilized in the reaction, products 35 and 37 were obtained, which possess additional substitution on the cyclobutyl ring (entries 3 and 4, respectively). Notably, the trapping with 34 proceeds with moderate stereospecificity82,86 and results in cyclobutane 35, containing three contiguous stereocenters. Substitution on the ring of the dienamine trapping partner was also tolerated. Employing iodosubstituted substrate 38 in the cycloaddition led to product 39, whereas employment of 40 bearing a gemdimethyl group furnished 41 (entries 5 and 6, respectively). The erosion in stereoselectivity when comparing the formation of 41 to 25 is attributed to the increased steric bulk from the gem-dimethyl group present in 41.87 Of note, cycloadduct 41 contains vicinal quaternary carbons.

Figure 3.

Figure 3.

Scope of [2+2] cycloadditions with dienamine trapping partners 28. Reaction conditions unless otherwise stated: 12 (1.0 equiv, 0.15 mmol), trapping partners 28 (3.0 equiv), CsF (5.0 equiv), Bu4NOTf (50 mol%), CH2Cl2 (0.1 M), 23 °C, 24 h. Yields reflect an average of two isolation experiments; dr determined by 1H NMR analysis. aThe 2.2:1 ratio reflects the ratio of the major diastereomer (2.2) to the sum of all minor diastereomers (1). bReaction was performed with CsF (5.0 equiv) in MeCN (0.1 M) at 23 °C for 24 h.

The use of styrenyl trapping partners (Figure 4, i.e., 42 or 43) in the methodology, including heterocyclic substrates, delivered structurally complex spiro cyclobutyl products bearing four or more interconnected rings. Using similar reaction conditions to those mentioned earlier (see Figure 3), employment of styrene 26 provided cycloadduct 27 in 80% yield and 4.0:1 dr (entry 1). Substrates 46 and 48, which possess six-membered heterocyclic rings, could be employed, thus giving rise to cycloadducts 47 and 49 (entries 2 and 3). Moreover, employment of substrate 50 furnished pentacyclic product 51 containing a spiro tetrahydrocarbazole in 86% yield and 15:1 dr (entry 4). Substitution on the exocyclic alkene was also tolerated, as demonstrated by the cycloaddition of tri- and tetra-substituted alkenes 52, 54, 56, and 58. The use of these substrates generated cycloadducts 53, 55, 57, and 59 (entries 5–8, respectively), which possess interesting features, such as three contiguous stereocenters (i.e., 53), as well as vicinal fully substituted carbons (i.e., 55, 57 and 59). Lastly, we were delighted to find that trapping of the strained cyclic allene intermediate with pyridyl-substituted alkene 60 bearing a gem-dimethyl unit afforded cycloadduct 61 in high yield and excellent diastereoselectivity (entry 9). The structure of this product, bearing vicinal quaternary carbons, was verified by X-ray crystallography.

Figure 4.

Figure 4.

Scope of [2+2] cycloadditions with (het)aryl-substituted alkenes 42 and 43. Reaction conditions unless otherwise stated: 12 (1.0 equiv, 0.15 mmol), trapping partners 42 or 43 (3.0 equiv), CsF (5.0 equiv), Bu4NOTf (50 mol%), CH2Cl2 (0.1 M), 23 °C, 24 h. Yields reflect an average of two isolation experiments; dr determined by 1H NMR analysis. aTrapping partner (5.0 equiv). bBu4NOTf (20 mol%). cSee Supporting Information, section F for details and crystallographic data.

Although our primary focus was on the use of heterocyclic allenes to build medicinally-relevant heterocycles, we also examined carbocyclic allenes (Figure 5). Treatment of silyl triflate 62 with dienamine 24 under conditions previously developed for carbocyclic allene generation,45 delivered 64 in 75% yield and 7.1:1 dr, presumably via 1,2-cyclohexadiene (63). We were also intrigued by the possibility of using a larger, less strained cyclic allene in this methodology.41 As such, silyl tosylate 65 was subjected to conditions previously developed to generate the corresponding 1,2-cycloheptadiene (66),59 again using dienamine 24 as the trapping partner. These conditions resulted in the formation of cycloadduct 67 in 24% yield and 5.9:1 dr. Dimerization of the presumed allene intermediate 66 occurred competitively in this case.51 1,2-Dienes in smaller rings (i.e., 5-membered rings) are not yet accessible for use in synthesis,88,89 but remain an opportunity for future discovery.

