Nitrone Cycloadditions of 1,2-Cyclohexadiene - Journal of the

Feb 8, 2016 - ... Chemical Society. *[email protected], *[email protected] .... Published online 8 February 2016. Published in print 2 March 201...
0 downloads 0 Views 4MB Size
Subscriber access provided by UNIV OSNABRUECK

Communication

Nitrone Cycloadditions of 1,2-Cyclohexadiene Joyann S. Barber, Evan D. Styduhar, Hung V. Pham, Travis C. McMahon, K. N. Houk, and Neil K. Garg J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.5b13304 • Publication Date (Web): 08 Feb 2016 Downloaded from http://pubs.acs.org on February 8, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Nitrone Cycloadditions of 1,2-Cyclohexadiene Joyann S. Barber, Evan D. Styduhar, Hung V. Pham, Travis C. McMahon, K. N. Houk,* and Neil K. Garg* Department of Chemistry and Biochemistry, University of California, Los Angeles, California 90095, United States E-Mail: [email protected], [email protected]

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 (1–6, 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.3a,3d–i,3m Likewise, our understanding of the structure and reactivity of arynes and cyclic alkynes has evolved considerably.3l,3m,4 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 1966,5 but has seen little synthetic use since, especially compared to its aryne and alkyne counterparts. Seminal efforts include theoretical studies,6 and the two C–C bond forming reactions: [2+2]6a,6c,7 and [4+2]6a,6d,7c,7d,7e,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. Similarly, 7 has not previously been used to prepare heterocycles. We now report the first dipolar cycloadditions of 1,2-cyclohexadiene, which allow for a facile entryway to

isoxazolidine products (Figure 1, 7 + 10 à 11). 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. Figure 1. Strained intermediates, known cycloadditions of 1,2cyclohexadiene (7), and nitrone cycloadditions of 7 described in the present study. Synthetically Useful Strained Intermediates R N N

N

R

1

2

3

4

R

N

5

6

In situ generated building blocks for the construction of heterocycles and natural product scaffolds

Various Cycloadditions of 1,2-Cyclohexadiene O

O

Ph H

8

Ph

[2+2] cycloadditions

[4+2] cycloadditions

7 R

Nitrone [3+2] cycloadditions

O

(Present Study) R First 1,3-dipolar cycloaddition of 1,2cyclohexadiene

R' N

10 R''

R'

*

N R"

*

9

O

H

Unconventional route to stereochemicallyrich heterocycles

11 Regio- and stereochemical observations explained by insight from computational studies

We initiated our study toward the trapping of 1,2cyclohexadiene (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 equivalents 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

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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 to 3 d for entry 3). Finally, to probe the necessary stoichiometry, we tested the cycloaddition using just 1 equivalent of nitrone 13. We were delighted to find that cycloadduct 14 could be formed in comparable yields under these conditions (entry 6).

Figure 2. Scope of reaction methodology.a R OTf SiEt3

SiEt3

12 entry

O

H N

N

CH3CN, 80 °C

R''

11 Me

t-Bu

O

N

O

t-Bu

O H

14

15

16

88% yield (8.9 : 1 dr)

63% yieldb (11.4 : 1 dr)

76% yield (>20 : 1 dr)

S

S

N

t-Bu

H

N

t-Bu

O H

19

18

17

N

t-Bu

O H

88% yield (8.4 : 1 dr)

77% yieldb (11.9 : 1 dr)

67% yieldb (12.7 : 1 dr)

O

t-Bu

H

O

t-Bu

t-Bu

N

H

R'

O H

O

N

N R"

10

N

R'

CsF

Ph

conditions

t-Bu

R'

O

12

N

R Ph +

R +

Table 1. Optimization of nitrone cycloaddition. OTf

Page 2 of 5

H

13

7

14a: R = Ph, R' = H 14b: R = H, R' = Ph yield equiv of 13 (14a:14b dr) a

Ts N N

conditions

temp (°C)

conc. (M)

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)

20

4

CsF, CH3CN

80

0.025

5.0

90% (9.5:1)

92% yield (2.5 : 1 dr)

5

CsF, CH3CN

80

0.1

5.0

92% (9.3:1)

6

CsF, CH3CN

80

0.1

1.0

84% (8.5:1)

N

t-Bu

N

O

Ph

Yields and diastereomeric ratios were determined by 1H NMR analysis using 1,3,5-trimethoxybenzene as an external standard.

