Asymmetric Diels–Alder Reaction Involving Dynamic

A Practical and Chemoselective Ammonia-Free Birch Reduction. Organic Letters ... Picazo, Anthony, Giroud, Simon, Miller, Houk, and Garg. 2018 140 (24)...
2 downloads 0 Views 1MB Size
Article pubs.acs.org/joc

Cite This: J. Org. Chem. 2018, 83, 9300−9304

Asymmetric Diels−Alder Reaction Involving Dynamic Enantioselective Crystallization Naohiro Uemura,† Seiya Toyoda,† Hiroki Ishikawa,† Yasushi Yoshida,† Takashi Mino,† Yoshio Kasashima,‡ and Masami Sakamoto*,† †

Department of Applied Chemistry and Biotechnology, Graduate School of Engineering, and Molecular Chirality Research Center, Yayoi-cho, Inage-ku, Chiba 265-8522, Japan ‡ Education Center, Faculty of Creative Engineering, Chiba Institute of Technology, Shibazono, Narashino, Chiba 275-0023, Japan

J. Org. Chem. 2018.83:9300-9304. Downloaded from pubs.acs.org by UNIV OF SUSSEX on 08/17/18. For personal use only.

S Supporting Information *

ABSTRACT: Asymmetric Diels−Alder reaction was achieved under achiral conditions. Reaction of prochiral 2-methylfuran and N-phenylmaleimide in heptane or hexane solution at 80 °C efficiently gave a conglomerate crystal of exo-type Diels−Alder adduct selectively, and continuous suspension of the reaction mixture with glass beads promoted attrition-enhanced deracemization, leading to an optically active exo-adduct in 90% ee.



INTRODUCTION The Diels−Alder (DA) reaction is one of the most fundamental and illustrious reactions in organic chemistry, providing access to a variety of polycyclic heterocycles.1 Numerous important skeletons have been conveniently constructed by the reaction of various dienes and alkenes.2 Furthermore, this reaction applies to many asymmetric reactions using catalytic asymmetric synthesis3 and chiral auxiliaries,4 leading to optically active cyclic compounds. Naturally, all of these synthetic methodologies require optically active sources such as chiral catalysts or precursors. Recently, asymmetric synthesis involving a reversible reaction and crystal chirality has been developed. Stereoselective crystallization via stereoisomerization from meso to dl-isomers,5 azaMichael addition reactions,6 photoisomerization followed by ring-closing−opening reactions,7 Strecker-type reactions,8 and reversible photodimerization reactions9 have been reported. These reactions were initiated from prochiral precursors, and high optically active crystalline products were conveniently provided under absolutely achiral conditions. These types of reactions all involve dynamic stereoselective crystallization, which is well documented as a total optical resolution method or a crystallization-induced transformation (CIET).10−15 The combination of chiral center generation from prochiral starting materials followed by CIET makes it possible to design absolute asymmetric synthesis involving many types of organic reactions. Up to now, absolute asymmetric synthesis has been limited to the application of physical chiral forces such as circularly polarized light16 and chiral magnetic fields.17 Of course, Soai’s reaction18 and asymmetric synthesis using the chirality of crystals generated by the crystallization of prochiral materials is also considered to be an absolute asymmetric synthesis.19,20 These processes are of great interest to many scientists in a variety of areas because absolute asymmetric © 2018 American Chemical Society

synthesis could support the hypothesis of the origin of homochirality on the primitive Earth.21 Here, we developed a novel absolute asymmetric Diels−Alder reaction by using prochiral maleimide and furan, which is a well-known reversible reaction.22 Specifically, the reaction of asymmetrical 2-methylfuran and symmetrical N-phenylmaleimide was selected (Scheme 1).23 Scheme 1. Diels−Alder Reaction of N-Phenylmaleimide and 2-Methylfuran



RESULTS AND DISCUSSION Jarosz et al. reported that their DA reaction gave a mixture of exo- and endo-adducts in a ratio of 1.6:1.0 under high-pressure conditions at room temperature.24 Furthermore, the crystal structure of the exo-adduct revealed that the space group was chiral P212121. We reanalyzed the crystal structure both to eliminate the possibility of a polymorphic crystal system and to determine the crystal system of the endo-adduct. When a homogeneous toluene solution of N-phenylmaleimide 2 and 2-methylfuran 1 (1.5 equiv) was left standing overnight, a colorless solid of the exo-adduct was obtained. The filtered exo-3 and endo-4 isolated from the mother liquor were analyzed Received: May 17, 2018 Published: May 28, 2018 9300

DOI: 10.1021/acs.joc.8b01273 J. Org. Chem. 2018, 83, 9300−9304

Article

The Journal of Organic Chemistry by X-ray crystallography. The exo-3 was indeed obtained as a conglomerate of the P212121 space group. The absolute configuration of (−)-exo-3 was determined as (2aS,1R,7S,6aR) (Figure 1), whereas the space group of endo-4 was racemic Pbca (Figure 2).

