Next Article in Journal
Symmetry and Beauty in Plato
Next Article in Special Issue
Fluctuating Asymmetry: Methods, Theory, and Applications
Previous Article in Journal
Hartee Fock Symmetry Breaking Effects in La2CuO4: Hints for connecting the Mott and Slater Pictures and Pseudogap Prediction
Previous Article in Special Issue
Inverse Supersymmetry Breaking in S1 × R3
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diastereoselective Desymmetrization of Symmetric Dienes and its Synthetic Application

Graduate School of Pharmaceutical Sciences, Osaka University, 1-6 Yamada-oka, Suita, Osaka 565-0871, Japan
*
Author to whom correspondence should be addressed.
Symmetry 2010, 2(2), 437-454; https://doi.org/10.3390/sym2020437
Submission received: 30 January 2010 / Revised: 4 March 2010 / Accepted: 4 March 2010 / Published: 25 March 2010
(This article belongs to the Special Issue Feature Papers: Symmetry Concepts and Applications)

Abstract

:
The desymmetrization of symmetric compounds is a useful approach to obtain chiral building blocks. Readily available precursors with a prochiral unit could be converted into complex molecules with multiple stereogenic centers in a single step. In this review, recent advances in the desymmetrization of symmetric dienes in the diastereotopic group differentiating reaction and its synthetic application are presented.

Graphical Abstract

1. Introduction

The desymmetrization of meso compounds has become one of the most powerful strategies in organic synthesis [1]. It allows the formation of multiple chiral centers in a single step and offers an entry to a wide range of stereochemically complex molecules including natural products. The main advantage of the desymmetrization reactions is that symmetric precursors are readily accessible and easily synthesized. Therefore, there are numerous reports on asymmetric synthesis through desymmetrization reactions. Among them, the desymmetrization of diene substrates has been stereoselectively achieved by cycloaddition, Michael addition, halocyclization, radical cyclization, ring-closing metathesis, etc. A remaining olefin unit in the reaction products could be used for further transformations. Substrate-based asymmetric inductions from the latent stereogenic centers, namely, diastereotopic group differentiating reactions during the asymmetric syntheses have been explored [2]. In particular, substituted 1,4-cyclohexadienes have a great potential for the desymmetrization reactions [3,4]. These compounds are easily prepared by the Birch reduction/alkylation sequence of aromatic compounds. A broad range of reactions have been developed for the desymmetrization of cyclohexadienes, and synthetic applications to natural product synthesis have been reported. The reviews on the desymmetrization of dienes before 2005 have already been reported [1,2,3,4]. We will then highlight the recent reports for the desymmetrization of dienes by diastereotopic group differentiating reactions in a stereoselective manner and its applications (Figure 1).

2. Desymmetrization of diene substrates by diastereotopic group differentiating reactions

2.1. Cycloaddition

Stockman et al. reported the tandem intramolecular cycloaddition/triazoline fragmentation reaction [5]. The treatment of the mesylate 1 with sodium azide in DMF at 50 °C produced the azide 2, and the subsequent [3+2] cycloaddition resulted in 3. Fragmentation of the cycloadduct 3 followed by the ring closing reaction afforded the quinolizidine derivative 4 in 52% (Scheme 1). This reaction features a novel self-desymmetrizing cascade process.
Linclau et al. developed the formation method of a C9-substituted trans-hydrindene ring by the intramolecular Diels-Alder reaction (IMDA) [6]. The reaction was conducted under thermal conditions (150 °C in toluene) and the bis(1,3-diene) unit of 5 was desymmetrized leading to a 70:30 mixture of 6 and 7. The use of a Lewis acid (EtAlCl2 in CH2Cl2 at -78 °C) as a catalyst afforded 6 and 7 with a higher selectivity (79:21) (Scheme 2).

2.2. Michael addition

The sulfoxide-directed desymmetrization of cyclohexadienes was studied by Elliott et al. [7]. When the sulfoxide 8 was treated with KOt-Bu, the bicyclic compounds 9 and 10 were obtained in different yields (Scheme 3).
Carreño et al. reported the diastereoselective Michael addition on the cyclohexadiene with the chiral sulfoxide [8]. The chiral sulfoxide 11 was reacted with aryl organoaluminum reagents, leading to the addition products 12 (up to 98% yield, up to 96:4 dr) (Scheme 4). They envisaged that the hydroxyl group has an essential role in the asymmetric induction. They also applied this strategy to synthesize the aryl cyclitol derivatives.

2.3. Iodocyclization

Elliott et al. employed the regio- and diastereoselective iodoetherification of 1,4-cyclohexadiene with a chiral tether [9]. Iodocyclization of the cyclohexadienes 13 having stereogenic centers adjacent to the oxygen atom with I2 and NaHCO3 or Na2CO3 smoothly proceeded to give 5-endo cyclization products 14 with a high selectivity (up to >99:1 dr) and up to 91% yield (Scheme 5). They also studied the competition between the 6-exo and 7-endo cyclization.
Crich et al. developed the desymmetrization of aryl cyclohexadienes [10]. Iodocyclization of the aryl cyclohexadiene 15 provided the lactone 16 in 71% yield (Scheme 6). Compound 16 is a key intermediate in the synthsis of (±)-pancratistatin reported by Danishefsky et al. [11].

