Diastereoselective and Enantioselective Synthesis of Barbiturate

Jul 18, 2018 - Under the catalysis of chiral palladium(0)/ligand complex, the [4 + 2] .... from 64% to 70% and the enantioselectivities from 93% to 96...
1 downloads 0 Views 1MB Size
Article pubs.acs.org/joc

Cite This: J. Org. Chem. 2018, 83, 9291−9299

Diastereoselective and Enantioselective Synthesis of Barbiturate-Fused Spirotetrahydroquinolines via Chiral Palladium(0)/Ligand Complex Catalyzed [4 + 2] Cycloaddition of Vinyl Benzoxazinanones with Barbiturate-Based Olefins

Downloaded via UNIV OF SOUTH DAKOTA on August 17, 2018 at 06:29:30 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Hong-Wu Zhao,* Ning-Ning Feng, Jia-Ming Guo, Juan Du, Wan-Qiu Ding, Li-Ru Wang, and Xiu-Qing Song College of Life Science and Bio-engineering, Beijing University of Technology, No.100 Pingleyuan, Chaoyang District, Beijing 100124, P. R. China S Supporting Information *

ABSTRACT: Under the catalysis of chiral palladium(0)/ ligand complex, the [4 + 2] cycloaddition between vinyl benzoxazinanones and barbiturate-based olefins proceeded readily and provided barbiturate-fused spirotetrahydroquinolines in up to 96% chemical yield with up to >99:1 dr and 97% ee. The absolute configuration of barbiturate-fused spirotetrahydroquinolines was clearly identified by X-ray single crystal structure analysis. The reaction mechanism was proposed to shed light on the enantioselective formation of barbituratefused spirotetrahydroquinolines.



INTRODUCTION

Figure 1. Representative bioactive chiral barbiturate-fused spirotetrahydroquinolines

Over the past years, intrigued by the biological significances of these chemical entities, many efforts have been invested on their racemic4 and enantiomeric3b,5 synthesis. The previous enantioselective synthetic methodologies for chiral barbituratefused spirotetrahydroquinolines mainly involve using chiral substrates and chiral resolutions. However, in this content, the asymmetric catalytic protocols have fully been unexplored to date. So, developing an asymmetric catalytic synthetic methodology is highly urgent to synthesize chiral barbiturate-fused spirotetrahydroquinolines enantioselectively. Being robust and versatile building blocks, vinyl benzoxazinanones participate in the catalytic enantioselective construction of structurally diverse and complex nitrogen-containing chiral heterocycles.6 Upon treatment with palladium(0)/chiral ligand or palladium(0)/chiral N-heterocyclic carbene, vinyl benzoxazinanones decarboxylate into the reactive chiral zwitterionic palladium-stabilized-π-allyl intermediates easily and efficiently.7 Regarding these zwitterionic intermediates in situ formed, they can be readily and enantioselectively captured by structurally diverse reactants such as sulfur ylides, electrondeficient olefins, isatines, and enals in the [4 + 1],8 [4 + 2],9 and [4 + 3]10 cycloadditions, respectively. Surely, these previously disclosed enantioselective catalytic cycloadditions of vinyl benzoxazinanones have functioned as alternative and important tools for the synthesis of optically active nitrogenous heterocycles (Scheme 1). Despite the significant advances in the catalytic enantioselective cycloaddition of vinyl benzoxazinanones,

PUN-2866073 have been determined as clinic antibacterial candidates.

Received: May 17, 2018 Published: July 18, 2018

Chiral barbiturate-fused spirotetrahydroquinolines constitute a class of structurally diverse and complex nitrogen-containing spiroheterocyclic scaffolds and possess biologically and medicinally important properties such as antibacterial and anticancer activities as illustrated in Figure 1.1 Especially, ETX09142 and

© 2018 American Chemical Society

9291

DOI: 10.1021/acs.joc.8b01268 J. Org. Chem. 2018, 83, 9291−9299

Article

The Journal of Organic Chemistry Scheme 1. Representative Enantioselective Catalytic Vinyl Benzoxazinanone-Involved Cycloadditions

Figure 2. Chiral ligands screened in this work.

Table 1. Screening of Chiral Ligandsa

the design and development of novel enantioselective catalytic cycloadditions of vinyl benzoxazinanones remains highly needed for the construction of potentially bioactive chiral barbituratefused spirotetrahydroquinolines. Motivated by the previous works, we designed the novel chiral palladium(0)/ligand complex catalyzed [4 + 2] cycloaddition of vinyl benzoxazinanones with synthetically valuable and versatile barbiturate-based olefins11 for the easily and efficiently enantioselective synthesis of chiral barbiturate-fused spirotetrahydroquinolines. Pleasantly, under the catalysis of the chiral palladium(0)/ligand complex, the [4 + 2] cycloaddition between vinyl benzoxazinanones and barbiturate-based olefins proceeded smoothly and furnished the novel chiral barbituratefused spirotetrahydroquinolines in the reasonable chemical yields with high diastereoselectivities and enantioselectivities. To the best of our knowledge, such a work has not been reported in the literature to date.

entry

ligand

time (h)

yieldb (%)

drc

eec

1 2 3 4 5 6

L1 L2 L3 L4 L5

3 3 3 3 3 3

66 56 57 64 trace 19

>99:1 >99:1 >99:1 >99:1

0 88 50 60

>99:1

80

a

Unless otherwise noted, reactions were carried out with 1a (0.1 mmol), 2a (0.1 mmol), Pd(PPh3)4 (2.5 mol %), and chiral ligands L1−L5 (10 mol %) in CH2Cl2 (1.0 mL) at room temperature. bIsolated chemical yield. cDetermined by chiral HPLC analysis.