Scheme 2. Absolute Asymmetric Synthesis by Diels−Alder Reaction Involving Dynamic Stereoselective Crystallization

Figure 1. Perspective view of (−)-exo-3. Ellipsoids were drawn in 50% probability.

Figure 3. Time-dependent reaction course as monitored by 1H NMR spectroscopy. N-Phenylmaleimide (0.02 M) and 2-methylfuran (0.2 M) were reacted in CDCl3 at 60 °C. Blue triangles = unreacted 2; red squares = exo-3; green diamonds = endo-4.

Figure 2. Perspective view of endo-4. Ellipsoids were drawn in 50% probability.

of the reverse reaction under the same conditions. The reverse reaction without catalyst was quite slow; thus, we sought out the best catalyst for activating the reverse reaction using various catalysts such as Zn(OAc)2, TfOH, TsOH, HCl, AcOH, (TfO)3Yb, (TfO)3Eu, p-CF3C6H4B(OH)2, (i-PrO)2TiCl2, TMSNTf2, TMSOTf, and trifluoroacetic acid (TFA). When TMSNTf2 or TMSOTf was used to activate the reverse reaction, 3-methyl-N-phenylphthalimide was obtained quantitatively by C−O bond cleavage followed by aromatization. Other Lewis acids and Brönsted acids were not effective for the reverse DA reaction. Of these catalysts, TFA was the best for accelerating the reversible DA reaction. Figure 4 shows the time course of the retro-DA reaction in which a 0.02 M CDCl3 solution of the exo-adduct was heated at 60 °C with or without TFA (0.2 equiv). Additionally, it was confirmed that TFA did not promote other side reactions. Next, we examined the asymmetric DA reaction involving dynamic enantioselective crystallization. When a heptane solution of N-phenylmaleimide, 2-methylfuran (15.0 equiv), and TFA (0−1.0 equiv) was stirred in the presence of glass beads (Φ = 0.2 mm) at 80 °C in a sealed tube, the crystalline exoadduct appeared quickly. The solution was kept in suspension by stirring for several days at the same temperature, and the change in ee value of exo-3 by attrition-enhanced deracemization was monitored by HPLC using a chiral column, CIRALPAK IA (Daicel Ind.). When we stirred achiral N-phenylmaleimide and 2-methylfuran without TFA in a heptane solution, the induction period until

The DA reaction is well-known to be a concerted electrocycloaddition reaction and to involve a reversible process. We applied this reaction system to the asymmetric synthesis involving dynamic stereoselective crystallization as shown in Scheme 2. The primary DA reaction provides asymmetric chiral materials from prochiral starting substrates, and the racemization of (±)-3 might be promoted by a reversible process. If dynamic stereoselective crystallization by the property of conglomerate crystals is promoted, deracemization will occur, and optically active crystals should be obtained. While the DA reaction generally proceeds endo-selectively, the exo-adduct is more thermodynamically stable. We monitored the reaction course by NMR spectroscopy. Figure 3 shows the reaction course in which 0.02 M of N-phenylmaleimide and 10.0 equiv of 2-methylfuran was reacted at 60 °C in CDCl3. At the early stage of the reaction, both exo- and endo-adducts formed equally; however, after half an hour the endo-adduct gradually decreased, and the ratio of exo to endo reached 8:1 after 6 h. When the reaction was performed on a preparative scale and concentration, the crystalline exo-adduct precipitated and was removed from the reaction system. Finally, the products converged to the exo-adduct, which was isolated in 90% yield accompanied by a trace amount of endo-adduct. These results indicated that the conglomerate crystal of the exo-adduct was formed stereoselectively in this DA reaction system. Another important requirement to apply this asymmetric DA reaction using dynamic stereoselective crystallization is the rate 9301