2.4. Radical cyclization / Epoxidation-cyclization

Recently, Elliott et al. reported the synthetic studies of lycopodium alkaloids, lycoposerramine A and S, by the desymmetrization of cyclohexadienes under radical conditions [12,13]. The radical cyclization of the mono-TBS compound 17 gave 18 and 19 as a separable mixture. The major isomer 18 underwent further transformations to give the heterocycle compound 20, which is a model for the tetracyclic core of lycoposerramine A [12] (Scheme 7). They also carried out a synthesis towards lycoposerramine S by the same strategy [13].
They also described the stereoselective epoxidation and cyclization of 1,4-cyclohexadienes [14]. Epoxidation of the cyclohexadienes 21 with m-CPBA gave the cyclized products 22 (up to 75% yield) as single diastereoisomers, via the bis-epoxide intermediates (Scheme 8).

2.5. Ring-closing metathesis

Wallace et al. studied the double ring-closing metathesis reaction of the tetraenes 23 with the first generation Grubbs’ catalyst to stereoselectively form the spirocyclic rings 24 (up to 87% yield, up to 92:8 dr) [15]. They also applied this reaction to the enantioselective synthesis of the NK-1 receptor antagonist in 3 steps from 24 (R = Ph) [16] (Scheme 9). Recently, they revealed the reaction mechanism [17,18].
Lee et al. developed the dienyne ring-closing metathesis reaction of the alkynyl silaketals 25 with the second generation Grubbs’ catalyst to afford the bicyclic siloxanes 26 in up to 96% yield. Removal of the silicon tether was achieved with TBAF to give the corresponding 1,3-dienes 27 [19,20] (Scheme 10). This methodology is an efficient tool to synthesize the 1,4-disubstituted 1,3-dienes.
The diastereoselective ring-closing metathesis of phosphorus-containing trienes was reported by Gouverneur et al. [21]. The prochiral trienes 28 reacted with the second generation Grubbs’ catalyst to give the P-stereogenic six-membered heterocycles 29 in up to 100% yield with a high diastereoselectivity (up to 94:6 dr) (Scheme 11).
Blechert et al. employed the diastereoselective ring-rearrangement metathesis (RRM) [22]. Treatment of the cyclodiene 30 with the second generation Grubbs’ catalyst gave the dihydropyran 31 in favor of the 2,6-cis product (>98% yield, 7:2 dr). They supposed that compound 30 reacted the precatalyst to produce the carbene complex and then RCM occurred to give the cyclopentene intermediate. This cyclopentene would rearrange to afford the isolated dihydropyran 31. They also conducted the reactions of the cyclooctadienes 32a,b under the same conditions, and obtained the dihydropyrrole 33a (83% yield, 6:1 dr) and the piperidine derivative 33b (95% yield, 14:1 dr), both of which are trans isomers (Scheme 12).
Pickett et al. explored the concise synthesis of the sex pheromone of Sitodiplosis mosellana by the silicon tethered ring-closing metathesis [23]. The di-t-butyl silaketal 35, prepared from the chiral alcohol, prochiral 1,4-pentadien-3-ol and di-t-butylsilyl bis(trifluoromethanesulfonate) 34, was subjected to metathesis conditions to give 36 in 70% yield from 34 over 2 steps. Removal of the silyl group and further transformations furnished the desired product 37 in 22% overall yield with 97:3 dr (Scheme 13).

2.6. Miscellaneous reactions

Studer et al. found that the chiral cyclohexadienyl-Ti complex reacted with aldehydes to afford chiral 1,3-cyclohexadienyl compounds with a high stereoselectivity [24,25]. They recently reported the desymmetrization of the metallated cyclohexadienes 38 and 39 with chiral N-tert-butanesulfinyl imines 40 [26]. The cyclohexadienyl-MgCl compound 38 provided the 1,3-dienes 41 with a high diastereoseletivity (>99:1 dr). When the metal on the cyclohexadiene was changed from Mg to Zn, the 1,4-dienes 42 were obtained (up to 99:1 dr) (Scheme 14). The 1,4-dienes 42 were readily oxidized to the corresponding diarylmethylamine derivatives, which are found in biologically active compounds.
Gromov and Mulzer et al. developed a desymmetrization approach to construct the functionalized cis-decaline core of branimycin [27]. The treatment of the diepoxynaphthalene 43 with Me2PhSiCH2MgCl/CuCl/Ph3P resulted in an anti SN´ opening one of the oxa-bridges to stereoselectively give 44 in 75% yield (82% based on the recovered starting material 43). Compound 44 underwent further transformations including the opening of the remaining oxa-bridge by hydride attack to synthesize the tricycle structure 45 (Scheme 15).
Landais et al. reported that the highly regio- and stereoselective intramolecular hydroamination of the cyclohexadienes 46 bearing the chiral secondary amine moieties led to the bicyclic allylic amines 47 (up to 95% yield). This cascade reaction is a useful method to provide a key structure, the azabicyclo[4.3.0]nonane system, found in various alkaloids [28] (Scheme 16).
The same group described two different approaches to the tetracyclic core of the aspidosperma alkaloids by a double oxidative amination and double conjugate addition process [29]. Treating the cyclohexadiene 48 with Pd(OAc)2 and O2 in DMSO provided the tetracyclic compound 49 as a single isomer in 75% yield. On the other hand, the allylic oxidation using Pd/C-t-BuOOH and subsequent double 1,4-addition in the presence of DBU led to 50 in 54% yield (Scheme 17).
Prins reaction was also used for the desymmetrization of the 1,8-diene [30]. Elliott et al. reported the Prins desymmetrization of 1,4-cyclohexadienes [31,32]. The cyclohexadienes 51 were treated with TiCl4 or TfOH to form the benzo[c]furan carboxaldehyde derivatives 52 via the oxocarbenium ion (up to 66% yield) (Scheme 18). TfOH is effective for this reaction due to its strong acidity with a non-nucleophilic counter-ion.
Landais et al. studied the rearrangement of spirocyclohexadienyl oxindoles [33,34]. Upon treatment of the spirooxindole 53 with LDA at -40 °C, the dienes rearranged to deliver the lithium enolate intermediate. The intermediate was then trapped by alkyl halides at the C3 center to afford 54 (up to 67%). On the other hand, the aldehydes were reacted at the C5 center to give 55 (up to 63%) with a high stereoselectivity (up to >95:5 dr) (Scheme 19).
Crich et al. reported the dihydroxylation of the aryl cyclohexadienes by the desymmetrization strategy [35]. Dihydroxylation of the cyclohexadiene 56 occurred exclusively with anti-selectivity to the substituted aromatic ring to minimize allylic strain to afford the triol 57 in 65% yield (Scheme 20).

