ACS Publications. Most Trusted. Most Cited. Most Read
Indium(III)-Catalyzed Stereoselective Synthesis of Tricyclic Frameworks by Cascade Cycloisomerization Reactions of Aryl 1,5-Enynes
My Activity
  • Open Access
Article

Indium(III)-Catalyzed Stereoselective Synthesis of Tricyclic Frameworks by Cascade Cycloisomerization Reactions of Aryl 1,5-Enynes
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

The Journal of Organic Chemistry

Cite this: J. Org. Chem. 2021, 86, 14, 9515–9529
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.joc.1c00825
Published June 25, 2021

Copyright © 2021 American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

The indium(III)-catalyzed cascade cycloisomerization reaction of 1,5-enynes with pendant aryl nucleophiles is reported. The reaction proceeds in cascade under mild reaction conditions, using InI3 (5 mol %) as a catalyst with a range of 1,5-enynes furnished with aryl groups (phenyl and phenol) at alkene (E and Z isomers) and with terminal and internal alkynes. Using 1-bromo-1,5-enynes, a one-pot sequential indium-catalyzed cycloisomerization and palladium-catalyzed cross-coupling with triorganoindium reagents were developed. The double cyclization is stereospecific and operates via a biomimetic cascade cation-olefin through 1,5-enyne cyclization (6-endo-dig) and subsequent C–C hydroarylation or C–O phenoxycyclization. Density functional theory (DFT) computational studies on 1,5-enynyl aryl ethers support a two-step mechanism where the first stereoselective 1,5-enyne cyclization produces a nonclassical carbocation intermediate that evolves to the tricyclic reaction product through a SEAr mechanism. Using this approach, a variety of tricyclic heterocycles such as benzo[b]chromenes, phenanthridines, xanthenes, and spiroheterocyclic compounds are efficiently synthesized with high atom economy.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2021 American Chemical Society

Introduction

Click to copy section linkSection link copied!

The design of synthetic methodologies based on catalytic cascade reactions constitutes an ideal tool for the construction of complex molecules with high chemo-, regio-, and stereoselectivity. (1) In particular, catalytic cascade cycloisomerization reactions allow the synthesis of a large structural diversity of molecules under complete atom economy and mild reaction conditions. As an example, cascade polyene cyclizations are one of the most impressive biosynthetic transformations known, and their chemical emulation represents a major challenge in modern synthetic chemistry. (2) Usually, these transformations involve the epoxide activation in a polyenic compound using oxophilic Lewis acids under stoichiometric or catalytic conditions. (3) Alternatively, electrophilic alkyne activation under metal catalysis has been recently envisaged as a different synthetic approach to promote catalytic cascade polyenynic π-cyclizations. (4) The catalytic electrophilic activation of alkynes promotes the addition of nucleophiles and allows the formation of new carbon–carbon and carbon–heteroatom bonds in an intermolecular and intramolecular manner. Although this methodology has been associated with the use of carbophilic late precious transition metals such as platinum (5) or gold (6) as catalysts, main group metals such as gallium (7) or indium (8) have been shown as valuable alternatives (Scheme 1a).

Scheme 1

Scheme 1. Indium(III)-Catalyzed Electrophilic Activation of Alkynes
Indium(III) is a soft Lewis acid that exhibits a dual-mode catalytic activity as σ-acid and π-acid, enabling both the electrophilic activation of carbon–heteroatom and to carbon–carbon unsaturated bonds. (9) In addition, reactions involving indium are used to provide high chemoselectivity and substantial economical, environmental, and safety advantages. (10) Along the years, these attractive chemical properties have been exploited in classical synthetic transformations as oxophilic Lewis acid either by using stoichiometric or catalytic conditions. (11) Recent contributions have demonstrated the synthetic utility of indium(III) salts as carbophilic π-acid catalysts in the electrophilic activation of alkynes. (12)
Metal-catalyzed 1,n-enyne cycloisomerization reactions constitute a straightforward methodology for the synthesis of polycyclic structures. (13) In particular, gold(I)-catalyzed cycloisomerization reactions of functionalized 1,5- and 1,6-enynes have been widely explored for the synthesis of polycyclic organic compounds. (14) On the other hand, indium(III)-catalyzed enyne cycloisomerization reactions were first reported by Chatani, (15) and more recently, Corey described the stereoselective synthesis of complex chiral polycyclic molecules by cascade cycloisomerization of chiral aryl polyenynes under indium(III) catalysis. (16a) In this contribution, the catalytic activity of In(III) as π-acid for the electrophilic activation of alkynes is remarked and attributed to the vacant 5s and 5p orbitals, which might lead to coordinate the C–C triple bond by bidentate complexation. Later on, Corey also reported the superior catalytic activity of diiodoindium(III) cation (InI2+), generated by the addition of Ag(I) salts to InI3. (16b)
As part of a long-term research on indium chemistry, (17) our group has recently reported intramolecular hydroarylation and hydroalkoxylation reactions and sequential indium-catalyzed polyyne reactions under indium(III) catalysis (Scheme 1b). (18) Herein, we report the In(III)-catalyzed cascade cycloisomerization reactions of 1,5-enynes (E and Z isomers) furnished with aryl nucleophiles and density functional theory (DFT) studies about the mechanism of the reaction and the nature of the organoindium intermediates involved (Scheme 1c).

Results and Discussion

Click to copy section linkSection link copied!

Our investigation started with the cycloisomerization reaction of (E)-1,5-enynyl aryl ether 1a under In(III) catalysis. This substrate was chosen to measure the indium alkynophilicity, the regioselectivity of the reaction (6-endo vs 5-exo), and to establish comparisons with other metal catalysts. (14a) In our first experiments, we found that the treatment of (E)-1a with 5 mol % InI3 in 1,2-dichloroethane (DCE) at 60 °C resulted in the stereoselective formation of the tricyclic compound 2a in 58% yield after 5 h as the only isolated product as a separable mixture of diastereoisomers (cis/trans = 25:75, Table 1, entry 1). Alternatively, the use of InBr3 (5 mol %) also provided 2a in similar yield and diastereoselectivity, whilst InCl3 (5 mol %) resulted to be ineffective (entries 2 and 3, respectively). In all cases, trans-fused 2a was identified as the major diastereoisomer and its formation can be explained by a double regioselective 6-endo-dig/endo-trig cyclization in cascade.
Table 1. Indium-Catalyzed Reactions with (E)-1,3-Dimethoxy-5-[(3-methylhept-2-en-6-yn-1-yl)oxy]benzene (1a)
entryInX3solventT (°C)t (h)ayield (%)b2a (cis/trans)c
1InI3DCE6055825:75
2InBr3DCE60205333:67
3InCl3DCE6048d 
4InI3eDCE60245531:69
5InI3toluene6026511:89
6InI3ftoluenert66220:80
7InI3fDCMrt76319:81
8InI3fDCM–2024d 
9InI3gDCE60572h 
10In(NTf2)3DCE60563i 
11In(OTf)3DCE6024j 
12In(acac)3DCE6048d 
a

Monitored by thin-layer chromatography (TLC).

b

Isolated yield.

c

Determined by gas chromatography–mass spectrometry (GCMS).

d

No reaction observed.

e

2 mol %.

f

20 mol %.

g

AgSbF6 (5 mol %) as a cocatalyst.

h

Mixture of products I:2a (4:1 ratio).

i

Compound II was isolated.

j

Decomposition.

The stereoselectivity of this In(III)-catalyzed 1,5-enyne cycloisomerization was then studied under different reaction conditions. In this endeavor, we found that the reaction also takes place with 2 mol % of InI3 in similar stereoselectivity but longer reaction times (55%, cis/trans = 31:69, entry 4). Interestingly, the yield, stereoselectivity, and reaction time were slightly improved using toluene as a solvent (65%, cis/trans = 11:89, entry 5). The reaction also proceeds at room temperature using 20 mol % of InI3 in toluene or dichloromethane (DCM) with similar results (entries 6 and 7), but no reaction was observed at −20 °C (entry 8). Surprisingly, the combination of InI3 (5 mol %) with the halide abstractor AgSbF6 (5 mol %) to generate the diiodonium cation (InI2+) gave a new major tricyclic product (I) concomitant with 2a (I:2a = 4:1 ratio) as a separable mixture in a combined 72% isolated yield (entry 9). The formation of compound I, as a mixture of cis/trans (11:89) diastereoisomers, can be explained by ether cleavage, Friedel–Crafts C-alkylation, and 1,5-enyne phenoxycyclization reaction. In a similar pathway, treatment with In(NTf2)3 (5 mol %) resulted in the formation of the bicyclic compound II (63% yield, entry 10) with an alkene phenoxycyclization as the last step. The use of In(OTf)3 as a catalyst resulted in decomposition (entry 11). No reaction was observed employing In(acac)3 as the catalyst, recovering the starting 1,5-enyne 1a (entry 12). In this optimization process, the important role of the catalyst counteranion is noteworthy, confirming InI3 as the most efficient π-catalyst. Although different solvents can be used, toluene was the solvent of choice for this 1,5-enyne cyclization. The high chemoselectivity of InI3 in the electrophilic activation of the alkyne over the ether cleavage is remarkable.
To further study the chemo-, regio-, and stereoselectivity and access to novel functionalized benzo[b]chromenes, our investigation was extended to 1,5-enynes functionalized at the arene and alkyne units and with different alkene geometries. Under the previously optimized reaction conditions, we were pleased to find that the reaction of 1,5-enyne (E)-1b furnished with a methyl group at the terminal alkyne proceeded successfully to give trans-2b in 68% yield as a separable diastereoisomeric mixture (cis/trans= 22:78, Table 2, entry 2). This synthetic transformation was more efficient and selective than using Hg(II), probably due to the softer Lewis acid character of In(III). (14a) Interestingly, the reaction using 1-bromo-1,5-enyne (E)-1c also provided 7-bromobenzo[b]chromene trans-2c diastereoselectively in good yield (62%, cis/trans= 27:73, entry 3). The synthesis of trans-2c should allow access to a variety of substituted benzo[b]chromenes by metal-catalyzed cross-coupling reactions. (17)
Table 2. Indium-Catalyzed Reactions with 1,5-Enynyl Aryl Ethers 1a–dc
a

Major isolated diastereomer.

b

Isolated yield.

c

Measured by 1H NMR.

The role of electronic effects was examined using phenyl 1,5-enynyl ether (E)-1d. Interestingly, under the previously developed reaction conditions, the double cycloisomerization proceeded with complete stereochemical inversion giving rise to the cis-benzo[b]chromene cis-2d as a single diastereoisomer by 1H NMR (61% yield, Table 2, entry 4). The isomerization suggests a two-step mechanism, where the lower nucleophilicity of the phenyl group facilitates the stereochemical inversion after the first 6-endo-dig cycloisomerization reaction to produce the most thermodynamically stable diastereoisomer.
Prompted by these interesting results, we studied the influence of alkene stereochemistry on the reactivity, regio-, and stereoselectivity. Previous metal-catalyzed 1,5-enyne cycloisomerization reactions have shown a divergent outcome based on the alkene geometry. (14c) In our case, the In(III)-catalyzed cycloisomerization with (Z)-1a also proceeded with 6-endo regioselectivity without isomerization providing the cis-2a as a single diastereoisomer as determined by 1H NMR in 64% yield (Table 2, entry 5). The same regiochemistry and stereochemical outcomes were observed using 1-bromo-1,5-enyne (Z)-1c (entry 6). Interestingly, the reaction with 1,5-enyne (Z)-1d also proceeded with full retention of the stereochemistry to obtain cis-2d as the only diastereoisomer as established by 1H NMR (entry 7). These results point out at least three reactivity patterns: the reaction is stereospecific through a two-step mechanism; electronic effects affect the stereochemical outcome; and the cis stereoisomer is thermodynamically more favorable.
After these interesting results, we explored the reaction with 1,5-enynyl phenyl N-tosylamines, substrates of interest for the synthesis of nitrogen heterocycles such as phenanthridines or indoles. In this venture, we found that the reaction of 3,5-dimethoxyphenyl 1,5-enynyl N-tosylamine (E)-3a with InI3 (5 mol %) in toluene at 60 °C gave the phenanthridine trans-4a in 90% yield in just 2 h as the only diastereoisomer detected by 1H NMR (Table 3, entry 1). Analogously, the reaction with 1,5-enynes (E)-3b and (E)-3c furnished with a methyl and a phenyl group at the alkyne afforded phenanthridines trans-4b and trans-4c stereoselectively as the only isolated products in 85 and 82% yield (Table 3, entries 2 and 3, respectively). Furthermore, the reaction with 1-bromo-1,5-enyne (E)-3d led to the 7-bromo-hexahydrophenanthridine trans-4d in 73% yield (entry 4). Although no isomerization was detected with these 1,5-enynes, the reaction of (E)-1,5-enyne 3e, furnished with an unsubstituted phenyl group, proceeded with complete inversion and the cis-4e was obtained as a single diastereoisomer (49% yield, entry 5). As previously reported, the isomerization can be explained by a stepwise mechanism where after the first 6-endo-dig cyclization, the synthetic intermediate could isomerize toward the most thermodynamically stable polycyclic compound. Therefore, we can conclude that aryl 1,5-enynes 3a–d furnished with the N-tosylamine moiety are more reactive than the ethers 1a–d and the cycloisomerization reaction takes place with complete retention of the alkene configuration.
Table 3. Indium-Catalyzed Reactions of 1,5-Enynyl Aryl N-Tosylamines 3af
a

Isolated yield.

The stereospecificity of the cycloisomerization reaction was also explored with the Z-alkene analogues. In this case, we observed that the reaction with 1,5-enyne (Z)-3a afforded phenanthridine cis-4a as a single diastereoisomer in 88% yield (entry 6). Analogously, the reaction with bromoalkyne (Z)-3d provided phenanthridine cis-4d in 72% yield (entry 7) and the reaction with 1,5-enyne (Z)-3e without methoxy groups at the phenyl unit gave cis-4e in 63% yield (entry 8). As previously observed with the 1,5-enynyl ether (Z)-1d, in this case, the reaction proceeded with full retention of the alkene stereochemistry. Finally, we also explored the reaction with (E)-1,5-enyne 3f, where the alkene moiety is disubstituted. Gratifyingly, the cascade cycloisomerization proceeded with complete regio- and stereospecificity to afford trans-4f in 79% yield as the only detected and isolated product. (19) This result demonstrates that the cascade cycloisomerization reaction is not limited to trisubstituted alkenes and resembles a biomimetic cascade olefin process.
The synthetic utility of the In(III)-catalyzed double cycloisomerization was then explored using 1,5-enynes equipped with a phenol moiety. Although InI3 showed as an efficient catalyst for the synthesis of benzo[b]furans from ortho-alkynylphenols, (18d) this cascade cycloisomerization process found some synthetic limitations using other transition metal catalysts. (20) In addition, the regioselective phenoxycyclization should provide access to xanthenes, a tricyclic skeleton of a relevant class of natural products. (21)
Using 1,5-enyne (E)-5a as a model substrate, (20) we found that InI3 (5 mol %) catalyzes the double cycloisomerization reaction in toluene at room temperature to afford the tricyclic 6-endo-dig/endo-trig product trans-6a in 86% isolated yield as a single diastereoisomer in 5 h (Table 4, entry 1). It is interesting to note the higher reactivity exhibited compared to the previous aryl 1,5-enynyl ethers 1a–d and N-tosylamines 3a–f as well as the chemical compatibility of the In(III) catalysis with the free hydroxyl group of the phenol. Furthermore, the reaction with (Z)-5a provided the cis-fused xanthene cis-6a in an excellent yield of 87% as the only isolated product (entry 2). These experimental results could be explained either by a stereospecific concerted or by a stepwise mechanism.
Table 4. Indium-Catalyzed Phenoxycyclization of 1,5-Enynes 5a–b
a

Isolated yield.

As the next step, we also tested the reaction of aryl 1,5-enynes 5 substituted at the alkyne. However, the complex synthesis of these substrates led us to consider a sequential procedure based on indium-catalyzed cascade cycloisomerization of the 1-bromo-1,5-enyne (E)-5b (14c) and subsequent functionalization by the cross-coupling reaction. With this approach in mind, we found that the reaction of 1-bromo-1,5-enyne (E)-5b and (Z)-5b with InI3 (5 mol %) results in a stereospecific manner, affording the expected 1-bromo-tetrahydroxanthenes trans-6b and cis-6b in 89 and 92% yield, respectively, as the only isolated products (Table 4, entries 3 and 4).
Having demonstrated the feasibility of both alkene isomers of 1-bromo-1,5-enyne 5b in the double cycloisomerization reaction, we assayed the one-pot sequential indium(III)-catalyzed 1,5-enyne cyclization and palladium-catalyzed cross-coupling reaction using triorganoindium reagents. (17,22) Gratifyingly, the treatment of (E)-5b with InI3 (5 mol %) in toluene at room temperature followed by addition of a solution of Ph3In (70 mol %) and Pd(PPh3)4 (5 mol %) in tetrahydrofuran (THF) at 80 °C afforded the 4-phenyltetrahydroxhantene trans-6c in 76% yield (Scheme 2). The sequential protocol using Me3In and n-Bu3In also gave the xanthene derivatives trans-6d and trans-6e in 83 and 81% yield, respectively. These results show the versatility of the In-catalyzed cascade cycloisomerization reaction using aryl 1,5-enynes and its chemical compatibility with Pd-catalyzed cross-coupling reactions.

Scheme 2

Scheme 2. Sequential One-Pot In-Catalyzed 1,5-Enyne Cycloisomerization and Pd-Catalyzed Cross-Coupling Reaction with (E)-5b
Finally, the indium-catalyzed cycloisomerization reaction of 1,5-enynes with aryl nucleophiles at the C-5 alkene unit (7a–c) was also briefly studied (Table 5). These substrates should allow the synthesis of spiroheterocycles if the cycloisomerization proceeds with 6-endo regioselectivity according to the previously described cation-olefin mechanism. (14a) Interestingly, the treatment of 1,5-enyne 7a with InI3 (5 mol %) in toluene at 60 °C afforded oxaspirane 8a as the only isolated product in an excellent yield of 84% (Table 5, entry 1). As expected, the cascade cycloisomerization reaction proceeded with 6-endo-dig/endo-trig regioselectivity to afford the Markovnikov product exclusively. Analogously, the reaction using 1,5-enynyl benzyl N-tosylamine 7b provided the azaspirane 8b in 92% yield (entry 2). The reaction with 1-bromo-1,5-enyne 7c afforded the corresponding spirane 8c in 76% yield (entry 3). In addition, the one-pot sequential indium-catalyzed cycloisomerization of 7c followed by the cross-coupling reaction with Ph3In using Pd(PPh3)4 as the catalyst provided the phenyl-substituted azaspirane 8d in 61% overall yield (two steps, entry 4).
Table 5. Synthesis of Spiroheterocycles by Indium-Catalyzed Reaction with 1,5-Enynes 7a–cc
a

Isolated yield.

b

Obtained from 7c by sequential In-catalyzed cycloisomerization and Pd-catalyzed cross-coupling.

c

Overall yield (two steps).

Mechanistic Studies

According to the experimental results, we postulate that the course of the indium(III)-catalyzed double cascade cycloisomerization could be viewed as either a concerted process or a stepwise route depending on the arene nucleophilicity (Scheme 3). In a two-step mechanism, we postulate that the initial η2 coordination of the indium(III) halide with the alkyne moiety (C) would trigger 1,5-enyne cyclization to form the intermediate D. (2) This intermediate should not be a pure carbocation and could be seen as a resonance hybrid of two resonance structures, an indium-stabilized homoallylic carbocation (D) and a cyclopropylindium ylide. Once there, the second cyclization should proceed through a Friedel–Crafts type alkylation reaction, subsequent aromatization and protodemetallation should provide the corresponding tricyclic compound, regenerating the catalytic species (Scheme 3). The mechanistic pathway and the nature of the transition states and synthetic intermediates should also depend on the substituents at the alkene, alkyne, or arene units.

