Next Article in Journal
Discovery of Novel Inhibitors of Aspergillus fumigatus DHODH via Virtual Screening, MD Simulation, and In Vitro Activity Assay
Previous Article in Journal
Dynamic Study on Synergy Mechanism and Characteristics of Particle Removal in Electrostatic Atomization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Case of Competitive Aromatization vs. Sigmatropic [1,5]-Hydrogen Atom Migration in a 1,2,4-Cyclohexatriene Intermediate Derived from a Bis-Enyne Cyclization

Department of Chemistry, University of Minnesota, Minneapolis, MN 55455, USA
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(12), 2610; https://doi.org/10.3390/molecules30122610
Submission received: 21 May 2025 / Revised: 10 June 2025 / Accepted: 12 June 2025 / Published: 16 June 2025

Abstract

1,2,4-Cyclohexatrienes are strained, reactive intermediates often formed by the tetradehydro-Diels–Alder (TDDA) reaction of a conjugated enyne bearing a tethered alkyne as the enynophile. The ene component is commonly the π-bond of an aromatic group. In this focused study, we investigated the reactivity of a symmetrical substrate in which the pair of terminal ene moieties were simple 2-propenyl groups. The intermediate 1,2,4-cyclohexatriene, now bearing a 5-isopropenyl group, underwent competitive aromatization (the most usual outcome of the strain-relieving event of the cyclohexatriene), along with an intramolecular [1,5]-hydrogen atom migration, ultimately producing a non-benzenoid, pyrrole derivative. This represents a previously unknown process for a 1,2,4-cyclohexatriene derivative. Mechanistic aspects of the competitive processes were revealed by experiments performed in the presence of various protic additives (MeOD and BHT).

Graphical Abstract

1. Introduction

1,2,3-Cyclohexatriene (1, Figure 1a), a reactive intermediate and isomer of benzene, was first formed and intercepted in 1990 [1]. The isomeric 1,2,4-cyclohexatriene (2) was first generated and trapped in 1992 [2]. These species and their derivatives remain of contemporary interest [3,4]. Members of this class of compounds are highly reactive because they house a highly strained cyclic allene within a small ring. The strain energies in 1 and 2 have been evaluated computationally and are estimated to be 50 and 34 kcal mol−1, respectively [5]. The two pairs of terminal substituents on acyclic allenes (1,2-propadiene derivatives, 3) ideally, of course, have an orthogonal relationship. The constraints of any rings (with the exception of very large ones) that house a cyclic allene dictate that the 1,2-propadiene unit is unable to achieve that ideal geometry, hence introducing increasing amounts of strain for lower homologs within smaller rings [5,6].
1,2,4-Cyclohexatriene derivatives are often formed as transient intermediates in, for example, net (4 + 2) cycloadditions between an alkyne and a 1,3-enyne in a process that can be categorized as a tetradehydro-Diels–Alder (TDDA) reaction [7,8] (cf. 4 to 5, Figure 1b [9,10]). The transient 1,2,4-triene is known to undergo several types of processes to relieve the strain of the significantly twisted allene that it houses. Hydrogen atom migration that results in a benzenoid product (5 to 6, Figure 1b) is most common. Rarer are processes that result in non-aromatized products (7 to 8, [11] Figure 1c; or 9 to 10, [12] Figure 1d).
Figure 1. (a) Strained, cyclic cyclohexatrienes (1 and 2) vs. the strain-free parent, propa-1,2-diene (also known as allene, 3), whose most stable geometry has orthogonal (90°) a-C-b and c-C-d planes. (b) Formation of 1,2,4-cyclohexatriene derivatives by a tetradehydro-Diels–Alder (TDDA) reaction and their aromatization to benzene derivatives by formal [1,5]-hydrogen atom migration [9,10]. (c) Relief of ring-strain by skeletal rearrangement [11]. (d) Relief of ring-strain by migration of an exocyclic group [12]. (e) Computed lowest energy pathway for the net [1,5]-hydrogen atom migration via two, sequential (blue then green) 1,2-H-atom migrations [13,14,15].
Figure 1. (a) Strained, cyclic cyclohexatrienes (1 and 2) vs. the strain-free parent, propa-1,2-diene (also known as allene, 3), whose most stable geometry has orthogonal (90°) a-C-b and c-C-d planes. (b) Formation of 1,2,4-cyclohexatriene derivatives by a tetradehydro-Diels–Alder (TDDA) reaction and their aromatization to benzene derivatives by formal [1,5]-hydrogen atom migration [9,10]. (c) Relief of ring-strain by skeletal rearrangement [11]. (d) Relief of ring-strain by migration of an exocyclic group [12]. (e) Computed lowest energy pathway for the net [1,5]-hydrogen atom migration via two, sequential (blue then green) 1,2-H-atom migrations [13,14,15].
Molecules 30 02610 g001