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 equivalents 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 17–21, 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 22–24.14 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 DielsAlder 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

H

21 83% yield (12.2 : 1 dr) Ph

Me

H

CF3 N Me

N Me

a

t-Bu

O

H

N

O

O

O

H

H

H

22

23

24

62% yieldc (4.9 : 1 dr)

80% yieldb (6.2 : 1 dr)

71% yieldb (9.3 : 1 dr)

a

Major diastereomer is shown. Conditions unless otherwise stated: 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. b Nitrone (2.0 equiv) was used. c Nitrone (2.0 equiv) and THF (0.1 M) at 60 °C were used.

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). Figure 3. DFT calculations (B3LYP/6-31G*) were used to examine regioselectivites and diastereoselectivities. ΔG‡ = 15.1– 18.2 kcal/mol

Ph 2 1

Path A 7 2

2

O N

O

H N

ΔG‡ = 14.4– 14.5 kcal/mol

t-Bu

Ph t-Bu

N

1

H

H

Ph

25 endo

Ph

26 exo Disfavored

Path B

13

O N

t-Bu

+

Ph t-Bu

O

N

1

t-Bu

O

H

H

14a endo

14b exo Favored

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

ACS Paragon Plus Environment

Page 3 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

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 S2 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 phenomena 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. Figure 4. DFT calculations (B3LYP/6-31G*) of concerted and stepwise mechanistic pathways. Concerted Pathway

Stepwise Pathway t-Bu

t-Bu

H Ph

H

N

t-Bu N δ+ δ– O

Ph

14.5

13

O

Ph

N δ+ O

14.4

H

t-Bu

7 TS1

H

TS2

Ph

0.0

–7.4

t-Bu N

Ph

Ph

O

δ–

–16.5

t-Bu H

O H

H

TS3a or TS3b

δ+

27

t-Bu N

O

–6.2

H

H

N

O

Ph

H

N

Ph

2 1

3 Strain induced predistorted internal angle

132.8°

7

t-Bu N

Ph

2

H

3

129.8o' 129.8°

1

O

H

2.70 Å 2.28 Å

t-Bu N

Ph

O

14a

TS1

nitrone cycloaddition with 12. Next, palladiumcatalyzed Suzuki 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. Figure 6. Strategic manipulation of nitrone cycloadducts.

t-Bu H

N

Figure 5. Comparison of C2 internal angle for 7 and TS1.

Sequential Cycloaddition and Cross-Coupling

O H

N

14a

14a

14b

endo product (major)

endo product (major)

exo product (minor)

Regardless of the reaction trajectory, the 1,3-dipolar cycloaddition of 1,2-cyclohexadiene occurs quickly with mild heating. The low ~15 kcal/mol barriers for these reactions, compared to the >30 kcal/mol barriers for cycloaddition with linear allenes,17,18 can be attributed to strain and pre-distortion. The geometry-optimized structure 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 1,2cyclohexadiene. 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

N

O

O

Cl

(HO)2B 29 Pd(OAc) 2 (5 mol%) SPhos (12.5 mol%)

N

t-Bu K 3PO 4 PhMe, 100 °C

O H

28

N

t-Bu

O

(65% yield)

from nitrone cycloaddition

H

30

Unconventional Route to Polycyclic Heterocyclic Scaffold NHt-Bu OTf

3 steps

SiEt3

12

H

3 Bonds & 2 Stereocenters Formed

31

N F

O

N

N

t-Bu

F

CsF, CH3CN, 80 °C (87% yield)