DOI: 10.1021/acs.joc.8b01273 J. Org. Chem. 2018, 83, 9300−9304

Article

The Journal of Organic Chemistry

However, the maximum ee value of the solid obtained from the reaction in hexane was slightly lower than that in heptane (86% vs 90%). The solubility of the DA adduct may play an important role in the deracemization with its repeated dissolution and crystallization processes. Hexane had a slightly higher solubility for the DA adduct than heptane at 80 °C in a sealed tube, resulting in the shorter deracemization time and lower ee value of the solid. Chemical yields of the solid in both solvents were 80% after cooling to room temperature. The optical purity of the solid formed by attrition-enhanced deracemization at 80 °C may be much higher than 90% ee. The reaction vessel, a sealed tube, was cooled to room temperature before it was opened because the boiling point of 2-methylfuran was 63 °C. Cooling the reaction mixture introduced crystallization of the racemic DA adduct dissolved in the mother liquor, thus lowering the ee value of the crystals and explaining the capped ee values. However, a chance of a cocrystal or other factor could not be excluded. We were unable to control the handedness of the chirality in the spontaneous crystallization; solids of both types of handedness were obtained in each experiment in almost the same ratio. Primary nucleation gave one small crystal of one enantiomer (conglomerate behavior), followed by rapid propagation due to secondary nucleation of the enantiomer. However, we were able to control the handedness by a seeding method, in which a small amount of crystals was added during the induction period. Furthermore, when we started the attrition-enhanced deracemization from a DA adduct with low ee (5% ee) in heptane with TFA (0.2 equiv) at 80 °C, deracemization started immediately, and the same handedness of the enantiomer as the slightly major stereoisomer could be obtained in 90% ee after 8 days. The use of hexane as the solvent shortened the time for deracemization. We could achieve deracemization up to 86% ee from 5% ee of the starting enantiomer after only 4 days under the same conditions.

Figure 4. Time-dependent course of retro-Diels−Alder reaction. The reaction was applied to the exo-adduct (0.02 M) in CDCl3 solution at 60 °C. Red squares = exo-3; black circles = 2-methylfuran 1; blue triangles = 2. (a) Without TFA and (b) in the presence of TFA (0.2 equiv).

deracemization started was several days. Deracemization began gradually after 15 days, but once initiated, the ee value sharply increased up to 90% ee as shown in Figure 5, line 1. The



CONCLUSIONS The first example of asymmetric DA reaction was developed involving stereoselective dynamic crystallization via a reversible process. This second-order asymmetric reaction process can be applied to many types of reversible reactions or reactions combined with a racemization process for conglomerate crystals.

Figure 5. Asymmetric Diels−Alder reaction involving attritionenhanced deracemization using glass beads applied to achiral N-phenylmaleimide (100 mg, 0.578 mmol) and 2-methylfuran (710 mg, 8.67 mmol) in hexane or heptane (1.00 mL) at 80 °C with or without TFA (0−1.0 equiv).



EXPERIMENTAL SECTION

General Information. NMR spectra were recorded in CDCl3 solutions on a Bruker 300 spectrometer operating at 300 and 75 MHz, respectively, for 1H and 13C NMR. Chemical shifts are reported in parts per million (ppm) relative to TMS as an internal standard. IR spectra were recorded on a JASCO FT/IR-230 spectrometer. HPLC analyses were performed on a JASCO HPLC system (JASCO PU-1580 pump, DG-1580-53, LG-2080-02, MD-2015, UV-2075, and CD-2095 detector). X-ray single crystallographic analysis was conducted using a SMART APEX II (Bruker AXS) and APEX II ULTRA (Bruker AXS). Commercially available N-phenylmaleimide and 2-methylfuran were used without further purification. Reaction Conditions for Asymmetric DA Reaction Involving Dynamic Crystallization. In a sealed tube (L = 200 mm, Φ = 25 mm), N-phenylmaleimide (100 mg, 0.578 mmol), 2-methylfuran (710 mg, 8.67 mmol), TFA (0−1.0 equiv), and hexane or heptane (1.0 mL) were stirred with or without glass beads (250 mg) using a stir bar at 80 °C. The crystalline exo-adduct 3 appeared after a few minutes, and the solution was kept in suspension by stirring at 600 rpm for several days at the same temperature. The change of ee value of crystalline 3 was monitored by HPLC using a chiral column, CHIRALPAK IA