2. Results and Discussion

Of the three modes of reaction shown in Figure 1b–d, the isomerization of the cyclic allene to its isomeric benzenoid product (cf. 5 to 6) is by far the most common and, therefore, well studied. The generally accepted mechanism of the aromatization of 1,2,4-cyclohexatriene (and its derivatives, at least under conditions where there is no proton source in the reaction medium), is portrayed in Figure 1e [11 to benzene (13)]. Computations [13,14,15] suggest that this thermal-only pathway involves two consecutive 1,2-H-atom shifts rather than a single [1,5]-H-atom migration. The first 1,2-shift gives a species (carbene or orthogonal diradical, cf. 12/12′) that undergoes a rapid, second C–H insertion/migration event.
In the course of studying the thermal cyclization chemistry of bis-enyne 14, we observed, in addition to the normal aromatized benzenoid TDDA product, 16, an unexpected side product [16], 17 (Figure 2). The ratio of 16:17 under these conditions was ca. 2:1. Although this ratio changed under other reaction conditions (see below), 17 was never the major product. The assignment of the structure of the expected 16 was straightforward from analysis of its 1H and 13C 1D and 2D NMR spectra (see Supporting Information (SI)). However, the structure of 17 was not at all clear. Only after determining its structure would we be in a position to hypothesize a pathway for its formation.
It was straightforward to confirm (GCMS) that 17 had the same mass as 16 (i.e., they were isomers). The chromatographic behaviors on silica gel of 16 and 17 were nearly the same, but by shaving the fractions eluting from an MPLC separation, we were able to obtain 17 in a 96:4 ratio with 16. In the proton NMR spectrum of 17, it was clear that the NTs moiety was intact. It also contained three singlets, whose chemical shifts (δ 7.04, 6.83, and 6.07 ppm) were indicative of protons attached to sp2 hybridized carbon atoms. Another significant difference in the spectrum (vs. that of 16) was the presence of three, rather than two, methyl singlets (in addition to the Ts-Me), judged to be either allylic or benzylic in nature. The remaining protons were a singlet methylene pair (2.99 ppm) that, among other things, indicated that 17 possessed an achiral structure.
The NOESY NMR spectrum was particularly informative; enhancements (double-headed arrows in Figure 3a) led us to propose structure 17 as the identity of the unknown. More specifically, the NOEs between what are assigned there as H1/H7 and H3/H1′′ (red) were quite diagnostic. To further support this assignment, we located in the literature compound 18 [17], a structurally analogous, fused N-tosyl pyrrole that possesses quite similar chemical shifts of the indicated protons in 18 vs. the analogous protons in 17 (Figure 3b). We also carried out a DP4+ analysis [18] of the NMR chemical shifts of the protons and carbons in 17 vs. an isomeric alkene having the endocyclic cyclohexene double bond between C5 and C6; this analysis also supported the assignment of 17 (see SI). Finally, we performed resolution enhancement of the (many) “singlets” in 17, mentioned in the above paragraph; this revealed (many) small coupling constants that further reinforced the structure assignment (see J values in the 1H NMR line listings in Section 3).
With the structure of the unusual side product 17 established, we were obliged to propose a mechanism to account for its formation, especially so because, as far as we are aware, it represents an unprecedented type of transformation for a product arising from a TDDA reaction via a 1,2,4-cyclohexatriene-containing intermediate such as 5. We show in Figure 3c a proposed pathway that branches following the initial formation of the intermediate 15 from an initial TDDA cycloisomerization. The 1,2-hydrogen atom migration to 19 and final carbene C–H insertion account for the major product, 16. But in addition, the isopropenyl alkene in intermediate 15 is poised to undergo a sigmatropic rearrangement, namely [1,5]-hydrogen atom migration to produce 20. A final net 1,3-hydrogen atom migration (tautomerization) of the strained cyclic allene in 20 would establish the pyrrole ring in the minor product, 17. As indicated by the arrows in structure 20, there is the possibility that this aromatization could be mediated (catalyzed) by a protic molecule (e.g., water) acting as a proton shuttle; this conversion could be either concerted or stepwise, with the latter following an initial protonation of the central carbon of the allene.
It should be noted that allenes are known to undergo reaction with Brønsted acids via kinetic protonation at either their terminal or central atoms to give vinylic or allylic carbocations, respectively [19]. Particularly relevant here are the differences seen between unstrained, acyclic allenes vs. strained, cyclic allenes. The former allenes tend to prefer kinetic protonation at a terminal carbon atom, although even for this class of unstrained allenes, this selectivity can often be reversed by the nature of substituents on the terminal atoms. In contrast, strained, cyclic allenes, tend to protonate preferentially at their central carbon atom, with there being a reduced amount of twist needed to attain the more stable allylic cation. An early report that brought this difference to light described the addition of HCl to cyclonona-1,2-diene to give only 3-chlorocyclononene [20]. The 1983 review by Smadja [19] contains a collection of many dozens of examples of “protonation of allenic derivatives” from which these general trends can be gleaned.
Several experiments were performed to probe the possible role that a hydroxyl-containing species might be playing. Saá and coworkers have used CD3OD (or CH3OD) as an additive and observed incorporation of a deuterium atom at the central carbon of TDDA-derived cyclic allene intermediates generated in their experiments (further discussed below) [15,21]. This established that a protic additive provided a pathway for aromatization that was faster than the unimolecular series of 1,2-H migrations computed for the parent system (cf. Figure 1e). We performed an analogous experiment with our substrate 14. Heating the dienyne 14 in a mixture of DCE and CD3OD (90:10, v:v) at 130 °C for 16 h led to the formation of 16-d and 17-d, deuterated to ca. 91% and 90% (1H NMR) and in a ratio of 2:1, that ratio being the same as the reaction performed without added methanol. An analogous experiment using CH3OD as the additive also showed that both 16-d and 17-d were produced, indicating that the hydroxyl group, and not the methyl group, was the source of the hydrogen incorporated into each of those products.
These labeling experiments do not differentiate between the possible stepwise (via 15-d+ or 20-d+) vs. concerted net 1,3-tautomerizations shown in Figure 4a. A fuller discussion of the Saá experiments is instructive because they shed light on this question. As outlined in Figure 4b, heating the TDDA substrate 22 in toluene alone produced solely the naphthalene product 29-angular. This can be rationalized by the sequence of electrocyclic and isomerization events commencing from the initial cyclic allene 23 and proceeding through 2426. Computations supported the main features of this pathway, although whether the final aromatization occurred via sequential intramolecular 1,2-H shifts (cf. Figure 1e) or a protic molecule-assisted event (e.g., via 27) was not discerned. However, the role of the protic additive phenol was clearly demonstrated because, when present, it resulted in the formation of 29-linear as the only product. This is explained by protonation of allene 23 by PhOH more quickly than the rate of unimolecular ring-opening of 23 to 24.
We performed a similar experiment with 14 using butylated hydroxytoluene (BHT) as a phenolic, protic additive. In the presence of 0.1 equivalent of BHT, the ratio of 16 to 17 changed to 6:1. Using 1.0 equivalent of BHT, we observed formation of only the aromatic product 16. This suggests that even with the use of a small amount of BHT, the cyclic allene was protonated to give 15-h+ at a rate competitive with its [1,5]-H atom migration, and that the larger amount of BHT promoter fully suppressed the [1,5]-H atom migration. Recall that even in the presence of 10 vol% of methanol, the ratio of products 16 to 17 was the same as when no methanol was present. The different effect of the BHT additive is most likely a reflection of its greater acidity relative to that of methanol (MeOH pKa(DMSO) = 29.0 [22], PhOH pKa(DMSO) = 18.0 [23]), thereby accelerating the production of 15-h+.
Finally, we prepared the ynoate/enyne substrate 30 (Figure 5). When heated, it reacted at a similar rate to that of 14. It produced the aromatized compound 33 as the major product. As expected, when this reaction was performed in the presence of CD3OD, a deuterium atom was incorporated into the product 33-d. In principle, the carbonyl group of the ester in the intermediate cyclic allene 31 could have engaged in a [1,5]-hydrogen atom migration to give 32. However, the relative instability of the enol tautomer of an ester makes this a sufficiently high-barrier process, so there is little opportunity for that to occur enroute to a pyrrole derivative analogous to 17. This further emphasizes the unique role that the isopropenyl substituent is playing in diverting the reactivity of 15.