Scheme 3

Scheme 3. General Plausible Mechanism for the In(III)-Catalyzed Cascade Cycloisomerization Reaction of 1,5-Enynes 1a–d and 3a–f
Our experimental data show that the In(III)-catalyzed double cycloisomerization reaction of aryl-substituted (Z)-1,5-enynes is stereoselective, yielding in all cases the cis adducts with complete selectivity. For example, independently of the electronic nature of the aromatic ring, the final cis-products are obtained, cis-2a, cis-2c, and cis-2d from (Z)-1a, (Z)-1c, and (Z)-1d, respectively (Table 2, entries 5–7), as for the 1,5-enyne cycloisomerizations of Table 3 (entries 6–8) and Table 4. Accordingly, one would expect that using (E)-1,5-enynes, the tricyclic products would belong to the trans series. In fact, this is true for all of the dimethoxyphenyl, electron-rich arenes, 1,5-enynes like (E)-1a (Table 2, entry 1, cis/trans = 11:89), and others (Table 2, entries 1–3; Table 3, entries 1–4; Table 4). However, surprisingly, the unsubstituted phenyl 1,5-enynes (E)-1d and (E)-3e break this rule, affording the opposite cis diastereoisomers of the final adduct cis-2d (Table 2, entry 4) and cis-4e (Table 3, entry 5). In other words, phenyl ether 1,5-enynes 1d and 3e lead diastereoselectively to the cis final adducts, regardless of the E or Z configuration of the initial double bonds. To gain insights into the intriguing behavior of aryl 1,5-enynyl ethers (E)-1d and (E)-3e, we set out to study the reaction theoretically. (23)
Computational studies (24) for (Z)- and (E)-1,5-enynyl aryl ether 1a (Scheme 4) showed that the formation of the first cycle occurs viaTS1-Z or TS1-E, two transition states originated by the nucleophilic 6-endo-dig alkene attack to the indium-activated electrophilic triple bond. The activation energy is very similar for the two compounds, 17.3 and 18.9 kcal/mol, respectively. As expected, the aromatic ring does not participate in the formation of the first ring, which affords intermediates INT1 stereospecifically, INT1-cis from Z, and INT1-trans from E starting materials. These intermediates are low in energy, 0.3–1.8 kcal/mol, presenting a bicyclic structure [4.1.0] and a partial indium carbene character. The computational data shows that the first step (TS1) is rate limiting since the formation of the second cycle by an electrophilic aromatic substitution in TS2-type transition states proceeds with lower activation energies, 8.9 for TS2-cis and 12.6 kcal/mol for TS2-trans. As expected for this type of reaction (SEAr), electron-rich arenes are positively affected by the stabilization of the incipient positive charge in the aryl ring by the electron-releasing methoxy groups. For the same reason, the arenium intermediates INT2a (cis and trans) are very stable in the presence of methoxy substituents (−12.3 and −7.6 kcal/mol). Although we did not compute them, the easy final deprotonation and hydrolysis (protodemetallation) of the C–In bond after INT2a intermediates would render exclusive adducts cis-2a (from Z-1a, black pathway in Scheme 4) and trans-2a (from E-1a, red line), in complete agreement with the experimental results.

Scheme 4

Scheme 4. DFT-Calculated Mechanism of the Reaction of (Z)-1a and (E)-1a for the Selective Formation of INT2-cis and -trans
Intermediate INT1 presents very interesting structures, as they can be described by two resonance forms, I and II (Scheme 5, top). The former is a fused bicyclic skeleton with a polarized C–In bond, whilst II shows the zwitterionic character, with the positive charge on the methylated tertiary carbon and an anionic alkenyl-indium motif. The NBO analysis of INT1(OMe) and INT1(H) affords Wiberg bond orders showing the mixed character of both structures, slightly more akin to structure I, since the internal C–C bond b is advanced but not completely formed (BO = 0.67–0.81), and the C–C bond a presents slight but not complete double bond character (BO = 1.33–1.43). The bonding distances also show mixed characteristics of both resonance forms, with computed values of δa = 1.40 Å and δb = 1.68 Å. Thus, resonance form II also has a significant participation in the actual structure of these intermediates. We rationalized that, given the weakness of bond b and the partial carbocationic character at the C-3 position (Scheme 5, bottom), the bicyclic species INT1-cis and INT1-trans could potentially interconvert by an isomerization equilibrium. Indeed, a transition state was located (TS-isom) bearing an almost planar tertiary carbocation, where the adjacent carbon (labeled as C-3) is flipping between the upper and lower faces of the cyclohexene plane (Scheme 5). The activation energy of this process is quite low, ca. 13.5–15.1 kcal/mol, making the isomerization plausible, at least in some circumstances. However, as mentioned before for, the 1a starting materials, the second step (TS2) is low enough in energy (8.9 and 12.6 kcal/mol, Scheme 4) to outcompete the isomerization, providing a complete selectivity (experimental > 95:5) for the cis and trans final products.

Scheme 5

Scheme 5. DFT-Calculated Isomerization Process between the Intermediates INT1-cis and INT1-trans
The scenario is quite different for 1,5-enynes (E)- and (Z)-1d without methoxy groups (Scheme 6). The initial cyclization through TS1 is also rate limiting, with values of 17.4 and 19.6 kcal/mol. After the first transition state, the structure and energies of INT1 intermediates are very similar to the previous ones. However, due to the absence of stabilizing methoxy groups in the aromatic ring, the second cyclization (TS2) increases its energy significantly in ca. 5–7 kcal/mol (12.8 and 19.1 kcal/mol, Scheme 6) above the values noted for 1a. This fact is especially important in the case of the trans cyclization, which also shows larger ring strain, producing a sluggish cyclization (19.1 kcal/mol), which becomes slower than the isomerization (15.1 kcal/mol) between the two INT1 isomers. Therefore, intermediate INT1-trans prefers to isomerize to INT1-cis rather than cyclize, and the reaction follows the red line in Scheme 6 to cis isomer. Meanwhile, the double cyclization of the Z isomer proceeds via the black line to lead stereospecifically to the cis isomer. These observations can explain how both starting materials converge in INT1-cis to give the same cis final isomer under Curtin–Hammett conditions.

Scheme 6

Scheme 6. DFT-Calculated Mechanism of the Reaction of (Z)-1d and (E)-1d under Curtin–Hammett Conditions

Conclusions

Click to copy section linkSection link copied!

Indium(III) iodide is an efficient catalyst to promote a double cycloisomerization reaction of 1,5-enynes with pendant aryl nucleophiles. The reaction can be performed under mild reaction conditions using 5 mol % of catalyst and proceeds in cascade through alkyne electrophilic activation with complete 6-endo regioselectivity via a biomimetic cascade cation-olefin process. In some cases, the double cycloisomerization is stereospecific with the retention of the alkene configuration. In addition, the synthetic transformation is highly versatile, allowing 1,5-enynes substituted at the alkyne and alkene units and phenyl groups and phenol derivatives as nucleophiles. Accordingly, a diverse group of polycyclic heterocycles such as benzo[b]chromenes, phenanthridines, xanthenes, and spiroheterocyclic compounds was synthesized. Computational studies on aryl 1,5-enynyl ethers support a mechanism consisting of two consecutive cyclizations: the first one is a 6-endo-dig process catalyzed by a regioselective alkyne electrophilic activation and a second cyclization through a nonstereospecific SEAr process.

Experimental Section

Click to copy section linkSection link copied!

General Methods

All reactions were carried out in flame-dried glassware under argon using standard gastight syringes, cannulae, and septa. Toluene and THF were distilled from sodium/benzophenone. Dry MeOH, DCE, Et3N, and other commercially available reagents were used as received. Reaction temperatures refer to external bath temperatures. Butyllithium was titrated prior to use. Indium(III) iodide (99.998%) and indium(III) bromide (99.999%) were purchased from Aldrich and used as received under argon. The reactions were monitored by TLC using precoated silica gel plates (Alugram Xtra SIL G/UV254, 0.20 mm thick), UV light as the visualizing agent, and ethanolic phosphomolybdic acid as the developing agent. Flash column chromatography was performed with 230–400 mesh silica gel. 1H and 13C NMR spectra were recorded in CDCl3 at 300 K using a Bruker Advance 300 MHz, 400 MHz or a Bruker Advance 500 MHz spectrometer and calibrated to the solvent peak. Mass spectra were obtained with a MAT 95XP Magnetic Sector EI spectrometer or with a QSTAR Elite hybrid quadrupole time-of-flight (TOF) ESI mass spectrometer, both operating in the positive ionization mode. Gas chromatography (GC) was performed on a Trace 1300 autosampling GC with a TG-5SILMS capillary column and equipped with an ISQ QD mass spectrometer.

(E)-1,3-Dimethoxy-5-[(3-methyloct-2-en-6-yn-1-yl)oxy]benzene [(E)-1b] (14a)

To a cooled solution of enyne (E)-1a (14c) (202.8 mg, 0.78 mmol) in dry THF (10 mL) at 0 °C, n-BuLi (0.38 mL, 0.82 mmol, 2.19 M in hexanes) was added dropwise. After 30 min, MeI (0,06 mL, 0.94 mmol) was added, and the reaction mixture was stirred for 2 h. The reaction was quenched with EtOH (1 mL), the solvent was evaporated, and the corresponding residue was purified by flash chromatography (2% EtOAc/hexanes) to afford (E)-1b (162.6 mg, 76%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 6.11–6.09 (m, 3H), 5.54–5.50 (m, 1H), 4.51 (d, J = 6.5 Hz, 2H), 3.77 (s, 6H), 2.26 (m, 4H), 1.77 (s, 3H), 1.74 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 161.6, 160.9, 139.9, 120.5, 93.7, 93.1, 78.6, 76.2, 65.0, 55.5, 39.0, 17.7, 16.7, 3.6; HRMS (ESI) m/z: [M + Na]+ calcd for C17H22O3Na 297.1461; found 297.1458.

(E)-1-[(7-Bromo-3-methylhept-2-en-6-yn-1-yl)oxy]-3,5-dimethoxybenzene [(E)-1c]

To a room temperature solution of (E)-3-methylhept-2-en-6-yn-1-ol (25) (105.7 mg, 0.85 mmol) in acetone (5 mL), N-bromosuccinimide (NBS) (166.7 mg, 0.93 mmol) and AgNO3 (15.8 mg, 0.09 mmol) were added and the reaction mixture was stirred for 1 h. The reaction mixture was diluted with Et2O (10 mL), washed with H2O (10 mL) and brine (10 mL), dried with anhydrous MgSO4, filtered, and concentrated in vacuo to afford (E)-7-bromo-3-methylhept-2-en-6-yn-1-ol as an orange oil, which was used in the next step without further purification. The crude product was added to a solution of triphenylphosphine (329.4 mg, 1.25 mmol) and 3,5-dimethoxyphenol (193.2 mg, 1.25 mmol) in THF (10 mL). To this solution at 0 °C, diisopropyl azodicarboxylate (DIAD) (0.25 mL, 1.25 mmol) was added dropwise and the reaction mixture was stirred for 5 h at 60 °C. Then, the solvent was evaporated in vacuo and the corresponding residue was purified by flash chromatography (5% EtOAc/hexanes) to afford (E)-1c (161.0 mg, 56% in two steps) as a yellow oil; 1H NMR (300 MHz, CDCl3) δ 6.10 (m, 3H), 5.54–5.50 (m, 1H), 4.50 (d, J = 6.6 Hz, 2H), 3.76 (s, 6H), 2.36–2.29 (m, 4H), 1.74 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 161.6, 160.8, 139.1, 120.9, 93.7, 93.1, 79.6, 64.8, 55.4, 38.6, 38.0, 18.6, 16.6; IR (neat) νmax 2929, 2841, 1594, 1460, 1204, 1150, 1062, 819 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C16H19BrO3Na 361.0409; found 361.0416.

(E)-[(3-Methylhept-2-en-6-yn-1-yl)oxy]benzene [(E)-1d]

To a 0°C solution of (E)-3-methylhept-2-en-6-yn-1-ol (25) (201.9 mg, 1.63 mmol), triphenylphosphine (640.0 mg, 2.44 mmol) and phenol (229.6 mg, 2.44 mmol) in THF (10 mL), DIAD (0.48 mL, 2.44 mmol) was added dropwise. The reaction mixture was stirred for 5 h at 60 °C in an oil bath and the solvent was evaporated in vacuo. Then, the corresponding residue was purified by flash chromatography (5% EtOAc/hexanes) to afford (E)-1d (251.4 mg, 77%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 7.31–7.28 (m, 2H), 6.94–6.93 (m, 3H), 5.58–5.53 (m, 1H), 4.56 (d, J = 6.5 Hz, 2H), 2.35–2.31 (m, 4H), 1.95 (t, J = 2.3 Hz, 1H), 1.75 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 158.9, 139.1, 129.5, 121.1, 120.8, 114.8, 83.9, 68.9, 64.7, 38.3, 17.3, 16.6; HRMS (EI) m/z: [M]+ calcd for C14H16O 200.1201; found 200.1196.

(Z)-1,3-Dimethoxy-5-[(3-methylhept-2-en-6-yn-1-yl)oxy]benzene [(Z)-1a]

DIAD (0.51 mL, 2.57 mmol) was added dropwise at 0 °C to a solution of (Z)-3-methylhept-2-en-6-yn-1-ol (26) (212.6 mg, 1.71 mmol), triphenylphosphine (674.1 mg, 2.57 mmol), and 3,5-dimethoxyphenol (396.2 mg, 2.57 mmol) in THF (10 mL). The reaction mixture was stirred for 3 h at 60 °C in an oil bath and the solvent was evaporated in vacuo. Then, the corresponding residue was purified by flash chromatography (5% EtOAc/hexanes) to afford (Z)-1a (227.3 mg, 51%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 6.11–6.09 (m, 3H), 5.60 (td, J = 6.8, 1.6 Hz, 1H), 4.51 (d, J = 6.8 Hz, 2H), 3.76 (s, 6H), 2.36–2.34 (m, 4H), 1.97 (t, J = 2.4 Hz, 1H), 1.81 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 161.6, 160.7, 139.7, 122.0, 93.6, 93.0, 83.7, 69.0, 64.5, 55.4, 31.2, 23.3, 17.5; IR (neat) νmax 3289, 2935, 2840, 1593, 1474, 1204, 1147, 1060, 818 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C16H20O3Na 283.1304; found 283.1313.

(Z)-1-[(7-Bromo-3-methylhept-2-en-6-yn-1-yl)oxy]-3,5-dimethoxybenzene [(Z)-1c]

To a room temperature solution of (Z)-3-methylhept-2-en-6-yn-1-ol (26) (115.7 mg, 0.93 mmol) in acetone (5 mL), NBS (182.4 mg, 1.02 mmol) and AgNO3 (17.3 mg, 0.10 mmol) were added and the reaction mixture was stirred for 1 h. The reaction mixture was diluted with Et2O (10 mL) and washed with H2O (10 mL) and brine (10 mL), dried with MgSO4 (anhydrous), filtered, and concentrated in vacuo to afford (Z)-7-bromo-3-methylhept-2-en-6-yn-1-ol as an orange oil, which was used in the next step without further purification. The crude product was added to a solution of triphenylphosphine (354.0 mg, 1.35 mmol) and 3,5-dimethoxyphenol (208.9 mg, 1.35 mmol) in THF (10 mL), DIAD (0.26 mL, 1.35 mmol) at 0 °C was added dropwise, and the reaction mixture was heated at 60 °C for 5 h. Then, the solvent was evaporated in vacuo and the corresponding residue was purified by flash chromatography (5% EtOAc/hexanes) to afford (Z)-1c (149.9 mg, 47% in two steps) as a yellow oil; 1H NMR (300 MHz, CDCl3) δ 6.11–6.09 (m, 3H), 5.61 (m, 1H), 4.49 (d, J = 6.7 Hz, 2H), 3.77 (s, 6H), 2.34 (m, 4H), 1.80 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 161.5, 160.7, 139.5, 122.0, 93.6, 93.1, 79.5, 64.4, 55.4, 38.8, 31.0, 23.3, 18.8; IR (neat) νmax 2927, 2840, 1593, 1474, 1204, 1148, 1061, 818 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C16H19BrO3Na 361.0409; found 361.0403.

(Z)-[(3-Methylhept-2-en-6-yn-1-yl)oxy]benzene [(Z)-1d]

To a cooled solution of (Z)-3-methylhept-2-en-6-yn-1-ol (26) (215.8 mg, 1.74 mmol), triphenylphosphine (684.6 mg, 2.61 mmol) and phenol (245.6 mg, 2.61 mmol) in THF (10 mL) at 0 °C, DIAD (0.52 mL, 2.61 mmol) was added dropwise and the reaction mixture was stirred for 5 h at 60 °C in an oil bath. The solvent was evaporated in vacuo and the corresponding residue was purified by flash chromatography (5% EtOAc/hexanes) to afford (Z)-1d (236.7 mg, 68%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 7.31–7.25 (m, 2H), 6.95–6.91 (m, 3H), 5.62–5.60 (m, 1H), 4.55 (d, J = 6.8, 2H), 2.38–2.33 (m, 4H), 1.98–1.96 (m, 1H), 1.82 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 158.8, 139.5, 129.5, 122.2, 120.8, 114.8, 83.8, 69.0, 64.3, 31.3, 23.3, 17.5; IR (neat) νmax 3292, 2917, 2868, 1598, 1494, 1235, 1172, 1006, 752 cm–1; HRMS (EI) m/z: [M]+ calcd for C14H16O 200.1201; found 200.1197.

(E)-N-(3,5-Dimethoxyphenyl)-4-methyl-N-(3-methyloct-2-en-6-yn-1-yl)benzenesulfonamide [(E)-3b]

To a 0 °C solution of 1,5-enyne (E)-3a (14c) (205.6 mg, 0.50 mmol) in dry THF (10 mL), n-BuLi (0.21 mL, 0.52 mmol, 2.5 M in hexanes) was added dropwise. After 30 min, MeI (0,04 mL, 0.60 mmol) was added and the reaction mixture was left stirring for 2 h. The reaction was quenched with EtOH (3 mL), the solvent was evaporated, and the corresponding residue was purified by flash chromatography (10% EtOAc/hexanes) to afford (E)-3b (108.4 mg, 51%) as a white solid; 1H NMR (300 MHz, CDCl3) δ 7.56 (d, J = 8.4 Hz, 2H), 7.28–7.25 (m, 2H), 6.36 (d, J = 2.1 Hz, 1H), 6.20 (d, J = 2.1 Hz, 2H), 5.15 (t, J = 7.0 Hz, 1H), 4.13 (d, J = 6.9 Hz, 2H), 3.71 (s, 6H), 2.42 (s, 3H), 2.07 (m, 4H), 1.72 (s, 3H), 1.52 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 160.5, 143.4, 141.3, 139.1, 135.8, 129.4, 127.9, 119.7, 107.0, 100.1, 78.5, 76.0, 55.5, 48.7, 38.8, 21.6, 17.7, 16.2, 3.5; HRMS (ESI) m/z: [M + H]+ calcd for C24H30NO4S 428.1890; found 428.1893.