attrition of crystals using glass beads was quite effective for deracemization since deracemization was not observed under suspension without glass beads even after 30 days. Furthermore, the induction period could be reduced by the use of TFA as an acid catalyst (lines 2−4). TFA accelerated not only the DA cycloaddition reaction but also the retro-DA reaction without promoting other side reactions. As the amount of TFA increased, there was a tendency toward both a reduction in the induction period and an acceleration in the deracemization, with deracemization starting at the fourth day leading to 90% ee after 11 days (line 4). When using hexane instead of heptane, deracemization proceeded much faster (lines 5−8). Deracemization started after 10 days and reached 86% ee after 19 days without TFA (line 5). TFA influenced both the reduction of induction and deracemization periods, affording an 86% ee of exo-3 after only 5 days by the use of 1.0 equiv of TFA (line 8). 9302

DOI: 10.1021/acs.joc.8b01273 J. Org. Chem. 2018, 83, 9300−9304

The Journal of Organic Chemistry (Daicel Ind.): eluent: n-hexane/EtOH = 80/20 (v/v). Finally, crystalline exo-3 was obtained in 80% (118 mg, 0.46 mmol) chemical yield by filtration. The structures of both exo-3 and endo-4 were confirmed by comparison with the reported spectral data.24 (2aS*,1R*,7S*,6aR*)-1-Methyl-4-phenyl-10-oxa-4-azatricyclo[5.2.1.02,6]dec-8-ene-3,5-dione (exo-3). Colorless prism, Mp: 144− 146 °C; 1H NMR: δ 1.79 (s, 3 H, CH3), 2.86 (d, J = 6.5 Hz, 1 H, 2-H), 3.12 (d, 1 H, J = 6.5 Hz, 6-H), 5.30 (d, J = 1.8 Hz, 1 H, 7-H), 6.37 (d, J = 5.6 Hz, 1 H, 9-H), 6.56 (dd, J = 1.8 and 5.6 Hz, 1 H, 8-H); 7.26−7.29 (m, 2H, Ph), 7.37−7.48 (m, 3H, Ph) 13C{1H} NMR: δ 15.7 (CH3), 49.6, 50.6 (C-2, C-6), 81.1 (C-7), 88.5 (C-1), 126.5, 128.6, 129.0, 131.7, 137.0, 140.7, 174.0, 175.3. (2aS*,1S*,7R*,6aR*)-1-Methyl-4-phenyl-10-oxa-4-azatricyclo[5.2.1.02,6]dec-8-ene-3,5-dione (endo-4). Colorless prism, Mp: 115−116 °C; 1H NMR: δ 1.88, (s, 3 H, CH3), 3.27 (d, J = 7.7 Hz, 1 H, 2-H), 3.81 (dd, J = 5.5 and 7.7 Hz, 1 H, 6-H), 5.43 (dd, J = 1.6 and 5.5 Hz, 1 H, 7-H), 6.38 (d, J = 5.7 Hz, 1 H, 9-H), 6.54 (dd, J = 1.6 and 5.7 Hz, 1 H, 8-H); 7.11−7.13 (m, 2H, Ph), 7.35−7.46 (m, 3H, Ph) 13C{1H} NMR: δ 18.4, 48.5, 50.5, 79.4, 88.9, 126.3, 128.7, 129.1, 131.4, 135.1, 137.7, 174.0, 174.1. Single-Crystal X-ray Crystallographic Analysis of (2aS,1R,7S,6aR)1-Methyl-4-phenyl-10-oxa-4-azatricyclo[5.2.1.02,6]dec-8-ene-3,5dione ((−)-exo-3). Colorless prism (0.20 × 0.10 × 0.10 mm3), orthorhombic space group P212121, a = 6.7516(3) Å, b = 11.9488(5) Å, c = 15.6306(7) Å, V = 1260.98(10) Å3, Z = 4, λ (Cu Kα) = 1.54178 Å, ρ = 1.345 g/cm3, μ (CuKα) = 0.775 cm, 24869 reflections measured (T = 173 K, 4.658° < θ < 68.227°), nb of independent data collected: 2303; nb of independent data used for refinement: 2302 in the final least-squares refinement cycles on F2; the model converged at R1 = 0.0463, wR2 = 0.1263 [I > 2s(I)], R1 = 0.0464, wR2 = 0.1263 (all data), and GOF = 1.114, H-atom parameters constrained, absolute Flack parameter = 0.09(3) (CCDC 1827990). Single Crystal X-ray Crystallographic Analysis of (2aS*,1S*,7R*,6aR*)-1-Methyl-4-phenyl-10-oxa-4-azatricyclo[5.2.1.02,6]dec-8ene-3,5-dione (endo-4). Colorless prism (0.20 × 0.20 × 0.10 mm3), orthorhombic space group Pbca, a = 12.7547(17) Å, b = 11.3387(15) Å, c = 16.800(2) Å, V = 2429.6(6) Å3, Z = 8, λ (Mo Kα) = 0.71073 Å, ρ = 1.396 g/cm3, μ(Mo Kα) = 0.098 cm, 13062 reflections measured (T = 173 K, 2.4246° < θ < 20.9203°), nb of independent data collected: 2758; nb of independent data used for refinement: 1797 in the final least-squares refinement cycles on F2; the model converged at R1 = 0.0450, wR2 = 0.1001 [I > 2s(I)], R1 = 0.0800, wR2 = 0.1121 (all data), and GOF = 0.968, H-atom parameters constrained (CCDC 1827991).