3. Experimental Section

3.1. Chemistry

3.1.1. General Experimental Protocols

13C and 1H NMR spectra were obtained using a Bruker HD-500 spectrometer. All spectra were recorded in CDCl3. The reported proton chemical shifts are referenced to CHCl3 in (δ = 7.26 ppm) in CDCl3. Resonances in the line listings of 1H NMR spectral data are given in this format: chemical shift (δ, ppm) [multiplicity, coupling constant(s) J (in Hz), integral value (to the nearest integer)]. First-order coupling constants were extracted using methods we have described elsewhere [24,25]. An nfom, nfod, or nfot refers to a non-first order multiplet, doublet, or triplet in a 1H NMR spectrum. Coupling constants for these resonances are given as apparent values (Japp); for example, the observed spacing between the two most intense lines in an nfod is the value of Jo + Jp and not the true doublet coupling constant. The reported carbon chemical shifts in 13C spectra are referenced to CDCl3 (δ = 77.16 ppm). Infrared spectra were recorded using a using a Bruker Alpha II Spectrometer. The data were collected in attenuated total reflectance (ATR) mode. Samples were measured as thin films that were deposited onto a diamond ATR window by evaporative loss of solvent, often CDCl3. Absorption maxima are given as cm−1. High-resolution mass spectrometry (HRMS) data were collected in the ESI mode with a Thermo Orbitrap Velos instrument having mass accuracy of ≤3 ppm. PierceTM LTQ was used as an external calibrant. Samples were injected directly into the ion source. Melting points were measured on a Köfler hot-stage and polarizing microscope and are uncorrected. New substances were often purified by medium-pressure liquid chromatography (MPLC) at 25–200 psi. Silica gel columns (hand-packed, normal-phase, Teledyne RediSep Rf Gold®, 20–40 μm, 60 Å pore size) were used. The unit was assembled from a Waters HPLC pump (model 510) affixed to a Waters differential refractive index detector (R401). Preparative flash column chromatography was performed on columns hand-packed with Agela silica gel (230–400 mesh). Thin-layer chromatography (TLC) was carried out on silica gel-coated plates (plastic-backed), visualized by UV light and/or by staining with a solution of potassium permanganate. Reaction temperatures refer to the temperature of a pre-equilibrated and external heating oil bath. Any reaction performed at a temperature higher than that of the reaction solvent’s boiling point was carried out in a threaded culture tube, which was closed with an inert Teflon®-lined screw-cap.

3.1.2. Synthesis of Precursors 14 and 30 (Structures of Intermediates Not Specifically Shown in Any of Figure 1, Figure 2, Figure 3, Figure 4 and Figure 5 Are Given S#s Here)