Figure 5.

Figure 5.

[2+2] cycloadditions of carbocyclic allenes 63 and 66 with dienamine 24. Reaction conditions: 62 or 65 (1.0 equiv, 0.1 mmol), dienamine 24 (3.0–5.0 equiv), CsF (5.0 equiv), MeCN (0.1 M), 80 °C, 4–20 h.

Many products obtained from our studies possess strained alkylidene cyclobutanes with an adjacent π-system (i.e., alkene or aromatic ring), which could be susceptible to thermal isomerizations based on earlier studies by Christl and co-workers.72 Figure 6 highlights several key experimental findings, as well as computed ΔG values for each reaction shown. Cycloadduct 25, which was accessed in high diastereoselectivity (see Figure 3, entry 1) was heated at 110 °C. We observed an erosion of dr, affording epi-25 as the major isomer. This result was consistent with a computationally-determined ΔG value of −0.3 kcal/mol. Moreover, this finding suggested that comparing the ground state energetics of cycloadduct diastereomers could be used to predict the results of thermal isomerization studies. Thus, we considered cycloadducts 41 and 61 obtained in our earlier experiments. Computational analysis indicated that epi-41 and epi-61 were each thermodynamically more stable relative to the major cycloadducts obtained in our cyclic allene trapping experiments (ΔG = −3.8 and −2.7 kcal/mol, respectively). Gratifyingly, experimental results were consistent with computational predictions, allowing for the efficient conversion of 41 (1.5:1 dr) to epi-41 (>19:1 dr) and 61 (>19:1 dr) to epi-61 (>19:1 dr) under thermal conditions. Mechanistically, these isomerizations may occur by thermal homolytic cleavage of the strained bond between C6 and C4’ indicated (see substrate 25), followed by radical recombination, ultimately reaching a thermodynamic equilibrium of the two isomers. Our findings suggest that the selectivities observed in the aforementioned cyclic allene [2+2] trappings (see Figures 3 and 4) reflect kinetic distributions of products.

Figure 6.

Figure 6.

Thermal isomerization of cycloadducts 25, 41, and 61. Yields and dr determined by 1H NMR analysis using 1,3,5-trimethoxybenzene as an external standard. Free energies calculated using density functional theory (B3LYP/6–31G(d)).

The [2+2] cycloadditions of strained cyclic allenes with highly substituted alkenes provides a platform to quickly access exceedingly complex cyclobutane-containing structures. Figure 7 showcases several such examples via the formation of polycycles 73, 78, and 82, as well as their precursors. Ester-substituted cyclic allene precursor 68 was treated with trapping partner 38 under our usual reaction conditions to provide cycloadduct 70 in good yield and >19:1 dr. Of note, the reaction is also regioselective, favoring reaction of the alkene distal to the ester in the presumed cyclic allene intermediate 69 shown.90 The α,β-unsaturated ester and vinyl iodide functional handles provide opportunities for further elaboration. Treatment of 70 with 1,8-diazabicyclo(5.4.0)undec-7-ene (DBU) in nitromethane delivered 71 in 51% yield and >19:1 dr, introducing a quaternary stereocenter. By cross-coupling 71 with heterocyclic boronate 72, highly decorated product 73 was obtained bearing four contiguous stereocenters, two of which are quaternary, and five rings in the core scaffold. In another example, azacyclic allene 75, generated from silyl triflate 74, was intercepted with cyclobutyl methylidene 76 to afford cycloadduct 77. Subsequent epoxidation with meta-chloroperoxybenzoic acid (m-CPBA) yielded 78, a polyspiro compound possessing four small rings and three contiguous stereocenters. Lastly, oxacyclic allene 13 was trapped with trisubstituted alkene 7991 to provide cycloadduct 80 in 84% yield, introducing vicinal quaternary carbons and three contiguous stereocenters with a high degree of stereospecificity with respect to alkene geometry of 79. Cycloadduct 80 was then subjected to nitrile oxide (3+2) cycloaddition to deliver 82. This hexacyclic product possesses five contiguous stereocenters and three fully substituted carbons, including two quaternary centers. Cyclobutane-containing compounds 73, 78, and 82 are each accessible in either one or two steps from [2+2] cycloadducts, underscoring the structural complexity now readily accessible using strained cyclic allene methodology.