(E)-N-(3,5-Dimethoxyphenyl)-4-methyl-N-(3-methyl-7-phenylhept-2-en-6-yn-1-yl)benzenesulfonamide [(E)-3c]

To a solution of 1,5-enyne (E)-3a (14c) (163.0 mg, 0.39 mmol) in Et3N (4 mL), CuI (3.75 mg, 0.02 mmol), Pd(PPh3)2Cl2 (13.83 mg, 0.02 mmol), and iodobenzene (103.4 mg, 0.51 mmol) were added and the reaction mixture was stirred overnight at room temperature. Then, the reaction was quenched with H2O (15 mL) and the aqueous phase was extracted with Et2O (3 × 15 mL). The combined organic phase was washed with brine (50 mL), dried with MgSO4 (anhyd.), filtered, and concentrated in vacuo to afford (E)-3c (107.9 g, 57%) as an amorphous white solid after purification by column chromatography (10% EtOAc/hexanes); 1H NMR (300 MHz, CDCl3) δ 7.56 (d, J = 8.2 Hz, 2H), 7.34 (dd, J = 6.7, 3.1 Hz, 2H), 7.27–7.24 (m, 5H), 6.35 (t, J = 2.3 Hz, 1H), 6.20 (d, J = 2.3 Hz, 2H), 5.22 (t, J = 6.7 Hz, 1H), 4.15 (d, J = 6.8 Hz, 2H), 3.68 (s, 6H), 2.42 (s, 3H), 2.36 (t, J = 7.5 Hz, 2H), 2.20 (t, J = 7.5 Hz, 2H), 1.58 (s, J = 1.3 Hz, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 160.6, 143.5, 141.3, 138.9, 135.8, 131.6, 129.5, 128.3, 127.9, 127.7, 123.9, 120.1, 107.1, 100.2, 89.6, 82.2, 81.1, 55.5, 48.7, 38.5, 21.7, 18.6, 16.4; IR (neat) νmax 2925, 2841, 1596, 1458, 1346, 1205, 1154, 1066, 663 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C29H31NO4SNa: 512.1866; found 512.1853.

(E)-N-(7-Bromo-3-methylhept-2-en-6-yn-1-yl)-N-(3,5-dimethoxyphenyl)-4-methylbenzenesulfonamide [(E)-3d] (14c)

NBS (125.0 mg, 0.70 mmol) and AgNO3 (12.0 mg, 0.07 mmol) were added to a solution of the enyne (E)-3a (266.3 mg, 0.64 mmol) in acetone (6 mL), and the reaction mixture was stirred for 1 h at room temperature. The reaction mixture was diluted with Et2O (10 mL), the organic phase was washed with H2O (10 mL) and brine (10 mL), dried with MgSO4 (anhyd.), filtered, and concentrated in vacuo to afford, after purification by column chromatography (15% EtOAc/hexanes), (E)-3d (198.0 mg, 74%) as a yellow oil; 1H NMR (300 MHz, CDCl3) δ 7.56 (d, J = 8.0 Hz, 2H), 7.28 (m, 2H), 6.37 (m, J = 2.3 Hz, 1H), 6.20 (d, J = 2.3 Hz, 2H), 5.16 (t, J = 6.7 Hz, 1H), 4.13 (d, J = 6.9 Hz, 2H), 3.71 (s, 6H), 2.42 (s, 3H), 2.16–2.10 (m, 4H), 1.55 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 160.6, 143.5, 141.3, 138.5, 135.9, 129.5, 127.9, 120.2, 107.1, 100.2, 79.6, 55.5, 48.6, 38.4, 38.0, 21.7, 18.7, 16.3; HRMS (ESI) m/z: [M + Na]+ calcd for C23H26BrNO4SNa 514.0658; found: 514.0645.

(E)-4-Methyl-N-(3-methylhept-2-en-6-yn-1-yl)-N-phenylbenzenesulfonamide [(E)-3e]

DIAD (0.36 mL, 1.81 mmol) was added dropwise at 0 °C to a solution of (E)-3-methylhept-2-en-6-yn-1-ol (25) (150.0 mg, 1.21 mmol), triphenylphosphine (474.7 mg, 1.81 mmol), and p-toluenesulfonanilide (447.6 mg, 1.81 mmol) in THF (10 mL). The reaction mixture was stirred for 5 h at 60 °C in an oil bath and the solvent was evaporated in vacuo. The corresponding residue was purified by flash chromatography (20% EtOAc/hexanes) to afford (E)-3e (303.7 mg, 71%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 7.51 (d, J = 8.3 Hz, 2H), 7.27–7.24 (m, 5H), 7.06–7.04 (m, 2H), 5.18 (ddt, J = 8.2, 6.9, 1.3 Hz, 1H), 4.19 (d, J = 7.0 Hz, 2H), 2.42 (s, 3H), 2.10 (m, 4H), 1.84 (m, 1H), 1.49 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 143.4, 139.4, 138.7, 135.8, 129.5, 129.0, 128.9, 127.9, 127.8, 120.1, 83.7, 68.8, 48.5, 38.1, 21.7, 17.3, 16.1; IR (neat) νmax 3295, 2972, 2922, 1598, 1494, 1345, 1156, 1091, 654 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C21H23NO2SNa: 376.1341; found: 376.1338.

(Z)-N-(3,5-Dimethoxyphenyl)-4-methyl-N-(3-methylhept-2-en-6-yn-1-yl)benzenesulfonamide [(Z)-3a]

DIAD (0.34 mL, 1.74 mmol) was added dropwise at 0 °C to a solution of (Z)-3-methylhept-2-en-6-yn-1-ol (26) (144.2 mg, 1.16 mmol), triphenylphosphine (456.4 mg, 1.74 mmol), and N-(3,5-dimethoxybenzyl)-4-methylbenzenesulfonamide (27) (534.4 mg, 1.74 mmol) in THF (10 mL). The reaction mixture was stirred for 5 h at 60 °C in an oil bath and the solvent was evaporated in vacuo. The corresponding crude reaction product was purified by flash chromatography (15% EtOAc/hexanes) to afford (Z)-3a (359.8 mg, 75%) as a white solid; mp 100–102 °C; 1H NMR (300 MHz, CDCl3) δ 7.55 (d, J = 8.3 Hz, 2H), 7.28–7.25 (m, 2H), 6.37 (t, J = 2.3 Hz, 1H), 6.20 (d, J = 2.3 Hz, 2H), 5.24–5.20 (m, 1H), 4.15 (dd, J = 7.0, 1.2 Hz, 2H), 3.71 (s, 6H), 2.43 (s, 3H), 2.14–2.09 (m, 4H), 1.90 (m, 1H), 1.63 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 160.6, 143.4, 141.2, 138.2, 135.7, 129.4, 127.8, 121.4, 107.1, 100.1, 83.8, 68.8, 55.4, 48.5, 30.7, 23.0, 21.6, 17.1; IR (neat) νmax 3294, 2927, 2840, 1594, 1459, 1348, 1154, 1065, 663 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C23H27NO4SNa 436.1553; found 436.1555.

(Z)-N-(7-Bromo-3-methylhept-2-en-6-yn-1-yl)-N-(3,5-dimethoxyphenyl)-4-methylbenzenesulfonamide [(Z)-3d]

NBS (75.0 mg, 0.42 mmol) and AgNO3 (7.1 mg, 0.04 mmol) were added to a solution of 1,5-enyne (Z)-3a (158.3 mg, 0.38 mmol) in acetone (5 mL), and the reaction mixture was stirred for 1 h at room temperature. The reaction mixture was diluted with Et2O (10 mL), the organic phase was washed with H2O (10 mL) and brine (10 mL), dried with MgSO4 (anhyd.), filtered, and concentrated in vacuo. After purification by column chromatography (15% EtOAc/hexanes), (Z)-3d was obtained (128.9 mg, 69%) as a yellow oil; 1H NMR (300 MHz, CDCl3) δ 7.55 (d, J = 8.0 Hz, 2H), 7.28–7.25 (m, 2H), 6.37 (t, J = 2.3 Hz, 1H), 6.19 (dt, J = 2.3, 1.3 Hz, 2H), 5.24–5.20 (m, 1H), 4.13 (d, J = 7.0 Hz, 2H), 3.71 (s, 6H), 2.42 (s, 3H), 2.10 (m, 4H), 1.62 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 160.7, 143.5, 141.2, 138.2, 135.8, 129.5, 127.9, 121.5, 107.1, 100.2, 79.7, 55.5, 48.6, 38.5, 30.6, 23.2, 21.7, 18.5; IR (neat) νmax 2925, 2839, 1592, 1457, 1347, 1204, 1152, 1065, 662 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C23H26BrNO4SNa 514.0658; found 514.0651.

(Z)-4-Methyl-N-(3-methylhept-2-en-6-yn-1-yl)-N-phenylbenzenesulfonamide [(Z)-3e]

To a 0 °C solution of (Z)-3-methylhept-2-en-6-yn-1-ol (26) (151.4 mg, 1.21 mmol), triphenylphosphine (479.2 mg, 1.81 mmol) and p-toluene-sulfonanilide (451.8 mg, 1.81 mmol) in THF (15 mL), DIAD (0.36 mL, 1.81 mmol) was added dropwise. The reaction mixture was stirred for 5 h at 60 °C in an oil bath and the solvent was evaporated in vacuo. The corresponding residue was purified by flash chromatography (10% EtOAc/hexanes) to afford (Z)-3e as a colorless oil (329.3 mg, 77%); 1H NMR (300 MHz, CDCl3) δ 7.49 (d, J = 8.3 Hz, 2H), 7.29–7.25 (m, 5H), 7.05–7.02 (m, 2H), 5.24–5.19 (m, 1H), 4.19 (d, J = 7.0 Hz, 2H), 2.43 (s, 3H), 2.10–2.05 (m, 4H), 1.89 (t, J = 2.4 Hz, 1H), 1.61 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 143.4, 139.4, 138.3, 135.8, 129.5, 129.0, 129.0, 127.9, 127.8, 121.4, 83.8, 68.8, 48.5, 30.7, 23.0, 21.6, 17.1; IR (neat) νmax 3290, 2920, 2869, 1596, 1493, 1345, 1162, 1092, 656 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C21H23NO2SNa 376.1341; found 376.1334.

(E)-N-(3,5-Dimethoxyphenyl)-N-(hept-2-en-6-yn-1-yl)-4-methylbenzenesulfonamide [(E)-3f]

To a 0 °C solution of (E)-hept-2-en-6-yn-1-ol (28) (150.0 mg, 1.36 mmol), triphenylphosphine (535.1 mg, 2.04 mmol), and N-(3,5-dimethoxybenzyl)-4-methylbenzenesulfonamide (27) (627.0 mg, 2.04 mmol) in THF (10 mL), DIAD (0.40 mL, 2.04 mmol) was added dropwise and the reaction mixture was stirred for 5 h at 60 °C in an oil bath. The solvent was evaporated in vacuo and the corresponding residue was purified by flash chromatography (20% EtOAc/hexanes) to afford (E)-3f (396.6 g, 73%) as a white solid; mp 90–92 °C; 1H NMR (300 MHz, CDCl3) δ 7.53 (d, J = 8.2 Hz, 2H), 7.24 (d, J = 8.2 Hz, 2H), 6.35 (t, J = 2.3 Hz, 1H), 6.18 (d, J = 2.3 Hz, 2H), 5.56–5.40 (m, 2H), 4.07 (d, J = 5.9 Hz, 2H), 3.68 (s, 6H), 2.40 (s, 3H), 2.10 (m, J = 3.1, 2.6 Hz, 4H), 1.87 (m, 1H); 13C{1H} NMR (75 MHz, CDCl3) δ 160.5, 143.5, 140.9, 135.6, 133.2, 129.4, 127.8, 125.8, 107.2, 100.1, 83.5, 68.8, 55.4, 52.9, 31.0, 21.6, 18.3; IR (neat) νmax 3289, 2934, 2840, 1593, 1458, 1345, 1153, 1090, 662 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C22H25NO4SNa 422.1396; found 422.1385.

(Z)-4-Methoxy-2-(3-methylhept-2-en-6-yn-1-yl)phenol [(Z)-5a]

PBr3 (162.4 mg, 0.60 mmol) was added dropwise to a solution of (Z)-3-methylhept-2-en-6-yn-1-ol (26) (150.2 mg, 1.21 mmol) in Et2O (10 mL) at 0 °C and was stirred for 30 min. The reaction was quenched with H2O (10 mL) and the aqueous phase was extracted with Et2O (3 × 10 mL). The combined organic phase was washed with H2O (30 mL), NaHCO3 (30 mL, satd. sol.), and brine (10 mL), dried with MgSO4 (anhyd.), filtered, and concentrated in vacuo to afford (Z)-7-bromo-5-methylhept-5-en-1-yne as an orange oil, which was used in the next step without further purification. The crude product was dissolved in toluene (10 mL) and NaH 95% (35.5 mg, 1.33 mmol) and 4-methoxyphenol (165.2 mg, 1.33 mmol) was added. The reaction mixture was stirred overnight at room temperature, quenched with a NH4Cl saturated solution, and the aqueous phase was extracted with Et2O (3 × 10 mL). The resulting organic phase was washed with H2O (30 mL) and brine (30 mL), dried with MgSO4 (anhyd.), filtered, and concentrated in vacuo to yield, after purification by column chromatography (10% EtOAc/hexanes), (Z)-5a (158.8 mg, 57%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 6.73 (d, J = 8.5 Hz, 1H), 6.68–6.64 (m, 2H), 5.41–5.36 (m, 1H), 4.68 (s, 1H), 3.75 (s, 3H), 3.37 (d, J = 7.1 Hz, 2H), 2.42–2.40 (m, 2H), 2.38–2.35 (m, 2H), 1.99 (t, J = 2.5 Hz, 1H), 1.78 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 153.7, 148.1, 136.2, 128.2, 124.2, 116.3, 115.9, 112.1, 84.1, 68.9, 55.9, 30.8, 29.6, 23.1, 17.2; IR (neat) νmax 3405, 3290, 2913, 2834, 1497, 1430, 1199, 1039, 803 cm–1; HRMS (EI) m/z: [M]+ calcd for C15H18O2 230.1301; found 230.1296.

(Z)-2-(7-Bromo-3-methylhept-2-en-6-yn-1-yl)-4-methoxyphenol [(Z)-5b]

To a room temperature solution of (Z)-3-methylhept-2-en-6-yn-1-ol (26) (84.8 mg, 0.68 mmol) in acetone (5 mL), NBS (133.8 mg, 0.75 mmol) and AgNO3 (12.7 mg, 0.07 mmol) were added, and the reaction mixture was stirred for 1 h. The reaction mixture was diluted with Et2O (10 mL), washed with H2O (10 mL) and brine (10 mL), dried with MgSO4 (anhyd.), filtered, and concentrated in vacuo to afford (Z)-7-bromo-3-methylhept-2-en-6-yn-1-ol as an orange oil, which was used in the next step without further purification. The crude product was dissolved in Et2O (10 mL) and cooled at 0 °C, PBr3 (92.4 mg, 0.34 mmol) was then added dropwise, and the reaction mixture was stirred for 30 min. The reaction was quenched with H2O (10 mL), and the aqueous phase was extracted with Et2O (3 × 10 mL). The resulting organic phase was washed with H2O (30 mL), a NaHCO3 saturated solution (30 mL), and brine (10 mL), dried with MgSO4 (anhyd.), filtered, and concentrated in vacuo to afford (Z)-1,7-dibromo-5-methylhept-5-en-1-yne as an orange oil, which was used in the next step without further purification.
The crude product in toluene (10 mL) was added to a solution of NaH 95% (20.0 mg, 0.75 mmol) and 4-methoxyphenol (92.9 mg, 0.75 mmol) at room temperature, and the reaction mixture was stirred overnight. The reaction mixture was quenched with a NH4Cl saturated solution (2 mL) and was extracted with Et2O (3 × 10 mL). The combined organic phase was washed with H2O (30 mL) and brine (30 mL), dried with MgSO4 (anhyd.), filtered, and concentrated in vacuo to afford, after purification by column chromatography (10% EtOAc/hexanes), (Z)-5b (90.4 mg, 43%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 6.74–6.63 (m, 3H), 5.42–5.38 (m, 1H), 4.89 (s, 1H), 3.76 (s, 3H), 3.36 (d, J = 7.2 Hz, 2H), 2.40–2.35 (m, 4H), 1.77 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 153.7, 148.0, 135.9, 128.3, 124.2, 116.2, 115.8, 112.1, 79.9, 55.9, 38.5, 30.7, 29.3, 23.3, 18.6; IR (neat) νmax 3408, 2915, 2835, 1504, 1432, 1201, 1041, 805 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C15H17BrO2Na: 331.0304; found 331.0315.

General Procedure for In(III)-Catalyzed Cascade Cycloisomerization Reactions of 1,5-Enynes

In a Schlenk tube filled with argon, InI3 (5 mol %) was placed and a solution of the corresponding 1,5-enyne (∼0.07 M) in toluene was added. The reaction mixture was stirred at 60 °C in an oil bath (for 1a–d, 3a–f, and 7a–c) or at room temperature (for 5a–b) until the starting material is consumed (TLC). The reaction was quenched with NH4Cl (10 mL, satd. sol.), poured into a separatory funnel, and the aqueous phase was extracted with Et2O (3 × 10 mL). The resulting combined organic phase was washed with brine (15 mL), dried with anhydrous MgSO4, filtered, and concentrated under reduced pressure to afford, after purification by column chromatography, the corresponding tricyclic product.

(6aS*,10aS*)-1,3-Dimethoxy-10a-methyl-6a,9,10,10a-tetrahydro-6H-benzo[c]chromene [trans-2a] (14c)

According to the general procedure, the reaction of 1,5-enyne (E)-1a (87.4 mg, 0.34 mmol) with InI3 (8.5 mg, 0.017 mmol) gave 2a (56.8 mg, 65%; cis/trans = 11:89) as a colorless oil. Purification by column chromatography (2% EtOAc/hexanes) provided pure trans-2a: 1H NMR (300 MHz, CDCl3) δ 6.05–6.02 (m, 2H), 5.82–5.76 (m, 1H), 5.36 (dq, J = 9.8, 2.1 Hz, 1H), 4.08 (dd, J = 10.3, 4.1 Hz, 1H), 3.99 (dd, J = 12.4, 10.3 Hz, 1H), 3.77 (s, 3H), 3.74 (s, 3H), 3.07–3.00 (m, 1H), 2.71–2.67 (m, 1H), 2.22 (m, 2H), 1.57–1.53 (m, 1H), 1.17 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 160.3, 159.3, 155.8, 129.3, 123.8, 113.8, 94.1, 92.3, 66.4, 55.3, 55.3, 41.7, 33.3, 32.0, 24.4, 18.1. HRMS (ESI) m/z: [M + Na]+ calcd for C16H20O3Na 283.1304; found 283.1317.

Scale-Up Experiment for (trans)-2a

In a Schlenk tube filled with argon, InI3 (38.1 mg, 0.077 mmol) was placed and a solution of the enyne (E)-1a (400.5 mg, 1.54 mmol) in toluene (22 mL) was added. The reaction mixture was stirred at 60 °C in an oil bath for 2 h, quenched with a NH4Cl (30 mL, satd. sol.), poured into a separatory funnel, and extracted with Et2O (3 × 30 mL). The combined organic phase was washed with brine (35 mL), dried with anhydrous MgSO4, filtered, and concentrated under reduced pressure. The residue was purified by column chromatography (2% EtOAc/hexanes) to afford 2a (252.4 mg, 63%; cis/trans = 16:84) as a colorless oil.

(4aS*,9aS*)-6,8-Dimethoxy-4a-methyl-4,4a,9,9a-tetrahydro-3H-xanthene (I)

According to the general procedure, the reaction of 1,5-enyne (E)-1a (150.2 mg, 0.58 mmol) with InI3 (14.3 mg, 0.029 mmol) in the presence of AgSbF6 (10.0 mg, 0.029 mmol) afforded a mixture of I:2a (4:1) by 1H NMR. After purification by column chromatography (1% EtOAc/hexanes), compound I was isolated as a colorless oil (81.1 mg, 54%, cis/trans = 19:81); 1H NMR (500 MHz, CDCl3) δ 6.05 (d, J = 0.9 Hz, 2H), 5.66–5.64 (m, 1H), 5.52 (ddt, J = 9.7, 2.5, 1.7 Hz, 1H), 3.79 (s, 3H), 3.75 (s, 3H), 2.72 (dd, J = 16.2, 5.5 Hz, 1H), 2.50–2.42 (m, 1H), 2.35–2.28 (m, 1H), 2.27–2.18 (m, 1H), 2.11 (dd, J = 16.2, 13.7 Hz, 1H), 1.98–1.92 (m, 1H), 1.86 (td, J = 12.0, 6.9 Hz, 1H), 1.11 (s, 3H). 13C{1H} NMR (126 MHz, CDCl3) δ 159.5, 158.5, 154.9, 128.8, 126.5, 104.0, 93.8, 91.0, 75.9, 55.4, 55.3, 38.6, 35.2, 25.1, 22.3, 15.9. HRMS (ESI) m/z: [M + H]+ calcd for C16H21O3 261.1491; found 261.1486.