N

O

N

33

t-Bu

1. Zn, AcOH, H 2O CH2Cl2, 23 °C (82% yield) 2. NaH, THF, 50 °C (58% yield)

O H

32

In summary, we have discovered the first 1,3dipolar 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

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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. ASSOCIATED CONTENT Supporting Information Available. Detailed experimental and computational procedures, compound characterization, Cartesian coordinates, electronic energies, entropies, enthalpies, Gibbs free energies and lowest frequencies of the calculated structures. This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION Corresponding Authors

[email protected] and [email protected] ACKNOWLEDGMENT 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).

REFERENCES 1

In 2013, 15 of the top 20 small molecule pharmaceuticals contain heterocycles. For discussions, see: (a) Vitaku, E.; Smith, B. R.; Smith, D. T.; Njardarson, J. T. http://njardarson.lab.arizona.edu/sites/njardarson.lab.arizona.edu/files/Top %20US%20Pharmaceutical%20Products%20of%202013.pdf (accessed December 13, 2015). (b) McGrath, N. A.; Brichacek, M.; Njardarson, J. T. J. Chem. Educ. 2010, 87, 1348–1349. 2 Heterocycles are also prevalent motifs in agrochemicals and materials chemistry; see: (a) Dinges, J., Lamberth, C., Eds. Bioactive Heterocyclic Compounds Classes: Agrochemicals; Wiley-VCH: Weinheim, 2012. (b) Gilchrist, T. L. J. Chem. Soc., Perkin Trans. 1 1999, 2849–2866. 3 For reviews of arynes and hetarynes, see: (a) Pellissier, H.; Santelli, M. Tetrahedron 2003, 59, 701–730. (b) Wenk, H. H.; Winkler, M.; Sander, W. Angew. Chem., Int. Ed. 2003, 42, 502–528. (c) Sanz, R. Org. Prep. Proced. Int. 2008, 40, 215–291. (d) Bronner, S. M.; Goetz, A. E.; Garg, N. K. Synlett 2011, 2599–2604. (e) Tadross, P. M.; Stoltz, B. M. Chem. Rev. 2012, 112, 3550–3557. (f) Gampe, C. M.; Carreira, E. M. Angew. Chem., Int. Ed. 2012, 51, 3766–3778. (g) Bhunia, A.; Yetra, S. R.; Biju, A. T. Chem. Soc. Rev. 2012, 41, 3140–3152. (h) Yoshida, H.; Takaki, K. Synlett 2012, 1725–1732. (i) Dubrovskiy, A. V.; Markina, N. A.; Larock, R. C. Org. Biomol. Chem. 2013, 11, 191–218. (j) Wu, C.; Shi, F. Asian J. Org. Chem. 2013, 2, 116–125. (k) Hoffmann, R. W.; Suzuki, K. Angew. Chem., Int. Ed. 2013, 52, 2655–2656. (l) Goetz, A. E.; Garg, N. K. J. Org. Chem.