the chemical yield and enantioselectivity of 3aa as we expected (entries 2 vs 3). In the case of ligand L3, it contained the opposite axial chirality to that of ligand L2, and delivered 3aa in the comparable chemical yield and enantioselectivity to those obtained with ligand L2 (entries 3 vs 4). Badly, using BINAP-based L4 gave product 3aa only in a trace amount after 3 h (entry 5). Delightfully, biphenyl bisphosphine L5 could furnish 3aa in a higher enantioselectivity though the chemical yield of 3aa remained lower (entry 6). In view of the chemical yield and enantioselectivity obtained with these tested axially chiral ligands, L1 proved to be the most suitable for the cycloaddition (entry 2). Then, we explored the effect of the molar ratio of Pd(PPh3)4/ L1 on the [4 + 2] cycloaddition of vinyl benzoxazinanone 1a and barbiturate-based olefin 2a as depicted in Table 2. Basically, the different ratios tended to give 3aa in >99% dr (entries 1−6). By comparison, the chemical yield and enantioselectivity highly depended on the used molar ratios. Using 0.0025:0.01 ratio gave 3aa in 56% chemical yield and 88% ee (entry 1). Meanwhile, the choice of 0.005:0.01 ratio increased the chemical yield but decreased the enantioselctivity (entries 1 vs 2). Delightfully, the 0.005:0.02 ratio delivered 3aa in excellent enantioselectivity even with the moderate chemical yield (entry 3). In the case of the 0.01:0.02 ratio, it decreased both chemical yield and enantioselectivity as compared with the former case



RESULTS AND DISCUSSION Initially, in the presence of Pd(PPh3)4, we checked the reactivity and stereoselectivity of axially chiral BINOL-based ligands L1−L3, BINAP-based L4 and biphenyl bisphosphine L5 (Figure 2) in the [4 + 2] cycloaddition of vinyl benzoxazinanone 1a and barbiturate-based olefin 2a as shown in Table 1. Generally, most of the chiral ligands delivered >99:1 dr in the cycloaddition. By comparison, the chemical yield and enantioselectivity of the cycloaddition significantly changed with the tested chiral ligands (entries 2−4 and 6). Without a chiral ligand, the [4 + 2] cycloaddition furnished product 3aa in 66% chemical yield as a racemate (entry 1). BINOL-based phosphoramidite ligand L1 delivered 3aa in 56% chemical yield with 88% ee (entry 2). Use of ligand L2 bearing the combined axial and central chiralities did not greatly improve 9292

DOI: 10.1021/acs.joc.8b01268 J. Org. Chem. 2018, 83, 9291−9299

Article

The Journal of Organic Chemistry Table 2. Screening of Ratio of Pd(PPh3)4/L1a

entry

Pd(PPh3)4 /L1(mmol/mmol) Pd(PPh3)4/L2

time (h)

yieldb (%)

drc

eec

1 2 3 4 5 6

0.0025/0.01 0.005/0.01 0.005/0.02 0.01/0.02 0.005/0.02 0.005/0

3 3 3 3 17 3

56 68 66 61 59 66

>99:1 >99:1 >99:1 >99:1 >99:1 >99:1

88 78 90 80 93

increased chemical yield and enantioselectivity (entries 1 vs 2). Moreover, as compared with the former two cases, the employment of 3aa enhanced chemical yield, but lowered the enantioselectivity (entries 1−2 vs 3). Disappointedly, the cycloaddition did not occur at all in the presence of Pd(PPh3)2Cl2 (entry 4). With the use of [Pd2(dba)3]·CHCl3 as the palladium catalyst, the cycloaddition proceeded in 76% chemical yield with 96% ee (entry 5). Unfortunately, under the 1:4 ratio of Pd(OAc)2 and L1, increasing the actual loading of Pd(OAc)2 and L1 lowered the chemical yield and enantioselectivity of 3aa (entries 2 vs 6). Finally, we attempted the increased 1:5 ratio of Pd2(dba)3 and L1 in the cycloaddition, and discovered the chemical yield of 3aa decreased dramatically and the ennatioselectivity with it increased on some extents (entries 3 vs 7). Overall, [Pd2(dba)3]·CHCl3 behaved most efficiently with respect to the enantioselectivity achieved in the cycloaddition. Meanwhile, we investigated the solvent effect of a wide range of solvents possessing various palorities on the cycloaddition of vinyl benzoxazinanone 1a and barbiturate-based olefin 2a as outlined in Table 4. Except for the protic MeOH solvent, all of

a

Unless otherwise noted, reactions were carried out with 1a (0.1 mmol), 2a (0.1 mmol), Pd(PPh3)4, and L1 in CH2Cl2 (1.0 mL) at room temperature. bIsolated chemical yield. cDetermined by chiral HPLC analysis.