2.7. Asymmetric desymmetrization using C2-symmetric acetal or aminal

We have developed an asymmetric synthesis of the chiral 1,4- and 1,5-diols by the diastereoselective intramolecular haloetherification reaction of the ene acetals 58, easily prepared from the C2-symmetric diols and ene aldehydes, involving the remote asymmetric induction [36,37]. This reaction stereoselectively proceeded via the chiral bicyclic cationic intermediates, and the subsequent nucleophilic addition of an alcohol to the oxonium ion intermediate afforded 59, which have two newly formed chiral centers on the prochiral atom and the acetal center of the starting materials 58. The chiral auxiliary unit derived from the C2-symmetric diol could be removed by various procedures, i.e., Birch reduction, hydrogenolysis, and fragmentation promoted by CAN [38,39] (Scheme 21).
We then studied the intramolecular bromoetherification reaction of the substituted cyclohexadiene having the C2-symmetric acetal 60 and found that the reaction successfully discriminated the two prochiral olefins in a stereoselective manner to provide 61 in 63% yield [40,41] (Scheme 22). This reaction was useful for constructing complex molecules because it could produce optically active cyclohexane derivatives bearing multiple chiral centers and the remaining olefin, which promised further transformations. The sterically hindered acetal moiety derived from the C2-symmetric acetal could fix the cyclohexene ring conformation promising high regio- and stereoselective transformations. For example, the dihydroxylation of 62 with OsO4 stereoselectively proceeded from the less hindered face to give the corresponding diol 63 in 85% yield as a single product [41] (Scheme 23).
Enantioselective reactions with chiral catalysts are well-known as more powerful methods than diastereoselective approaches with chiral auxiliaries for asymmetric synthesis. The fatal drawback in diastereoselective reaction is that at least an equimolar amount of the auxiliary is necessary. However, for the synthesis of complex molecules including natural products, various reactions such as the regio-, stereo-, and chemoselective transformations must be considered. Recently, we have proposed a novel concept, the chiral auxiliary multiple-use methodology [42]. This is a new perspective in asymmetric synthesis utilizing a chiral auxiliary several times, namely, asymmetric induction, regio-, stereoselective transformation and protective group of the functional group (Figure 2).
We achieved the concise synthesis of (+)-Sch 642305 by this methodology starting from the bromo acetal 61, obtained by the bromoetherification of the cyclohexadiene acetal 60 [42]. First, the acetal 61 was subjected to the hydroboration-oxidation condition. The hydroboration reaction regioselectively proceeded and the subsequent oxidation gave the bromo ketone intermediate. This intermediate spontaneously converted to the enone 64 by elimination of HBr in 53% yield. The structure of 64 showed that the upper side of the cyclohexene ring was shielded by the axial C-C bond and methylene group of the acetal ring. Furthermore, the 8-membered acetal ring derived from chiral auxiliary fixed the conformation to prevent the cyclohexane ring inversion (Scheme 24).
The aldol reaction of 64 with the chiral aldehyde proceeded from the sterically less hindered face to afford the alcohol 65 as a single diastereomer. Further transformations including the radical reduction and macrolactonization gave the macrolactone 66. The chiral auxiliary was finally removed under oxidative conditions to furnish (+)-Sch 642305 via the ester intermediate (Scheme 25).
We also achieved the asymmetric total synthesis of scyphostatin, a potent inhibitor of neutral sphingomyelinase [43]. The radical reduction of 61 and stereoselective introduction of the tertiary alcohol with SeO2 gave 68. Further transformations afforded the dimethylacetal 69. Deprotection of the acetal function by combination with silyl triflate/2,4,6-collidine under basic conditions gave 70 [44,45]. With 71 in hand, the deprotection of 2,4-dimethoxyphenylmethyl (2,4DMPM) group with Ph3C+BF4- then completed the total synthesis of scyphostatin (Scheme 26).
We have found that the novel double iodoetherification reaction of acyclic σ-symmetric diene acetals from the C2-symmetric diol, i.e., the chiral hydrobenzoin, afforded tetrahydrofurans with multiple chiral centers in a single operation [46,47]. Although discrimination of the two olefins of the σ-symmetric dienes with chiral auxiliaries by asymmetric intramolecular halolactonization have been reported, they do not have auxiliaries in the reaction products [48,49,50,51,52,53]. Compared with other reports, our method features that the chiral auxiliary remained in the product. Therefore, we could utilize the product for further regio- and stereoselective transformations. The acyclic diene acetals 72 with NIS and H2O led to the optically active tetrahydrofuran derivatives 75 involving the newly installed four stereogenic centers (up to 78% yield, up to 11:1 dr). The reaction proceeded via the chiral hemiacetal intermediates 74 formed by iodoetherification of the acetal oxygen atoms and subsequent nucleophilic addition of H2O to 73, followed by a second iodoetherication to give 75 (Scheme 27). It is known that hemiacetal compounds could convert to hydroxyl aldehydes through a ring-opning reaction. Alternatively, the hemiacetal intermediates 74 have another double bond in the proper position and the second iodoetherification smoothly occurs.
Since the acetals 75 have the iodomethyl functions, we tried to perform the regioselective nucleophilic substitution of iodine. The treatment of 75 with sodium cyanide or sodium malonate gave 76 (88% yield) and 77 (73% yield), the regioselective substitution products in the less hindered position (b) (Scheme 28). This result indicates that 75 could be a useful chiral synthon for the synthesis of optically active tetrahydrofurans or γ-lactones.
We applied this method to the asymmetric synthesis of rubrenolide and rubrynolide, which have the remote asymmetric center and substituted the γ-lactone ring. The 8-membered acetal 78 was converted to the lactone 79. Removal of the chiral auxiliary with CAN, developed in our laboratory [38,39], followed by epoxidation of the resulting alcohol with Ag2O afforded 80. The treatment of the epoxide 80 with the corresponding Grignard reagent, and further transformation to furnish the synthesis of rubrenolide and rubrynolide (Scheme 29).
Our groups also employed the desymmetrization reaction using C2-symmetric diene aminals in place of acetals to give compounds containing the nitrogen atom [54]. This reaction involves three steps, i.e., the formation of aminal (8182), stereoselective bromoamination (8283), and oxidation of the aminal (8384) in a cascade process (57% yield) (Scheme 30).
Furthermore, we applied this reaction to the concise synthesis of (–)-γ-lycorane. We proposed that the chiral imidazoline 85 derived from 1,2-di(4-methoxyphenyl)-1,2-diamine was an useful chiral synthon for achievement of the total synthesis. The compound 85 was converted to 86 via lactam formation and condensation of the aromatic unit. Finally, Friedel-Crafts reaction of 86 and subsequent reduction furnished (–)-γ-lycorane (Scheme 31).