N,N-bis(4-hydroxy-4-methylpent-2-yn-1-yl)-4-methylbenzenesulfonamide (S2). In a 100 mL round-bottomed flask equipped with a magnetic stir bar, 4-methyl-N,N-di(prop-2-yn-1-yl)benzenesulfonamide (S1 [26], 2.3 g, 9.4 mmol, 1 equiv) was added. THF (47 mL) was added at −78 °C under an atmosphere of N2. n-BuLi (7.9 mL, 2.5 M in hexanes, 2.1 equiv) was added dropwise. After 1 h of stirring at −78 °C, acetone (2.1 mL, 28 mmol, 3 equiv) was added dropwise. The reaction mixture was stirred at −78 °C for another 1 h. The reaction was quenched with saturated NH4Cl aqueous solution. THF was removed under vacuum, and the aqueous layer was extracted with DCM, washed with brine, dried (Na2SO4), concentrated, and purified by flash column chromatography (10:1 to 1:1 Hex:EtOAc) to give S2 (2.9 g, 8.0 mmol, 85%) as a white solid. mp: 98–100 °C. 1H-NMR (500 MHz, CDCl3): δ 7.76 (nfod, Japp = 8.3 Hz, 2H), 7.33 (nfod, Japp = 8.0 Hz, 2H), 4.17 (s, 4H), 2.43 (s, 3H), 1.89 (s, 2H), and 1.37 (s, 12H). 13C-NMR{1H} (126 MHz, CDCl3): δ 144.0, 136.1, 129.7, 128.3, 90.5, 74.9, 65.1, 37.1, 31.2, and 21.6. IR (neat): νmax 3524 (br O-H), 3265 (br O-H), 2981 (CH3), and 2924 (CH2) cm−1. HRMS (ESI-TOF): Calcd for C19H26NO4S+ [M + H]+, 364.1577; found, 364.1560 (9%). Calcd. for 346.1471 C19H24NO3S+ [M + H+ – H2O]; found, 346.1455 (100%).
4-Methyl-N,N-bis(4-methylpent-4-en-2-yn-1-yl)benzenesulfonamide (14). In a 100 mL round-bottomed flask, equipped with a magnetic stir bar, S2 (790 mg, 2.2 mmol, 1 equiv) was added. DCM (13 mL) was added under a N2 atmosphere. To this solution, triethylamine (10.6 mL, 76 mmol, 35 equiv) was added at 0 °C. Methanesulfonic anhydride (6.7 g, 38 mmol, 18 equiv) dissolved in DCM (26 mL) was added dropwise. The reaction mixture was stirred at RT for 2.5 h. The reaction mixture was quenched with water and extracted with DCM. The combined organic layers were washed with brine, dried (Na2SO4), filtered, and concentrated; and the residue was purified by flash column chromatography (10:1 Hex:EtOAc) to give 14 (511 mg, 1.6 mmol, 72%) as a pale-yellow oil. 1H-NMR (500 MHz, CDCl3): δ 7.76 (nfod, Japp = 8.0 Hz, 2H), 7.31 (nfod, Japp = 8.2 Hz, 2H), 5.19 (br s, 2H), 5.13 (br s, 2H), 4.29 (s, 4H), 2.42 (s, 3H), and 1.76 (br s, J = 1.3 Hz, 6H). 13C-NMR{1H} (126 MHz, CDCl3): δ 143.8, 135.6, 129.7, 128.1, 126.1, 122.5, 87.0, 80.8, 37.3, 23.2, and 21.6. IR (neat): νmax 3003 (Csp2-H), 2925 (H2Csp3-H), and 1598 (C=C) cm−1. HRMS (ESI-TOF): Calcd for C19H22NO2S+ [M + H]+ 328.1366; found 328.1353 (100%).
4-Methyl-N-(4-methylpent-4-en-2-yn-1-yl)-N-(prop-2-yn-1-yl)benzenesulfonamide (S4). In a culture tube, equipped with a magnetic stir bar, N-(4-hydroxy-4-methylpent-2-yn-1-yl)-4-methyl-N-(prop-2-yn-1-yl)benzenesulfonamide (S3 [27], 305 mg, 1 mmol, 1 equiv) was added. DCM (5.9 mL) was added under an atmosphere of N2. Triethylamine (2.4 mL, 17.6 mmol, 17.6 equiv) was added at 0 °C. Methanesulfonic anhydride (1.5 g, 8.8 mmol, 8.8 equiv) was added dropwise. The reaction mixture was stirred at RT for 1 h and then poured into ice-cold concentrated aqueous HCl (2 mL). The aqueous layer was extracted with DCM. The combined organic layers were washed with brine, dried (Na2SO4), filtered, concentrated, and purified by MPLC (15:1 Hex:EtOAc) to give S4 (162.88 mg, 0.56 mmol, 57%) as a pale-yellow oil. 1H-NMR (500 MHz, CDCl3): δ 7.72 (nfod, Japp = 8.4 Hz, 2H), 7.29 (nfod, Japp = 8.1 Hz, 2H), 5.15 (dq, J = 1.6, 1.6 Hz, 1H), 5.07 (dq, J = 1.8, 1.0 Hz, 1H), 4.29 (s, 2H), 4.13 (dd, J = 2.5, 0.7 Hz, 2H), 2.41 (t, J = 0.9 Hz, 3H), 2.16 (t, J = 2.5 Hz, 1H), and 1.71 (dd, J = 1.6, 1.1 Hz, 3H). 13C{1H} (126 MHz, CDCl3): δ 144.0, 135.4, 129.7, 128.1, 126.0, 122.5, 87.2, 80.4, 76.6, 74.0, 37.1, 36.5, 23.2, and 21.7. IR (neat): νmax 3278 (Csp-H), 2982 (H2Csp3-H), 2927 (RHCsp3-H), and 1597 (C=C) cm−1. HRMS (ESI-TOF): Calcd for C16H18NO2S+ [M + H]+ 288.1053; found 288.1060 (100%).
Methyl 4-{[4-Methyl-N-(4-methylpent-4-en-2-yn-1-yl)phenyl]sulfonamido}but-2-ynoate (30). In a 10 mL round-bottomed flask, equipped with a magnetic stir bar, S4 (100 mg, 0.35 mmol, 1 equiv) was added. THF (3.5 mL) was added under a N2 atmosphere. nBuLi (0.17 mL, 0.42 mmol, 1.2 equiv) was added dropwise at −78 °C. The reaction mixture was stirred at −78 °C for 1 h. Methyl chloroformate (0.08 mL, 1.0 mmol, 3 equiv) was added dropwise. The reaction mixture was stirred at −78 °C for another 1.5 h. The reaction mixture was quenched with NH4Cl and extracted with EtOAc. The combined organic layers were washed with brine, dried (Na2SO4), and concentrated; and the residue was purified by flash column chromatography (30:1 to 10:1 Hex:EtOAc) to give 30 (21.3 mg, 0.0615 mmol, 18%) as a pale-yellow oil. 1H-NMR (500 MHz, CDCl3): δ 7.72 (nfod, Japp = 8.4 Hz, 2H), 7.31 (nfod, Japp = 8.3 Hz, 2H), 5.17 (m, 1H), 5.09 (m, 1H), 4.27 (s, 2H), 4.26 (s, 2H), 3.73 (s, 3H), 2.41 (s, 3H), and 1.71 (dd, J = 1.3, 1.4 Hz, 3H). 13C-NMR{1H} (126 MHz, CDCl3): δ 153.2, 144.3, 135.0, 129.9, 128.0, 125.9, 122.9, 87.7, 80.6, 80.0, 77.1, 52.9, 37.7, 36.5, 23.1, and 21.7. IR (neat): νmax 2954 (Csp3-H), 2922 (Csp3-H), 2847 (Csp3-H), 2242 (C≡C), and 1717 (C=O) cm−1. HRMS (ESI-TOF): Calcd for C18H20NO4S+ [M + H]+ 346.1108; found 346.1094 (100%).