2-(But-3-yn-1-yl)-5,7-dimethoxy-2-methylchromane (II)

According to the general procedure, the reaction of 1,5-enyne (E)-1a (90.1 mg, 0.35 mmol) with In(NTf2)3 (16.7 mg, 0.018 mmol) in DCE (5 mL) at 60 °C in an oil bath for 5 h afforded, after purification by column chromatography (5% EtOAc/hexanes), compound II (56.7 mg, 63%) as a yellow oil; 1H NMR (400 MHz, CDCl3) δ 6.03 (d, J = 2.4 Hz, 1H), 5.99 (d, J = 2.4 Hz, 1H), 3.78 (s, 3H), 3.75 (s, 3H), 2.64–2.49 (m, 2H), 2.37–2.32 (m, 2H), 1.97–1.91 (m, 2H), 1.86–1.82 (m, 1H), 1.79–1.73 (m, 2H), 1.28 (s, 3H); 13C{1H} NMR (101 MHz, CDCl3) δ 159.6, 158.6, 154.8, 102.5, 93.8, 91.1, 84.8, 75.3, 68.3, 55.5, 55.4, 38.4, 30.8, 23.8, 16.3, 13.1; IR (neat) νmax 3289, 2926, 2853, 1616, 1590, 1202, 1143, 1105, 811 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C16H20NaO3 283.1304; found 283.1317.

(6aS*,10aS*)-1,3-Dimethoxy-7,10a-dimethyl-6a,9,10,10a-tetrahydro-6H-benzo[c]chromene [trans-2b] (14a)

According to the general procedure, the reaction of 1,5-enyne (E)-1b (99.2 mg, 0.36 mmol) with InI3 (8.9 mg, 0.018 mmol) afforded 2b (67.5 mg, 68%; cis/trans = 22:78) as a colorless oil. Purification by column chromatography (2% EtOAc/hexanes) afforded pure trans-2b: 1H NMR (300 MHz, CDCl3) δ 6.04–6.01 (m, 2H), 5.46 (m, 1H), 4.38 (dd, J = 10.4, 3.6 Hz, 1H), 4.01–3.94 (m, 1H), 3.76 (s, 3H), 3.74 (s, 3H), 3.00 (dd, J = 13.1, 6.1 Hz, 1H), 2.66 (d, J = 12.1 Hz, 1H), 2.14 (m, 2H), 1.67 (s, 3H), 1.49–1.41 (m, 1H) 1.18 (s, 3H); 13C{1H} NMR (101 MHz, CDCl3) δ 160.3, 159.3, 155.7, 130.0, 123.5, 113.8, 93.9, 92.4, 64.4, 55.3, 55.3, 44.7, 33.7, 31.9, 23.7, 20.8, 18.4; HRMS (ESI) m/z: [M + Na]+ calcd for C17H22O3Na 297.1461; found 297.1454.

(6aS*,10aS*)-7-Bromo-1,3-dimethoxy-10a-methyl-6a,9,10,10a-tetrahydro-6H-benzo[c]chromene [trans-2c]

According to the general procedure, the reaction of 1,5-enyne (E)-1c (68.6 mg, 0.20 mmol) with InI3 (5.0 mg, 0.010 mmol) afforded 2c (42.5 mg, 62%; cis/trans = 27:73) as a white solid. Purification by column chromatography (2% EtOAc/hexanes) afforded pure trans-2c: 1H NMR (300 MHz, CDCl3) δ 6.18–6.16 (m, 1H), 6.05–6.02 (m, 2H), 4.57 (dd, J = 10.4, 3.5 Hz, 1H), 3.95 (dd, J = 11.8, 10.4 Hz, 1H), 3.76 (s, 3H), 3.74 (s, 3H), 3.14–3.07 (m, 1H), 2.99–2.93 (m, 1H), 2.26–2.22 (m, 2H), 1.49 (m, 1H), 1.25 (s, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 159.9, 159.5, 155.8, 131.1, 119.7, 112.4, 93.9, 92.5, 66.2, 55.4, 55.3, 46.8, 36.3, 31.3, 26.2, 18.5; IR (neat) νmax 2933, 2837, 1612, 1583, 1463, 1201, 1151, 1104, 814 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C16H19BrO3Na 361.0409; found 361.0418.

(6aR*,10aS*)-10a-Methyl-6a,9,10,10a-tetrahydro-6H-benzo[c]chromene [cis-2d]

According to the general procedure, the reaction of 1,5-enyne (E)-1d (105.4 mg, 0.53 mmol) with InI3 (13.0 mg, 0.027 mmol) or 1,5-enyne (Z)-1d (99.2 mg, 0.49) with InI3 (11.8 mg, 0.025 mmol) afforded cis-2d [64.3 mg, 61% from (E)-1d and 64.5 mg, 65% from (Z)-1d] as a yellow oil after purification by column chromatography (2% EtOAc/hexanes); 1H NMR (300 MHz, CDCl3) δ 7.23 (dd, J = 7.8, 1.7 Hz, 1H), 7.08 (ddd, J = 8.0, 7.2, 1.7 Hz, 1H), 6.90 (td, J = 7.5, 1.3 Hz, 1H), 6.79 (dd, J = 8.1, 1.4 Hz, 1H), 5.82–5.79 (m, 1H), 5.59 (ddt, J = 10.0, 3.8, 2.0 Hz, 1H), 4.22 (dd, J = 11.0, 3.0 Hz, 1H), 3.90 (dd, J = 11.0, 6.6 Hz, 1H), 2.34 (m, 1H), 2.02–1.98 (m, 2H), 1.88–1.68 (m, 2H), 1.35 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 154.3, 130.4, 129.9, 127.3, 127.1, 126.1, 120.9, 117.0, 66.8, 41.3, 33.9, 32.9, 29.1, 22.6; IR (neat) νmax 2922, 2855, 1579, 1488, 1447, 1218, 751 cm–1; HRMS (EI) m/z: [M]+ calcd for C14H16O 200.1201; found 200.1182.

(6aR*,10aS*)-1,3-Dimethoxy-10a-methyl-6a,9,10,10a-tetrahydro-6H-benzo[c]chromene [cis-2a]

According to the general procedure, the reaction of 1,5-enyne (Z)-1a (91.5 mg, 0.35 mmol) with InI3 (9.1 mg, 0.018 mmol) afforded cis-2a (58.6 mg, 64%) as a colorless oil after purification by column chromatography (2% EtOAc/hexanes); 1H NMR (300 MHz, CDCl3) δ 6.07 (d, J = 2.6 Hz, 1H), 6.01 (d, J = 2.6 Hz, 1H), 5.84–5.81 (m, 1H), 5.51 (ddt, J = 9.9, 3.6, 1.8 Hz, 1H), 4.13 (dd, J = 10.7, 2.8 Hz, 1H), 3.84 (dd, J = 10.7, 6.8 Hz, 1H), 3.77 (s, 3H), 3.74 (s, 3H), 2.26–2.24 (m, 1H), 2.22–2.18 (m, 1H), 1.99–1.78 (m, 3H), 1.39 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 160.7, 159.0, 156.5, 130.5, 125.9, 111.0, 93.7, 92.9, 66.8, 55.3, 55.3, 43.4, 33.2, 30.8, 25.7, 23.0; IR (neat) νmax 2922, 2837, 1612, 1583, 1463, 1200, 1153, 814 cm–1; HRMS (EI) m/z: [M]+ calcd for C16H20O3 260.1407; found 260.1402.

(6aR*,10aS*)-7-Bromo-1,3-dimethoxy-10a-methyl-6a,9,10,10a-tetrahydro-6H-benzo[c]chromene [cis-2c]

According to the general procedure, the reaction of 1,5-enyne (Z)-1c (84.0 mg, 0.25 mmol) with InI3 (6.1 mg, 0.012 mmol) afforded cis-2c (47.9 mg, 57%) as a yellow oil after purification by column chromatography (2% EtOAc/hexanes); 1H NMR (300 MHz, CDCl3) δ 6.21(m, 1H), 6.07 (d, J = 2.6 Hz, 1H), 6.05 (d, J = 2.6 Hz, 1H), 4.32 (dd, J = 10.9, 2.9 Hz, 1H), 4.06 (dd, J = 10.9, 8.0 Hz, 1H), 3.77 (s, 3H), 3.74 (s, 3H), 2.52–2.49 (m, 1H), 2.03 (m, 3H), 1.89 (m, 1H), 1.44 (s, 3H); 13C{1H} NMR (101 MHz, CDCl3) δ 160.3, 159.4, 156.7, 132.5, 121.6, 110.9, 93.9, 93.3, 65.8, 55.3, 51.3, 36.2, 29.2, 25.1, 24.8; IR (neat) νmax 2935, 2837, 1612, 1584, 1464, 1201, 1153, 815 cm–1; HRMS (EI) m/z: [M]+ calcd for C16H19BrO3 338.0512; found 338.0511.

(6aS*,10aS*)-1,3-Dimethoxy-10a-methyl-5-tosyl-5,6,6a,9,10,10a-hexahydrophenanthridine [trans-4a] (14c)

According to the general procedure, the reaction of 1,5-enyne (E)-3a (120.0 mg, 0.29 mmol) with InI3 (7.2 mg, 0.015 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), trans-4a (108.1 mg, 90%) as a white solid; 1H NMR (300 MHz, CDCl3) δ 7.51 (d, J = 8.3, 2H), 7.21–7.17 (m, 3H), 6.22 (d, J = 2.5, 1H), 5.71 (dq, J = 9.9, 3.3 Hz, 1H), 5.33 (dq, J = 9.8, 2.1 Hz, 1H), 4.00 (dd, J = 11.9, 4.2 Hz, 1H), 3.80 (s, 3H), 3.72 (s, 3H), 3.22 (dd, J = 13.6, 11.8 Hz, 1H), 2.99 (dt, J = 13.3, 4.1 Hz, 1H), 2.36 (s, 3H), 2.30 (m, 1H), 2.10–2.05 (m, 2H), 1.30–1.29 (m, 1H), 0.66 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 159.3, 158.2, 143.8, 137.9, 135.8, 129.6, 128.8, 127.4, 125.0, 120.4, 100.4, 96.3, 55.5, 55.4, 47.9, 41.0, 35.2, 31.9, 23.9, 21.6, 16.4; HRMS (ESI) m/z: [M + Na]+ calcd for C23H27NO4SNa 436.1553; found 436.1559.

(6aS*,10aS*)-1,3-Dimethoxy-7,10a-dimethyl-5-tosyl-5,6,6a,9,10,10a-hexahydrophenanthridine [trans-4b]

According to the general procedure, the reaction of 1,5-enyne (E)-3b (48.8 mg, 0.11 mmol) with InI3 (2.8 mg, 0.006 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), trans-4b (41.5 mg, 85%) as a white solid; 1H NMR (300 MHz, CDCl3) δ 7.53 (d, J = 8.3 Hz, 2H), 7.21–7.17 (m, 3H), 6.24 (d, J = 2.5 Hz, 1H), 5.39 (m, 1H), 4.38 (dd, J = 12.6, 3.6 Hz, 1H), 3.80 (s, 3H), 3.72 (s, 3H), 3.20 (t, J = 12.8 Hz, 1H), 2.93 (dd, J = 13.1, 6.2 Hz, 1H), 2.36 (s, 3H), 2.22–2.18 (m, 1H), 2.05–1.98 (m, 2H), 1.70 (s, 3H), 1.15–1.07 (m, 1H), 0.80 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 159.5, 158.1, 143.8, 137.7, 136.3, 130.8, 129.6, 127.4, 123.3, 120.4, 100.4, 96.6, 55.5, 55.4, 45.2, 44.1, 35.9, 31.9, 23.5, 21.7, 21.2, 17.1; IR (neat) νmax 2926, 2839, 1606, 1581, 1455, 1350, 1187, 1164, 671 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C24H29NO4SNa 450.1709; found 450.1705.

(6aR*,10aS*)-1,3-Dimethoxy-10a-methyl-7-phenyl-5-tosyl-5,6,6a,9,10,10a-hexahydrophenanthridine [trans-4c]

According to the general procedure, the reaction of 1,5-enyne (E)-3c (80.9 mg, 0.17 mmol) with InI3 (4.2 mg, 0.008 mmol) afforded trans-4c (66.3 mg, 82%) as a white solid after purification by column chromatography (5% EtOAc/hexanes); 1H NMR (300 MHz, CDCl3) δ 7.51 (d, J = 8.3 Hz, 2H), 7.36–7.28 (m, 3H), 7.21–7.11 (m, 4H), 7.13 (d, J = 2.5 Hz, 1H), 6.26 (d, J = 2.5 Hz, 1H), 5.67 (q, J = 3.4 Hz, 1H), 4.09 (dd, J = 12.7, 3.4 Hz, 1H), 3.79 (s, 3H), 3.76 (s, 3H), 3.09–3.04 (m, 1H), 2.98 (t, J = 12.8 Hz, 1H), 2.79 (dq, J = 12.9, 3.1 Hz, 1H), 2.37 (s, 3H), 2.23–2.21 (m, 2H), 1.41–1.30 (m, 1H), 0.94 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 159.4, 158.2, 143.7, 141.6, 137.9, 137.5, 136.6, 129.6, 128.3, 127.9, 127.5, 127.3, 126.9, 120.3, 100.5, 96.5, 55.5, 45.7, 43.4, 36.2, 31.7, 23.9, 21.7, 17.3; IR (neat) νmax 2935, 2837, 1604, 1579, 1418, 1349, 1201, 1163, 1152, 664 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C29H31NO4SNa 512.1866; found: 512.1860.

(6aS*,10aS*)-7-Bromo-1,3-dimethoxy-10a-methyl-5-tosyl-5,6,6a,9,10,10a-hexahydrophenanthridine [trans-4d] (14c)

According to the general procedure, the reaction of 1,5-enyne (E)-3d (79.8 mg, 0.17 mmol) with InI3 (4.1 mg, 0.008 mg) afforded, after purification by column chromatography (5% EtOAc/hexanes), trans-4d (58.3 mg, 73%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 7.60–7.57 (m, 2H), 7.24–7.21 (m, 2H), 7.20 (d, J = 2.5 Hz, 1H), 6.25 (d, J = 2.5 Hz, 1H), 6.08 (dd, J = 4.8, 2.8 Hz, 1H), 4.65 (dd, J = 13.0, 3.5 Hz, 1H), 3.81 (s, 3H), 3.73 (s, 3H), 3.24 (t, J = 12.8 Hz, 1H), 3.00 (dd, J = 13.3, 6.0 Hz, 1H), 2.41 (m, 1H), 2.38 (s, 3H), 2.14–2.05 (m, 2H), 1.20–1.12 (m, 1H), 0.93 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 159.2, 158.5, 143.9, 137.9, 136.3, 130.8, 129.8, 127.5, 121.0, 119.1, 100.8, 96.8, 55.5, 55.4, 47.2, 46.3, 38.2, 31.4, 25.9, 21.7, 17.3; HRMS (ESI) m/z: [M + H]+ calcd for C23H27BrNO4S 492.0838; found 492.0839.

(6aR*,10aS*)-10a-Methyl-5-tosyl-5,6,6a,9,10,10a-hexahydrophenanthridine [cis-4e]

According to the general procedure, reactions of 1,5-enyne (E)-3e (85.5 mg, 0.24 mmol) with InI3 (5.9 mg, 0.012 mmol) or 1,5-enyne (Z)-3e (89.0 mg, 0.25) with InI3 (6.3 mg, 0.013 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), cis-4e [41.9 mg, 49% from (E)-3e and 56.1 mg, 63% from (Z)-3e] as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 7.75–7.72 (m, 1H), 7.51–7.48 (m, 2H), 7.26–7.13 (m, 5H), 5.75 (ddd, J = 9.4, 4.7, 2.4 Hz, 1H), 5.45 (ddt, J = 9.2, 4.9, 2.3 Hz, 1H), 4.16 (dd, J = 14.1, 3.4 Hz, 1H), 3.23 (ddd, J = 15.2, 10.8, 4.4 Hz, 1H), 2.37 (s, 3H), 1.99–1.93 (m, 2H), 1.81 (d, J = 11.3 Hz, 1H), 1.57 (dt, J = 9.4, 2.3 Hz, 2H), 0.91 (s, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 143.7, 139.0, 137.4, 135.7, 129.7, 128.9, 127.6, 127.3, 126.4, 125.7, 125.5, 124.5, 48.4, 39.3, 34.7, 33.3, 25.7, 22.4, 21.6; IR (neat) νmax 2925, 1486, 1447, 1347, 1163, 1090, 658 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C21H23NO2SNa: 376.1341; found: 376.1348.

(6aR*,10aS*)-1,3-Dimethoxy-10a-methyl-5-tosyl-5,6,6a,9,10,10a-hexahydrophenanthridine [cis-4a]

According to the general procedure, the reaction of 1,5-enyne (Z)-3a (95.3 mg, 0.23 mmol) with InI3 (5.7 mg, 0.011 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), cis-4a (83.9 mg, 88%) as a white solid; mp 101–103 °C; 1H NMR (300 MHz, CDCl3) δ 7.52 (d, J = 8.1 Hz, 2H), 7.20 (d, J = 8.1 Hz, 2H), 6.97 (d, J = 2.6 Hz, 1H), 6.31 (d, J = 2.6 Hz, 1H), 5.79–5.75 (m, 1H), 5.41–5.38 (m, 1H), 4.14 (dd, J = 14.0, 3.2 Hz, 1H), 3.79 (s, 3H), 3.75 (s, 3H), 3.17 (dd, J = 14.0, 11.2 Hz, 1H), 2.37 (s, 3H), 2.15 (dt, J = 13.2, 4.2 Hz, 1H), 1.93–1.87 (m, 2H), 1.74 (d, J = 11.0 Hz, 1H), 1.42 (ddd, J = 13.1, 10.6, 5.4 Hz, 1H), 0.95 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 159.9, 158.0, 143.7, 137.9, 137.6, 129.7, 129.5, 127.3, 125.6, 120.2, 101.2, 97.8, 55.5, 55.3, 48.3, 41.5, 34.9, 28.0, 22.6, 22.3, 21.6; IR (neat) νmax 2928, 2837, 1607, 1579, 1460, 1349, 1161, 672 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C23H27NO4SNa 436.1553; found 436.1555.

(6aR*,10aS*)-7-Bromo-1,3-dimethoxy-10a-methyl-5-tosyl-5,6,6a,9,10,10a-hexahydrophenanthridine [cis-4d]

According to the general procedure, the reaction of 1,5-enyne (Z)-3d (90.2 mg, 0.18 mmol) with InI3 (4.6 mg, 0.009 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), cis-4d (64.9 mg, 72%) as a white solid; mp 144–146 °C; 1H NMR (300 MHz, CDCl3) δ 7.53 (d, J = 8.3 Hz, 2H), 7.23 (d, J = 8.3 Hz, 2H), 7.12 (d, J = 2.6 Hz, 1H), 6.34 (d, J = 2.6 Hz, 1H), 6.12 (t, J = 4.0 Hz, 1H), 4.65 (dd, J = 14.3, 3.2 Hz, 1H), 3.84 (s, 3H), 3.75 (s, 3H), 3.06 (dd, J = 14.3, 11.9 Hz, 1H), 2.37 (s, 3H), 2.30–2.24 (m, 1H), 1.99–1.95 (m, 2H), 1.79–1.75 (m, 1H), 1.25 (ddd, J = 13.6, 10.6, 6.7 Hz, 1H), 0.92 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 159.4, 158.3, 143.9, 137.8, 136.8, 131.1, 129.7, 127.7, 121.5, 120.1, 102.1, 98.3, 55.6, 55.3, 49.0, 47.4, 37.3, 26.1, 24.7, 21.7, 21.4; IR (neat) νmax 2932, 1608, 1579, 1461, 1352, 1203, 1164, 672 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C23H26BrNO4SNa 514.0658; found: 514.0657.