3. Conclusions

The desymmetrization reaction have enjoyed much popularity in organic synthetic chemistry. Among the many desymmetrization reactions, we focused on the symmetric dienes as a precusor. Differentiation of the diastereotopic diene group could be attained by cycloaddition, Michael addition, halocyclization, radical cyclization, ring-closing metathesis, etc. The alkene moiety in the reaction products could be used for further transformation. Therefore, the cyclic and acyclic dienes are promising precusors to obtain complex molecules with several chiral centers including biologically active natural products.

References

  1. Willis, M.C. Enantioselective desymmetrization. J. Chem. Soc. Perkin Trans. 1 1999, 1765–1784. [Google Scholar] [CrossRef]
  2. Hoffmann, R.W. Stereoselective synthesis using diastereotopic groups. Synthesis 2004, 2075–2090. [Google Scholar] [CrossRef]
  3. Abd Rahman, N.; Landais, Y. Desymmetrization of cyclic dienes. An efficient strategy for natural products synthesis. Curr. Org. Chem. 2002, 6, 1369–1395. [Google Scholar] [CrossRef]
  4. Studer, A.; Schleth, F. Desymmetrization and diastereotopic group selection in 1,4-cyclohexadienes. Synlett 2005, 3033–3041. [Google Scholar] [CrossRef]
  5. Rejzek, M.; Stockman, R.A.; van Maarseveen, J.H.; Hughes, D.L. Combining two-directional synthesis and tandem reactions: Desymmetrisation by intramolecular cycloaddition/triazoline fragmentation. Chem. Commun. 2005, 4661–4662. [Google Scholar] [CrossRef] [PubMed]
  6. Azzi, N.; Griffen, E.; Light, M.; Linclau, B. An enantioselective desymmterisation approach to C9-substituted trans-hydrindene rings based on a diastereotopic group-selective intramolecular Diels-Alder reaction. Chem. Commun. 2006, 4909–4911. [Google Scholar] [CrossRef] [PubMed]
  7. Elliott, M.C.; El Sayed, N.N.E.; Ooi, L.-l. Sulfoxide-directed desymmetrisation of cyclohexa-1,4-dienes. Tetrahedron Lett. 2007, 48, 4561–4564. [Google Scholar] [CrossRef]
  8. Merino, E.; Melo, R.P.A.; Ortega-Guerra, M.; Ribagorda, M.; Carreño, M.C. Stereocontrolled approach to phenyl cyclitols from (SR)-[(p-tolylsulfinyl)methyl]-p-quinol. J. Org. Chem. 2009, 74, 2824–2831. [Google Scholar] [CrossRef]
  9. Butters, M.; Elliott, M.C.; Hill-Cousins, J.; Paine, J.S.; Walker, J.K.E. Iodocyclization reactions for the desymmetrization of cyclohexa-1,4-dienes. Org. Lett. 2007, 9, 3635–3638. [Google Scholar] [CrossRef]
  10. Crich, D.; Krishnamurthy, V. Radical dearomatization of benzene leading to phenanthridine and phenanthridinone derivatives related to (±)-pancratistatin. Tetrahedron 2006, 62, 6830–6840. [Google Scholar] [CrossRef]
  11. Danishefsky, S.; Lee, J.Y. Total synthesis of (±)-pancratistatin. J. Am. Chem. Soc. 1989, 111, 4829–4837. [Google Scholar] [CrossRef]
  12. Elliott, M.C.; Paine, J.S. Studies towards the total synthesis of lycoposerramine A. Synthesis of a model for the tetracyclic core. Org. Biomol. Chem. 2009, 7, 3455–3462. [Google Scholar] [CrossRef] [PubMed]
  13. Elliott, M.C.; El Sayed, N.N.E.; Paine, J.S. Concise synthesis of the tricyclic core of lycoposerramine S. Org. Biomol. Chem. 2008, 6, 2611–2618. [Google Scholar] [CrossRef] [PubMed]
  14. Butters, M.; Beetstra, D.J.; Elliott, M.C.; Hill-Cousins, J.; Kariuki, B.M. Directed epoxidation of cyclohexa-1,4-dienes–stereoselective formation of up to six contiguous stereogenic centres. Org. Biomol. Chem. 2008, 6, 4426–4434. [Google Scholar] [CrossRef] [PubMed]
  15. Wallace, D.J.; Cowden, C.J.; Kennedy, D.J.; Ashwood, M.S.; Cottrell, I.F.; Dolling, U.-H. Stereoselective double ring closing metathesis reactions in the synthesis of spirocyclic compounds. Tetrahedron Lett. 2000, 41, 2027–2029. [Google Scholar] [CrossRef]