3.1.3. Synthesis of Products of the TDDA Reactions

6-Methyl-4-(prop-1-en-2-yl)-2-tosylisoindoline (16) and 6-methyl-4-(propan-2-ylidene)-2-tosyl-4,5-dihydro-2H-isoindole (17). In a 10 mL culture tube, equipped with a magnetic stir bar, 14 (21 mg, 0.064 mmol) was added. 1,2-DCE (2 mL) was then added. The reaction mixture was heated at 130 °C for 16 h. The reaction mixture was concentrated and purified by flash column chromatography (30:1 Hex:EtOAc) to give, in order of elution, a small amount of the pyrrole derivative 17, contaminated with a small amount of co-eluting remaining starting material 14, and the major product, 16 (6.3 mg, 0.019 mmol, 30%) as a white crystalline solid. The ratio of 17:16 in this experiment was ca. 1:2 (analysis of the 1H NMR spectrum of the crude product mixture prior to chromatographic purification). The pyrrole derivative showed evidence of slowly giving rise to several new compounds over time and handling (TLC and NMR). The experiment was repeated on a larger scale [50 mg 14 in 1.5 mL of DCE in the presence of added BHT (0.1 equiv); 130 °C, ca. 16 h]. The ratio of products 17:16 was now 1:6 (crude NMR). MPLC (5:1 Hex:EtOAc) provided a purer sample of the pyrrole derivative 17 (10 mg, 10%, corrected for residual EtOAc and hexanes, which were purposely left to minimize loss of the somewhat volatile product; this is the sample for which the characterization data were recorded) and the isoindoline 16 (30 mg, 60%), respectively. 16: mp: 138–142 °C. 1H-NMR (500 MHz, CDCl3): δ 7.76 (nfod, Japp = 8.2 Hz, 2H), 7.31 (nfod, Japp = 8.0 Hz, 2H), 6.95 (dq, J = 1.6, 0.8 Hz, 1H), 6.88 (br dq, J = 1.6, 0.8 Hz, 1H), 5.18 (dq, J = 1.5, 1.5 Hz, 1H), 4.92 (dq, J = 1.4, 1.0 Hz, 1H), 4.60 (nfom, 2H), 4.57 (nfom, 2H), 2.40 (s, 3H), 2.30 (dd, J = 0.8, 0.8 Hz, 3H), and 2.04 (dd, J = 1.5, 1.0 Hz, 3H). 13C-NMR{1H} (126 MHz, CDCl3): δ 143.8, 143.1, 138.4, 138.0, 136.9, 133.8, 130.4, 129.9, 127.7, 127.2, 121.9, 115.5, 53.70, 53.69, 23.4, 21.6, and 21.3. IR (neat): νmax 3016 (Csp2-H), 2920 (Csp3-H), and 2847 (Csp3-H) cm−1. HRMS (ESI-TOF): Calcd for C19H22NO2S+ [M + H]+ 328.1366; found 328.1351 (100%). 17: 1H-NMR (500 MHz, CDCl3): δ 7.72 (nfod, Japp = 8.3 Hz, 2H), 7.25 (nfod, Japp = 8.2 Hz, 2H), 7.03 (br d, J = 2.3 Hz, 1H), 6.82 (d, J = 2.1 Hz, 1H), 6.07 (tqd, simulated as J = 2.1, 1.3, 0.8 Hz, 1H), 3.05–2.97 (m, ΣJs = ~14 Hz, 2H), 2.38 (t, J = 0.8 Hz, 3H), 1.93 (br t, J = 2.2 Hz, 3H), 1.82 (br m, 3H), and 1.81 (dt, J = 1.3, 1.3 Hz, 3H). 13C-NMR{1H} (126 MHz, CDCl3): δ 144.7, 136.6, 136.3, 130.0, 129.4, 126.84, 126.78, 124.0, 120.6, 117.8, 114.0, 113.3, 35.7, 24.1, 23.4, 22.3, and 21.7. HRMS (ESI-TOF): Calcd for C19H22NO2S+ [M + H]+ 328.1366; found 328.1357. This compound showed signs of oxidative decomposition upon handling and storage, and the HRMS data showed ions consistent with such processes (M + H+–H2, M + H+ + O, M + H+ – H2 + O).
6-Methyl-4-(prop-1-en-2-yl)-2-tosylisoindoline-7-d (16-d) and 6-Methyl-4-(propan-2-ylidene)-2-tosyl-4,5-dihydro-2H-isoindole-7-d (17-d). In a 10 mL culture tube, equipped with a magnetic stir bar, 14 (21 mg, 0.064 mmol) was added. 1,2-DCE (1.8 mL) and CD3OD (0.2 mL) were then added. The reaction mixture was heated at 130 °C for 16 h. The reaction mixture was concentrated and purified by flash column chromatography (30:1 Hex:EtOAc) to give 16-d (6.12 mg, 0.019 mmol, 29%) as a white crystalline solid. The ratio of 17-d:16-d in this experiment was ca. 1:2 (analysis of the 1H NMR spectrum of the crude product mixture prior to chromatographic purification). 16-d: mp: 136–140 °C. 1H-NMR (500 MHz, CDCl3): δ 7.76 (nfod, Japp = 8.3 Hz, 2H), 7.31 (nfod, Japp = 8.1 Hz, 2H), 6.95 (q, J = 0.8 Hz, 1H), 6.88 (s, 0.09H), 5.18 (dq, J = 1.6, 1.6 Hz, 1H), 4.92 (dq, J = 1.4, 1.0 Hz, 1H), 4.60 (nfom 2H), 4.57 (nfom 2H), 2.40 (s, 3H), 2.30 (d, J = 0.8 Hz, 3H), and 2.04 (dd, J = 1.5, 1.1 Hz, 3H). 13C-NMR{1H} (126 MHz, CDCl3): δ 143.7, 143.1, 138.4, 137.9, 136.8, 133.8, 130.5, 129.9, 127.8, 127.2, 115.5, 53.71, 53.68, 23.4, 21.6, and 21.3. (No resonance was observed at 122.0 ppm for the now-deuterated carbon.) IR (neat): νmax 3016 (Csp2-H), 2919 (Csp3-H), and 2848 (Csp3-H) cm−1. HRMS (ESI-TOF): Calcd for C19H21DNO2S+ [M + H]+ 329.1429; found 329.1414 (100%). 17-d: 1H-NMR (500 MHz, CDCl3): δ 7.72 (nfod, Japp = 8.3 Hz, 2H), 7.25 (nfod, Japp = 8.2 Hz, 2H), 7.04 (br d, J = 2.1 Hz, 1H), 6.82 (d, J = 2.1 Hz, 1H), 6.07 (m, 0.1H), 3.00–2.97 (m, ΣJs = ~14 Hz, 2H), 2.38 (t, J = 0.9 Hz, 3H), 1.93 (td, J = 2.0, 0.9 Hz, 3H), 1.82 (br m, 3H), and 1.81 (t, J = 1.2 Hz, 3H).
Methyl 6-Methyl-2-tosylisoindoline-4-carboxylate (33). In a 10 mL culture tube, equipped with a magnetic stir bar, 30 (30 mg, 0.087 mmol) was added. 1,2-DCE (2 mL) was added. The reaction mixture was heated to 130 °C for 16 h. The reaction mixture was concentrated and purified by flash column chromatography (10:1 Hex:EtOAc) to give 33 (12.61 mg, 0.037 mmol, 42%) as a white crystalline solid. mp: 175–180 °C. 1H-NMR (500 MHz, CDCl3): δ 7.78 (nfod, Japp = 8.0 Hz, 2H), 7.72 (s, 1H), 7.30 (nfod, Japp = 7.9 Hz, 2H), 7.16 (s, 1H), 4.87 (s, 2H), 4.60 (s, 2H), 3.89 (s, 3H), 2.39 (s, 3H), and 2.35 (s, 3H). 13C-NMR{1H} (126 MHz, CDCl3): δ 166.4, 143.8, 138.3, 137.9, 135.7, 134.0, 130.2, 130.0, 127.73, 127.72, 125.2, 54.9, 53.2, 52.3, 21.6, and 21.2. IR (neat): νmax 2952 (Csp3-H), 2923 (Csp3-H), 2856 (Csp3-H), and 1721(C=O) cm−1. HRMS (ESI-TOF): Calcd for C18H20NO4S+ [M + H]+ 346.1108; found 346.1094 (100%).
Methyl 6-Methyl-2-tosylisoindoline-4-carboxylate-7-d (33-d). In a 10 mL culture tube, equipped with a magnetic stir bar, 30 (21 mg, 0.064 mmol) was added. 1,2-DCE (1.8 mL) and CD3OD (0.2 mL) were then added. The reaction mixture was heated at 130 °C for 16 h. The reaction mixture was concentrated and purified by flash column chromatography (10:1 Hex:EtOAc) to give methyl 6-methyl-2-tosylisoindoline-4-carboxylate-7-d (33-d, 5.90 mg, 0.017 mmol, 27%) as a white crystalline solid: mp: 186–190 °C. 1H-NMR (500 MHz, CDCl3): δ 7.78 (nfod, Japp = 8.3 Hz, 2H, H2), 7.72 (q, J = 0.8 Hz, 1H, H5′), 7.30 (nfod, Japp = 8.2 Hz, 2H, H3), 7.16 (s, 0.1H, H7′), 4.87 (nfot, Japp = 2.1 Hz, 2H, H2′), 4.60 (nfot, Japp = 2.2 Hz, 2H, H9′), 3.89 (s, 3H, CO2CH3), 2.39 (s, 3H, C4CH3), and 2.35 (d, J = 0.9 Hz, 3H, C6′CH3). 13C-NMR{1H} (126 MHz, CDCl3): δ 166.4, 143.8, 138.2, 137.8, 135.7, 134.0, 130.2, 130.0, 127.74, 125.3, 54.9, 53.2, 52.3, 21.7, and 21.1. (No resonance was observed at 127.72, the largely deuterated C7′). IR (neat): νmax 2952 (Csp3-H), 2921 (Csp3-H), 2850 (Csp3-H), and 1721(C=O) cm−1. HRMS (ESI-TOF): [M + H]+ Calcd for C18H19DNO4S+ 347.1170; Found 347.1155 (100%).