Page 4 of 5

2014, 79, 846–851. (m) Goetz, A. E.; Shah, T. K.; Garg, N. K. Chem. Commun. 2015, 51, 34–45. 4 For select recent studies of cyclic alkynes, see: (a) Tlais, S. F.; Danheiser, R. L. J. Am. Chem. Soc. 2014, 136, 15489–15492. (b) Medina, J. M.; McMahon, T. C.; Jiménez-Osés, G.; Houk, K. N.; Garg, N. K. J. Am. Chem. Soc. 2014, 136, 14706–14709. (c) McMahon, T. C.; Medina, J. M.; Yang, Y.-F.; Simmons, B. J.; Houk, K. N.; Garg, N. K. J. Am. Chem. Soc. 2015, 137, 4082–4085. 5 Wittig, G.; Fritze, P. Angew. Chem., Int. Ed. 1966, 5, 846. 6 (a) Bottini, A. T.; Hilton, L. L.; Plott, J. Tetrahedron 1975, 31, 1997– 2001. (b) Schmidt, M. W.; Angus, R. O.; Johnson, R. P. J. Am. Chem. Soc. 1982, 104, 6838–6839. (c) Wentrup, C.; Gross, G.; Maquestiau, A.; Flammang, R. Angew. Chem., Int. Ed. Engl. 1983, 22, 542–543. (d) Nendel, M.; Tolbert, L. M.; Herring, L. E.; Islam, M. N.; Houk, K. N. J. Org. Chem. 1999, 64, 976–983. (e) Hänninen, M. M.; Peuronen, A.; Tuononen, H. M. Chem. Eur. J. 2009, 15, 7287–7291. 7 (a) Moore, W. R.; Moser, W. R. J. Org. Chem. 1970, 35, 908–912. (b) Bottini, A. T.; Corson, F. P.; Fitzgerald, R.; Frost, K. A., II; Tetrahedron 1972, 28, 4883–4904. (c) Christl, M.; Schreck, M. Angew. Chem., Int. Ed. Engl. 1987, 26, 449–451. (d) Christl, M.; Fischer, H.; Arnone, M.; Engels, B. Chem. Eur. J. 2009, 15, 11266–11272. (e) Quintana, I.; Peña, D.; Pérez, D.; Guitián, E. Eur. J. Org. Chem. 2009, 5519–5524. 8 (a) Balci, M.; Jones, W. M. J. Am. Chem. Soc. 1980, 102, 7607–7608. (b) Tolbert, L. M.; Islam, M. N.; Johnson, R. P.; Loiselle, P. M.; Shakespeare, W. C. J. Am. Chem. Soc. 1990, 112, 6416–6417. 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 medical chemistry; see: a) Lovering, F.; Bikker, J.; Humblet, C. J. Med. Chem. 2009, 52, 6752–6756. b) Ritchie, T. J.; Macdonald, S. J. F. Drug Discov. Today 2009, 14, 1011–1020. c) Lovering, F. Med. Chem. Commun. 2013, 4, 515–519. 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’s, 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. 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 6311+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. 17 Schmidt, Y.; Lam, J. K.; Pham, H. V.; Houk, K. N.; Vanderwal, C. D. J. Am. Chem. Soc. 2013, 135, 7339–7348. 18 Pham, H. V.; Houk, K. N. J. Org. Chem. 2014, 79, 8968-8976. 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: (a) Gordon, C. G.; Mackey, J. L.; Jewett, J. C.; Sletten, E. M.; Houk, K. N.; Bertozzi, C. R. J. Am. Chem. Soc. 2012, 134, 9199–9208. (b) Lopez, S. A.; Houk, K. N. J. Org. Chem. 2013, 78, 1778–1783. (c) Hong, X.; Liang, Y.; Griffith, A. K.; Lambert, T. H.; Houk, K. N. Chem. Sci. 2014, 5, 471-475. 21 Billingsley, K.; Buchwald, S. L. J. Am. Chem. Soc. 2007, 129, 3358– 3366. 22 Swift, P. A.; Stagnito, M. L.; Mullen, G. B.; Palmer, G. C.; Ceorgiev, V. S. Eur. J. Med. Chem. 1988, 23, 465–471. 23 Saitton, S.; Kihlberg, J.; Luthman, K. Tetrahedron 2004, 6113–6120.

ACS Paragon Plus Environment

Page 5 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

SYNOPSIS TOC

+

R O

R

R' N

R'

*

N R"

*

R''

O

13 examples 62–92% yield

H

Experimental and Computational Studies of First 1,3-Dipolar Cycloaddition of 1,2-Cyclohexadiene NHt-Bu

3 Steps

New Tool to Access Stereochemically-rich polycyclic heterocycles

* * H

O

N

ACS Paragon Plus Environment