(entries 3 vs 4). In addition, by using the ratio of 0.005:0.02, we prolonged the reaction time for 17 h, and the chemical yield lowered and enantioselectivity improved a little (entries 3 vs 5). In addition, in the absence of L1, loading Pd(PPh3)4 in 0.005 mmol afforded 3aa in 66% chemical yield (entry 6). Obviously, among all of the tested ratios, as shown in entry 5, the ratio of 0.005:0.02 behaved more efficiently in terms of enantioselectivity in the cycloaddition. Moreover, in the presence of L1, we attempted several structurally different palladium catalysts in the [4 + 2] cycloaddition of vinyl benzoxazinanone 1a with barbiturate-based olefin 2a in CH2Cl2 as outlined in Table 3. With the exception of

Table 4. Screening of Solventsa

Table 3. Screening of Palladium Catalystsa

entry

solvent

time (h)

yieldb (%)

drc

eec

1 2 3 4 5 6 7

CH2Cl2 THF DCE toluene MeCN CHCl3 MeOH

17 17 17 17 17 17 17

76 76 70 64 70 70 nrd

>99:1 >99:1 >99:1 >99:1 >99:1 >99:1

96 93 93 94 96 95

a

entry

palladium catalysts

time (h)

yieldb (%)

drc

eec

1 2 3 4 5 6e 7f

Pd(PPh3)4 Pd(OAc)2 Pd2(dba)3 Pd(PPh3)2Cl2 [Pd2(dba)3]·CHCl3 Pd(OAc)2 Pd2(dba)3

17 17 17 17 17 17 17

59 72 85 nrd 76 58 60

>99:1 >99:1 >99:1

93 96 91

>99:1 >99:1 >99:1

96 92 96

Unless otherwise noted, reactions were carried out with 1a (0.1 mmol), 2a (0.1 mmol), [Pd2(dba)3]·CHCl3 (5 mol %), and L1 (20 mol %) in solvent (1.0 mL) at room temperature. bIsolated chemical yield. c Determined by chiral HPLC analysis. dNo reaction.

the other examined aprotic solvents resulted in excellent diastereoselectivities in the cycloaddition (entries 1−6 vs 7). Using CH2Cl2 as solvent produced 3aa in 76% chemical yield with 96% ee (entry 1). In the case of THF solvent, the chemical yield maintained the same as the previous case, but the enantioselectivity decreased (entries 1 vs 2). With regard to DCE, toluene, MeCN, and CHCl3 as solvents, the chemical yields ranged from 64% to 70% and the enantioselectivities from 93% to 96% (entries 3−6). Badly, in the very polar and protic MeOH solvent, the cycloaddition did not take place at all (entry 7). So, in CH2Cl2 solvent, the cycloaddition performed most efficiently, and furnished product 3aa in the highest chemical yield with excellent diastereoselectivity and enantioselectivity (entry 1). Currently, for the cycloaddition, we determined the optimal reaction conditions as below: 5 mol % [Pd2(dba)3]·CHCl3, 20 mol % L1, 1a/2a/[Pd2(dba)3]· CHCl3/L1 = 1:1:0.05:0.2, CH2Cl2, rt. Finally, under the optimal reaction conditions, we widely broaden the reaction scope by diversifying vinyl benzoxazinanones

a

Unless otherwise noted, reactions were carried out with 1a (0.1 mmol), 2a (0.1 mmol), palladium catalyst (5 mol %), and L1 (20 mol %) in CH2Cl2 (1.0 mL) at room temperature. bIsolated chemical yield. c Determined by chiral HPLC analysis. dNo reaction. ePd(OAc)2 (10 mmol %) and L1 (40 mmol %). fPd2(dba)3 (5 mmol %) and L1 (25 mmol %).

Pd(PPh3)2Cl2, all the other tested palladium catalysts delivered >99% dr in the cycloaddition (entries 1−3 and 5−7). Normally, the used palladium catalysts influenced the chemical yield and enantioselectivity quite differently. In case of Pd(PPh3)4, it provided 3aa in 59% chemical yield with 93% ee (entry 1). By comparison, use of Pd(OAc)2 led to the 9293

DOI: 10.1021/acs.joc.8b01268 J. Org. Chem. 2018, 83, 9291−9299

Article

The Journal of Organic Chemistry Table 5. Extension of Reaction Scopea

a

Unless otherwise noted, reactions were carried out with 1 (0.1 mmol), 2 (0.1 mmol), [Pd2(dba)3]·CHCl3 (5 mol %), and L1 (20 mol %) in CH2Cl2 (1.0 mL) at room temperature. bIsolated chemical yield. cDetermined by chiral HPLC analysis. dNo reaction. eAt 40 °C

a bulky iso-propyl as R4 group and 2m without a substituent at R5 position exhibited much lower reactivities and did not undergo the cycloaddition with 1a at all (entries 12−13). So, the chemical nature of R4 and R5 groups in substrates 2 significantly affected their chemical reactivities in the cycloaddition with 1a. Also, we attempted the [4 + 2] cycloaddition between 1b−1i and 2a as summarized in Table 5 (entries 14−21). In this context, the substrate 1d with an electron-donating methyl as R1 group behaved better than the substrates 1b−1c bearing an

1 and barbiturate-based olefins 2 as shown in Table 5. In the most cases, the diastereoselectivity of the cycloaddition reached >99:1 (entries 1−11, 14−19, and 22−28). In contrast, the chemical yield and enantioselectivity of the cycloaddition noticeably varied with the used substrates 1 and 2. Concerning the cycloaddition with 1a, most substrates 2 bearing an electron-rich or electron-poor aromatic ring as R4 group and a methyl as R5 group proceeded readily, and afforded products 3aa−3ak in 45−90% chemical yields with 88−96% enantioselectivities (entries 1−11). Remarkably, the substrates 2l with 9294