4. Conclusions

The symmetrical bis-enyne 14 represents a rare type of substrate used for study of a thermally driven tetradehydro-Diels–Alder reaction. As such, the initially produced strained 1,2,4-cyclohexatriene intermediate 15 uniquely had an alkene (a 2-propenyl group) attached to C5. Heating 14 gave rise to not only the expected aromatic benzenoid product, 16, but a second, unusual isomeric side product, 17, whose structure was not immediately apparent. After detailed NMR analyses, it was concluded that 17 was a pyrrole derivative. Its formation is best rationalized by a [1,5]-hydrogen atom migration within 15 (or 21; cf. Figure 3). Deuteration studies using CD3OD or (CH3OD) showed that the presence of a protic additive was involved in forming both products 16 and 17. When the more acidic protic additive BHT was used, formation of the pyrrole side product 17 was completely suppressed, suggesting that the acidity of the OH additive played an important role in determining the fate of the 1,2,4-cyclohexatriene derivative 15.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30122610/s1: a PDF containing copies of 1H and 13C NMR spectra for all new compounds and the data resulting from the DP4+ analysis; a zip file of the Gaussian out files used in the DP4+ analysis; and a zip file of Mnova files containing raw NMR fid data of all new structures.

Author Contributions

Conceptualization, R.T. and Q.X.; methodology, R.T. and Q.X.; validation, R.T. and T.R.H.; formal analysis, R.T. and T.R.H.; resources, T.R.H.; data R.T. and T.R.H.; writing—original draft preparation, R.T. and T.R.H.; writing—review and editing, R.T., Q.X. and T.R.H.; funding acquisition, T.R.H. All authors have read and agreed to the published version of the manuscript.

Funding

This study was generously supported by the U.S. National Institutes of Health (NIH) (R35 GM127097). NMR spectra were obtained using an instrument partially funded through an NIH Shared Instrumentation Grant program (S10 OD011952). High-resolution mass spectrometric (HRMS) measurements were performed with instrumentation funded, in part, by an NIH Cancer Center Support Grant (P30 CA77598) at the University of Minnesota Masonic Cancer Center. Computations were performed using facilities made available at the Minnesota Supercomputing Institute (MSI).