  16. Wallace, D.J.; Goodman, J.M.; Kennedy, D.J.; Davies, A.J.; Cowden, C.J.; Ashwood, M.S.; Cottrell, I.F.; Dolling, U.-H.; Reider, P.J. A double ring closing metathesis reaction in the rapid, enantioselective synthesis of NK-1 receptor antagonists. Org. Lett. 2001, 3, 671–674. [Google Scholar] [CrossRef] [PubMed]
  17. Wallace, D.J. On the mechanism of a double ring-closing metathesis reaction. Tetrahedron Lett. 2005, 46, 591–594. [Google Scholar] [CrossRef]
  18. Wallace, D.J. Exploiting catalyst characteristics: A protocol for increasing diastereoselectivity in a double ring-closing metathesis reaction. J. Mol. Catal. A: Chem. 2006, 254, 78–84. [Google Scholar] [CrossRef]
  19. Grimm, J.B.; Lee, D. Efficient synthesis of alkynylsilyl ethers and silaketals via base-induced alkynylsilane alcoholysis. J. Org. Chem. 2004, 69, 8967–8970. [Google Scholar] [CrossRef]
  20. Grimm, J.B.; Otte, R.D.; Lee, D. Tandem dienyne ring-closing metathesis of alkynyl silaketals for the formation of bicyclic siloxanes. J. Organomet. Chem. 2005, 690, 5508–5516. [Google Scholar] [CrossRef]
  21. Dunne, K.S.; Bisaro, F.; Odell, B.; Paris, J.-M.; Gouverneur, V. Diastereoselective ring-closing metathesis: Synthesis of P-stereogenic phosphinates from prochiral phosphinic acid derivatives. J. Org. Chem. 2005, 70, 10803–10809. [Google Scholar] [CrossRef] [PubMed]
  22. Böhrsch, V.; Neidhöfer, J.; Blechert, S. Diastereoselective ring-rearrangement metathesis. Angew. Chem., Int. Ed. 2006, 45, 1302–1305. [Google Scholar] [CrossRef] [PubMed]
  23. Hooper, A.M.; Dufour, S.; Willaert, S.; Pouvreau, S.; Pickett, J.A. Synthesis of (2S,7S)-dibutyroxynonane, the sex pheromone of the orange wheat blossom midge, Sitodiplosis mosellana (Géhin) (Diptera: Cecidomyiidae), by diastereoselective silicon-tethered ring-closing metathesis. Tetrahedron Lett. 2007, 48, 5991–5994. [Google Scholar] [CrossRef]
  24. Schleth, F.; Studer, A. Desymmetrization of metalated cyclohexadienes and application to the synthesis of nephrosteranic acid. Angew. Chem., Int. Ed. 2004, 43, 313–315. [Google Scholar] [CrossRef] [PubMed]
  25. Schleth, F.; Vogler, T.; Harms, K.; Studer, A. Stereoselective synthesis of (+)-nephrosteranic acid, (+)-trans-cognac lactone, and (+)-trans-whisky lactone using a chiral cyclohexadienyl Ti compound. Chem. Eur. J. 2004, 10, 4171–4185. [Google Scholar] [CrossRef] [PubMed]
  26. Maji, M.S.; Fröhlich, R.; Studer, A. Desymmetrization of metallated cyclohexadienes with chiral N-tert-butanesulfinyl imines. Org. Lett. 2008, 10, 1847–1850. [Google Scholar] [CrossRef] [PubMed]
  27. Gromov, A.; Enev, V.; Mulzer, J. A desymmetrization approach toward highly oxygenated cis-decalins. Org. Lett. 2009, 11, 2884–2886. [Google Scholar] [CrossRef]
  28. Lebeuf, R.; Robert, F.; Schenk, K.; Landais, Y. Desymmetrization of cyclohexa-2,5-dienes through a diastereoselective protonation-hydroamination cascade. Org. Lett. 2006, 8, 4755–4758. [Google Scholar] [CrossRef]
  29. Beniazza, R.; Dunet, J.; Robert, F.; Schenk, K.; Landais, Y. Efficient synthetic approaches to the common scaffold of indole alkaloids. Org. Lett. 2007, 9, 3913–3916. [Google Scholar] [CrossRef]
  30. Rychnovsky, S.D.; Yang, G.; Hu, Y.; Khire, U.R. Prins desymmetrization of a C2-symmetric diol: Application to the synthesis of 17-deoxyroflamycoin. J. Org. Chem. 1997, 62, 3022–3023. [Google Scholar] [CrossRef]
  31. Elliott, M.C.; El Sayed, N.N.E.; Paine, J.S. Diastereospecific tandem Prins cyclisation/rearrangement reactions for the desymmetrisation of cyclohexa-1,4-dienes. Eur. J. Org. Chem. 2007, 792–803. [Google Scholar] [CrossRef]
  32. Butters, M.; Elliott, M.C.; Hill-Cousins, J.; Paine, J.S.; Westwood, A.W.J. An improved protocol for the Prins desymmetrisation of cyclohexa-1,4-dienes. Tetrahedron Lett. 2008, 49, 4446–4448. [Google Scholar] [CrossRef]