Figure 7.

Figure 7.

Synthetic elaborations of cycloadducts 70, 77, and 80 to access complex products 73, 78, and 82, respectively. aThe 13:1 ratio reflects the ratio of the major diastereomer (13) to the sum of all minor diastereomers (1).

CONCLUSION

Cyclobutanes are desirable motifs typically introduced using photochemical cycloadditions. As an alternative approach, we utilize the strain-driven reactivity of cyclic allenes, which are emerging synthetic building blocks, to promote [2+2] cycloadditions. We have shown that exceedingly complex structures can be accessed by intercepting cyclic allenes with highly substituted exocyclic alkenes at ambient temperature. The products obtained directly from the methodology, along with their isomers and derivatives, contain the desirable cyclobutane motif and a number of attractive features, such as heterocycles, spiro centers, contiguous stereocenters, and vicinal quaternary centers. We expect this study should not only provide a versatile tool to access highly substituted cyclobutanes but should also prompt the further strategic use of unconventional strained intermediates for the rapid assembly of complex architectures.

Supplementary Material

SI

ACKNOWLEDGMENT

The authors are grateful to the NIH-NIGMS (R35 GM139593 for N.K.G. and T32 GM136614 for M.S.M.), the Trueblood Family (for N.K.G.), and the Foote Family (for M.S.M.). We thank Dr. Saeed Khan (UCLA) for X-ray analysis studies, Dr. Ta-Chung Ong (UCLA) for help with NMR experiments, and Dr. James Logan Bachman (UCLA), Paris Dee (UCLA), and Madeline Ruos (UCLA) for experimental assistance. 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

Complete contact information is at: the ACS Publications website.

Supporting Information.

The Supporting Information is available free of charge on the ACS Publications website.

Detailed experimental procedures, compounds characterization data, crystallographic data for 61, Cartesian coordinates, electronic energies, entropies, enthalpies, Gibbs free energies, and lowest frequencies of the calculated structures (PDF)

Accession Codes.

CCDC 2323505 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.

No competing financial interests have been declared.