Data Availability Statement

Data are contained within the article or its Supplementary Materials.

Conflicts of Interest

The authors have no conflicts of interest to declare. No generative artificial intelligence was used in any aspect of this study.

Abbreviations

The following abbreviations are used in this manuscript:
ATRattenuated total reflectance
BHTbutylated hydroxytoluene
DCEdichloroethane
DCMdichloromethane
ESIelectrospray ionization
EtOAcethyl acetate
GCMSgas chromatography–mass spectrometry
Hexhexanes
IRinfrared
mpmelting point
MPLCmedium pressure liquid chromatography
NMRnuclear magnetic resonance
NOEsnuclear Overhauser effects
NOESYnuclear Overhauser effect spectroscopy
NTsN-tosyl
SIsupporting information
TDDAtetradehydro-Diels–Alder
THFtetrahydrofuran
TLCthin layer chromatography

References

  1. Shakespeare, W.C.; Johnson, R.P. 1,2,3-Cyclohexatriene and cyclohexen-3-yne: Two new highly strained C6H6 isomers. J. Am. Chem. Soc. 1990, 112, 8578–8579. [Google Scholar] [CrossRef]
  2. Christl, M.; Braun, M.; Müller, G. 1,2,4-Cyclohexatriene, an isobenzene, and bicyclo[4.4.0]deca-1,3,5,7,8-pentaene, an isonaphthalene: Generation and trapping reactions. Angew. Chem. Int. Ed. Engl. 1992, 31, 473–476. [Google Scholar] [CrossRef]
  3. Kelleghan, A.V.; Bulger, A.S.; Witkowski, D.C.; Garg, N.K. Strain-promoted reactions of 1,2,3-cyclohexatriene and its derivatives. Nature 2023, 618, 748–754. [Google Scholar] [CrossRef] [PubMed]
  4. Witkowski, D.C.; Turner, D.W.; Bulger, A.S.; Houk, K.N.; Garg, N.K. Synthesis and reactivity of nitrogen-containing strained cyclic 1,2,3-trienes. Nat. Synth 2025. [Google Scholar] [CrossRef]
  5. Daoust, K.J.; Hernandez, S.M.; Konrad, K.M.; Mackie, I.D.; Winstanley, J., Jr.; Johnson, R.P. Strain estimates for small-ring cyclic allenes and butatrienes. J. Org. Chem. 2006, 71, 5708–5714. [Google Scholar] [CrossRef]
  6. Angus, R.O., Jr.; Johnson, R.P. Small-ring cyclic cumulenes: The structure and energetics of cyclic butatrienes and the synthesis of 1,2,3-cyclononatriene. J. Org. Chem. 1984, 49, 2880–2883. [Google Scholar] [CrossRef]
  7. Wessig, P.; Müller, G. The dehydro-Diels-Alder reaction. Chem. Rev. 2008, 108, 2051–2063. [Google Scholar] [CrossRef]
  8. Hoye, T.R.; Baire, B.; Niu, D.; Willoughby, P.H.; Woods, B.P. The hexadehydro-Diels-Alder reaction. Nature. 2012, 490, 208–212. [Google Scholar] [CrossRef]
  9. Danheiser, R.L.; Gould, A.E.; de la Pradilla, R.F.; Helgason, A.L. Intramolecular [4 + 2] cycloaddition reactions of conjugated enynes. J. Org. Chem. 1994, 59, 5514–5515. [Google Scholar] [CrossRef]
  10. Burrell, R.C.; Daoust, K.J.; Bradley, A.Z.; DiRico, K.J.; Johnson, R.P. Strained cyclic cumulene intermediates in Diels-Alder cycloaddition of enynes and diynes. J. Am. Chem. Soc. 1996, 118, 4218–4219. [Google Scholar] [CrossRef]
  11. Wills, M.S.B.; Danheiser, R.L. Intramolecular [4 + 2] cycloaddition reactions of conjugated ynones. Formation of polycyclic furans via the generation and rearrangement of strained heterocyclic allenes. J. Am. Chem. Soc. 1998, 120, 9378–9379. [Google Scholar] [CrossRef]
  12. Xu, Q.; Hoye, T.R. A cascade of strain-driven events converting benzynes to alkynylbenzocyclobutenes to 1,3-dien-5-ynes to cyclic allenes to benzocyclohexadienones. J. Am. Chem. Soc. 2024, 146, 6438–6443. [Google Scholar] [CrossRef] [PubMed]
  13. Ananikov, V.P. Ab initio study of the mechanisms of intermolecular and intramolecular [4 + 2] cycloaddition reactions of conjugated enynes. J. Phys. Org. Chem. 2001, 14, 109–121. [Google Scholar] [CrossRef]
  14. Prall, M.; Krüger, A.; Schreiner, P.R.; Hopf, H. The cyclization of parent and cyclic hexa-1,3-dien-5-ynes—A combined theoretical and experimental study. Chem. Eur. J. 2001, 7, 4386–4394. [Google Scholar] [CrossRef]
  15. Rodríguez, D.; Navarro-Vázquez, A.; Castedo, L.; Domínguez, D.; Saá, C. Cyclic allene intermediates in intramolecular dehydro Diels-Alder reactions: Labeling and theoretical cycloaromatization studies. J. Org. Chem. 2003, 68, 1938–1946. [Google Scholar] [CrossRef]
  16. Watson, W. On byproducts and side products. Org. Process Res. Dev. 2012, 16, 1877. [Google Scholar] [CrossRef]
  17. Tokimizu, Y.; Wieteck, M.; Rudolph, M.; Oishi, S.; Fujii, N.; Hashmi, A.S.K.; Ohno, H. Dual gold catalysis: A novel synthesis of bicyclic and tricyclic pyrroles from N-propargyl ynamides. Org. Lett. 2015, 17, 604–607. [Google Scholar] [CrossRef]
  18. Zanardi, M.M.; Sarotti, A.M. Sensitivity analysis of DP4+ with the probability distribution terms: Development of a universal and customizable method. J. Org. Chem. 2021, 86, 8544–8548. [Google Scholar] [CrossRef]
  19. Smadja, W. Electrophilic addition to allenic derivatives: Chemo-, regio-, and stereochemistry and mechanisms. Chem. Rev. 1983, 83, 263–320. [Google Scholar] [CrossRef]