  33. Rousseau, G.; Robert, F.; Schenk, K.; Landais, Y. Rearrangement of spirocyclic oxindoles with lithium amide bases. Org. Lett. 2008, 10, 4441–4444. [Google Scholar] [CrossRef] [PubMed]
  34. Rousseau, G.; Robert, F.; Landais, Y. Functionalization and rearrangement of spirocyclohexadienyl oxindoles: Experimental and theoretical investigations. Chem. Eur. J. 2009, 15, 11160–11173. [Google Scholar] [CrossRef] [PubMed]
  35. Crich, D.; Grant, D.; Wink, D.J. Expedient two-step synthesis of phenolic cyclitols from benzene. J. Org. Chem. 2006, 71, 4521–4524. [Google Scholar] [CrossRef] [PubMed]
  36. Fujioka, H.; Kitagawa, H.; Matsunaga, N.; Nagatomi, Y.; Kita, Y. Remote asymmetric induction for chiral 1,4-diols using a chiral acetal derived from chiral hydrobenzoin. Tetrahedron Lett. 1996, 37, 2245–2248. [Google Scholar] [CrossRef]
  37. Fujioka, H.; Kitagawa, H.; Nagatomi, Y.; Kita, Y. Asymmetric induction via an intramolecular haloetherification reaction of chiral ene acetals: A novel approach to optically active 1,4- and 1,5-diols. J. Org. Chem. 1996, 61, 7309–7315. [Google Scholar] [CrossRef]
  38. Fujioka, H.; Ohba, Y.; Hirose, H.; Murai, K.; Kita, Y. Mild and efficient removal of hydroxyethyl unit from 2-hydroxyethyl ether derivatives leading to alcohols. Org. Lett. 2005, 7, 3303–3306. [Google Scholar] [CrossRef]
  39. Fujioka, H.; Hirose, H.; Ohba, Y.; Murai, K.; Nakahara, K.; Kita, Y. Cerium ammonium nitrate (CAN) for mild and efficient reagent to remove hydroxyethyl units from 2-hydroxyethyl ethers and 2-hydroxyethyl amines. Tetrahedron 2007, 63, 625–637. [Google Scholar] [CrossRef]
  40. Fujioka, H.; Kotoku, N.; Sawama, Y.; Nagatomi, Y.; Kita, Y. Concise asymmetric synthesis of a model compound, (4S,5S,6S)-6-(2,2-dimethoxy)ethyl-4,5-epoxy-6-hydroxy-2-cyclohexenone, for the cyclohexenone core of scyphostatin. Tetrahedron Lett. 2002, 43, 4825–4828. [Google Scholar] [CrossRef]
  41. Fujioka, H.; Kotoku, N.; Sawama, Y.; Kitagawa, H.; Ohba, Y.; Wang, T.-L.; Nagatomi, Y.; Kita, Y. Asymmetric synthesis by the intramolecular haloetherification reaction of ene acetal: Discrimination of prochiral dienes in cyclohexane systems. Chem. Pharm. Bull. 2005, 53, 952–957. [Google Scholar] [CrossRef] [PubMed]
  42. Fujioka, H.; Ohba, Y.; Nakahara, K.; Takatsuji, M.; Murai, K.; Ito, M.; Kita, Y. New perspective for natural products synthesis: Concise synthesis of (+)-Sch 642305 by chiral auxiliary multiuse methodology. Org. Lett. 2007, 9, 5605–5608. [Google Scholar] [CrossRef] [PubMed]
  43. Fujioka, H.; Sawama, Y.; Kotoku, N.; Ohnaka, T.; Okitsu, T.; Murata, N.; Kubo, O.; Li, R.; Kita, Y. Concise asymmetric total synthesis of scyphostatin, a potent inhibitor of neutral sphingomyelinase. Chem. Eur. J. 2007, 13, 10225–10238. [Google Scholar] [CrossRef] [PubMed]
  44. Fujioka, H.; Sawama, Y.; Murata, N.; Okitsu, T.; Kubo, O.; Matsuda, S.; Kita, Y. Unexpected highly chemoselective deprotection of the acetals from aldehydes and not ketones: TESOTf–2,6-lutidine combination. J. Am. Chem. Soc. 2004, 126, 11800–11801. [Google Scholar] [CrossRef] [PubMed]
  45. Fujioka, H.; Okitsu, T.; Sawama, Y.; Murata, N.; Li, R.; Kita, Y. Reaction of the acetals with TESOTf–base combination; Speculation of the intermediates and efficient mixed acetal formation. J. Am. Chem. Soc. 2006, 128, 5930–5938. [Google Scholar] [CrossRef] [PubMed]
  46. Fujioka, H.; Ohba, Y.; Hirose, H.; Murai, K.; Kita, Y. A double iodoetherification of σ-symmetric diene acetals for installing four stereogenic centers in a single operation: Short asymmetric total synthesis of rubrenolide. Angew. Chem., Int. Ed. 2005, 44, 734–737. [Google Scholar] [CrossRef] [PubMed]
  47. Fujioka, H.; Ohba, Y.; Hirose, H.; Nakahara, K.; Murai, K.; Kita, Y. Facile formation of tetrahydrofurans with multiple chiral centers using double iodoetherification of σ-symmetric diene acetals: Short asymmetric total synthesis of rubrenolide and rubrynolide. Tetrahedron 2008, 64, 4233–4245. [Google Scholar] [CrossRef]