REFERENCES

  • (1).Lovering F; Bikker J; Humblet C Escape from flatland: Increasing saturation as an approach to improving clinical success. J. Med. Chem 2009, 52, 6752–6756. [DOI] [PubMed] [Google Scholar]
  • (2).Lovering F Escape from flatland 2: Complexity and promiscuity. Med. Chem. Commun 2013, 4, 515–519. [Google Scholar]
  • (3).van der Kolk MR; Janssen MACH; Rutjes FPJT; Blanco-Ania D Cyclobutanes in small-molecule drug candidates. ChemMedChem 2022, 17, e202200020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (4).Li J; Gao K; Bian M; Ding H Recent advances in the total synthesis of cyclobutane-containing natural products. Org. Chem. Front 2020, 7, 136–154. [Google Scholar]
  • (5).Hui C; Liu Y; Jiang M; Wu P Cyclobutane-containing scaffolds in bioactive small molecules. Trends Chem 2022, 4, 677–681. [Google Scholar]
  • (6).Carreira EM; Fessard TC Four-membered ring-containing spirocycles: Synthetic strategies and opportunities. Chem. Rev 2014, 114, 8257–8322. [DOI] [PubMed] [Google Scholar]
  • (7).Singh J; Bisacchi GS; Ahmad S; Godfrey JD; Kissick TP; Mitt T; Kocy O; Vu T; Papaioannou CG; Wong MK; Heikes JE; Zahler R; Mueller RH A practical asymmetric synthesis of the antiviral agent Lobucavir, BMS-180194. Org. Proc. Res. Dev 1998, 2, 393–399. [Google Scholar]
  • (8).Hopkins B; Ma B; Marx I; Schulz J; Vandeveer G; Prince R; Nevalainen M; Chen T; Yousaf Z; Himmelbauer M; Pattaropong V; Jones J; Lin E; Gonzalez-Lopez de Turiso F; Purgett T; Capacci A; Sciabola S Pyrazolo[1,5-A]pyrazine derivatives as BTK inhibitors WO 2022104079, May 05, 2022.
  • (9).Sharma R; Nikas SP; Paronis CA; Wood JT; Halikhedkar A; Guo JJ; Thakur GA; Kulkarni S; Benchama O; Raghav JG; Gifford RS; Järbe TUC; Bergman J; Makriyannis A Controlled-deactivation cannabinergic ligands. J. Med. Chem 2013, 56, 10142–10157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (10).Reisman SE; Ready JM; Weiss MM; Hasuoka A; Hirata M; Tamaki K; Ovaska TV; Smith CJ; Wood JL Evolution of a synthetic strategy: Total synthesis of (±)-welwitindolinone A isonitrile. J. Am. Chem. Soc 2008, 130, 2087–2100. [DOI] [PubMed] [Google Scholar]
  • (11).Baran PS; Richter JM Enantioselective total syntheses of welwitindolinone A and fischerindoles I and G. J. Am. Chem. Soc 2005, 127, 15394–15396. [DOI] [PubMed] [Google Scholar]
  • (12).Misiek M; Williams J; Schmich K; Hüttel W; Merfort I; Salomon CE; Aldrich CC; Hoffmeister D Structure and cytotoxicity of arnamial and related fungal sesquiterpene aryl esters. J. Nat. Prod 2009, 72, 1888–1891. [DOI] [PubMed] [Google Scholar]
  • (13).Zhou M; Li X-R; Tang J-W; Liu Y; Li X-N; Wu B; Qin H-B; Du X; Li L-M; Wang W-G; Pu J-X; Sun H-D Scopariusicides, novel unsymmetrical cyclobutanes: Structural elucidation and concise synthesis by a combination of intermolecular [2 + 2] cycloaddition and C–H functionalization. Org. Lett 2015, 17, 6062–6065. [DOI] [PubMed] [Google Scholar]
  • (14).Namyslo JC; Kaufmann DE The application of cyclobutane derivatives in organic synthesis. Chem. Rev 2003, 103, 1485–1538. [DOI] [PubMed] [Google Scholar]
  • (15).Secci F; Frongia A; Piras PP Stereocontrolled synthesis and functionalization of cyclobutanes and cyclobutanones. Molecules 2013, 18, 15541–15572. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (16).Eckart-Frank IK; Wilkerson-Hill SM Palladium-catalyzed trans-selective synthesis of spirocyclic cyclobutanes using α,α-dialkylcrotyl- and allylhydrazones. J. Am. Chem. Soc 2023, 145, 18591–18597. [DOI] [PubMed] [Google Scholar]