(6aS*,10aS*)-1,3-Dimethoxy-5-tosyl-5,6,6a,9,10,10a-hexahydrophenanthridine [trans-4f]

According to the general procedure, the reaction of 1,5-enyne (E)-3f (85.0 mg, 0.22 mmol) with InI3 (5.5 mg, 0.011 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), trans-4f (67.2 mg, 79%,) as a white solid; mp 101–103 °C; 1H NMR (500 MHz, CDCl3) δ 7.55 (d, J = 8.3 Hz, 2H), 7.21 (d, J = 8.3 Hz, 2H), 7.07 (d, J = 2.5 Hz, 1H), 6.26 (d, J = 2.5 Hz, 1H), 5.66–5.63 (m, 1H), 5.52 (dq, J = 9.7, 2.0 Hz, 1H), 4.22 (dd, J = 12.8, 3.7 Hz, 1H), 3.80 (s, 3H), 3.73 (s, 3H), 3.05 (dd, J = 12.8, 12.1 Hz, 1H), 2.82 (ddt, J = 13.0, 5.5, 2.5 Hz, 1H), 2.38 (s, 3H), 2.21 (td, J = 11.2, 11.2, 2.2 Hz, 1H), 2.19–1.99 (m, 2H), 2.02 (bt, J = 11.5 Hz, 1H), 1.03–0.96 (m, 1H); 13C{1H} NMR (126 MHz, CDCl3) δ 159.6, 158.5, 143.7, 139.1, 137.0, 129.7, 128.8, 127.7, 127.3, 114.8, 101.3, 96.6, 55.6, 55.4, 51.2, 39.2, 38.6, 27.2, 27.1, 21.7; IR (neat) νmax 2929, 2836, 1607, 1579, 1455, 1346, 1185, 1162, 665 cm–1; HRMS (ESI) m/z: [M + Na]+ calcd for C22H25NO4SNa 422.1396; found 422.1397.

(4S*,9S*)-7-Methoxy-4-methyl-4,9-tetrahydro-3H-xanthene [trans-6a] (20)

According to the general procedure, the reaction of 1,5-enyne (E)-5a (20) (97.6 mg, 0.42 mmol) with InI3 (10.5 mg, 0.021 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), trans-6a (83.9 mg, 86%) as a white solid; 1H NMR (400 MHz, CDCl3) δ 6.76 (d, J = 8.8 Hz, 1H), 6.70 (dd, J = 8.9, 2.9 Hz, 1H), 6.63 (d, J = 2.9 Hz, 1H), 5.68–5.65 (m, 1H), 5.49 (ddt, J = 9.7, 2.9, 1.6 Hz, 1H), 3.76 (s, 3H), 2.69 (dd, J = 13.9, 3.4 Hz, 1H), 2.67–2.47 (m, 2H), 2.34–2.29 (m, 1H), 2.26–2.22 (m, 1H), 1.98–1.93 (m, 1H), 1.86 (td, J = 11.9, 6.9 Hz, 1H), 1.10 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 153.1, 148.0, 128.5, 127.0, 123.1, 118.1, 114.5, 113.7, 75.5, 55.8, 39.1, 35.4, 28.5, 25.2, 16.0; HRMS (EI) m/z: [M]+ calcd for C15H18O2 [M]+: 230.1301; found: 230.1294.

(4S*,9R*)-1-Bromo-7-methoxy-4-methyl-4,9-tetrahydro-3H-xanthene [trans-6b] (14c)

According to the general procedure, the reaction of 1,5-enyne (E)-5b (14c) (105.4 mg, 0.31 mmol) with InI3 (7.7 mg, 0.016 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), trans-6b (93.8 mg, 89%) as a white solid; 1H NMR (300 MHz, CDCl3) δ 6.78–6.67 (m, 3H), 6.11–6.08 (m, 1H), 3.76 (s, 3H), 3.05 (dd, J = 16.1, 5.2 Hz, 1H), 2.86–2.82 (m, 1H), 2.57 (ddt, J = 16.1, 13.4, 1.0 Hz, 1H), 2.29–2.24 (m, 2H), 1.99–1.89 (m, 2H), 1.15 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 153.4, 147.2, 129.2, 123.8, 122.6, 117.9, 114.3, 114.1, 76.1, 55.8, 45.1, 34.9, 29.0, 25.9, 16.5; HRMS (ESI) m/z: [M + H]+ calcd for C15H18BrO2 309.0484; found: 309.0481.

(4S*,9R*)-7-Methoxy-4-methyl-4,9-tetrahydro-3H-xanthene [cis-6a]

According to the general procedure, the reaction of 1,5-enyne (Z)-5a (82.7 mg, 0.35 mmol) with InI3 (8.9 mg, 0.017 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), cis-6a (71.9 mg, 87%) as a white solid; mp 75–77 °C; 1H NMR (300 MHz, CDCl3) δ 6.69–6.67 (m, 2H), 6.58 (d, J = 2.6 Hz, 1H), 5.72–5.68 (m, 1H), 5.35 (dq, J = 9.9, 2.2 Hz, 1H), 3.74 (s, 3H), 3.08 (dd, J = 16.3, 6.1 Hz, 1H), 2.52 (dd, J = 16.3, 3.7 Hz, 1H), 2.43 (dt, J = 5.6, 2.8 Hz, 1H), 2.31 (ddq, J = 15.7, 9.5, 2.9 Hz, 1H), 2.06–1.95 (m, 2H), 1.70 (ddd, J = 13.4, 9.6, 6.6 Hz, 1H), 1.33 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 152.9, 148.1, 128.8, 128.5, 120.0, 117.2, 114.1, 113.5, 73.7, 55.8, 36.8, 33.8, 29.7, 26.0, 22.9; IR (neat) νmax 2921, 2832, 1494, 1236, 1213, 1101, 1041 cm–1; HRMS (EI) m/z: [M]+ calcd for C15H18O2 230.1301; found 230.1293.

(4S*,9S*)-1-Bromo-7-methoxy-4-methyl-4,9-tetrahydro-3H-xanthene [cis-6b]

According to the general procedure, the reaction of 1,5-enyne (Z)-5b (76.0 mg, 0.25 mmol) with InI3 (6.1 mg, 0.012 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), cis-6b (69.9 mg, 92%) as a white solid; 1H NMR (300 MHz, CDCl3) δ 6.70–6.69 (m, 2H), 6.61 (d, J = 2.6 Hz, 1H), 6.05 (td, J = 4.1, 1.6 Hz, 1H), 3.76 (s, 3H), 3.13 (dd, J = 16.6, 6.0 Hz, 1H), 2.94 (dd, J = 16.6, 6.0 Hz, 1H), 2.62 (m, 1H), 2.36–2.30 (m, 1H), 2.11–2.02 (m, 1H), 1.94 (dt, J = 12.4, 6.1 Hz, 1H), 1.74–1.65 (m, 1H), 1.42 (s, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 153.3, 147.6, 129.9, 124.4, 120.2, 117.1, 113.8, 113.7, 75.9, 55.9, 45.5, 31.6, 29.1, 26.4, 25.0; HRMS (EI) m/z: [M]+ calcd for C15H17BrO2 308.0406; found: 308.0395.

5′,7′-Dimethoxyspiro[cyclohexane-1,4′-isochroman]-3-ene (8a) (14c)

According to the general procedure, the reaction of 1,5-enyne 7a (98.3 mg, 0.38 mmol) with InI3 (9.5 mg, 0.019 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), 8a (82.6 mg, 84%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 6.35 (d, J = 2.5 Hz, 1H), 6.12 (d, J = 2.5 Hz, 1H), 5.70 (d, J = 2.6 Hz, 2H), 4.69 (d, J = 0.8 Hz, 2H), 3.92 (d, J = 11.4 Hz, 1H), 3.79 (s, 3H), 3.78 (s, 3H), 3.59 (dd, J = 11.4, 1.3 Hz, 1H), 2.88–2.68 (m, 2H), 2.08–2.01 (m, 3H), 1.46–1.42 (m, 1H); 13C{1H} NMR (75 MHz, CDCl3) δ 159.8, 158.6, 137.5, 126.4, 125.5, 122.5, 99.9, 98.1, 73.9, 70.0, 55.3, 55.1, 34.5, 31.6, 27.5, 22.1; HRMS (ESI) m/z: [M + H]+ calcd for C16H21O3 261.1485; found: 261.1479.

5′,7′-Dimethoxy-2′-tosyl-2′,3′-dihydro-1′H-spiro[cyclohexane-1,4′-isoquinolin]-3-ene (8b) (14c)

According to the general procedure, the reaction of 1,5-enyne 7b (88.5 mg, 0.21 mmol) with InI3 (5.3 mg, 0.011 mmol) afforded, after purification by column chromatography (10% EtOAc/hexanes), 8b (81.4, 92%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 7.72 (d, J = 8.0 Hz, 2H), 7.34 (d, J = 8.0 Hz, 2H), 6.32 (d, J = 2.5 Hz, 1H), 6.15 (d, J = 2.5 Hz, 1H), 5.76–5.65 (m, 2H), 4.16 (d, J = 14.5 Hz, 1H), 4.00 (d, J = 14.5 Hz, 1H), 3.76 (s, 3H), 3.75 (s, 3H), 3.16 (d, J = 11.7 Hz, 1H), 3.00–2.93 (m, 2H), 2.69 (td, J = 13.1, 12.4, 6.8 Hz, 1H), 2.44 (s, 3H), 2.12 (m, 2H), 1.90 (d, J = 18.2 Hz, 1H), 1.45 (dd, J = 13.5, 5.5 Hz, 1H); 13C{1H} NMR (126 MHz, CDCl3) δ 159.8, 158.7, 143.7, 134.4, 133.0, 129.8, 128.0, 126.0, 125.5, 122.9, 102.1, 98.5, 55.4, 55.2, 52.5, 50.1, 37.0, 31.9, 27.8, 22.0, 21.7; HRMS (EI) m/z: [M]+ calcd for C23H27NO4S 413.1655; found 413.1643.

3-Bromo-5′,7′-dimethoxy-2′-tosyl-2′,3′-dihydro-1′H-spiro[cyclohexane-1,4′-isoquinolin]-3-ene (8c) (14c)

According to the general procedure, the reaction of 1,5-enyne 7c (95.0 mg, 0.19 mmol) with InI3 (4.8 mg, 0.010 mmol) afforded, after purification by column chromatography (10% EtOAc/hexanes), 8c (72.2 mg, 76%) as a colorless oil; 1H NMR (300 MHz, CDCl3) δ 7.73 (d, J = 8.1 Hz, 2H), 7.35 (d, J = 8.1 Hz, 2H), 6.35 (d, J = 2.5 Hz, 1H), 6.15 (d, J = 2.5 Hz, 1H), 6.08 (m, 1H), 4.10 (q, J = 14.6 Hz, 2H), 3.78 (s, 3H), 3.74 (s, 3H), 3.51–3.44 (m, 1H), 3.19 (d, J = 12.0 Hz, 1H), 2.95 (d, J = 11.9 Hz, 1H), 2.55 (td, J = 12.3, 6.5 Hz, 1H), 2.44 (s, 3H), 2.26–2.09 (m, 3H), 1.64–1.53 (m, 1H); 13C{1H} NMR (75 MHz, CDCl3) δ 159.4, 159.0, 143.8, 134.4, 132.9, 129.8, 127.8, 127.0, 121.0, 120.8, 102.2, 98.4, 55.3, 55.2, 52.2, 49.8, 41.3, 39.6, 26.3, 24.0, 21.6; HRMS (EI) m/z: [M]+ calcd for C23H26BrNO4S: 491.0760; found 491.0770.

General Procedure for the One-Pot Sequential Indium-Catalyzed Cycloisomerization and Palladium-Catalyzed Cross-Coupling Reactions of (E)-5b and 7c

In a Schlenk tube filled with argon, InI3 (5 mol %) was placed and a solution of 1,5-enyne (E)-5b or 7c (∼0.07 M) in toluene was stirred at room temperature (for 5b) or 60°C in an oil bath (for 7c) until the starting material was consumed (TLC). Then, Pd(PPh3)4 (5 mol %) and a solution of R3In (70 mol %, 0.45 M in dry THF) were added and the mixture was stirred at 80 °C in an oil bath for 10 h. The reaction was quenched by the addition of a few drops of MeOH and the mixture was concentrated in vacuo. H2O (10 mL) was added and the aqueous phase was extracted with EtOAc (3 × 10 mL). The combined organic phase was washed with brine (15 mL), dried with anhydrous MgSO4, filtered, and concentrated in vacuo. The residue was purified by flash chromatography to afford, after concentration and high vacuo drying, the corresponding products trans-6c–e and 8d.

(4S*,9S*)-7-Methoxy-4-methyl-1-phenyl-4,9-tetrahydro-3H-xanthene [trans-6c] (20)

According to the general procedure, the reaction of 1,5-enyne (E)-5b (14c) (96.0 mg, 0.31 mmol) with InI3 (7.1 mg, 0.016 mmol), triphenylindium (0.217 mmol), and Pd(PPh3)4 (17.5 mg, 0.016 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), trans-6c (72.2 mg, 76% in two steps) as a white solid; 1H NMR (300 MHz, CDCl3) δ 7.36–7.28 (m, 3H), 7.26–7.20 (m, 2H), 6.78 (d, J = 8.9 Hz, 1H), 6.69 (dd, J = 8.9, 2.9 Hz, 1H), 6.51 (d, J = 3.0 Hz, 1H), 5.73 (dt, J = 5.1, 2.7 Hz, 1H), 3.70 (s, 3H), 3.08–3.03 (m, 1H), 2.70 (dd, J = 16.7, 5.2 Hz, 1H), 2.45–2.35 (m, 2H), 2.27–2.16 (m, 1H), 2.03–1.99 (m, 2H), 1.19 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 153.1, 147.5, 141.4, 140.0, 128.2, 127.5, 126.9, 126.1, 123.3, 118.0, 114.2, 114.0, 76.0, 55.8, 40.6, 35.3, 27.3, 24.7, 16.6; HRMS (EI) m/z: [M]+ calcd for C21H22O2 306.1614; found: 306.1587.

(4S*,9S*)-7-Methoxy-1,4-dimethyl-4,9-tetrahydro-3H-xanthene [trans-6d] (20)

According to the general procedure, the reaction of 1,5-enyne (E)-5b (96.3 mg, 0.31 mmol) with InI3 (7.1 mg, 0.016 mmol), trimethylindium (0.217 mmol), and Pd(PPh3)4 (17.5 mg, 0.016 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), trans-6d (62.9 mg, 83% in two steps) as a white solid; 1H NMR (300 MHz, CDCl3) δ 6.77–6.65 (m, 3H), 5.39 (m, 1H), 3.76 (s, 3H), 2.87–2.79 (m, 1H), 2.53–2.50 (m, 2H), 2.21–2.18 (m, 2H), 1.92–1.82 (m, 2H), 1.72 (s, 3H), 1.09 (s, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 153.1, 147.7, 133.7, 123.1, 121.7, 117.9, 114.5, 113.7, 75.9, 55.8, 42.1, 35.6, 26.4, 24.2, 20.1, 16.5; HRMS (EI) m/z: [M]+ calcd for C16H20O2 244.1458; found 244.1439.

(4S*,9S*)-1-Butyl-7-methoxy-4-methyl-4,9-tetrahydro-3H-xanthene [trans-6e]

According to the general procedure, the reaction of 1,5-enyne (E)-5b (14c) (95.7 mg, 0.31 mmol) with InI3 (7.1 mg, 0.016 mmol), tributylindium (0.217 mmol), and Pd(PPh3)4 (17.5 mg, 0.016 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), trans-6e (71.9 mg, 81% two steps) as a white solid; mp 48–50 °C; 1H NMR (300 MHz, CDCl3) δ 6.78–6.72 (m, 2H), 6.66 (d, J = 2.7 Hz, 1H), 5.39 (m, 1H), 3.76 (s, 3H), 2.85 (d, J = 11.0 Hz, 1H), 2.54–2.50 (m, 2H), 2.21 (m, 2H), 2.06–1.81 (m, 4H), 1.41–1.26 (m, 4H), 1.09 (s, 3H), 0.92 (t, J = 7.0 Hz, 3H); 13C{1H} NMR (75 MHz, CDCl3) δ 153.0, 147.7, 137.7, 123.1, 120.8, 117.9, 114.4, 113.8, 76.1, 55.8, 40.7, 35.6, 33.5, 30.7, 26.2, 24.2, 22.6, 16.6, 14.2; IR (neat) νmax 2929, 2856, 1493, 1224, 1148, 1081, 1041 cm–1; HRMS (EI) m/z: [M]+ calcd for C19H26O2 286.1927; found: 286.1923.

5′,7′-Dimethoxy-3-phenyl-2′-tosyl-2′,3′-dihydro-1′H-spiro[cyclohexane-1,4′-isoquinolin]-3-ene (8d)

According to the general procedure, the reaction of 1,5-enyne 7c (95.2 mg, 0.19 mmol) with InI3 (4.9 mg, 0.010 mmol), triphenylindium (0.13 mmol), and Pd(PPh3)4 (11.0 mg, 0.010 mmol) afforded, after purification by column chromatography (5% EtOAc/hexanes), 8d (56.7 mg, 61% in two steps) as a white solid; mp 125–127 °C; 1H NMR (300 MHz, CDCl3) δ 7.70–7.67 (m, 2H), 7.41–7.39 (m, 2H), 7.32–7.21 (m, 5H), 6.35 (d, J = 2.5 Hz, 1H), 6.18 (d, J = 2.5 Hz, 1H), 6.13 (q, J = 3.4, 2.4 Hz, 1H), 4.21–4.08 (m, 2H), 3.76 (s, 6H), 3.37 (dd, J = 17.2, 3.3 Hz, 1H), 3.12 (q, J = 11.9 Hz, 2H), 2.73–2.67 (m, 1H), 2.39 (s, 3H), 2.30 (d, J = 17.3 Hz, 3H), 1.26 (m, 1H); 13C{1H} NMR (126 MHz, CDCl3) δ 159.8, 158.9, 143.7, 142.6, 135.7, 134.6, 133.2, 129.8, 128.3, 127.9, 126.8, 125.5, 123.0, 122.6, 102.2, 98.6, 55.4, 55.3, 52.5, 50.0, 37.7, 34.0, 27.2, 22.8, 21.7; IR (neat) νmax 2932, 2841, 1608, 1460, 1340, 1164, 1055, 831 cm–1; HRMS (EI) m/z: [M]+ calcd for C29H31NO4S 489.1968; found 489.1976.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.1c00825.

  • Copies of 1H and 13C{1H} NMR spectra (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Ramón E. Millán - Centro de Investigacións Científicas Avanzadas (CICA) and Departamento de Química, Universidade da Coruña, E-15071 A Coruña, Spain
    • Jaime Rodríguez - Centro de Investigacións Científicas Avanzadas (CICA) and Departamento de Química, Universidade da Coruña, E-15071 A Coruña, SpainOrcidhttps://orcid.org/0000-0001-5348-6970
    • Luis A. Sarandeses - Centro de Investigacións Científicas Avanzadas (CICA) and Departamento de Química, Universidade da Coruña, E-15071 A Coruña, SpainOrcidhttps://orcid.org/0000-0003-1114-7107
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

Dedicated to the memory of Prof. Kilian Muñiz. We thank the Spanish Ministerio de Ciencia, Innovación y Universidades (PGC2018-097792-B-I00 and PID 2019-110008GB-I00), Xunta de Galicia (GRC2018/039), IZO-SGI SGIker of UPV/EHU, and EDRF funds for financial and human support.