  48. Takano, S.; Murakata, C.; Imamura, Y.; Tamura, N.; Ogasawara, K. Chiral route to some alkaloids through asymmetric iodolactonization. Heterocycles 1981, 16, 1291–1294. [Google Scholar] [CrossRef]
  49. Hart, D.J.; Huang, H.-C.; Krishnamurthy, R.; Schwartz, T. Free-radical cyclizations: Application to the total synthesis of dl-pleurotin and dl-dihydropleurotin acid. J. Am. Chem. Soc. 1989, 111, 7507–7519. [Google Scholar] [CrossRef]
  50. Fuji, K.; Node, M.; Naniwa, Y.; Kawabata, T. Enantioselective iodolactonization through diastereotopic group differentiation. Tetrahedron Lett. 1990, 31, 3175–3178. [Google Scholar]
  51. Yokomatsu, T.; Iwasawa, H.; Shibuya, S. Highly enantioselective creation of quaternary carbons in a halolactonization of bis-γ,δ-unsaturated carboxylic imides derived from a camphorsultam: Enantioselective synthesis of (+)-mesembrine. Tetrahedron Lett. 1992, 33, 6999–7002. [Google Scholar] [CrossRef]
  52. Yokomatsu, T.; Iwasawa, H.; Shibuya, S. Enantioselective halolactonisation of bis-γ,δ-unsaturated carboxylic acid derivatives: Use of a sultam and oxazolidine-2-ones as chiral auxiliary. J. Chem. Soc. Chem. Commun. 1992, 728–729. [Google Scholar] [CrossRef]
  53. Kitagawa, O.; Momose, S.; Fushimi, Y.; Taguchi, T. Efficient synthesis of various atropisomeric amides in optically pure forms and their application to asymmetric reactions. Tetrahedron Lett. 1999, 40, 8827–8831. [Google Scholar] [CrossRef]
  54. Fujioka, H.; Murai, K.; Ohba, Y.; Hirose, H.; Kita, Y. Intramolecular bromo-aminaion of 1,4-cyclohexadiene aminal: one-pot discrimination of two olefins and concise asymmetric synthesis of (–)-γ-lycorane. Chem. Commun. 2006, 832–834. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Diastereoselective desymmetrization of symmetric dienes.
Figure 1. Diastereoselective desymmetrization of symmetric dienes.
Symmetry 02 00437 g001
Scheme 1. Intramolecular cycloaddition/triazoline fragmentation.
Scheme 1. Intramolecular cycloaddition/triazoline fragmentation.
Symmetry 02 00437 sch001
Scheme 2. Diastereotopic group selective intramolecular Diels-Alder reaction.
Scheme 2. Diastereotopic group selective intramolecular Diels-Alder reaction.
Symmetry 02 00437 sch002
Scheme 3. Sulfoxide-directed desymmetrization of cyclohexadiene.
Scheme 3. Sulfoxide-directed desymmetrization of cyclohexadiene.
Symmetry 02 00437 sch003
Scheme 4. Diastereoselective Michael addition on cyclohexadiene with chiral sulfoxide.
Scheme 4. Diastereoselective Michael addition on cyclohexadiene with chiral sulfoxide.
Symmetry 02 00437 sch004
Scheme 5. Iodocyclization of 1,4-cyclohexadiene with a chiral tether.
Scheme 5. Iodocyclization of 1,4-cyclohexadiene with a chiral tether.
Symmetry 02 00437 sch005
Scheme 6. Iodocyclization of aryl cyclohexadiene.
Scheme 6. Iodocyclization of aryl cyclohexadiene.
Symmetry 02 00437 sch006
Scheme 7. Synthesis of the tetracyclic core of lycoposerramine A.
Scheme 7. Synthesis of the tetracyclic core of lycoposerramine A.
Symmetry 02 00437 sch007
Scheme 8. Stereoselective epoxidation and cyclization of cyclohexadiene.
Scheme 8. Stereoselective epoxidation and cyclization of cyclohexadiene.
Symmetry 02 00437 sch008
Scheme 9. Double ring-closing metathesis of tetraenes.
Scheme 9. Double ring-closing metathesis of tetraenes.
Symmetry 02 00437 sch009
Scheme 10. Dienyne ring-closing metathesis of alkynyl silaketals.
Scheme 10. Dienyne ring-closing metathesis of alkynyl silaketals.
Symmetry 02 00437 sch010
Scheme 11. Ring-closing metathesis of P-containing trienes.
Scheme 11. Ring-closing metathesis of P-containing trienes.
Symmetry 02 00437 sch011
Scheme 12. Diastereoselective ring-rearrangement metathesis.
Scheme 12. Diastereoselective ring-rearrangement metathesis.