DOI: 10.1021/acs.joc.8b01268 J. Org. Chem. 2018, 83, 9291−9299

Article

The Journal of Organic Chemistry electron-withdrawing halogen group in the cycloaddition and gave higher chemical yield and enantioselectivity (entries 14− 15 vs 16). Moreover, the substitution position of R1 group in substrates 1 also influenced the chemical yield and enantioselectivity of the cycloaddition. For example, the substrates 1e−1f having R1 group at C7 position behaved better than the substrates 1b−1d containing R1 group at C6 position and furnished the higher chemical yields and enantioselectivities (entries 14−16 vs 17−18). Regarding the substrate 1g without any substituent at R2 position, it reacted with 2a to provide 3ga in excellent chemical yield with very low enantioselectivity (entry 19). Clearly, the replacement of Ts group with benzyl group in the substrates 1 rendered the cycloaddition unsuccessful (entries 1 vs 20). Similarly, the substrate 1i with a methyl at R3 position also failed to react with 2a (entry 21). Therefore, the substrates 1 did not tolerate the wide variation of R2 and R3 groups in the cycloaddition. Lastly, we conducted some crossed cycloadditions by utilizing more reactive substrates 1d−1f and 2c or 2f−2g, and the chemical yield ranged from 62−83%, and enantioselectivity changed from 67−97% (entries 22−28). Lastly, we examined the cycloaddition between 1i and 2a at 40 °C and still no reaction took place (entry 29). Moreover, the single crystal of 3ac was obtained, and its absolute configuration was determined to be (C-14S, C-18S) by X-ray analysis as presented in Figure 3.12 On the basis of the determined

Scheme 2. Proposed Reaction Mechanism for the Formation of 3ac

In conclusion, we first have established the novel chiral palladium(0)/ligand complex catalyzed [4 + 2] cycloaddition between vinyl benzoxazinanones and barbiturate-based olefins for the enantioselective synthesis of the potentially bioactive barbiturate-fused spirotetrahydroquinolines. The [4 + 2] cycloaddition proceeded readily and provided barbituratefused spirotetrahydroquinolines in the reasonable chemical yields and enantioselectivities.

Figure 3. X-ray single crystal structure of 3ac (with thermal ellipsoid shown at the 50% probability level).



absolute configuration of 3ac, the absolute configurations of the other desired chiral barbiturate-fused spirotetrahydroquinolines were assigned similarly as shown in Table 5. To account for the enantioselective formation of 3ac, we predicted a reaction mechanism for the [4 + 2] cycloaddition between 1a and 2c as depicted in Scheme 2 according to previous works.9a,b,d Both pathway A and pathway B include a reversible aza-michael addition and an intramolecular cyclization sequence. Catalyzed by in situ formed chiral palladium(0)/L1 complex, 1a decarboxylates into a chiral palladium-stabilized zwitterionic intermediate 4. Then, the intermediate 4 conducts the reversible aza-michael addition with 2c to furnish two zwitterionic intermediates Int-I and Int-II. Regarding the intramolecular cyclization, Int-1 in pathway A is kinetically favored than that of Int-2 in pathway B because of the strong steric repulsion between Ar and Ts groups caused during the intramolecular cyclization of Int-2. Finally, the intramolecular cyclization of the intermediate Int-1 gives cis-3ac as the major product by using the carbon anion on the barbiturate moiety to attack the si face of the palladium-stabilized π-allyl subunit. Commonly, as described in the previous works, the initial reversible aza-michael addition is fast, and the subsequent intramolecular cyclization is slow and works as a rate- and stereochemistrydetermining step for the similar [4 + 2] cycloadditions.7b,9a

EXPERIMENTAL SECTION

General Information. Unless noted otherwise, all reagents were commercially available and used without further purification. All solvents were distilled from the appropriate drying agents immediately before use. Reactions were monitored by TLC carried out on 0.25 mm SDS silica gel coated glass plates (60F254) and compounds were detected with UV light. The melting points of compounds were determined with a melting point instrument. NMR spectra were recorded on 400 MHz instrument and calibrated using tetramethylsilane (TMS) as internal reference. High resolution mass spectra (HRMS) were recorded under electrospray ionization (ESI) conditions on an Orbitrap mass analyzer. Specific optical rotations were measured with a polarimeter. HPLC analysis was performed on Waters equipment using Chiralpak AD (25 cm × 0.46 cm), AD-H (25 cm × 0.46 cm), AS-H (25 cm × 0.46 cm), IC (25 cm × 0.46 cm), and OD-H (25 cm × 0.46 cm) columns. Vinyl benzoxazinanones 1a−1i9a,d,10a and barbiturate-based olefins 2a−2m11b,d were prepared according to literature procedures. General Procedure for Enantioselective Synthesis of Products 3. A mixture of vinyl benzoxazinanone 1 (0.1 mmol, 1.0 equiv), barbiturate-based olefin 2 (0.1 mmol, 1.0 equiv), [Pd2(dba)3]· CHCl3 (0.005 mmol, 5 mol %), and L1 (0.02 mmol, 20 mol %) in 1.0 mL of CH2Cl2 was stirred at room temperature. After the reaction was completed as indicated by TLC plate, the solvent was removed by evaporation and the crude product was purified by flash column 9295