References

Click to copy section linkSection link copied!

This article references 28 other publications.

  1. 1
    (a) Ye, F.; Ge, Y.; Spannenberg, A.; Neumann, H.; Beller, M. The Role of Allyl Ammonium Salts in Palladium-Catalyzed Cascade Reactions Towards the Synthesis of Spiro-Fused Heterocycles. Nat. Commun. 2020, 11, 5383  DOI: 10.1038/s41467-020-19110-3

    and references therein.

    (b) Catalytic Cascade Reactions; Xu, P. F.; Wang, W., Eds; Wiley: Hoboken, New Jersey, 2014.
  2. 2

    For some reviews see:

    (a) Barrett, A. G. M.; Ma, T.-K.; Mies, T. Recent Developments in Polyene Cyclizations and Their Applications in Natural Product Synthesis. Synthesis 2019, 51, 6782,  DOI: 10.1055/s-0037-1610382
    (b) Wendt, K. U.; Schulz, G. E.; Corey, E. J.; Liu, D. R. Enzyme Mechanisms for Polycyclic Triterpene Formation. Angew. Chem., Int. Ed. 2000, 39, 28122833,  DOI: 10.1002/1521-3773(20000818)39:16<2812::AID-ANIE2812>3.0.CO;2-#
    (c) Yoder, R. A.; Johnston, J. N. A Case Study in Biomimetic Total Synthesis: Polyolefin Carbocyclizations to Terpenes and Steroids. Chem. Rev. 2005, 105, 47304756,  DOI: 10.1021/cr040623l
  3. 3

    Representative examples:

    (a) Corey, E. J.; Lee, J. Enantioselective Total Synthesis of Oleanolic Acid, Erythrodiol, β-Amyrin, and Other Pentacyclic Triterpenes From a Common Intermediate. J. Am. Chem. Soc. 1993, 115, 88738874,  DOI: 10.1021/ja00072a064
    (b) Corey, E. J.; Lin, S. A Short Enantioselective Total Synthesis of Dammarenediol II. J. Am. Chem. Soc. 1996, 118, 87658766,  DOI: 10.1021/ja9620806
    (c) Huang, A. X.; Xiong, Z.; Corey, E. J. An Exceptionally Short and Simple Enantioselective Total Synthesis of Pentacyclic Triterpenes of the β-Amyrin Family. J. Am. Chem. Soc. 1999, 121, 999910003,  DOI: 10.1021/ja992411p
    (d) Surendra, K.; Corey, E. J. Rapid and Enantioselective Synthetic Approaches to Germanicol and Other Pentacyclic Triterpenes. J. Am. Chem. Soc. 2008, 130, 88658869,  DOI: 10.1021/ja802730a
    (e) Surendra, K.; Corey, E. J. A Short Enantioselective Total Synthesis of the Fundamental Pentacyclic Triterpene Lupeol. J. Am. Chem. Soc. 2009, 131, 1392813929,  DOI: 10.1021/ja906335u
    (f) Bartels, F.; Hong, Y. J.; Ueda, D.; Weber, M.; Sato, T.; Tantillo, D. J.; Christmann, M. Bioinspired Synthesis of Pentacyclic Onocerane Triterpenoids. Chem. Sci. 2017, 8, 82858290,  DOI: 10.1039/C7SC03903D
  4. 4
    (a) Fürstner, A.; Davies, P. W. Catalytic Carbophilic Activation: Catalysis by Platinum and Gold π Acids. Angew. Chem., Int. Ed. 2007, 46, 34103449,  DOI: 10.1002/anie.200604335
    (b) Yamamoto, Y. From σ- to π-Electrophilic Lewis Acids. Application to Selective Organic Transformations. J. Org. Chem. 2007, 72, 78177831,  DOI: 10.1021/jo070579k
    (c) Sethofer, S. G.; Mayer, T.; Toste, F. D. Gold(I)-Catalyzed Enantioselective Polycyclization Reactions. J. Am. Chem. Soc. 2010, 132, 82768277,  DOI: 10.1021/ja103544p
  5. 5

    For recent key feature article, see:

    (a) Ríos, P.; Rodríguez, A.; Conejero, S. Enhancing the Catalytic Properties of Well-Defined Electrophilic Platinum Complexes. Chem. Commun. 2020, 56, 53335349,  DOI: 10.1039/D0CC01438A

    Some representative references:

    (b) Chatani, N.; Furukawa, N.; Sakurai, H.; Murai, S. PtCl2-Catalyzed Conversion of 1,6- and 1,7-Enynes to 1-Vinylcycloalkenes. Anomalous Bond Connection in Skeletal Reorganization of Enynes. Organometallics 1996, 15, 901903,  DOI: 10.1021/om950832j
    (c) Oi, S.; Tsukamoto, I.; Miyano, S.; Inoue, Y. Cationic Platinum-Complex-Catalyzed Skeletal Reorganization of Enynes. Organometallics 2001, 20, 37043709,  DOI: 10.1021/om010316v
    (d) Fürstner, A. From Understanding to Prediction: Gold- and Platinum-Based Π-Acid Catalysis for Target Oriented Synthesis. Acc. Chem. Res. 2014, 47, 925938,  DOI: 10.1021/ar4001789
    (e) Geier, M. J.; Gagné, M. R. Diastereoselective Pt Catalyzed Cycloisomerization of Polyenes to Polycycles. J. Am. Chem. Soc. 2014, 136, 30323035,  DOI: 10.1021/ja500656k
    (f) Toullec, P. Y.; Michelet, V. Chiral Cationic Platinum Complexes: New Catalysts for the Activation of Carbon-Carbon Multiple Bonds Towards Nucleophilic Enantioselective Attack. Curr. Org. Chem. 2010, 14, 12451253,  DOI: 10.2174/138527210791330431
    (g) Mascareñas, J. L.; Varela, I.; López, F. Allenes and Derivatives in Gold(I)- and Platinum(II)-Catalyzed Formal Cycloadditions. Acc. Chem. Res. 2019, 52, 465479,  DOI: 10.1021/acs.accounts.8b00567
  6. 6

    For some leading references, see:

    (a) Zhang, L.; Kozmin, S. A. Gold-Catalyzed Assembly of Heterobicyclic Systems. J. Am. Chem. Soc. 2005, 127, 69626963,  DOI: 10.1021/ja051110e
    (b) Lim, C.; Kang, J.-E.; Lee, J.-E.; Shin, S. Gold-Catalyzed Tandem C–C and C–O Bond Formation: A Highly Diastereoselective Formation of Cyclohex-4-ene-1,2-diol Derivatives. Org. Lett. 2007, 9, 35393542,  DOI: 10.1021/ol071402f
    (c) Fürstner, A. Gold and Platinum Catalysis-a Convenient Tool for Generating Molecular Complexity. Chem. Soc. Rev. 2009, 38, 32083221,  DOI: 10.1039/b816696j
    (d) Dorel, R.; Echavarren, A. M. Gold(I)-Catalyzed Activation of Alkynes for the Construction of Molecular Complexity. Chem. Rev. 2015, 115, 90289072,  DOI: 10.1021/cr500691k
    (e) Li, Y.; Li, W.; Zhang, J. Gold-Catalyzed Enantioselective Annulations. Chem. – Eur. J. 2017, 23, 467512,  DOI: 10.1002/chem.201602822
    (f) Marín-Luna, M.; Nieto Faza, O.; Silva López, C. Gold-Catalyzed Homogeneous (Cyclo)isomerization Reactions. Front. Chem. 2019, 7, 296,  DOI: 10.3389/fchem.2019.00296
    (g) Virumbrales, C.; Suárez-Pantiga, S.; Marín-Luna, M.; Silva López, C.; Sanz, R. Unlocking the 5-exo Pathway with the AuI-Catalyzed Alkoxycyclization of 1,3-Dien-5-ynes. Chem. – Eur. J. 2020, 26, 84438451,  DOI: 10.1002/chem.202001296
    (h) Mies, T.; White, A. J. P.; Parsons, P. J.; Barrett, A. G. M. Biomimetic Syntheses of Analogs of Hongoquercin A and B by Late-Stage Derivatization. J. Org. Chem. 2021, 86, 18021817,  DOI: 10.1021/acs.joc.0c02638
  7. 7
    (a) Chatani, N.; Inoue, H.; Kotsuma, T.; Murai, S. Skeletal Reorganization of Enynes to 1-Vinylcycloalkenes Catalyzed by GaCl3. J. Am. Chem. Soc. 2002, 124, 1029410295,  DOI: 10.1021/ja0274554
    (b) Tang, S.; Monot, J.; El-Hellani, A.; Michelet, B.; Guillot, R.; Bour, C.; Gandon, V. Cationic Gallium(III) Halide Complexes: a New Generation of π-Lewis Acids. Chem. – Eur. J. 2012, 18, 1023910243,  DOI: 10.1002/chem.201201202
    (c) Strom, K. R.; Impastato, A. C.; Moy, K. J.; Landreth, A. J.; Snyder, J. K. Gallium(III)-Promoted Halocyclizations of 1,6-Diynes. Org. Lett. 2015, 17, 21262129,  DOI: 10.1021/acs.orglett.5b00716
  8. 8
    (a) Kita, Y.; Yata, T.; Nishimoto, Y.; Chiba, K.; Yasuda, M. Selective Oxymetalation of Terminal Alkynes Via 6-endo Cyclization: Mechanistic Investigation and Application to the Efficient Synthesis of 4-Substituted Isocoumarins. Chem. Sci. 2018, 9, 60416052,  DOI: 10.1039/C8SC01537F
    (b) Kang, K.; Nishimoto, Y.; Yasuda, M. Regio- and Stereoselective Carboindation of Internal Alkynyl Ethers with Organosilicon or -Stannane Nucleophiles. J. Org. Chem. 2019, 84, 1334513363,  DOI: 10.1021/acs.joc.9b01505
    (c) Tani, T.; Sohma, Y.; Tsuchimoto, T. Zinc/Indium Bimetallic Lewis Acid Relay Catalysis for Dehydrogenative Silylation/Hydrosilylation Reaction of Terminal Alkynes with Bis(hydrosilane)s. Adv. Synth. Catal. 2020, 362, 40984108,  DOI: 10.1002/adsc.202000501
    (d) Vayer, M.; Bour, C.; Gandon, V. Exploring the Versatility of 7-Alkynylcycloheptatriene Scaffolds Under π-Acid Catalysis. Eur. J. Org. Chem. 2020, 2020, 53505357,  DOI: 10.1002/ejoc.202000623
    (e) Tian, J.; Chen, Y.; Vayer, M.; Djurovic, A.; Guillot, R.; Guermazi, R.; Dagorne, S.; Bour, C.; Gandon, V. Exploring the Limits of π-Acid Catalysis Using Strongly Electrophilic Main Group Metal Complexes: the Case of Zinc and Aluminium. Chem. – Eur. J. 2020, 26, 1283112838,  DOI: 10.1002/chem.202001376
  9. 9

    For some revisions see:

    (a) Augé, J.; Lubin-Germain, N.; Uziel, J. Recent Advances in Indium-Promoted Organic Reactions. Synthesis 2007, 2007, 17391764,  DOI: 10.1055/s-2007-983703
    (b) Pathipati, S. R.; van der Werf, A.; Selander, N. Indium(III)-Catalyzed Transformations of Alkynes: Recent Advances in Carbo- and Heterocyclization Reactions. Synthesis 2017, 49, 49314941,  DOI: 10.1055/s-0036-1588555
    (c) Sestelo, J. P.; Sarandeses, L. A.; Martínez, M. M.; Alonso-Marañón, L. Indium(III) as π-Acid Catalyst for the Electrophilic Activation of Carbon–Carbon Unsaturated Systems. Org. Biomol. Chem. 2018, 16, 57335747,  DOI: 10.1039/C8OB01426D
  10. 10
    Araki, S.; Hirashita, T. Comprehensive Organometallic Chemistry III; Crabtree, R. H.; Mingos, D. M. P., Eds.; Elsevier: Oxford, 2007; Vol. 9, pp 649722.
  11. 11
    (a) Cintas, P. Synthetic Organoindium Chemistry: What Makes Indium So Appealing?. Synlett 1995, 1995, 10871096,  DOI: 10.1055/s-1995-5192
    (b) Frost, C. G.; Hartley, J. P. New Applications of Indium Catalysts in Organic Synthesis. Mini-Rev. Org. Chem. 2004, 1, 17,  DOI: 10.2174/1570193043489006
    (c) Fringuelli, F.; Piermatti, O.; Pizzo, F.; Vaccaro, L. Indium Salt-Promoted Organic Reactions. Curr. Org. Chem. 2003, 7, 16611689,  DOI: 10.2174/1385272033486251
    (d) Yadav, J. S.; Antony, A.; George, J.; Subba Reddy, B. V. Recent Developments in Indium Metal and Its Salts in Organic Synthesis. Eur. J. Org. Chem. 2010, 2010, 591605,  DOI: 10.1002/ejoc.200900895
    (f) Singh, M. S.; Raghuvanshi, K. Recent Advances in InCl3-Catalyzed One-Pot Organic Synthesis. Tetrahedron 2012, 68, 86838697,  DOI: 10.1016/j.tet.2012.06.099
  12. 12

    For representative examples, see:

    (a) Mamane, V.; Hannen, P.; Fürstner, A. Synthesis of Phenanthrenes and Polycyclic Heteroarenes by Transition-Metal Catalyzed Cycloisomerization Reactions. Chem. – Eur. J. 2004, 10, 45564575,  DOI: 10.1002/chem.200400220
    (b) Nishimoto, Y.; Moritoh, R.; Yasuda, M.; Baba, A. Regio- and Stereoselective Generation of Alkenylindium Compounds From Indium Tribromide, Alkynes, and Ketene Silyl Acetals. Angew. Chem., Int. Ed. 2009, 48, 45774580,  DOI: 10.1002/anie.200901417
    (c) Antoniotti, S.; Dalla, V.; Duñach, E. Metal Triflimidates: Better Than Metal Triflates as Catalysts in Organic Synthesis-the Effect of a Highly Delocalized Counteranion. Angew. Chem., Int. Ed. 2010, 49, 78607888,  DOI: 10.1002/anie.200906407
    (d) Tsuchimoto, T.; Kanbara, M. Reductive Alkylation of Indoles with Alkynes and Hydrosilanes Under Indium Catalysis. Org. Lett. 2011, 13, 912915,  DOI: 10.1021/ol1029673
    (e) Kumar, A.; Li, Z.; Sharma, S. K.; Parmar, V. S.; Van der Eycken, E. V. Switching the Regioselectivity via Indium(III) and Gold(I) Catalysis: a Post-Ugi Intramolecular Hydroarylation to Azepino- and Azocino-[c,d]indolones. Chem. Commun. 2013, 49, 68036805,  DOI: 10.1039/c3cc42704h
    (f) Michelet, B.; Colard-Itte, J.-R.; Thiery, G.; Guillot, R.; Bour, C.; Gandon, V. Dibromoindium(III) Cations as a π-Lewis Acid: Characterization of [IPr·InBr2][SbF6] and Its Catalytic Activity Towards Alkynes and Alkenes. Chem. Commun. 2015, 51, 74017404,  DOI: 10.1039/C5CC00740B
    (g) Yonekura, K.; Yoshimura, Y.; Akehi, M.; Tsuchimoto, T. A Heteroarylamine Library: Indium-Catalyzed Nucleophilic Aromatic Substitution of Alkoxyheteroarenes with Amines. Adv. Synth. Catal. 2018, 360, 11591181,  DOI: 10.1002/adsc.201701452
    (h) de Orbe, M. E.; Zanini, M.; Quinonero, O.; Echavarren, A. M. Gold- or Indium-Catalyzed Cross-Coupling of Bromoalkynes with Allylsilanes Through a Concealed Rearrangement. ACS Catal. 2019, 9, 78177822,  DOI: 10.1021/acscatal.9b02314
  13. 13

    For some reviews see:

    (a) Echavarren, A. M.; Nevado, C. Non-Stabilized Transition Metal Carbenes as Intermediates in Intramolecular Reactions of Alkynes with Alkenes. Chem. Soc. Rev. 2004, 33, 431436,  DOI: 10.1039/b308768a
    (b) Jiménez-Núñez, E.; Echavarren, A. M. Gold-Catalyzed Cycloisomerizations of Enynes: a Mechanistic Perspective. Chem. Rev. 2008, 108, 33263350,  DOI: 10.1021/cr0684319
    (c) Toullec, P. Y.; Michelet, V. Cycloisomerization of 1,n-Enynes via Carbophilic Activation. Top. Curr. Chem. 2011, 302, 3180
    (d) Aubert, C.; Fensterbank, L.; Garcia, P.; Malacria, M.; Simonneau, A. Transition Metal Catalyzed Cycloisomerizations of 1,n-Allenynes and -Allenenes. Chem. Rev. 2011, 111, 19541993,  DOI: 10.1021/cr100376w
  14. 14
    (a) Imagawa, H.; Iyenaga, T.; Nishizawa, M. Mercuric Triflate-Catalyzed Tandem Cyclization Leading to Polycarbocycles. Org. Lett. 2005, 7, 451453,  DOI: 10.1021/ol047472t
    (b) Pradal, A.; Chen, Q.; Faudot dit Bel, P.; Toullec, P. Y.; Michelet, V. Gold-Catalyzed Cycloisomerization of Functionalized 1,5-Enynes – an Entry to Polycyclic Framework. Synlett 2012, 2012, 7479,  DOI: 10.1055/s-0031-1289867
    (c) Rong, Z.; Echavarren, A. M. Broad Scope Gold(I)-Catalysed Polyenyne Cyclisations for the Formation of Up to Four Carbon-Carbon Bonds. Org. Biomol. Chem. 2017, 15, 21632167,  DOI: 10.1039/C7OB00235A
    (d) Lu, X.-L.; Lyu, M.-Y.; Peng, X.-S.; Wong, H. N. C. Gold(I)-Catalyzed Tandem Cycloisomerization of 1,5-Enyne Ethers by Hydride Transfer. Angew. Chem., Int. Ed. 2018, 57, 1136511368,  DOI: 10.1002/anie.201806842
  15. 15
    Miyanohana, Y.; Chatani, N. Skeletal Reorganization of Enynes Catalyzed by InCl3. Org. Lett. 2006, 8, 21552158,  DOI: 10.1021/ol060606d
  16. 16
    (a) Surendra, K.; Qiu, W.; Corey, E. J. A Powerful New Construction of Complex Chiral Polycycles by an Indium(III)-Catalyzed Cationic Cascade. J. Am. Chem. Soc. 2011, 133, 97249726,  DOI: 10.1021/ja204142n
    (b) Surendra, K.; Corey, E. J. Diiodoindium(III) Cation, InI2+, a Potent Yneophile. Generation and Application to Cationic Cyclization by Selective π-Activation of C≡C. J. Am. Chem. Soc. 2014, 136, 1091810920,  DOI: 10.1021/ja506502p
  17. 17

    For some representative references, see:

    (a) Pérez, I.; Pérez Sestelo, J.; Sarandeses, L. A. Atom-Efficient Metal-Catalyzed Cross-Coupling Reaction of Indium Organometallics with Organic Electrophiles. J. Am. Chem. Soc. 2001, 123, 41554160,  DOI: 10.1021/ja004195m
    (b) Caeiro, J.; Pérez Sestelo, J.; Sarandeses, L. A. Enantioselective Nickel-Catalyzed Cross-Coupling Reactions of Trialkynylindium Reagents with Racemic Secondary Benzyl Bromides. Chem. – Eur. J. 2008, 14, 741746,  DOI: 10.1002/chem.200701035
    (c) Mato, M.; Pérez-Caaveiro, C.; Sarandeses, L. A.; Pérez Sestelo, J. Ferrocenylindium Reagents in Palladium-Catalyzed Cross-Coupling Reactions: Asymmetric Synthesis of Planar Chiral 2-Aryl Oxazolyl and Sulfinyl Ferrocenes. Adv. Synth. Catal. 2017, 359, 13881393,  DOI: 10.1002/adsc.201601397
    (d) Gil-Negrete, J. M.; Pérez Sestelo, J.; Sarandeses, L. A. Synthesis of Bench-Stable Solid Triorganoindium Reagents and Reactivity in Palladium-Catalyzed Cross-Coupling Reactions. Chem. Commun. 2018, 54, 14531456,  DOI: 10.1039/C7CC09344F
    (e) Gil-Negrete, J. M.; Pérez Sestelo, J.; Sarandeses, L. A. Transition-Metal-Free Oxidative Cross-Coupling of Triorganoindium Reagents with Tetrahydroisoquinolines. J. Org. Chem. 2019, 84, 97789785,  DOI: 10.1021/acs.joc.9b00928
  18. 18
    (a) Alonso-Marañón, L.; Martínez, M. M.; Sarandeses, L. A.; Pérez Sestelo, J. Indium-Catalyzed Intramolecular Hydroarylation of Aryl Propargyl Ethers. Org. Biomol. Chem. 2015, 13, 379387,  DOI: 10.1039/C4OB02033B
    (b) Alonso-Marañón, L.; Sarandeses, L. A.; Martínez, M. M.; Pérez Sestelo, J. Sequential In-Catalyzed Intramolecular Hydroarylation and Pd-Catalyzed Cross-Coupling Reactions Using Bromopropargyl Aryl Ethers and Amines. Org. Chem. Front. 2017, 4, 500505,  DOI: 10.1039/C6QO00721J
    (c) Alonso-Marañón, L.; Sarandeses, L. A.; Martínez, M. M.; Pérez Sestelo, J. Synthesis of Fused Chromenes by the Indium(III)-Catalyzed Cascade Hydroarylation/Cycloisomerization Reactions of Polyyne-Type Aryl Propargyl Ethers. Org. Chem. Front. 2018, 5, 23082312,  DOI: 10.1039/C8QO00457A
    (d) Alonso-Marañón, L.; Martínez, M. M.; Sarandeses, L. A.; Gómez-Bengoa, E.; Pérez Sestelo, J. Indium(III)-Catalyzed Synthesis of Benzo[b]Furans by Intramolecular Hydroalkoxylation of ortho-Alkynylphenols: Scope and Mechanistic Insights. J. Org. Chem. 2018, 83, 79707980,  DOI: 10.1021/acs.joc.8b00829
  19. 19

    Using 1H-1H COSY, edited-HSQC and HMBC experiments, we were able to assign all protons and carbons for (trans)-4f. Key signals to deduce the trans-fused bicyclic system were protons H-6 (δH 2.21 ppm, td, J = 11.2, 11.2, 2.2 Hz) and H-5 (δH 2.02 ppm, bt, J = 11.5 Hz), which clearly show an antiperiplanar relationship between them (see the SI on page S40–S42).

  20. 20
    Toullec, P. Y.; Blarre, T.; Michelet, V. Mimicking Polyolefin Carbocyclization Reactions: Gold-Catalyzed Intramolecular Phenoxycyclization of 1,5-Enynes. Org. Lett. 2009, 11, 28882891,  DOI: 10.1021/ol900864n
  21. 21
    Capon, R. J. Studies in Natural Products Chemistry; Rahman, A.-u., Ed.; Elsevier: New York, 1995; Vol. 15, pp 289326.
  22. 22
    (a) Shen, Z.-L.; Wang, S.-Y.; Chok, Y.-K.; Xu, Y.-H.; Loh, T.-P. Organoindium Reagents: the Preparation and Application in Organic Synthesis. Chem. Rev. 2013, 113, 271401,  DOI: 10.1021/cr300051y
    (b) Zhao, K.; Shen, L.; Shen, Z.-L.; Loh, T.-P. Transition Metal-Catalyzed Cross-Coupling Reactions Using Organoindium Reagents. Chem. Soc. Rev. 2017, 46, 586602,  DOI: 10.1039/C6CS00465B
  23. 23
    Sotorríos, L.; Demertzidou, V. P.; Zografos, A. L.; Gómez-Bengoa, E. DFT Studies on Metal-Catalyzed Cycloisomerization of trans-1,5-Enynes to Cyclopropane Sesquiterpenoids. Org. Biomol. Chem. 2019, 17, 51125120,  DOI: 10.1039/C9OB00890J
  24. 24

    All structures were optimized using Gaussian 16 with the B3LYP/6-31G(d,p) method for C, H, and O and SDD basis set for In and I. Final energies were refined at the M06/def2tzvpp level of theory in toluene. For more details, see the Supporting Information.

  25. 25
    Surendra, K.; Rajendar, G.; Corey, E. J. Useful Catalytic Enantioselective Cationic Double Annulation Reactions Initiated at an Internal π-Bond: Method and Applications. J. Am. Chem. Soc. 2014, 136, 642645,  DOI: 10.1021/ja4125093
  26. 26
    Corey, E. J.; Seibel, W. L. First stekeospecific synthesis of E-γ-bisabolene. A method for the concurrent generation of a ring and a tetrasubstituted exocyclic double bond. Tetrahedron Lett. 1986, 27, 905908,  DOI: 10.1016/S0040-4039(00)84133-7
  27. 27
    Nicolaou, K. C.; Reingruber, R.; Sarlah, D.; Bräse, S. Enantioselective Intramolecular Friedel–Crafts-Type α-Arylation of Aldehydes. J. Am. Chem. Soc. 2009, 131, 20862087,  DOI: 10.1021/ja809405c
  28. 28
    Mostafa, M. A. B.; Grafton, M. W.; Wilson, C.; Sutherland, A. A one-pot, three-step process for the diastereoselective synthesis of aminobicyclo[4.3.0]nonanes using consecutive palladium(II)- and ruthenium(II)-catalysis. Org. Biomol. Chem. 2016, 14, 32843297,  DOI: 10.1039/C6OB00165C

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 18 publications.

  1. Eleonora Tufano, Euijae Lee, Matteo Barilli, Emanuele Casali, Andraž Oštrek, Hoimin Jung, Marta Morana, Jihye Kang, Dongwook Kim, Sukbok Chang, Giuseppe Zanoni. Iridium Acylnitrenoid-Initiated Biomimetic Cascade Cyclizations: Stereodefined Access to Polycyclic δ-Lactams. Journal of the American Chemical Society 2023, 145 (45) , 24724-24735. https://doi.org/10.1021/jacs.3c08331
  2. Galder Llorente, Maria Teresa Herrero, Garazi Urgoitia, Raul SanMartin. Regio‐ and Diastereoselective Cascade Synthesis of 5‐Alkylidene‐γ‐Lactams via Copper‐Catalyzed Hydro‐Oxycarbonylation. Advanced Synthesis & Catalysis 2025, 367 (8) https://doi.org/10.1002/adsc.202401535
  3. Mohanad A. Hussein, Ahmed Hassoon Mageed. Recent Advances in Direct Amidation Via Organocatalysis. Asian Journal of Organic Chemistry 2025, 14 (2) https://doi.org/10.1002/ajoc.202400528
  4. Fadhil Faez Sead, Vicky Jain, Anjan Kumar, Subbulakshmi Ganesan, Aman Shankhyan, Girish Chandra Sharma, Pushpa Negi Bhakuni, Mosstafa Kazemi, Ramin Javahershenas. One-Pot, for the Green Preparation of Benzimidazoles and 2,3-Dihydroquinazolin-4(1H)-ones Catalyzed by In2O3 NPs/Glycerol. Journal of Inorganic and Organometallic Polymers and Materials 2025, 25 https://doi.org/10.1007/s10904-025-03604-y
  5. Fadhil Faez Sead, Vicky Jain, Suhas Ballal, Abhayveer Singh, Anita Devi, Girish Chandra Sharma, Kamal Kant Joshi, Mosstafa Kazemi, Ramin Javahershenas. Research on transition metals for the multicomponent synthesis of benzo-fused γ-lactams. RSC Advances 2025, 15 (4) , 2334-2346. https://doi.org/10.1039/D4RA08798D
  6. Maryna M. Kornet, Thomas J. J. Müller. Recent Advances in Sequentially Pd-Catalyzed One-Pot Syntheses of Heterocycles. Molecules 2024, 29 (22) , 5265. https://doi.org/10.3390/molecules29225265
  7. Mohanad A. Hussein, Mohammed Hadi Ali Al‐Jumaili, Ali A. Sabi, Heider A. Abdulhussein. Indium‐Catalyzed Direct Amidation Reaction of Carboxylic Acids and In Silico Study for Screening the Activity of Potential Therapeutics of the Synthesized Products. ChemistrySelect 2024, 9 (37) https://doi.org/10.1002/slct.202403097
  8. Alyssa Daniels, Christoph Wölper, Gebhard Haberhauer. Indium(III)‐Catalyzed Haloalkynylation Reaction of Alkynes. Chemistry – A European Journal 2024, 30 (37) https://doi.org/10.1002/chem.202401070
  9. Timothy O. Ajiboye, Oluwaseun J. Ajala, Jerry O. Adeyemi, Subhendu Dhibar. Indium(III) complexes: application as organic catalyst, precursor for chalcogenides nanoparticles and starting materials in the industry. Chemical Papers 2024, 78 (8) , 4605-4622. https://doi.org/10.1007/s11696-024-03411-8
  10. Sana Jamshaid. Indium Triflate's Transformative Role in Modern Organic Synthesis: A Five‐Year Survey of Functionalization, Cyclization, and Multicomponent Reactions. ChemistrySelect 2024, 9 (18) https://doi.org/10.1002/slct.202400313
  11. Raquel Pérez‐Guevara, Luis A. Sarandeses, Rosana Álvarez, M. Montserrat Martínez, José Pérez Sestelo. Stereospecific Indium(III)‐Catalysed Tandem Cycloisomerization of Functionalized 1,6‐enynes: Scope and Mechanistic Insights. Advanced Synthesis & Catalysis 2024, 366 (4) , 852-861. https://doi.org/10.1002/adsc.202301329
  12. Aurélien Alix. Diastereoselective Transformation Using Group 2 and 13 Metal Salts. 2024, 357-431. https://doi.org/10.1016/B978-0-32-390644-9.00134-7
  13. Fabio Seoane-Carabel, Lorena Alonso-Marañón, Luis A. Sarandeses, José Pérez Sestelo. Synthesis of 1H-Isochromenes and 1,2-Dihydroisoquinolines by Indium(III)-Catalyzed Cycloisomerization of ortho-(Alkynyl)benzyl Derivatives. Synthesis 2023, 55 (11) , 1714-1723. https://doi.org/10.1055/s-0042-1751383
  14. María Herrero, Jokin Díaz de Sarralde, Nerea Conde, Aitor Herrán, Garazi Urgoitia, Raul SanMartin. Metal-Catalyzed Cascade Reactions between Alkynoic Acids and Dinucleophiles: A Review. Catalysts 2023, 13 (3) , 495. https://doi.org/10.3390/catal13030495
  15. Clementina M.M. Santos, Artur M.S. Silva. Six-membered ring systems: with O and/or S atoms. 2023, 487-557. https://doi.org/10.1016/B978-0-443-18939-5.00015-9
  16. Raquel Pérez-Guevara, Luis A. Sarandeses, M. Montserrat Martínez, José Pérez Sestelo. Indium-catalyzed synthesis of benzannulated spiroketals by intramolecular double hydroalkoxylation of ortho -(hydroxyalkynyl)benzyl alcohols. Organic Chemistry Frontiers 2022, 9 (24) , 6894-6901. https://doi.org/10.1039/D2QO01600A
  17. Somnath Ganguly, Sayantika Bhakta, Tapas Ghosh. Gold‐Catalyzed Synthesis of Spirocycles: Recent Advances. ChemistrySelect 2022, 7 (28) https://doi.org/10.1002/slct.202201407
  18. Dawrin Pech-Puch, Abel M. Forero, Juan Carlos Fuentes-Monteverde, Cristina Lasarte-Monterrubio, Marta Martinez-Guitian, Carlos González-Salas, Sergio Guillén-Hernández, Harold Villegas-Hernández, Alejandro Beceiro, Christian Griesinger, Jaime Rodríguez, Carlos Jiménez. Antimicrobial Diterpene Alkaloids from an Agelas citrina Sponge Collected in the Yucatán Peninsula. Marine Drugs 2022, 20 (5) , 298. https://doi.org/10.3390/md20050298

The Journal of Organic Chemistry

Cite this: J. Org. Chem. 2021, 86, 14, 9515–9529
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.joc.1c00825
Published June 25, 2021

Copyright © 2021 American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

2484

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 1. Indium(III)-Catalyzed Electrophilic Activation of Alkynes

    Scheme 2

    Scheme 2. Sequential One-Pot In-Catalyzed 1,5-Enyne Cycloisomerization and Pd-Catalyzed Cross-Coupling Reaction with (E)-5b

    Scheme 3

    Scheme 3. General Plausible Mechanism for the In(III)-Catalyzed Cascade Cycloisomerization Reaction of 1,5-Enynes 1a–d and 3a–f

    Scheme 4

    Scheme 4. DFT-Calculated Mechanism of the Reaction of (Z)-1a and (E)-1a for the Selective Formation of INT2-cis and -trans

    Scheme 5

    Scheme 5. DFT-Calculated Isomerization Process between the Intermediates INT1-cis and INT1-trans

    Scheme 6

    Scheme 6. DFT-Calculated Mechanism of the Reaction of (Z)-1d and (E)-1d under Curtin–Hammett Conditions
  • References


    This article references 28 other publications.

    1. 1
      (a) Ye, F.; Ge, Y.; Spannenberg, A.; Neumann, H.; Beller, M. The Role of Allyl Ammonium Salts in Palladium-Catalyzed Cascade Reactions Towards the Synthesis of Spiro-Fused Heterocycles. Nat. Commun. 2020, 11, 5383  DOI: 10.1038/s41467-020-19110-3

      and references therein.

      (b) Catalytic Cascade Reactions; Xu, P. F.; Wang, W., Eds; Wiley: Hoboken, New Jersey, 2014.
    2. 2

      For some reviews see:

      (a) Barrett, A. G. M.; Ma, T.-K.; Mies, T. Recent Developments in Polyene Cyclizations and Their Applications in Natural Product Synthesis. Synthesis 2019, 51, 6782,  DOI: 10.1055/s-0037-1610382
      (b) Wendt, K. U.; Schulz, G. E.; Corey, E. J.; Liu, D. R. Enzyme Mechanisms for Polycyclic Triterpene Formation. Angew. Chem., Int. Ed. 2000, 39, 28122833,  DOI: 10.1002/1521-3773(20000818)39:16<2812::AID-ANIE2812>3.0.CO;2-#
      (c) Yoder, R. A.; Johnston, J. N. A Case Study in Biomimetic Total Synthesis: Polyolefin Carbocyclizations to Terpenes and Steroids. Chem. Rev. 2005, 105, 47304756,  DOI: 10.1021/cr040623l
    3. 3

      Representative examples:

      (a) Corey, E. J.; Lee, J. Enantioselective Total Synthesis of Oleanolic Acid, Erythrodiol, β-Amyrin, and Other Pentacyclic Triterpenes From a Common Intermediate. J. Am. Chem. Soc. 1993, 115, 88738874,  DOI: 10.1021/ja00072a064
      (b) Corey, E. J.; Lin, S. A Short Enantioselective Total Synthesis of Dammarenediol II. J. Am. Chem. Soc. 1996, 118, 87658766,  DOI: 10.1021/ja9620806
      (c) Huang, A. X.; Xiong, Z.; Corey, E. J. An Exceptionally Short and Simple Enantioselective Total Synthesis of Pentacyclic Triterpenes of the β-Amyrin Family. J. Am. Chem. Soc. 1999, 121, 999910003,  DOI: 10.1021/ja992411p
      (d) Surendra, K.; Corey, E. J. Rapid and Enantioselective Synthetic Approaches to Germanicol and Other Pentacyclic Triterpenes. J. Am. Chem. Soc. 2008, 130, 88658869,  DOI: 10.1021/ja802730a
      (e) Surendra, K.; Corey, E. J. A Short Enantioselective Total Synthesis of the Fundamental Pentacyclic Triterpene Lupeol. J. Am. Chem. Soc. 2009, 131, 1392813929,  DOI: 10.1021/ja906335u
      (f) Bartels, F.; Hong, Y. J.; Ueda, D.; Weber, M.; Sato, T.; Tantillo, D. J.; Christmann, M. Bioinspired Synthesis of Pentacyclic Onocerane Triterpenoids. Chem. Sci. 2017, 8, 82858290,  DOI: 10.1039/C7SC03903D
    4. 4
      (a) Fürstner, A.; Davies, P. W. Catalytic Carbophilic Activation: Catalysis by Platinum and Gold π Acids. Angew. Chem., Int. Ed. 2007, 46, 34103449,  DOI: 10.1002/anie.200604335
      (b) Yamamoto, Y. From σ- to π-Electrophilic Lewis Acids. Application to Selective Organic Transformations. J. Org. Chem. 2007, 72, 78177831,  DOI: 10.1021/jo070579k
      (c) Sethofer, S. G.; Mayer, T.; Toste, F. D. Gold(I)-Catalyzed Enantioselective Polycyclization Reactions. J. Am. Chem. Soc. 2010, 132, 82768277,  DOI: 10.1021/ja103544p
    5. 5

      For recent key feature article, see:

      (a) Ríos, P.; Rodríguez, A.; Conejero, S. Enhancing the Catalytic Properties of Well-Defined Electrophilic Platinum Complexes. Chem. Commun. 2020, 56, 53335349,  DOI: 10.1039/D0CC01438A

      Some representative references:

      (b) Chatani, N.; Furukawa, N.; Sakurai, H.; Murai, S. PtCl2-Catalyzed Conversion of 1,6- and 1,7-Enynes to 1-Vinylcycloalkenes. Anomalous Bond Connection in Skeletal Reorganization of Enynes. Organometallics 1996, 15, 901903,  DOI: 10.1021/om950832j
      (c) Oi, S.; Tsukamoto, I.; Miyano, S.; Inoue, Y. Cationic Platinum-Complex-Catalyzed Skeletal Reorganization of Enynes. Organometallics 2001, 20, 37043709,  DOI: 10.1021/om010316v
      (d) Fürstner, A. From Understanding to Prediction: Gold- and Platinum-Based Π-Acid Catalysis for Target Oriented Synthesis. Acc. Chem. Res. 2014, 47, 925938,  DOI: 10.1021/ar4001789
      (e) Geier, M. J.; Gagné, M. R. Diastereoselective Pt Catalyzed Cycloisomerization of Polyenes to Polycycles. J. Am. Chem. Soc. 2014, 136, 30323035,  DOI: 10.1021/ja500656k
      (f) Toullec, P. Y.; Michelet, V. Chiral Cationic Platinum Complexes: New Catalysts for the Activation of Carbon-Carbon Multiple Bonds Towards Nucleophilic Enantioselective Attack. Curr. Org. Chem. 2010, 14, 12451253,  DOI: 10.2174/138527210791330431
      (g) Mascareñas, J. L.; Varela, I.; López, F. Allenes and Derivatives in Gold(I)- and Platinum(II)-Catalyzed Formal Cycloadditions. Acc. Chem. Res. 2019, 52, 465479,  DOI: 10.1021/acs.accounts.8b00567
    6. 6

