Free Pdfmcgrath

J Am Chem Soc. Author manuscript; available in PMC 2016 Jun 9.

Published in final edited form as:

PMCID: PMC4899974

NIHMSID: NIHMS791526

Nitrone Cycloadditions of 1,2-Cyclohexadiene

Abstract

We report the first 1,3-dipolar cycloadditions of 1,2-cyclohexadiene, a rarely exploited strained allene. 1,2-Cyclohexadiene is generated in situ under mild conditions and trapped with nitrones to give isoxazolidine products in synthetically useful yields. The reactions occur regioselectively and exhibit a notable endo preference, thus resulting in the controlled formation of two new bonds and two stereogenic centers. DFT calculations of stepwise and concerted reaction pathways are used to rationalize the observed selectivities. Moreover, the strategic manipulation of nitrone cycloadducts demonstrates the utility of this methodology for the assembly of compounds bearing multiple heterocyclic units. These studies showcase the exploitation of a traditionally avoided reactive intermediate in chemical synthesis.

The synthesis of heterocycles and natural product scaffolds remains a vital area of study.1,2 One approach to construct such important compounds involves the use of strained intermediates such as benzyne, hetarynes, and cyclic alkynes (16, Figure 1).3 Although initially viewed as undesirably reactive species several decades ago, these intermediates can now be used to prepare medicinally privileged scaffolds and stereochemically complex natural products.3,4 Likewise, our understanding of the structure and reactivity of arynes and cyclic alkynes has evolved considerably.3,4

An external file that holds a picture, illustration, etc.  Object name is nihms791526f1.jpg

Strained intermediates 16, known cycloadditions of 1,2-cyclohexadiene (7), and nitrone cycloadditions of 7 described in the present study.

A less well-studied class of highly strained intermediates is cyclic allenes, such as 1,2-cyclohexadiene (7, Figure 1). 1,2-Cyclohexadiene (7) was first reported by Wittig in 19665 but has seen little synthetic use since, especially compared with its aryne and alkyne counterparts. Seminal efforts include theoretical studies6 and the two C–C bond forming reactions: [2 + 2]6a,c,7 and [4 + 2]6a,d,7c–e,8 cycloadditions to form products such as 8 and 9, respectively. These transformations hint at the potential utility of 7 as a versatile synthetic building block, but no other cycloadditions involving 7 have been reported.

We now report the first dipolar cycloadditions of 1,2-cyclohexadiene, which allow for a facile entryway to isoxazolidine products (Figure 1, 7 + 1011). The resulting products contain significant sp3-character9 and are obtained with high regio- and diastereoselectivities. Computational studies suggest that stepwise and concerted reaction pathways are operative in the cycloadditions and predict the observed selectivity trends.

We initiated our study toward the trapping of 1,2-cyclohexadiene (7) by first accessing a suitable precursor that could allow for allene generation in situ. Encouraged by Guitián's synthesis of a trimethylsilyltriflate precursor to 7,7e which was used in Diels–Alder cycloadditions, we prepared the new triethylsilyltriflate species 12,10 which proved more readily accessible and could be obtained in gram quantities.11 We initially surveyed the reaction of silyltriflate 12 with 5 equiv of commercially available nitrone 13 using different fluoride sources at ambient temperature (Table 1). Although the use of TBAF gave low yields of cycloadduct 14 (entry 1), KF/18-crown-6 gave more promising results (entry 2). Additionally, the use of CsF cleanly gave product 14 in modest yield, although the remainder of the mass was unreacted silyl triflate 12 (entry 3). In all cases, it should be noted that 14 was formed regioselectively and in approximately 9:1 dr, with the major product being the endo isomer (14a). For practical reasons, we elected to further pursue the use of CsF and ultimately found that the reaction proceeded smoothly at elevated temperatures and higher concentration (entries 4 and 5) to give cycloadduct 14 in 4 h (entry 5; compared with 3 d for entry 3). Finally, to probe the necessary stoichiometry, we tested the cycloaddition using just 1 equiv of nitrone 13. We were delighted to find that cycloadduct 14 could be formed in comparable yields under these conditions (entry 6).