ACKNOWLEDGMENTS



REFERENCES

This work was supported by Grants-in-Aid for Scientific Research (Nos. 25288017, 26410124, and 16H04144) from the Ministry of Education, Culture, Sports, Science, and Technology (MEXT) of the Japanese Government. Mr. Uemura acknowledges financial support from the Frontier Science Program of Graduate School of Science and Engineering, Chiba University.

(1) (a) Diels, O.; Alder, K. Synthesen in der Hydroaromatischen Reihe. Justus Liebigs Ann. Chem. 1928, 460, 98−122. (b) Diels, O.; Alder, K. Synthesen in der Hydroaromatischen Reihe, IV. Mitteilung: Ü ber die Anlagerung von Maleinsäure-anhydrid an Arylierte Diene, Triene und Fulvene. Ber. Dtsch. Chem. Ges. B 1929, 62, 2081−2087. (2) For recent reviews on the application of DA reactions in total synthesis, see: (a) Nicolaou, K. C.; Snyder, S. A.; Montagnon, T.; Vassilikogiannakis, G. The Diels-Alder Reaction in Total Synthesis. Angew. Chem., Int. Ed. 2002, 41, 1668−1698. (b) Takao, K.-I.; Munakata, R.; Tadano, K. Recent advances in natural product synthesis by using intramolecular Diels-Alder reactions. Chem. Rev. 2005, 105, 4779−4807. (3) Examples for asymmetric DA reaction using chiral auxiliaries: (a) Walborsky, H.; Barash, L.; Davis, T. Communications- Partial Asymmetric Syntheses: The Diels-Alder Reaction,. J. Org. Chem. 1961, 26, 4778−4779. (b) Evans, D. A.; Chapman, K. T.; Bisaha, J. Asymmetric Diels-Alder Cycloaddition Reactions with Chiral α,βUnsaturated N-Acyloxazolidinones. J. Am. Chem. Soc. 1988, 110, 1238−1256. (4) Examples for catalytic asymmetric synthesis: (a) Corey, E. J.; Shibata, T.; Lee, T. W. Asymmetric Diels-Alder Reactions Catalyzed by a Triflic Acid Activated Chiral Oxazaborolidine. J. Am. Chem. Soc. 2002, 124, 3808−3809. (b) Choy, W.; Reed, L. A.; Masamune, S. Asymmetric Diels-Alder Reaction: Design of Chiral Dienophiles. J. Org. Chem. 1983, 48, 1137−1139. (c) Oppolzer, W. Asymmetric DielsAlder and Ene Reactions in Organic Synthesis. New Synthetic Methods (48). Angew. Chem., Int. Ed. Engl. 1984, 23, 876−889. (d) Kagan, H. B.; Riant, O. Catalytic Asymmetric Diels-Alder Reactions. Chem. Rev. 1992, 92, 1007−1019. (e) Mehta, G.; Uma, R. Stereoelectronic Control in Diels−Alder Reaction of Dissymmetric 1,3-Dienes. Acc. Chem. Res. 2000, 33, 278−286. (f) Wilson, R. M.; Jen, W. S.; MacMillan, D. W. C. Enantioselective Organocatalytic Intramolecular Diels−Alder Reactions. The Asymmetric Synthesis of Solanapyrone D,. J. Am. Chem. Soc. 2005, 127, 11616−11617. (5) Hachiya, S.; Kasashima, Y.; Yagishita, F.; Mino, T.; Masu, H.; Sakamoto, M. Asymmetric Transformation by Dynamic Crystallization of Achiral Succinimides. Chem. Commun. 2013, 49, 4776−4778. (6) (a) Steendam, R. R. E.; Verkade, J. M. M.; van Benthem, T. J. B.; Meekes, H.; van Enckevort, W. J. P.; Raap, J.; Rutjes, F. P. J. T.; Vlieg, E. Emergence of Single-molecular Chirality from Achiral Reactants. Nat. Commun. 2014, 5, 5543. (b) Kaji, Y.; Uemura, N.; Kasashima, Y.; Ishikawa, H.; Yoshida, Y.; Mino, T.; Sakamoto, M. Asymmetric Synthesis of an Amino Acid Derivative from Achiral Aroyl Acrylamide by Reversible Michael Addition and Preferential Crystallization. Chem. - Eur. J. 2016, 22, 16429−16432. (7) Sakamoto, M.; Shiratsuki, K.; Uemura, N.; Ishikawa, H.; Yoshida, Y.; Kasashima, Y.; Mino, T. Asymmetric Synthesis by Using Natural Sunlight under Absolute Achiral Conditions. Chem. - Eur. J. 2017, 23, 1717−1721. (8) Kawasaki, T.; Takamatsu, N.; Aiba, S.; Tokunaga, Y. Spontaneous Formation and Amplification of an Enantioenriched α-Amino Nitrile: a Chiral Precursor for Strecker Amino Acid Synthesis. Chem. Commun. 2015, 51, 14377−14380. (9) Ishikawa, H.; Uemura, N.; Yagishita, F.; Baba, N.; Yoshida, Y.; Mino, T.; Kasashima, Y.; Sakamoto, M. Asymmetric Synthesis Involving Reversible Photodimerization of a Prochiral Flavonoid Followed by Crystallization. Eur. J. Org. Chem. 2017, 2017, 6878− 6881.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.joc.8b01273. X-ray crystallographic analysis, HPLC analysis, and spectroscopic data for synthesized compounds (PDF) Exo 3 (CIF) Endo 4 (CIF)



Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Yasushi Yoshida: 0000-0002-3498-3696 Takashi Mino: 0000-0003-1588-1202 Masami Sakamoto: 0000-0001-9489-2641 Notes

The authors declare no competing financial interest. 9303

DOI: 10.1021/acs.joc.8b01273 J. Org. Chem. 2018, 83, 9300−9304

Article

The Journal of Organic Chemistry (10) (a) Havinga, E. Spontaneous Formation of Optically Active Substances. Biochim. Biophys. Acta 1954, 13, 171−174. (b) Frank, F. C. On Spontaneous Asymmetric Synthesis. Biochim. Biophys. Acta 1953, 11, 459−463. (11) (a) Jacques, J.; Collet, A.; Wilen, S. H. In Enantiomers, Racemates and Resolution; Krieger: FL, 1994. (b) Yoshioka, R. Racemization, Optical Resolution and Crystallization-induced Asymmetric Transformation of Amino Acids and Pharmaceutical Intermediates. Top. Curr. Chem. 2007, 269, 83−132. (c) Sakamoto, M.; Mino, T. Asymmetric Reaction Using Molecular Chirality Controlled by Spontaneous Crystallization. In Crystallization Processes; Mastai, Y., Ed.; InTech: 2012; pp 59−80. (12) Deracemization via enolate anione or enolate: (a) Reider, P. J.; Davis, P.; Hughes, D. L.; Grabowski, E. J. J. Crystallization-induced Asymmetric Transformation: Stereospecific Synthesis of a Potent Peripheral CCK Antagonist. J. Org. Chem. 1987, 52, 955−957. (b) Noorduin, W. L.; Izumi, T.; Millemaggi, A.; Leeman, M.; Meekes, H.; van Enckevort, W. J. P.; Kellogg, R. M.; Kaptein, B.; Vlieg, E.; Blackmond, D. G. Emergence of a Single Solid Chiral State from a Nearly Racemic Amino Acid Derivative. J. Am. Chem. Soc. 2008, 130, 1158−1159. (c) Noorduin, W. L.; Bode, A. A. C.; van der Meiden, M.; Meekes, H.; van Etteger, A. F.; van Enckevort, W. J. P.; Christianen, P. C. M.; Kaptein, B.; Kellogg, R. M.; Rasing, T.; Vlieg, E. Complete Chiral Symmetry Breaking of an Amino Acid Derivative Directed by Circularly Polarized Light. Nat. Chem. 2009, 1, 729−732. (d) Hein, J. E.; Huynh, C. B.; Viedma, C.; Kellogg, R. M.; Blackmond, D. G. Pasteur’s Tweezers Revisited: on the Mechanism of Attritionenhanced Deracemization and Resolution of Chiral Conglomerate Solids. J. Am. Chem. Soc. 2012, 134, 12629−12636. (e) Sogutoglu, L.C.; Steendam, R R. E.; Meekes, H.; Vlieg, E.; Rutjes, F. P. J. T. Viedma Ripening: a Reliable Crystallisation Method to Reach Single