DOI: 10.1021/acs.joc.8b01268 J. Org. Chem. 2018, 83, 9291−9299

Article

The Journal of Organic Chemistry

hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 8.68 min, tR (minor) = 15.20 min. 3af: White solid, yield: 40 mg (74%); mp 203−204 °C; dr > 99:1; ee = 91%; [α]18 D: 134 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.99 (d, J = 8.0 Hz, 1H), 7.50−7.46 (m, 3H), 7.28−7.24 (m, 3H), 7.06 (s, 4H), 6.88 (d, J = 7.2 Hz, 1H), 6.24 (s, 1H), 5.42− 5.36 (m, 1H), 5.22 (d, J = 10.0 Hz, 1H), 4.80 (d, J = 16.8 Hz, 1H), 3.32 (s, 3H), 2.79 (d, J = 10 Hz, 1H), 2.66 (s, 3H), 2.42 (s, 3H), 2.28 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 170.1, 165.6, 150.1, 143.9, 137.6, 136.6, 136.1, 134.6, 133.1, 130.6, 129.2, 128.1, 127.7, 127.4, 126.7, 126.1, 125.6, 123.0, 66.3, 65.8, 50.5, 29.0, 28.0, 21.5, 21.2 ppm; HRMS (ESI) calculated for C30H30N3O5S [M + H]+: 544.1901, found 544.1887; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 16.31 min, tR (major) = 18.86 min. 3ag: White solid, yield: 44 mg (78%); mp 191−192 °C; dr > 99:1; ee = 89%; [α]18 D: 126 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.99 (d, J = 8.0 Hz, 1H), 7.49 (d, J = 7.6 Hz, 3H), 7.28− 7.23 (m, 5H), 7.16 (d, J = 8.0 Hz, 2H), 6.86 (d, J = 7.6 Hz, 1H), 6.32 (s, 1H), 5.44−5.35 (m, 1H), 5.24 (d, J = 9.6 Hz, 1H), 4.78 (d, J = 16.4 Hz, 1H), 3.31 (s, 3H), 2.72 (d, J = 10.0 Hz, 4H), 2.42 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.8, 165.2, 150.0, 144.2, 137.8, 136.6, 134.3, 133.7, 132.4, 130.1, 129.3, 128.7, 128.4, 127.8, 127.7, 127.3, 126.9, 125.6, 123.3, 65.7, 65.2, 51.2, 29.1, 28.1, 21.5 ppm; HRMS (ESI) calculated for C29H27ClN3O5S [M + H]+: 564.1354, found 564.1350; HPLC separation (Chiralpak AD column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 6.69 min, tR (minor) = 9.42 min. 3ah: White solid, yield: 27 mg (45%); mp 175−176 °C; dr > 99:1; ee = 93%; [α]18 D: 134 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 8.03 (d, J = 8.0 Hz, 1H), 7.54−7.50 (m, 5H), 7.37 (d, J = 7.6 Hz, 2H), 7.29−7.25 (m, 3H), 6.86 (d, J = 7.2 Hz, 1H), 6.45 (s, 1H), 5.46−5.37 (m, 1H), 5.25 (d, J = 10.0 Hz, 1H), 4.79 (d, J = 16.8 Hz, 1H), 3.32 (s, 3H), 2.70 (d, J = 11.6 Hz, 4H), 2.42 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.7, 165.0, 150.0, 144.2, 143.4, 136.6, 134.3, 132.2, 129.9, 129.3, 128.5, 127.7, 127.4, 127.0, 126.9, 126.3 (q, J = 270.0 Hz), 125.5, 123.4, 65.9, 65.1, 51.5, 29.1, 28.1, 21.5 ppm; HRMS (ESI) calculated for C30H27F3N3O5S [M + H]+: 598.1618, found 598.1609; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 7.70 min, tR (major) = 8.22 min. 3ai: White solid, yield: 52 mg (84%); mp 213-214 °C; dr > 99:1; ee = 89%; [α]18 D: 166 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.97 (d, J = 8.0 Hz, 1H), 7.51−7.48 (m, 3H), 7.28−7.25 (m, 3H), 6.89 (d, J = 8.0 Hz, 1H), 6.40 (s, 2H), 6.20 (s, 1H), 5.44− 5.35 (m, 1H), 5.23 (d, J = 10.0 Hz, 1H), 4.81 (d, J = 16.8 Hz, 1H), 3.78 (d, J = 8.8 Hz, 9H), 3.32 (s, 3H), 2.81 (d, J = 10.0 Hz, 1H), 2.71 (s, 3H), 2.42 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 170.2, 165.4, 153.1, 150.1, 144.1, 136.5, 134.6, 134.5, 133.2, 130.5, 129.3, 128.1, 127.7, 127.0, 126.9, 125.7, 123.2, 103.4, 66.5, 65.8, 60.8, 55.9, 50.3, 29.0, 28.2, 21.5 ppm; HRMS (ESI) calculated for C32H34N3O8S [M + H]+: 620.2061, found 620.2068; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 28.15 min, tR (minor) = 35.81 min. 3aj: White solid, yield: 44 mg (76%); mp 248−249 °C; dr > 99:1; ee = 90%; [α]18 D: 126 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 8.11 (d, J = 8.0 Hz, 1H), 7.78−7.70 (m, 4H), 7.55 (d, J = 8.0 Hz, 3H), 7.47−7.45 (m, 2H), 7.33−7.26 (m, 4H), 6.91 (d, J = 7.6 Hz, 1H), 6.51 (s, 1H), 5.48−5.39 (m, 1H), 5.25 (d, J = 10.0 Hz, 1H), 4.83 (d, J = 16.8 Hz, 1H), 3.35 (s, 3H), 2.85 (d, J = 10.0 Hz, 1H), 2.55 (s, 3H), 2.43 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 170.1, 165.4, 150.0, 144.0, 136.8, 136.6, 134.6, 133.0, 132.9, 130.4, 129.3, 128.3, 128.3, 128.2, 127.8, 127.6, 127.5, 126.8, 126.3, 126.2, 125.6, 125.6, 124.1, 123.1, 66.3, 66.0, 50.9, 29.1, 28.0, 21.5 ppm; HRMS (ESI) calculated for C33H30N3O5S [M + H]+: 580.1901, found 580.1912; HPLC separation (Chiralpak IC column, solvent: hexane/ ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 17.32 min, tR (minor) = 18.56 min.