Table 1

Optimization of Nitrone Cycloaddition

An external file that holds a picture, illustration, etc.  Object name is nihms791526u1.jpg


entry conditions temp (°C) concn (M) equiv of 13 yield, % (14a/14b dr)a
1 TBAF, THF 23 0.025 5.0 35 (10.8:1)
2 KF, 18-crown-6, CH3CN 23 0.025 5.0 77 (12.1:1)
3 CsF, CH3CN 23 0.025 5.0 41 (13.5:1)
4 CsF, CH3CN 80 0.025 5.0 90 (9.5:1)
5 CsF, CH3CN 80 0.1 5.0 92 (9.3:1)
6 CsF, CH3CN 80 0.1 1.0 84 (8.5:1)

Having identified suitable reaction conditions, we assessed the scope of the methodology by varying the nitrone component and evaluating isolated yields and diastereoselectivities (Figure 2). All reactions were performed using CsF at 80 °C and either 1 or 2 equiv of the nitrone trapping agent, as indicated.12 In addition to the parent experiment, which gave 14 in 88% yield and roughly 9:1 dr, we were delighted to find that nitrones derived from acetaldehyde and pivaldehyde could be employed to give products 15 and 16, respectively. Nitrones derived from aldehydes bearing heterocycles were also tested. As shown by the formation of products 1721, furan, thiophene, thiazole, quinoline,13 and indole heterocycles are tolerated in this methodology. Finally, several nitrones prepared from ketone or cyclic amine precursors were tested, resulting in isoxazolidines 2224.14

An external file that holds a picture, illustration, etc.  Object name is nihms791526f2.jpg

Scope of reaction methodology. Major diastereomer is shown. Conditions unless otherwise stated are silyl triflate (1.0 equiv), nitrone (1.0 equiv), CsF (5.0 equiv), and CH3CN (0.1 M) at 80 °C. The yields reflect the average of two isolation experiments. a Nitrone (2.0 equiv) was used. b Nitrone (2.0 equiv) and THF (0.1 M) at 60 °C were used.

DFT calculations were performed to assess mechanistic aspects of the 1,2-cyclohexadiene/nitrone cycloaddition.15 We investigated both the regio- and diastereoselectivity of the cycloaddition. Although Diels–Alder and (2 + 2) cycloadditions of 1,2-cyclohexadiene (7) have been studied computationally, no reports of (3 + 2) cycloadditions of 7 are available. We assessed pathways that could lead to either of two possible constitutional isomers, in addition to all possible stereoisomers, as summarized in Figure 3. Cycloaddition via pathway A would give diastereomers 25 and 26, whereas pathway B would furnish 14a and 14b. Previous reports with linear allenes indicate the plausibility of both pathways. However, consistent with our experimental results, computations show that pathway A is disfavored for the reaction of 7 and nitrone 13.16 The lower energy pathway involves attack of the central allene carbon (C2) on the nitrone, with attack of the nitrone oxygen on C1 (pathway B).

An external file that holds a picture, illustration, etc.  Object name is nihms791526f3.jpg

DFT calculations (B3LYP/6-31G*) were used to examine regioselectivites and diastereoselectivities.

To explain the observed diastereoselectivities, we compared the possible transition states for the formation of both endo and exo products 14a and 14b (Figure 4). First examining a concerted reaction mechanism, we found that the activation energy for the endo reaction is 14.5 kcal/mol (TS1), which presents a plausible pathway for the formation of 14a. A concerted exo transition state could not be explicitly located; instead diradical transition state TS2 always resulted, suggesting the concerted exo reaction is higher in energy. As we do observe exo product 14b, we proceeded to examine the stepwise mechanism through TS2. We found that diradical 27 can be formed via initial formation of the C–C bond; an intrinsic reaction coordinate scan and a nonzero S 2 value confirms the diradical nature of this process. From diradical 27, cyclization furnishes 14a or 14b (via TS3a or TS3b); the formation of 14a is favored by 1.2 kcal/mol, which correlates to the observed 10:1 ratio of products. Therefore, we propose that both stepwise and concerted mechanisms for cycloaddition are in competition, a phenomenon we have previously observed in our studies on Diels–Alder reactions with allenes,17,18 which are known to be very reactive and form relatively stable allylic radical intermediates.