Symmetry 02 00437 sch012
Scheme 13. Diastereoselective silicon-tethered ring-closing metathesis and synthesis of the sex pheromone of the orange wheat blossom midge.
Scheme 13. Diastereoselective silicon-tethered ring-closing metathesis and synthesis of the sex pheromone of the orange wheat blossom midge.
Symmetry 02 00437 sch013
Scheme 14. Desymmetrization of metallated cyclohexadienes with N-tert-butanesulfinyl imines.
Scheme 14. Desymmetrization of metallated cyclohexadienes with N-tert-butanesulfinyl imines.
Symmetry 02 00437 sch014
Scheme 15. Desymmetrization of diepoxynaphthalene and synthesis of cis-decaline core of branimycin.
Scheme 15. Desymmetrization of diepoxynaphthalene and synthesis of cis-decaline core of branimycin.
Symmetry 02 00437 sch015
Scheme 16. Diastereoselective protonation-hydroamination cascade reaction.
Scheme 16. Diastereoselective protonation-hydroamination cascade reaction.
Symmetry 02 00437 sch016
Scheme 17. Two approaches to tetracyclic core of aspidosperma alkaloids.
Scheme 17. Two approaches to tetracyclic core of aspidosperma alkaloids.
Symmetry 02 00437 sch017
Scheme 18. Prins desymmetrization of 1,4-cyclohexadienes.
Scheme 18. Prins desymmetrization of 1,4-cyclohexadienes.
Symmetry 02 00437 sch018
Scheme 19. Rearrangement of spirocyclohexadienyl oxindoles followed by addition of alkyl halides or aldehydes.
Scheme 19. Rearrangement of spirocyclohexadienyl oxindoles followed by addition of alkyl halides or aldehydes.
Symmetry 02 00437 sch019
Scheme 20. Dihydroxylation of aryl cyclohexadiene.
Scheme 20. Dihydroxylation of aryl cyclohexadiene.
Symmetry 02 00437 sch020
Scheme 21. Intramolecular haloetherification of ene acetals.
Scheme 21. Intramolecular haloetherification of ene acetals.
Symmetry 02 00437 sch021
Scheme 22. Intramolecular bromoetherification of cyclohexadiene acetal.
Scheme 22. Intramolecular bromoetherification of cyclohexadiene acetal.
Symmetry 02 00437 sch022
Scheme 23. Stereoselective dihydroxylation.
Scheme 23. Stereoselective dihydroxylation.
Symmetry 02 00437 sch023
Figure 2. Chiral auxiliary multiple-use methodology.
Figure 2. Chiral auxiliary multiple-use methodology.
Symmetry 02 00437 g002
Scheme 24. Hydroboration-oxidation procedure.
Scheme 24. Hydroboration-oxidation procedure.
Symmetry 02 00437 sch024
Scheme 25. Total synthesis of (+)-Sch 642305.
Scheme 25. Total synthesis of (+)-Sch 642305.
Symmetry 02 00437 sch025
Scheme 26. Total synthesis of scyphostatin.
Scheme 26. Total synthesis of scyphostatin.
Symmetry 02 00437 sch026
Scheme 27. Double iodoetherification of diene acetals.
Scheme 27. Double iodoetherification of diene acetals.
Symmetry 02 00437 sch027
Scheme 28. Regioselective substitution.
Scheme 28. Regioselective substitution.
Symmetry 02 00437 sch028
Scheme 29. Total synthesis of rubrenolide and rubrynolide.
Scheme 29. Total synthesis of rubrenolide and rubrynolide.
Symmetry 02 00437 sch029
Scheme 30. Desymmetrization of diene aminal.
Scheme 30. Desymmetrization of diene aminal.
Symmetry 02 00437 sch030
Scheme 31. Total synthesis of (–)-γ-lycorane.
Scheme 31. Total synthesis of (–)-γ-lycorane.
Symmetry 02 00437 sch031

Share and Cite

MDPI and ACS Style

Nakahara, K.; Fujioka, H. Diastereoselective Desymmetrization of Symmetric Dienes and its Synthetic Application. Symmetry 2010, 2, 437-454. https://doi.org/10.3390/sym2020437

AMA Style

Nakahara K, Fujioka H. Diastereoselective Desymmetrization of Symmetric Dienes and its Synthetic Application. Symmetry. 2010; 2(2):437-454. https://doi.org/10.3390/sym2020437

Chicago/Turabian Style

Nakahara, Kenji, and Hiromichi Fujioka. 2010. "Diastereoselective Desymmetrization of Symmetric Dienes and its Synthetic Application" Symmetry 2, no. 2: 437-454. https://doi.org/10.3390/sym2020437

Article Metrics

Back to TopTop