Chirality. Chem. Soc. Rev. 2015, 44, 6723−6732. (13) Deracemization of axially chiral materials: (a) Pincock, R. E.; Perkins, R. R.; Ma, A. S.; Wilson, K. R. Probability Distribution of Enantiomorphous Forms in Spontaneous Generation of Optically Active Substances. Science 1971, 174, 1018−1020. (b) Kondepudi, D. K.; Kaufman, R. J.; Singh, N. Chiral Symmetry Breaking in Sodium Chlorate Crystallization. Science 1990, 250, 975−976. (c) Sakamoto, M.; Mino, T. Total Resolution of Racemates by Dynamic Preferential Crystallization. In Advances in Organic Crystal Chemistry, Comprehensive Reviews 2015; Tamura, R., Miyata, M., Eds.; Springer: 2015; pp 445−462. (14) Attrition-enhanced deracemization: (a) Viedma, C. Chiral Symmetry Breaking During Crystallization: Complete Chiral Purity Induced by Nonlinear Autocatalysis and Recycling. Phys. Rev. Lett. 2005, 94, 065504. (b) Coquerel, G. Crystallization of Molecular Systems from Solution: Phase Diagrams, Supersaturation and Other Basic Concepts. Chem. Soc. Rev. 2014, 43, 2286−2300. (c) Viedma, C.; Ortiz, J. E.; de Torres, T.; Izumi, T.; Blackmond, D. G. Evolution of Solid Phase Homochirality for a Proteinogenic Amino Acid. J. Am. Chem. Soc. 2008, 130, 15274−15275. (15) (15) Deracemization with reversible Mannich-type reaction: Tsogoeva, S. B.; Wei, S.; Freund, M.; Mauksch, M. Angew. Chem., Int. Ed. 2009, 48, 590−594. (16) Avalos, M.; Babiano, R.; Cintas, P.; Jiménez, J. L.; Palacios, J. C.; Barron, L. D. Generation of Highly Enantioenriched Crystalline Products in Reversible Asymmetric Reactions with Racemic or Achiral Catalysts. Chem. Rev. 1998, 98, 2391−2404. (17) Rikken, G. L. J. A.; Raupach, E. Enantioselective Magnetochiral Photochemistry. Nature 2000, 405, 932. (18) Soai, K.; Shibata, T.; Morioka, H.; Choji, K. Asymmetric Autocatalysis and Amplification of Enantiomeric Excess of a Chiral Molecule. Nature 1995, 378, 767−768. (19) Solid-state reaction using chiral crystals. (a) Schmidt, G. M. J. Photodimerization in the Solid State. Pure Appl. Chem. 1971, 27, 647− 678. (b) Green, B. S.; Lahav, M.; Rabinovich, D. Asymmetric Synthesis via Reactions in Chiral Crystals. Acc. Chem. Res. 1979, 12, 191−197. (c) Ramamurthy, V.; Venkatesan, K. Photochemical Reactions of Organic Crystals. Chem. Rev. 1987, 87, 433−481. (d) Scheffer, J. R.;