An external file that holds a picture, illustration, etc.  Object name is nihms791526f4.jpg

DFT calculations (B3LYP/6-31G*) of concerted and stepwise mechanistic pathways.

Regardless of the reaction mechanism, the 1,3-dipolar cycloaddition of 1,2-cyclohexadiene occurs quickly with mild heating. The low ~15 kcal/mol barriers for these reactions, compared with the >30 kcal/mol barriers for cycloaddition with linear allenes,17,18 can be attributed to the predistortion of 7 into geometries near those of the transition states for cycloadditions. The optimized structures of 1,2-cyclohexadiene (7) and TS1 (concerted pathway) are shown in Figure 5. Of note, the C1–C2–C3 angle is nearly identical in both cases (132.8° vs 129.8°).19 Thus, only 3° contraction is necessary to reach the transition state geometry. In contrast, linear allenes must bend by more than 30° in order to reach their optimal transition state geometries (see SI for details). Examples of distortion-accelerated reactions are known20 and explain the facility of these reactions with 1,2-cyclohexadiene.

An external file that holds a picture, illustration, etc.  Object name is nihms791526f5.jpg

Comparison of C2 internal angle for 7 and TS1.

The nitrone cycloaddition of 1,2-cyclohexadiene (7), when used sequentially with other transformations, provides new and unique strategies for preparing compounds bearing multiple heterocyclic units (Figure 6). For example, isoxazolidine 28 was synthesized via a nitrone cycloaddition from 12. Next, palladium-catalyzed Suzuki–Miyaura cross-coupling21 with boronic acid 29 afforded product 30 in 65% yield. Of note, 30 contains three distinct heterocycles (i.e., a furan, pyridine, and isoxazolidine). In another example, we questioned if silyltriflate 12 could be used as a building block for the assembly of polycyclic fused heterocycles, such as 33. Nitrone cycloaddition with 31 furnished isoxazolidine 32. Subsequent N–O bond cleavage,22 followed by SNAr cyclization23 gave tricycle 33. The conversion of 12 to 33 allows for the facile construction of three new bonds and two stereogenic centers. Moreover, the two synthetic applications shown in Figure 6 showcase how the unusual intermediate cyclohexadiene (7), despite its high reactivity, can be used to access varying types of polycyclic heterocycles in a controlled way.

An external file that holds a picture, illustration, etc.  Object name is nihms791526f6.jpg

Strategic manipulation of nitrone cycloadducts.

In summary, we have discovered the first 1,3-dipolar cycloadditions of 1,2-cyclohexadiene. The transformation involves in situ generation of the strained intermediate under mild conditions, and trapping with nitrones to give isoxazolidine products in synthetically useful yields. The reactions occur regioselectively, with a notable endo preference, thus resulting in the controlled formation of two new bonds and two stereogenic centers. DFT calculations suggest that stepwise and concerted reaction pathways are operative in the cycloadditions and predict the observed selectivity trends. Moreover, the strategic manipulation of nitrone cycloadducts demonstrates the utility of this methodology for the assembly of compounds bearing multiple heterocyclic units. These studies are expected to prompt the further exploitation of traditionally avoided reactive intermediates in chemical synthesis.

Supplementary Material

supporting info

Acknowledgments

The authors are grateful to the NIH-NIGMS (R01 GM090007 for N.K.G.), the NSF (CHE-1361104 for K.N.H.), Bristol–Myers Squibb, the A. P. Sloan Foundation, the Dreyfus Foundation, the University of California, Los Angeles, the Foote Family (H.V.P. and E.D.S.), and the Chemistry–Biology Interface training program (J.S.B., USPHS National Research Service Award 5T32GM008496-20) for financial support. These studies were supported by shared instrumentation grants from the NSF (CHE-1048804) and the NIH NCRR (S10RR025631). Computations were performed with resources made available from the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by the National Science Foundation (OCI-1053575), as well as the UCLA Institute of Digital Research and Education (IDRE).