  • (17).Scholz SO; Kidd JB; Capaldo L; Fikweert NE; Little-filed RM; Yoon TP Construction of complex cyclobutane building blocks by photosensitized [2 + 2] cycloaddition of vinyl boronate esters. Org. Lett 2021, 23, 3496–3501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Sherbrook EM; Jung H; Cho D; Baik M-H; Yoon TP Brønsted acid catalysis of photosensitized cycloadditions. Chem. Sci 2020, 11, 856–861. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (19).Conner ML; Brown MK Synthesis of 1,3-substituted cyclobutanes by allenoate-alkene [2+2] cycloaddition. J. Org. Chem 2016, 81, 8050–8060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).Liu Y; Ni D; Brown MK Boronic ester enabled [2 + 2]-cycloadditions by temporary coordination: Synthesis of artochamin J and piperarborenine B. J. Am. Chem. Soc 2022, 144, 18790–18796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (21).Liu Y; Brown MK Photosensitized [2 + 2]-cycloadditions of dioxaborole: Reactivity enabled by boron ring constraint strategy. J. Am. Chem. Soc 2023, 145, 25061–25067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Poplata S; Tröster A; Zou Y-Q; Bach T Recent advances in the synthesis of cyclobutanes by olefin [2 + 2] photocycloaddition reactions. Chem. Rev 2016, 116, 9748–9815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (23).Sarkar D; Bera N; Ghosh S [2 + 2] Photochemical cycloaddition in organic synthesis. Eur. J. Org. Chem 2020, 2020, 1310–1326. [Google Scholar]
  • (24). For a recent example of an intermolecular selective photochemical [2+2] cycloaddition, see:; Jiang Y; Wang C; Rogers CR; Kodaimati MS; Weiss EA Regio- and diastereoselective intermolecular [2+2] cycloadditions photocatalysed by quantum dots. Nat. Chem 2019, 11, 1034–1040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (25).Roberts JD; Simmons HE; Carlsmith LA; Vaughn CW Rearrangement in the reaction of chlorobenzene-1-C14 with potassium amide. J. Am. Chem. Soc 1953, 75, 3290–3291. [Google Scholar]
  • (26).Tadross PM; Stoltz BM A comprehensive history of arynes in natural product total synthesis. Chem. Rev 2012, 112, 3550–3577. [DOI] [PubMed] [Google Scholar]
  • (27).Gampe CM; Carreira EM Arynes and cyclohexyne in natural product synthesis. Angew. Chem., Int. Ed 2012, 51, 3766–3778. [DOI] [PubMed] [Google Scholar]
  • (28).Wenk HH; Winkler M; Sander W One century of aryne chemistry. Angew. Chem., Int. Ed 2003, 42, 502–528. [DOI] [PubMed] [Google Scholar]
  • (29).Goetz AE; Garg NK; Enabling the use of heterocyclic arynes in chemical synthesis. J. Org. Chem 2014, 79, 846–851. [DOI] [PubMed] [Google Scholar]
  • (30).Nakamura Y; Yoshida S; Hosya T Recent advances in synthetic hetaryne chemistry. Heterocycles 2019, 98, 1623–1677. [Google Scholar]
  • (31).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]
  • (32).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]
  • (33).Tlais SF; Danheiser RL N-Tosyl-3-azacyclohexyne. Synthesis and chemistry of a strained cyclic ynamide. J. Am. Chem. Soc 2014, 136, 15489–15492. [DOI] [PubMed] [Google Scholar]
  • (34).Medina JM; McMahon TC; Jimeńez-Oseś G; Houk KN; Garg NK Cycloadditions of cyclohexynes and cyclopentyne. J. Am. Chem. Soc 2014, 136, 14706–14709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (35). For a comprehensive review on the different modes of reactivity of strained intermediates, including [2+2] cycloadditions, 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]
  • (36).Liebman JF; Greenberg A A survey of strained organic molecules. Chem. Rev 1976, 76, 311–365. [Google Scholar]
  • (37).Moser WR The Reactions of gem-Dihalocyclopropanes with Organometallic Reagents Ph. D. Thesis. Massachusetts Institute of Technology, Cambridge, MA, 1964. [Google Scholar]
  • (38).Moore WR; Moser WR The reaction of 6,6-dibromobicyclo[3.1.0]hexane with methyllithium. Efficient trapping of 1,2-cyclohexadiene by styrene. J. Org. Chem 1970, 35, 908–912. [Google Scholar]
  • (39).Wittig G; Fritze P On the Intermediate Occurrence of 1,2-Cyclohexadiene. Angew. Chem., Int. Ed 1966, 5, 846. [Google Scholar]
  • (40).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]
  • (41).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]
  • (42).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]
  • (43).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]
  • (44).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]
  • (45).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]
  • (46).Inoue K; Nakura R; Okano K; Mori A One-pot synthesis of silylated enol triflates from silyl enol ethers for cyclohexynes and 1,2-cyclohexadienes. Eur. J. Org. Chem 2018, 2018, 3343–3347. [Google Scholar]