Garcia-Garibay, M.; Nalamasu, O. The Influence of the Molecular Crystalline Environment on Organic Photorearrangements. In Organic Photochemistry; Padwa, A., Ed.; Marcel Dekker: New York, Basel, 1987; Vol. 8, pp 249−338. (e) Vaida, M.; Popovitz-Biro, R.; Leiserowitz, L.; Lahav, M. Probing Reaction Pathways via Asymmetric Transformations in Chiral and Centrosymmetric Crystals. In Photochemistry in Organized and Constrained Media; Ramamurthy, V., Ed.; VCH: New York, 1991; pp 247−302. (f) Sakamoto, M. Absolute Asymmetric Photochemistry Using Spontaneous Chiral Crystallization. In Chiral Photochemistry; Inoue, Y., Ramamurthy, V., Eds.; Marcel Dekker: New York, 2004; pp 415−461. (g) Sakamoto, M. Spontaneous Chiral Crystallization of Achiral Materials and Absolute Asymmetric Photochemical Transformation Using the Chiral Crystalline Environment. J. Photochem. Photobiol., C 2007, 7, 183−196. (h) Weissbuch, I.; Lahav, M. Crystalline Architectures as Templates of Relevance to the Origins of Homochirality. Chem. Rev. 2011, 111, 3236−3267. (20) Asymmetric synthesis using chiral crystals in homogeneous conditions. (a) Sakamoto, M.; Unosawa, A.; Kobaru, S.; Saito, A.; Mino, T.; Fujita, T. Asymmetric Photocycloaddition in Solution of a Chiral Crystallized Naphthamide. Angew. Chem., Int. Ed. 2005, 44, 5523−5526. (b) Sakamoto, M.; Kato, M.; Aida, Y.; Fujita, K.; Mino, T.; Fujita, T. Photosensitized 2 + 2 Cycloaddition Reaction Using Homochirality Generated by Spontaneous Crystallization. J. Am. Chem. Soc. 2008, 130, 1132−1133. (c) Mai, T. T.; Branca, M.; Gori, D.; Guillot, R.; Kouklovsky, C.; Alezra, V. Absolute Asymmetric Synthesis of Tertiary α-Amino Acids. Angew. Chem., Int. Ed. 2012, 51, 4981−4984. (21) (a) Addadi, L.; Lahav, M. In Origin of Optical Activity in Nature; Walker, D. C., Ed.; Elsevier: New York, Basel, 1979. (b) Mason, S. F. Origins of Biomolecular Handedness. Nature 1984, 311, 19−23. (c) Elias, W. E. The Natural Origin of Optically Active Compounds. J. Chem. Educ. 1972, 49, 448−454. (d) Bonner, W. A. The Origin and Amplification of Biomolecular Chirality. Origins Life Evol. Biospheres 1991, 21, 59. (e) Salam, A. The Role of Chirality in the Origin of Life. J. Mol. Evol. 1991, 33, 105−113. (f) Bonner, W. A. Chirality of Life, Origins Life Evol. Origins Life Evol. Biospheres 1995, 25, 175−190. (g) Feringa, B. L.; van Delden, R. Absolute Asymmetric Synthesis: The Origin, Control, and Amplification of Chirality. Angew. Chem., Int. Ed. 1999, 38, 3418−3438. (22) (22) Recent review for retro-Diels−Alder reactions: Kotha, S.; Banerjee, S. Recent Developments in the Retro-Diels−Alder Reaction. RSC Adv. 2013, 3, 7642−7666. (23) These products were previously obtained in low yield (14% as a mixture of two isomers in ca. 2:1 ratio) and were not fully characterized: Tsuchiya, T.; Arai, H.; Igeta, H. Photochemistry-IX: Formation of Cyclopropenyl Ketones and Furans from Pyridazine NOxides by Irradiation. Tetrahedron 1973, 29, 2747−2751. (24) This reaction was performed under high pressure and gave a 1.6:1.0 mixture of exo:endo-adducts. See: Jarosz, S.; Mach, M.; Szewczyk, K.; Skóra, S.; Ciunik, Z. Synthesis of Sugar-derived 2’and 3′-Substituted Furans and Their Application in Diels-Alder Reactions. Eur. J. Org. Chem. 2001, 2001, 2955−2964.

9304

DOI: 10.1021/acs.joc.8b01273 J. Org. Chem. 2018, 83, 9300−9304