Footnotes

Notes

The authors declare no competing financial interest.

Supporting Information

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.5b13304.

Detailed experimental and computational procedures, compound characterization, Cartesian coordinates, electronic energies, entropies, enthalpies, Gibbs free energies, and lowest frequencies of the calculated structures (PDF)

References

2. Heterocycles are also prevalent motifs in agrochemicals and materials chemistry; see: Dinges J, Lamberth C, editors. Bioactive Heterocyclic Compounds Classes: Agrochemicals. Wiley-VCH: Weinheim; 2012. [Google Scholar] Gilchrist TL. J Chem Soc, Perkin Trans 1. 1999:2849–2866. [Google Scholar]

3. For reviews of arynes and hetarynes, see: Pellissier H, Santelli M. Tetrahedron. 2003;59:701–730. [Google Scholar] Wenk HH, Winkler M, Sander W. Angew Chem, Int Ed. 2003;42:502–528. [PubMed] [Google Scholar] Sanz R. Org Prep Proced Int. 2008;40:215–291. [Google Scholar] Bronner SM, Goetz AE, Garg NK. Synlett. 2011:2599–2604. [Google Scholar] Tadross PM, Stoltz BM. Chem Rev. 2012;112:3550–3557. [PubMed] [Google Scholar] Gampe CM, Carreira EM. Angew Chem, Int Ed. 2012;51:3766–3778. [PubMed] [Google Scholar] Bhunia A, Yetra SR, Biju AT. Chem Soc Rev. 2012;41:3140–3152. [PubMed] [Google Scholar] Yoshida H, Takaki K. Synlett. 2012;23:1725–1732. [Google Scholar] Dubrovskiy AV, Markina NA, Larock RC. Org Biomol Chem. 2013;11:191–218. [PubMed] [Google Scholar] Wu C, Shi F. Asian J Org Chem. 2013;2:116–125. [Google Scholar] Hoffmann RW, Suzuki K. Angew Chem, Int Ed. 2013;52:2655–2656. [PubMed] [Google Scholar] Goetz AE, Garg NK. J Org Chem. 2014;79:846–851. [PubMed] [Google Scholar] Goetz AE, Shah TK, Garg NK. Chem Commun. 2015;51:34–45. [PubMed] [Google Scholar]

4. For select recent studies of cyclic alkynes, see: Tlais SF, Danheiser RL. J Am Chem Soc. 2014;136:15489–15492. [PubMed] [Google Scholar] Medina JM, McMahon TC, Jiménez-Osés G, Houk KN, Garg NK. J Am Chem Soc. 2014;136:14706–14709. [PMC free article] [PubMed] [Google Scholar] McMahon TC, Medina JM, Yang YF, Simmons BJ, Houk KN, Garg NK. J Am Chem Soc. 2015;137:4082–4085. [PMC free article] [PubMed] [Google Scholar]

5. Wittig G, Fritze P. Angew Chem, Int Ed Engl. 1966;5:846. [Google Scholar]

6. (a) Bottini AT, Hilton LL, Plott J. Tetrahedron. 1975;31:1997–2001. [Google Scholar] (b) Schmidt MW, Angus RO, Johnson RP. J Am Chem Soc. 1982;104:6838–6839. [Google Scholar] (c) Wentrup C, Gross G, Maquestiau A, Flammang R. Angew Chem, Int Ed Engl. 1983;22:542–543. [Google Scholar] (d) Nendel M, Tolbert LM, Herring LE, Islam MN, Houk KN. J Org Chem. 1999;64:976–983. [PubMed] [Google Scholar] (e) Hänninen MM, Peuronen A, Tuononen HM. Chem - Eur J. 2009;15:7287–7291. [PubMed] [Google Scholar]