chromatography on silica gel (petroleum ether/ethyl acetate = 5:1−6:1) to afford products 3. 3aa: White solid, yield: 40 mg (76%); mp 235−236 °C; dr > 99:1; ee = 96%; [α]18 D: 160 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 8.01 (d, J = 7.6 Hz, 1H), 7.52−7.47 (m, 3H), 7.29−7.24 (m, 5H), 7.21−7.18 (m, 3H), 6.89 (d, J = 7.6 Hz, 1H), 6.31 (s, 1H), 5.46−5.36 (m, 1H), 5.22 (dd, J = 10.0, 1.2 Hz, 1H), 4.80 (d, J = 16.4 Hz, 1H), 3.32 (s, 3H), 2.80 (d, J = 10.0 Hz, 1H), 2.64 (s, 3H), 2.42 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 170.1, 165.5, 150.1, 144.0, 139.1, 136.6, 134.6, 133.2, 130.5, 129.3, 128.6, 128.1, 127.9, 127.7, 127.5, 126.8, 126.2, 125.6 123.1, 66.3, 65.9, 50.6, 29.1, 28.0, 21.5 ppm; HRMS (ESI) calculated for C29H28N3O5S [M + H]+: 530.1744, found 530.1751; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 14.06 min, tR (major) = 16.73 min. 3ab: White solid, yield: 38 mg (68%); mp 223−224 °C; dr > 99:1; ee = 91%; [α]18 D: 240 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 8.06 (d, J = 8.0 Hz, 1H), 7.62−7.56 (m, 3H), 7.48 (t, J = 7.6 Hz, 1H), 7.28−7.22 (m, 3H), 7.17 (t, J = 7.6 Hz, 1H), 6.97−6.93 (m, 1H), 6.88 (d, J = 7.6 Hz, 1H), 6.69 (d, J = 8.0 Hz, 1H), 6.63 (s, 1H), 5.48−5.39 (m, 1H), 5.19 (d, J = 9.6 Hz, 1H), 4.72 (d, J = 16.8 Hz, 1H), 3.64 (s, 3H), 3.34 (s, 3H), 2.81 (d, J = 10.0 Hz, 1H), 2.66 (s, 3H), 2.40 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 171.1, 166.1, 154.2, 150.9, 143.7, 136.0, 135.5, 134.6, 130.5, 129.1, 128.8, 128.6, 128.6, 128.1, 128.0, 127.8, 127.1, 125.6, 122.7, 121.5, 109.4, 65.7, 62.0, 55.1, 51.4, 28.8, 27.8, 21.5 ppm; HRMS (ESI) calculated for C30H30N3O6S [M + H]+: 560.1850, found 560.1847; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 8.68 min, tR (minor) = 15.20 min. 3ac: White solid, yield: 41 mg (68%); mp 250−251 °C; dr > 99:1; ee = 94%; [α]18 D: 188 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.94 (d, J = 7.6 Hz, 1H), 7.58 (d, J = 8.4 Hz, 2H), 7.51− 7.44 (m, 2H), 7.35−7.26(m, 4H), 7.22−7.19 (m, 1H), 7.10−7.06 (m, 1H), 7.00 (d, J = 7.6 Hz, 1H), 6.58 (s, 1H), 5.50−5.40 (m, 1H), 5.24 (dd, J = 10.0, 1.2 Hz, 1H), 4.83 (dd, J = 16.8, 1.2 Hz, 1H), 3.28 (s, 3H), 3.02 (d, J = 10.0 Hz, 1H), 2.58 (s, 3H), 2.43 (s, 3H) ppm; 13 C NMR (100 MHz, CDCl3): δ 169.5, 166.4, 150.0, 144.0, 139.0, 135.7, 134.9, 132.4, 131.4, 130.6, 129.6, 129.3, 127.8, 127.7, 127.5, 127.2, 125.8, 123.6, 120.8, 65.3, 65.1, 49.8, 29.2, 28.0, 21.6 ppm; HRMS (ESI) calculated for C29H27BrN3O5S [M + H]+: 608.0849, found 608.0857; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 12.36 min, tR (major) = 13.69 min. 3ad: White solid, yield: 49 mg (90%); mp 220−221 °C; dr > 99:1; ee = 88%; [α]18 D: 142 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 8.00 (d, J = 8.0 Hz, 1H), 7.49 (t, J = 8.0 Hz, 3H), 7.28− 7.24 (m, 3H), 7.13 (t, J = 7.6 Hz, 1H), 7.02−6.94 (m, 3H), 6.89 (d, J = 7.6 Hz, 1H), 6.22 (s, 1H), 5.45−5.36 (m, 1H), 5.22 (dd, J = 10.0, 1.2 Hz, 1H), 4.81 (d, J = 16.8 Hz, 1H), 3.32 (s, 3H), 2.82 (d, J = 10.0 Hz, 1H), 2.63 (s, 3H), 2.42 (s, 3H), 2.28 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 170.1, 165.6, 150.1, 144.0, 138.9, 138.2, 136.5, 134.6, 133.3, 130.7, 129.2, 128.7, 128.4, 128.0, 127.7, 127.4, 126.9, 126.7, 125.6, 123.3, 123.0, 66.6, 65.8, 50.3, 29.0, 28.0, 21.5, 21.4 ppm; HRMS (ESI) calculated for C30H30N3O5S [M + H]+: 544.1901, found 544.1899; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 11.72 min, tR (major) = 12.96 min. 3ae: White solid, yield: 45 mg (81%); mp 130−131 °C; dr > 99:1; ee = 92%; [α]18 D: 130 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.98 (d, J = 8.0 Hz, 1H), 7.50−7.46 (m, 3H), 7.28−7.24 (m, 3H), 7.11 (d, J = 8.4 Hz, 2H), 6.88 (d, J = 7.6 Hz, 1H), 6.78 (d, J = 8.8 Hz, 2H), 6.23 (s, 1H), 5.45−5.35 (m, 1H), 5.22 (d, J = 10 Hz, 1H), 4.79 (d, J = 16.8 Hz, 1H), 3.76 (s, 3H), 3.31 (s, 3H), 2.79 (d, J = 10.0 Hz, 1H), 2.68 (s, 3H), 2.41 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 170.1, 165.6, 159.1, 150.1, 144.0, 136.6, 134.6, 133.1, 131.1, 130.6, 129.2, 128.1, 127.7, 127.5, 127.4, 126.7, 125.6, 123.0, 113.9, 66.0, 65.9, 55.2, 50.5, 29.1, 28.1, 21.5 ppm; HRMS (ESI) calculated for C30H30N3O6S [M + H]+: 560.1850, found 560.1854; HPLC separation (Chiralpak AD column, solvent: 9296