  • (47).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, 131112. [Google Scholar]
  • (48).Xu Q; Hoye TR A distinct mode of strain-driven cyclic allene reactivity: Group migration to the central allene carbon atom. J. Am. Chem. Soc 2023, 145, 9867–9875. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (49).Miyasaka T Cycloaddition reaction using highly strained 6-membered allenes. J. Syn. Org. Chem. Jpn, 2024, 82, 63–64. [Google Scholar]
  • (50).Lofstrand VA; McIntosh KC; Almehmadi YA; West FG Strain-activated Diels–Alder trapping of 1,2-cyclohexadienes: Intramolecular capture by pendent furans. Org. Lett 2019, 21, 6231–6234. [DOI] [PubMed] [Google Scholar]
  • (51).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]
  • (52).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]
  • (53).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]
  • (54).Jankovic CL; McIntosh KC; Lofstrand VA; West FG Stereoselective intramolecular [2+2] trapping of 1,2-cyclohexadienes: A route to rigid, angularly fused tricyclic scaffolds. Chem. Eur. J 2023, 29, e202301668. [DOI] [PubMed] [Google Scholar]
  • (55).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]
  • (56).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]
  • (57).Yamano MM; Knapp RR; Ngamnithiporn A; Ramirez M; Houk KN; Stoltz BM; Garg NK Cycloadditions of oxacyclic allenes and a catalytic asymmetric entryway to enaniotenriched cyclic allenes. Angew. Chem., Int. Ed 2019, 58, 5653–5657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (58).Chari JV; Ippoliti FM; Garg NK Concise approach to cyclohexyne and 1,2-cyclohexadiene precursors. J. Org. Chem 2019, 84, 3652–3655. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (59).McVeigh MS; Kelleghan AV; Yamano MM; Knapp RR; Garg NK Silyl tosylate precursors to cyclohexynes, 1,2-cyclohexadienes, and 1,2-cycloheptadiene. Org. Lett 2020, 22, 4500–4504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (60).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]
  • (61).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]
  • (62).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]
  • (63).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]
  • (64).Witkowski DC; McVeigh MS; Scherer GM; Anthony SM; Garg NK Catalyst-controlled annulation of strained cyclic allenes with π-allylpalladium complexes. J. Am. Chem. Soc 2023, 145, 10491–10496. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (65).Mehta MM; Gonzalez JAM; Bachman JL; Garg NK Cyclic allene approach to the manzamine alkaloid keramaphidin B. Org. Lett 2023, 25, 5553–5557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (66).Kelleghan AV; Tena-Meza A; Garg NK Generation and reactivity of unsymmetrical strained heterocyclic allenes. Nat. Synth 2023, In Press, DOI: 10.1038/s44160-023-00432-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (67).Roberts CC Strained allenes for heterocyclic difunctionalization. Nat. Synth 2024, In Press, DOI: 10.1038/s44160-023-00455-8. [DOI] [Google Scholar]
  • (68).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]
  • (69).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]
  • (70).Bottini AT; Hilton LL; Plott J Relative reactivities of 1,2-cyclohexadiene with conjugated dienes and styrene. Tetrahedron 1975, 31, 1997–2001. [Google Scholar]
  • (71).Harnos S; Tivakornpannarai S; Waali EE The addition of 1,2-cyclohexadiene to substituted styrenes. Tetrahedron Lett 1986, 27, 3701–3704. [Google Scholar]
  • (72).Christl M; Schreck M 7-Arylbicyclo[4.2.0]oct-1-ene – Synthese durch [2 + 2]-cycloadditionen von 1,2-cyclohexadien sowie 1-methyl-1,2-cyclohexadien und thermische äquilibrierung der exo/endo-isomeren. Chem. Ber 1987, 120, 915–920. [Google Scholar]
  • (73).Elliot RL; Takle AK; Tyler JW; White J Cycloadditions of cephalosporins. A general synthesis of novel 2,3-fused cyclobutane and cyclobutene cephems. J. Org. Chem 1993, 58, 6954–6955. [Google Scholar]