7. (a) Moore WR, Moser WR. J Org Chem. 1970;35:908–912. [Google Scholar] (b) Bottini AT, Corson FP, Fitzgerald R, Frost KA., II Tetrahedron. 1972;28:4883–4904. [Google Scholar] (c) Christl M, Schreck M. Angew Chem, Int Ed Engl. 1987;26:449–451. [Google Scholar] (d) Christl M, Fischer H, Arnone M, Engels B. Chem - Eur J. 2009;15:11266–11272. [PubMed] [Google Scholar] (e) Quintana I, Peña D, Pérez D, Guitián E. Eur J Org Chem. 2009;5519–5524 [Google Scholar]

8. (a) Balci M, Jones WM. J Am Chem Soc. 1980;102:7607–7608. [Google Scholar] (b) Tolbert LM, Islam MN, Johnson RP, Loiselle PM, Shakespeare WC. J Am Chem Soc. 1990;112:6416–6417. [Google Scholar]

9. Another attractive aspect of using 1,2-cyclohexadiene (7) as a synthetic intermediate is the potential to access sp3-rich heterocycles, especially in comparison to benzyne cycloadducts. The synthesis of sp3-rich heterocycles is a current priority area in medicinal chemistry; see Lovering F, Bikker J, Humblet C. J Med Chem. 2009;52:6752–6756. [PubMed] [Google Scholar] Ritchie TJ, Macdonald SJF. Drug Discovery Today. 2009;14:1011–1020. [PubMed] [Google Scholar] Lovering F. MedChemComm. 2013;4:515–519. [Google Scholar]

10. Our efforts to prepare the known trimethylsilyl counterpart of 12 were thwarted by complications associated with the synthesis and purification of 2-(trimethylsilyl)cyclohexanone.

11. See Supporting Information for the synthesis of silyltriflate 12.

12. In some cases, improved yields were obtained using 2 equiv of the nitrone trapping agent. Optimal conditions used for each substrate are indicated.

13. The modest dr observed in the formation of 20 is currently not well understood.

14. We also explored the use of nitrones bearing chiral auxiliaries. Use of Vasella's mannose-derived nitrones led to modest dr, whereas the use of phenethylamine derivatives gave the corresponding cycloadducts in 1.5:1 dr. For Vasella's mannose derivative, see: Vasella A. Helv Chim Acta. 1977;60:1273–1295. [Google Scholar]

15. All geometries were optimized using the density functional B3LYP with a 6-31G(d) basis set. Free energies were then determined using B3LYP single point calculations with the D3 correction (with no Becke–Johnson damping) to account for dispersion, in conjunction with the larger 6-311+G(d,p) basis set. The conductor-like polarizable continuum model (CPCM) for acetonitrile was used to simulate implicit solvent. Minima and transition states were located and verified with 0 and 1 imaginary frequencies, respectively.

16. Stepwise and concerted reaction mechanisms were evaluated; see the Supporting Information for details.

19. The C1–C2–C3 angle in the diradical transition state (TS2) is very similar as well (i.e., 129.5°).

20. For examples of distortion-accelerated reactions, see: Gordon CG, Mackey JL, Jewett JC, Sletten EM, Houk KN, Bertozzi CR. J Am Chem Soc. 2012;134:9199–9208. [PMC free article] [PubMed] [Google Scholar] Lopez SA, Houk KN. J Org Chem. 2013;78:1778–1783. [PubMed] [Google Scholar] Hong X, Liang Y, Griffith AK, Lambert TH, Houk KN. Chem Sci. 2014;5:471–475. [Google Scholar]

21. Billingsley K, Buchwald SL. J Am Chem Soc. 2007;129:3358–3366. [PubMed] [Google Scholar]

22. Swift PA, Stagnito ML, Mullen GB, Palmer GC, Ceorgiev VS. Eur J Med Chem. 1988;23:465–471. [Google Scholar]

23. Saitton S, Kihlberg J, Luthman K. Tetrahedron. 2004;60:6113–6120. [Google Scholar]

cavazosyeand1973.blogspot.com

Source: https://europepmc.org/articles/pmc4899974/bin/nihms791526-supplement-supporting_info.pdf

0 Response to "Free Pdfmcgrath"

Post a Comment

Iklan Atas Artikel

Iklan Tengah Artikel 1

Iklan Tengah Artikel 2

Iklan Bawah Artikel