DOI: 10.1021/acs.joc.8b01268 J. Org. Chem. 2018, 83, 9291−9299

Article

The Journal of Organic Chemistry

3fa: White solid, yield: 45 mg (82%); mp 97−98 °C; dr > 99:1; ee = 96%; [α]18 D: 148 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.79−7.77 (m, 1H), 7.55 (d, J = 7.6 Hz, 2H), 7.28−7.16 (m, 7H), 6.97−6.94 (m, 1H), 6.85−6.81 (m, 1H), 6.28 (s, 1H), 5.42−5.33 (m, 1H), 5.24 (d, J = 10.0 Hz, 1H), 4.84 (d, J = 16.8 Hz, 1H), 3.32 (s, 3H), 2.80 (d, J = 10.0 Hz, 1H), 2.65 (s, 3H), 2.43 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.9, 165.5, 162.1 (d, J = 245.0 Hz), 149.9, 144.3, 138.8, 137.9 (d, J = 10.0 Hz), 134.5, 130.2, 129.4, 128.7 (d, J = 3.0 Hz), 128.6, 128.1, 127.7, 126.5 (d, J = 9.0 Hz), 126.1, 123.4, 114.9 (d, J = 25.0 Hz), 113.5 (d, J = 25.0 Hz), 66.4, 65.6, 50.2, 29.1, 28.0, 21.6 ppm; HRMS (ESI) calculated for C29H27FN3O5S [M + H]+: 548.1650, found 548.1650; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 9.31 min, tR (major) = 10.18 min. 3ga: White solid, yield: 36 mg (96%); mp 169−170 °C; dr > 99:1; ee = 35%; [α]18 D: 28 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.37−7.31 (m, 3H), 7.21−7.12 (m, 4H), 6.84−6.81 (m, 1H), 6.71 (d, J = 7.6 Hz, 1H), 5.70−5.61 (m, 1H), 5.39−5.29 (m, 2H), 4.81 (s, 1H), 4.56 (d, J = 9.6 Hz, 1H), 4.44 (s, 1H), 3.08 (s, 3H), 3.03 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.5, 165.1, 150.2, 142.4, 136.4, 135.4, 129.8, 128.8, 127.7, 127.3, 126.6, 122.0, 121.7, 118.6, 114.3, 64.1, 58.4, 48.9, 28.4, 27.8 ppm; HRMS (ESI) calculated for C22H22N3O3 [M + H]+: 376.1656, found 376.1656; HPLC separation (Chiralpak OD-H column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 7.36 min, tR (major) = 11.35 min. 3df: White solid, yield: 35 mg (63%); mp 194−195 °C; dr > 99:1; ee = 67%; [α]18 D: 118 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.86 (d, J = 8.0 Hz, 1H), 7.51 (d, J = 8.4 Hz, 2H), 7.28− 7.24 (m, 3H), 7.05 (s, 4H), 6.67 (s, 1H), 6.22 (s, 1H), 5.42−5.35 (m, 1H), 5.21 (dd, J = 10.0, 1.2 Hz, 1H), 4.78 (d, J = 16.8 Hz, 1H), 3.31 (s, 3H), 2.76 (d, J = 10.0 Hz, 1H), 2.66 (s, 3H), 2.42 (s, 3H), 2.36 (s, 3H), 2.27 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 170.2, 165.6, 150.2, 143.8, 137.5, 136.5, 136.2, 134.7, 133.9, 132.8, 130.8, 129.2, 129.2, 128.7, 127.7, 127.2, 126.3, 126.1, 122.8, 66.2, 65.9, 50.5, 29.0, 28.0, 21.5, 21.4, 21.0 ppm; HRMS (ESI) calculated for C31H32N3O5S [M + H]+: 558.2057, found 558.2054; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 17.67 min, tR (minor) = 19.94 min. 3dg: White solid, yield: 38 mg (66%); mp 111−112 °C; dr > 99:1; ee = 76%; [α]18 D: 108 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.86 (d, J = 8.4 Hz, 1H), 7.50 (d, J = 8.4 Hz, 2H), 7.30− 7.22 (m, 5H), 7.15 (d, J = 8.4 Hz, 2H), 6.65 (s, 1H), 6.30 (s, 1H), 5.43−5.34 (m, 1H), 5.22 (d, J = 10.0 Hz, 1H), 4.76 (d, J = 16.4 Hz, 1H), 3.30 (s, 3H), 2.74 (s, 3H), 2.66 (d, J = 10.0 Hz, 1H), 2.42 (s, 3H), 2.36 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.9, 165.1, 150.0, 144.1, 138.0, 136.7, 134.3, 133.9, 133.6, 132.1, 130.2, 129.3, 129.0, 128.6, 127.8, 127.7, 127.1, 126.3, 123.2, 65.8, 65.1, 51.2, 29.1, 28.1, 21.5, 21.4 ppm; HRMS (ESI) calculated for C30H29ClN3O5S [M + H]+: 578.1511, found 578.1490; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 11.57 min, tR (minor) = 13.98 min. 3dc: White solid, yield: 38.5 mg (62%); mp 100−101 °C; dr > 99:1; ee = 83%; [α]18 D: 146 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.81 (d, J = 8.0 Hz, 1H), 7.59 (d, J = 8.0 Hz, 2H), 7.44 (d, J = 8.0 Hz, 1H), 7.33 (d, J = 7.2 Hz, 1H) 7.29−7.26(m, 3H), 7.22−7.18 (m, 1H), 7.08−7.05 (m, 1H), 6.79 (s, 1H), 6.55 (s, 1H), 5.48−5.39 (m, 1H), 5.22 (d, J = 10.4 Hz, 1H), 4.80 (d, J = 16.8 Hz, 1H), 3.27 (s, 3H), 2.97 (d, J = 10.0 Hz, 1H), 2.59 (s, 3H), 2.41 (d, J = 12.8 Hz, 6H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.6, 166.4, 150.1, 143.9, 139.1, 137.0, 135.0, 134.5, 133.0, 132.4, 131.4, 130.8, 129.6, 129.2, 128.5, 127.8, 127.5, 127.4, 126.5, 123.4, 120.8, 65.3, 65.0, 49.9, 29.2, 28.0, 21.6, 21.5 ppm; HRMS (ESI) calculated for C30H29BrN3O5S [M + H]+: 622.1006, found 622.1010; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 12.67 min, tR (minor) = 14.34 min.