  • (74).Kelleghan AV; Bulger AS; Witkowski DC; Garg NK Strain-promoted reactions of 1,2,3-cyclohexatriene and its derivatives. Nature 2023, 618, 748–754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (75).Peterson EA; Overman LE Contiguous stereogenic quaternary carbons: A daunting challenge in natural product synthesis. Proc. Natl. Acad. Sci. U. S. A 2004, 101, 11943–11948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (76).Caille S; Cui S; Faul MM; Mennen SM; Tedrow JS; Walker SD Molecular complexity as a driver for chemical process innovation in the pharmaceutical industry. J. Org. Chem 2019, 84, 4583–4603. [DOI] [PubMed] [Google Scholar]
  • (77). Cyclic allene cycloadducts typically contain a high degree of saturation, as well an additional stereocenter when compared to analogous products of other strained intermediates, such as 7a–7d.
  • (78).Hiesinger K; Dar’in D; Proschak E; Krasavin M Spirocyclic scaffolds in medicinal chemistry. J. Med. Chem 2021, 64, 150–183. [DOI] [PubMed] [Google Scholar]
  • (79). Oxacyclic allene 13 was chosen for this study due to a number of factors, such as the ready availability of its precursor, 12, the ease of heterocyclic allene generation at ambient temperatures, and that its use results in heterocyclic products.
  • (80). The variation in observed diastereoselectivities is not presently understood but will be the subject of further investigation.
  • (81). Enamines and hetaryl-substituted alkenes have rarely been studied in reactions with strained cyclic allenes and therefore presented an opportunity for discovery. Furthermore, styrenyl substrates, such as 26 and heterocyclic variants, are readily accessible.
  • (82). We postulate that reactions proceed by a stepwise mechanism, analogous to what has been proposed for the trapping of 1,2-cyclohexadiene with styrenes (ref 69). The observed regioselectivity is consistent with a mechanism involving: a) initial C–C bond formation between the central carbon of the cyclic allene and terminal position of the alkene trapping partner, leading to a stabilized diradical intermediate and b) radical recombination to form the cyclobutyl ring. Although further studies are ongoing, kinetic isotope effect (KIE) experiments at natural abundance are consistent with this hypothesis.
  • (83). These conditions proved more efficient, as they enabled the use of fewer equivalents of trapping partners and led to their improved solubility in some instances; see Supporting Information, section D for reaction optimization details.
  • (84).Carpino LA; Sau AC Convenient source of ‘naked’ fluoride: Tetra-n-butylammonium chloride and potassium fluoride dihydrate. J. Chem. Soc., Chem. Commun 1979, 514–515. [Google Scholar]
  • (85).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]
  • (86). The observed stereospecificity may indicate that the relative rate of the second C–C bond formation is faster than a σ-bond rotation. However, we cannot rule out σ-bond rotation occurring rapidly, with the distribution of products correlating with the kinetic barriers for each cyclization pathway.
  • (87). The reaction is thought to proceed via a stepwise mechanism (see footnote 82). The presence of the gem-dimethyl group is thought to raise the kinetic barrier for formation of the second C–C bond (i.e., C6–C4’ forming bond of 41), leading to a less selective reaction when compared to the formation of 25.
  • (88).Algi F; Özen R; Balci M The first generation and trapping of a five-membered ring allene: 2-dehydro-3a,4,5,6,6a-pentahydropentalene. Tetrahedron Lett 2002, 43, 3129–3131. [Google Scholar]
  • (89).Ceylan M; Yalçin S; Seçen H; Sütbeyaz Y; Balci M Evidence for the formation of a new five-membered ring cyclic allene: generation of 1-cyclopenta-1,2-dien-1-ylbenzene. J. Chem. Research (S) 2003, 21–23. [Google Scholar]
  • (90). Similar regioselectivity in [2+2] cycloadditions of substituted cyclic allenes was observed by West; see ref. 53.
  • (91). Of note, 2.0 equivalents of trapping partner 38 and 1.7 equivalents of trapping partner 79 were used in their respective transformations, highlighting that large excesses of trapping partner are not strictly required for this methodology.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

SI

RESOURCES