      For some leading references, see:

      (a) Zhang, L.; Kozmin, S. A. Gold-Catalyzed Assembly of Heterobicyclic Systems. J. Am. Chem. Soc. 2005, 127, 69626963,  DOI: 10.1021/ja051110e
      (b) Lim, C.; Kang, J.-E.; Lee, J.-E.; Shin, S. Gold-Catalyzed Tandem C–C and C–O Bond Formation: A Highly Diastereoselective Formation of Cyclohex-4-ene-1,2-diol Derivatives. Org. Lett. 2007, 9, 35393542,  DOI: 10.1021/ol071402f
      (c) Fürstner, A. Gold and Platinum Catalysis-a Convenient Tool for Generating Molecular Complexity. Chem. Soc. Rev. 2009, 38, 32083221,  DOI: 10.1039/b816696j
      (d) Dorel, R.; Echavarren, A. M. Gold(I)-Catalyzed Activation of Alkynes for the Construction of Molecular Complexity. Chem. Rev. 2015, 115, 90289072,  DOI: 10.1021/cr500691k
      (e) Li, Y.; Li, W.; Zhang, J. Gold-Catalyzed Enantioselective Annulations. Chem. – Eur. J. 2017, 23, 467512,  DOI: 10.1002/chem.201602822
      (f) Marín-Luna, M.; Nieto Faza, O.; Silva López, C. Gold-Catalyzed Homogeneous (Cyclo)isomerization Reactions. Front. Chem. 2019, 7, 296,  DOI: 10.3389/fchem.2019.00296
      (g) Virumbrales, C.; Suárez-Pantiga, S.; Marín-Luna, M.; Silva López, C.; Sanz, R. Unlocking the 5-exo Pathway with the AuI-Catalyzed Alkoxycyclization of 1,3-Dien-5-ynes. Chem. – Eur. J. 2020, 26, 84438451,  DOI: 10.1002/chem.202001296
      (h) Mies, T.; White, A. J. P.; Parsons, P. J.; Barrett, A. G. M. Biomimetic Syntheses of Analogs of Hongoquercin A and B by Late-Stage Derivatization. J. Org. Chem. 2021, 86, 18021817,  DOI: 10.1021/acs.joc.0c02638
    7. 7
      (a) Chatani, N.; Inoue, H.; Kotsuma, T.; Murai, S. Skeletal Reorganization of Enynes to 1-Vinylcycloalkenes Catalyzed by GaCl3. J. Am. Chem. Soc. 2002, 124, 1029410295,  DOI: 10.1021/ja0274554
      (b) Tang, S.; Monot, J.; El-Hellani, A.; Michelet, B.; Guillot, R.; Bour, C.; Gandon, V. Cationic Gallium(III) Halide Complexes: a New Generation of π-Lewis Acids. Chem. – Eur. J. 2012, 18, 1023910243,  DOI: 10.1002/chem.201201202
      (c) Strom, K. R.; Impastato, A. C.; Moy, K. J.; Landreth, A. J.; Snyder, J. K. Gallium(III)-Promoted Halocyclizations of 1,6-Diynes. Org. Lett. 2015, 17, 21262129,  DOI: 10.1021/acs.orglett.5b00716
    8. 8
      (a) Kita, Y.; Yata, T.; Nishimoto, Y.; Chiba, K.; Yasuda, M. Selective Oxymetalation of Terminal Alkynes Via 6-endo Cyclization: Mechanistic Investigation and Application to the Efficient Synthesis of 4-Substituted Isocoumarins. Chem. Sci. 2018, 9, 60416052,  DOI: 10.1039/C8SC01537F
      (b) Kang, K.; Nishimoto, Y.; Yasuda, M. Regio- and Stereoselective Carboindation of Internal Alkynyl Ethers with Organosilicon or -Stannane Nucleophiles. J. Org. Chem. 2019, 84, 1334513363,  DOI: 10.1021/acs.joc.9b01505
      (c) Tani, T.; Sohma, Y.; Tsuchimoto, T. Zinc/Indium Bimetallic Lewis Acid Relay Catalysis for Dehydrogenative Silylation/Hydrosilylation Reaction of Terminal Alkynes with Bis(hydrosilane)s. Adv. Synth. Catal. 2020, 362, 40984108,  DOI: 10.1002/adsc.202000501
      (d) Vayer, M.; Bour, C.; Gandon, V. Exploring the Versatility of 7-Alkynylcycloheptatriene Scaffolds Under π-Acid Catalysis. Eur. J. Org. Chem. 2020, 2020, 53505357,  DOI: 10.1002/ejoc.202000623
      (e) Tian, J.; Chen, Y.; Vayer, M.; Djurovic, A.; Guillot, R.; Guermazi, R.; Dagorne, S.; Bour, C.; Gandon, V. Exploring the Limits of π-Acid Catalysis Using Strongly Electrophilic Main Group Metal Complexes: the Case of Zinc and Aluminium. Chem. – Eur. J. 2020, 26, 1283112838,  DOI: 10.1002/chem.202001376
    9. 9

      For some revisions see:

      (a) Augé, J.; Lubin-Germain, N.; Uziel, J. Recent Advances in Indium-Promoted Organic Reactions. Synthesis 2007, 2007, 17391764,  DOI: 10.1055/s-2007-983703
      (b) Pathipati, S. R.; van der Werf, A.; Selander, N. Indium(III)-Catalyzed Transformations of Alkynes: Recent Advances in Carbo- and Heterocyclization Reactions. Synthesis 2017, 49, 49314941,  DOI: 10.1055/s-0036-1588555
      (c) Sestelo, J. P.; Sarandeses, L. A.; Martínez, M. M.; Alonso-Marañón, L. Indium(III) as π-Acid Catalyst for the Electrophilic Activation of Carbon–Carbon Unsaturated Systems. Org. Biomol. Chem. 2018, 16, 57335747,  DOI: 10.1039/C8OB01426D
    10. 10
      Araki, S.; Hirashita, T. Comprehensive Organometallic Chemistry III; Crabtree, R. H.; Mingos, D. M. P., Eds.; Elsevier: Oxford, 2007; Vol. 9, pp 649722.
    11. 11
      (a) Cintas, P. Synthetic Organoindium Chemistry: What Makes Indium So Appealing?. Synlett 1995, 1995, 10871096,  DOI: 10.1055/s-1995-5192
      (b) Frost, C. G.; Hartley, J. P. New Applications of Indium Catalysts in Organic Synthesis. Mini-Rev. Org. Chem. 2004, 1, 17,  DOI: 10.2174/1570193043489006
      (c) Fringuelli, F.; Piermatti, O.; Pizzo, F.; Vaccaro, L. Indium Salt-Promoted Organic Reactions. Curr. Org. Chem. 2003, 7, 16611689,  DOI: 10.2174/1385272033486251
      (d) Yadav, J. S.; Antony, A.; George, J.; Subba Reddy, B. V. Recent Developments in Indium Metal and Its Salts in Organic Synthesis. Eur. J. Org. Chem. 2010, 2010, 591605,  DOI: 10.1002/ejoc.200900895
      (f) Singh, M. S.; Raghuvanshi, K. Recent Advances in InCl3-Catalyzed One-Pot Organic Synthesis. Tetrahedron 2012, 68, 86838697,  DOI: 10.1016/j.tet.2012.06.099
    12. 12

      For representative examples, see:

      (a) Mamane, V.; Hannen, P.; Fürstner, A. Synthesis of Phenanthrenes and Polycyclic Heteroarenes by Transition-Metal Catalyzed Cycloisomerization Reactions. Chem. – Eur. J. 2004, 10, 45564575,  DOI: 10.1002/chem.200400220
      (b) Nishimoto, Y.; Moritoh, R.; Yasuda, M.; Baba, A. Regio- and Stereoselective Generation of Alkenylindium Compounds From Indium Tribromide, Alkynes, and Ketene Silyl Acetals. Angew. Chem., Int. Ed. 2009, 48, 45774580,  DOI: 10.1002/anie.200901417
      (c) Antoniotti, S.; Dalla, V.; Duñach, E. Metal Triflimidates: Better Than Metal Triflates as Catalysts in Organic Synthesis-the Effect of a Highly Delocalized Counteranion. Angew. Chem., Int. Ed. 2010, 49, 78607888,  DOI: 10.1002/anie.200906407
      (d) Tsuchimoto, T.; Kanbara, M. Reductive Alkylation of Indoles with Alkynes and Hydrosilanes Under Indium Catalysis. Org. Lett. 2011, 13, 912915,  DOI: 10.1021/ol1029673
      (e) Kumar, A.; Li, Z.; Sharma, S. K.; Parmar, V. S.; Van der Eycken, E. V. Switching the Regioselectivity via Indium(III) and Gold(I) Catalysis: a Post-Ugi Intramolecular Hydroarylation to Azepino- and Azocino-[c,d]indolones. Chem. Commun. 2013, 49, 68036805,  DOI: 10.1039/c3cc42704h
      (f) Michelet, B.; Colard-Itte, J.-R.; Thiery, G.; Guillot, R.; Bour, C.; Gandon, V. Dibromoindium(III) Cations as a π-Lewis Acid: Characterization of [IPr·InBr2][SbF6] and Its Catalytic Activity Towards Alkynes and Alkenes. Chem. Commun. 2015, 51, 74017404,  DOI: 10.1039/C5CC00740B
      (g) Yonekura, K.; Yoshimura, Y.; Akehi, M.; Tsuchimoto, T. A Heteroarylamine Library: Indium-Catalyzed Nucleophilic Aromatic Substitution of Alkoxyheteroarenes with Amines. Adv. Synth. Catal. 2018, 360, 11591181,  DOI: 10.1002/adsc.201701452
      (h) de Orbe, M. E.; Zanini, M.; Quinonero, O.; Echavarren, A. M. Gold- or Indium-Catalyzed Cross-Coupling of Bromoalkynes with Allylsilanes Through a Concealed Rearrangement. ACS Catal. 2019, 9, 78177822,  DOI: 10.1021/acscatal.9b02314
    13. 13

      For some reviews see:

      (a) Echavarren, A. M.; Nevado, C. Non-Stabilized Transition Metal Carbenes as Intermediates in Intramolecular Reactions of Alkynes with Alkenes. Chem. Soc. Rev. 2004, 33, 431436,  DOI: 10.1039/b308768a
      (b) Jiménez-Núñez, E.; Echavarren, A. M. Gold-Catalyzed Cycloisomerizations of Enynes: a Mechanistic Perspective. Chem. Rev. 2008, 108, 33263350,  DOI: 10.1021/cr0684319
      (c) Toullec, P. Y.; Michelet, V. Cycloisomerization of 1,n-Enynes via Carbophilic Activation. Top. Curr. Chem. 2011, 302, 3180
      (d) Aubert, C.; Fensterbank, L.; Garcia, P.; Malacria, M.; Simonneau, A. Transition Metal Catalyzed Cycloisomerizations of 1,n-Allenynes and -Allenenes. Chem. Rev. 2011, 111, 19541993,  DOI: 10.1021/cr100376w
    14. 14
      (a) Imagawa, H.; Iyenaga, T.; Nishizawa, M. Mercuric Triflate-Catalyzed Tandem Cyclization Leading to Polycarbocycles. Org. Lett. 2005, 7, 451453,  DOI: 10.1021/ol047472t
      (b) Pradal, A.; Chen, Q.; Faudot dit Bel, P.; Toullec, P. Y.; Michelet, V. Gold-Catalyzed Cycloisomerization of Functionalized 1,5-Enynes – an Entry to Polycyclic Framework. Synlett 2012, 2012, 7479,  DOI: 10.1055/s-0031-1289867
      (c) Rong, Z.; Echavarren, A. M. Broad Scope Gold(I)-Catalysed Polyenyne Cyclisations for the Formation of Up to Four Carbon-Carbon Bonds. Org. Biomol. Chem. 2017, 15, 21632167,  DOI: 10.1039/C7OB00235A
      (d) Lu, X.-L.; Lyu, M.-Y.; Peng, X.-S.; Wong, H. N. C. Gold(I)-Catalyzed Tandem Cycloisomerization of 1,5-Enyne Ethers by Hydride Transfer. Angew. Chem., Int. Ed. 2018, 57, 1136511368,  DOI: 10.1002/anie.201806842
    15. 15
      Miyanohana, Y.; Chatani, N. Skeletal Reorganization of Enynes Catalyzed by InCl3. Org. Lett. 2006, 8, 21552158,  DOI: 10.1021/ol060606d
    16. 16
      (a) Surendra, K.; Qiu, W.; Corey, E. J. A Powerful New Construction of Complex Chiral Polycycles by an Indium(III)-Catalyzed Cationic Cascade. J. Am. Chem. Soc. 2011, 133, 97249726,  DOI: 10.1021/ja204142n
      (b) Surendra, K.; Corey, E. J. Diiodoindium(III) Cation, InI2+, a Potent Yneophile. Generation and Application to Cationic Cyclization by Selective π-Activation of C≡C. J. Am. Chem. Soc. 2014, 136, 1091810920,  DOI: 10.1021/ja506502p
    17. 17

      For some representative references, see:

      (a) Pérez, I.; Pérez Sestelo, J.; Sarandeses, L. A. Atom-Efficient Metal-Catalyzed Cross-Coupling Reaction of Indium Organometallics with Organic Electrophiles. J. Am. Chem. Soc. 2001, 123, 41554160,  DOI: 10.1021/ja004195m
      (b) Caeiro, J.; Pérez Sestelo, J.; Sarandeses, L. A. Enantioselective Nickel-Catalyzed Cross-Coupling Reactions of Trialkynylindium Reagents with Racemic Secondary Benzyl Bromides. Chem. – Eur. J. 2008, 14, 741746,  DOI: 10.1002/chem.200701035
      (c) Mato, M.; Pérez-Caaveiro, C.; Sarandeses, L. A.; Pérez Sestelo, J. Ferrocenylindium Reagents in Palladium-Catalyzed Cross-Coupling Reactions: Asymmetric Synthesis of Planar Chiral 2-Aryl Oxazolyl and Sulfinyl Ferrocenes. Adv. Synth. Catal. 2017, 359, 13881393,  DOI: 10.1002/adsc.201601397
      (d) Gil-Negrete, J. M.; Pérez Sestelo, J.; Sarandeses, L. A. Synthesis of Bench-Stable Solid Triorganoindium Reagents and Reactivity in Palladium-Catalyzed Cross-Coupling Reactions. Chem. Commun. 2018, 54, 14531456,  DOI: 10.1039/C7CC09344F
      (e) Gil-Negrete, J. M.; Pérez Sestelo, J.; Sarandeses, L. A. Transition-Metal-Free Oxidative Cross-Coupling of Triorganoindium Reagents with Tetrahydroisoquinolines. J. Org. Chem. 2019, 84, 97789785,  DOI: 10.1021/acs.joc.9b00928
    18. 18
      (a) Alonso-Marañón, L.; Martínez, M. M.; Sarandeses, L. A.; Pérez Sestelo, J. Indium-Catalyzed Intramolecular Hydroarylation of Aryl Propargyl Ethers. Org. Biomol. Chem. 2015, 13, 379387,  DOI: 10.1039/C4OB02033B
      (b) Alonso-Marañón, L.; Sarandeses, L. A.; Martínez, M. M.; Pérez Sestelo, J. Sequential In-Catalyzed Intramolecular Hydroarylation and Pd-Catalyzed Cross-Coupling Reactions Using Bromopropargyl Aryl Ethers and Amines. Org. Chem. Front. 2017, 4, 500505,  DOI: 10.1039/C6QO00721J
      (c) Alonso-Marañón, L.; Sarandeses, L. A.; Martínez, M. M.; Pérez Sestelo, J. Synthesis of Fused Chromenes by the Indium(III)-Catalyzed Cascade Hydroarylation/Cycloisomerization Reactions of Polyyne-Type Aryl Propargyl Ethers. Org. Chem. Front. 2018, 5, 23082312,  DOI: 10.1039/C8QO00457A
      (d) Alonso-Marañón, L.; Martínez, M. M.; Sarandeses, L. A.; Gómez-Bengoa, E.; Pérez Sestelo, J. Indium(III)-Catalyzed Synthesis of Benzo[b]Furans by Intramolecular Hydroalkoxylation of ortho-Alkynylphenols: Scope and Mechanistic Insights. J. Org. Chem. 2018, 83, 79707980,  DOI: 10.1021/acs.joc.8b00829
    19. 19

      Using 1H-1H COSY, edited-HSQC and HMBC experiments, we were able to assign all protons and carbons for (trans)-4f. Key signals to deduce the trans-fused bicyclic system were protons H-6 (δH 2.21 ppm, td, J = 11.2, 11.2, 2.2 Hz) and H-5 (δH 2.02 ppm, bt, J = 11.5 Hz), which clearly show an antiperiplanar relationship between them (see the SI on page S40–S42).

    20. 20
      Toullec, P. Y.; Blarre, T.; Michelet, V. Mimicking Polyolefin Carbocyclization Reactions: Gold-Catalyzed Intramolecular Phenoxycyclization of 1,5-Enynes. Org. Lett. 2009, 11, 28882891,  DOI: 10.1021/ol900864n
    21. 21
      Capon, R. J. Studies in Natural Products Chemistry; Rahman, A.-u., Ed.; Elsevier: New York, 1995; Vol. 15, pp 289326.
    22. 22
      (a) Shen, Z.-L.; Wang, S.-Y.; Chok, Y.-K.; Xu, Y.-H.; Loh, T.-P. Organoindium Reagents: the Preparation and Application in Organic Synthesis. Chem. Rev. 2013, 113, 271401,  DOI: 10.1021/cr300051y
      (b) Zhao, K.; Shen, L.; Shen, Z.-L.; Loh, T.-P. Transition Metal-Catalyzed Cross-Coupling Reactions Using Organoindium Reagents. Chem. Soc. Rev. 2017, 46, 586602,  DOI: 10.1039/C6CS00465B
    23. 23
      Sotorríos, L.; Demertzidou, V. P.; Zografos, A. L.; Gómez-Bengoa, E. DFT Studies on Metal-Catalyzed Cycloisomerization of trans-1,5-Enynes to Cyclopropane Sesquiterpenoids. Org. Biomol. Chem. 2019, 17, 51125120,  DOI: 10.1039/C9OB00890J
    24. 24

      All structures were optimized using Gaussian 16 with the B3LYP/6-31G(d,p) method for C, H, and O and SDD basis set for In and I. Final energies were refined at the M06/def2tzvpp level of theory in toluene. For more details, see the Supporting Information.

    25. 25
      Surendra, K.; Rajendar, G.; Corey, E. J. Useful Catalytic Enantioselective Cationic Double Annulation Reactions Initiated at an Internal π-Bond: Method and Applications. J. Am. Chem. Soc. 2014, 136, 642645,  DOI: 10.1021/ja4125093
    26. 26
      Corey, E. J.; Seibel, W. L. First stekeospecific synthesis of E-γ-bisabolene. A method for the concurrent generation of a ring and a tetrasubstituted exocyclic double bond. Tetrahedron Lett. 1986, 27, 905908,  DOI: 10.1016/S0040-4039(00)84133-7
    27. 27
      Nicolaou, K. C.; Reingruber, R.; Sarlah, D.; Bräse, S. Enantioselective Intramolecular Friedel–Crafts-Type α-Arylation of Aldehydes. J. Am. Chem. Soc. 2009, 131, 20862087,  DOI: 10.1021/ja809405c
    28. 28
      Mostafa, M. A. B.; Grafton, M. W.; Wilson, C.; Sutherland, A. A one-pot, three-step process for the diastereoselective synthesis of aminobicyclo[4.3.0]nonanes using consecutive palladium(II)- and ruthenium(II)-catalysis. Org. Biomol. Chem. 2016, 14, 32843297,  DOI: 10.1039/C6OB00165C
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.1c00825.

    • Copies of 1H and 13C{1H} NMR spectra (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.