  20. Sharma, R.K.; Shoulders, B.A.; Gardner, P.D. Oxymercuration of allenes. J. Org. Chem. 1967, 32, 241–244. [Google Scholar] [CrossRef]
  21. Rodríguez, D.; Navarro-Vázquez, A.; Castedo, L.; Domínguez, D.; Saá, C. Strained intermediates in intramolecular dehydro Diels-Alder reactions: Rearrangement of cyclic allenes via 1,2-dehydro[10]annulenes. J. Am. Chem. Soc. 2001, 123, 9178–9179. [Google Scholar] [CrossRef] [PubMed]
  22. Olmstead, W.N.; Margolin, Z.; Bordwell, F.G. Acidities of water and simple alcohols in dimethyl sulfoxide solution. J. Org. Chem. 1980, 45, 3295–3299. [Google Scholar] [CrossRef]
  23. Bordwell, F.G.; McCallum, R.J.; Olmstead, W.N. Acidities and hydrogen bonding of phenols in dimethyl sulfoxide. J. Org. Chem. 1984, 49, 1424–1427. [Google Scholar] [CrossRef]
  24. Hoye, T.R.; Hanson, P.R.; Vyvyan, J.R. A Practical guide to first-order multiplet analysis in 1H NMR spectroscopy. J. Org. Chem. 1994, 59, 4096–4103. [Google Scholar] [CrossRef]
  25. Hoye, T.R.; Zhao, H. A method for easily determining coupling constant values: An addendum to “A practical guide to first-order multiplet analysis in 1H NMR spectroscopy”. J. Org. Chem. 2002, 67, 4014–4016. [Google Scholar] [CrossRef]
  26. Parisot, W.; Haddad, M.; Phansavath, P.; Lefèvre, G.; Ratovelomanana-Vidal, V. A versatile, functional group-tolerant, and bench-stable iron precatalyst for building arene and triazine rings by [2+2+2] cycloadditions. Chem. Eur. J. 2024, 30, e202400096. [Google Scholar] [CrossRef]
  27. Trost, B.M.; Rudd, M.T. Ruthenium-catalyzed cycloisomerizations of diynols. J. Am. Chem. Soc. 2005, 127, 4763–4776. [Google Scholar] [CrossRef]
Figure 2. The unexpected formation of 17, an isomer of the aromatized TDDA product, 16.
Figure 2. The unexpected formation of 17, an isomer of the aromatized TDDA product, 16.
Molecules 30 02610 g002
Figure 3. (a) Nuclear Overhauser enhancements in 17. (b) Chemical shifts of Csp2 protons in 18 [17] vs. 17. (c) Proposed mechanism: branching from cyclic allene 15 to account for formation of 16 and 17.
Figure 3. (a) Nuclear Overhauser enhancements in 17. (b) Chemical shifts of Csp2 protons in 18 [17] vs. 17. (c) Proposed mechanism: branching from cyclic allene 15 to account for formation of 16 and 17.
Molecules 30 02610 g003
Figure 4. (a) Possible mechanisms for the key steps to produce the deuterated products 16-d and 17-d. (b) Relevant results from Saá and coworkers showing the effect of phenol product distribution from the TDDA reaction of 22. (c) The more acidic protic additive BHT quickly tautomerizes the 1,2,4-cyclohexatriene 15 to 16 before 15 has time to isomerize to 20 via [1,5]-hydrogen atom migration.
Figure 4. (a) Possible mechanisms for the key steps to produce the deuterated products 16-d and 17-d. (b) Relevant results from Saá and coworkers showing the effect of phenol product distribution from the TDDA reaction of 22. (c) The more acidic protic additive BHT quickly tautomerizes the 1,2,4-cyclohexatriene 15 to 16 before 15 has time to isomerize to 20 via [1,5]-hydrogen atom migration.
Molecules 30 02610 g004
Figure 5. Substrate 30 containing only one enyne moiety leads to 33 as the only isolable product, suggesting that the [1,5]-hydrogen atom migration in 31 (to give, transiently, 32) is not occurring to a significant extent.
Figure 5. Substrate 30 containing only one enyne moiety leads to 33 as the only isolable product, suggesting that the [1,5]-hydrogen atom migration in 31 (to give, transiently, 32) is not occurring to a significant extent.
Molecules 30 02610 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tang, R.; Xu, Q.; Hoye, T.R. A Case of Competitive Aromatization vs. Sigmatropic [1,5]-Hydrogen Atom Migration in a 1,2,4-Cyclohexatriene Intermediate Derived from a Bis-Enyne Cyclization. Molecules 2025, 30, 2610. https://doi.org/10.3390/molecules30122610

AMA Style

Tang R, Xu Q, Hoye TR. A Case of Competitive Aromatization vs. Sigmatropic [1,5]-Hydrogen Atom Migration in a 1,2,4-Cyclohexatriene Intermediate Derived from a Bis-Enyne Cyclization. Molecules. 2025; 30(12):2610. https://doi.org/10.3390/molecules30122610

Chicago/Turabian Style

Tang, Rong, Qian Xu, and Thomas R. Hoye. 2025. "A Case of Competitive Aromatization vs. Sigmatropic [1,5]-Hydrogen Atom Migration in a 1,2,4-Cyclohexatriene Intermediate Derived from a Bis-Enyne Cyclization" Molecules 30, no. 12: 2610. https://doi.org/10.3390/molecules30122610

APA Style

Tang, R., Xu, Q., & Hoye, T. R. (2025). A Case of Competitive Aromatization vs. Sigmatropic [1,5]-Hydrogen Atom Migration in a 1,2,4-Cyclohexatriene Intermediate Derived from a Bis-Enyne Cyclization. Molecules, 30(12), 2610. https://doi.org/10.3390/molecules30122610

Article Metrics

Back to TopTop