3ak: White solid, yield: 46 mg (89%); mp 166−167 °C; dr > 99:1; ee = 90%; [α]18 D: 120 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.88 (d, J = 7.6 Hz, 1H), 7.51−7.42 (m, 3H), 7.28−7.22 (m, 3H), 7.13 (s, 1H), 6.83 (d, J = 7.6 Hz, 1H), 6.55 (s, 1H), 6.38 (s, 1H), 6.30 (s, 1H), 5.48−5.39 (m, 1H), 5.23 (d, J = 10.0 Hz, 1H), 4.71 (d, J = 16.8 Hz, 1H), 3.34 (s, 3H), 2.92 (s, 3H), 2.56 (d, J = 10.0 Hz, 1H), 2.41 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.7, 165.3, 152.4, 150.7, 144.2, 141.6, 135.7, 134.4, 132.7, 129.8, 129.3, 128.3, 127.9, 127.8, 127.1, 125.4, 123.1, 111.3, 108.4, 64.0, 61.2, 50.9, 29.1, 28.3, 21.5 ppm; HRMS (ESI) calculated for C27H26N3O6S [M + H]+: 520.1537, found 520.1536; HPLC separation (Chiralpak AS-H column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 7.77 min, tR (major) = 9.88 min. 3ba: White solid, yield: 40 mg (66%); mp 195−196 °C; dr > 99:1; ee = 58%; [α]18 D: 94 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.87 (d, J = 8.8 Hz, 1H), 7.60 (dd, J = 8.4, 1.6 Hz, 1H), 7.54 (d, J = 8.0 Hz, 2H), 7.30−7.21 (m, 5H), 7.14 (d, J = 6.8 Hz, 2H), 7.02 (s, 1H), 6.22 (s, 1H), 5.40−5.31 (m, 1H), 5.26 (dd, J = 10.0, 1.2 Hz 1H),4.85−4.80 (m, 1H), 3.32 (s, 3H), 2.81 (d, J = 9.6 Hz, 1H), 2.64 (s, 3H), 2.44 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.8, 165.6, 149.9, 144.3, 138.7, 135.7, 135.4, 134.4, 131.1, 129.8, 129.5, 128.8, 128.7, 128.6, 128.2, 127.7, 126.1, 123.9, 120.5, 66.5, 65.6, 50.0, 29.1, 28.1, 21.5 ppm; HRMS (ESI) calculated for C29H27BrN3O5S [M + H]+: 608.0849, found 608.0833; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 10.24 min, tR (minor) = 11.45 min. 3ca: White solid, yield: 37 mg (66%); mp 174−175 °C; dr > 99:1; ee = 66%; [α]18 D: 84 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.93 (d, J = 8.4 Hz, 1H), 7.53 (d, J = 8.0 Hz, 2H), 7.45 (d, J = 7.2 Hz, 1H), 7.30−7.21 (m, 5H), 7.14 (d, J = 6.8 Hz, 2H), 6.87 (s, 1H), 6.23 (s, 1H), 5.40−5.31 (m, 1H) 5.26 (d, J = 9.6 Hz, 1H), 4.82 (d, J = 16.4 Hz 1H), 3.32 (s, 3H), 2.78 (d, J = 9.6 Hz, 1H), 2.64 (s, 3H), 2.44 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.9, 165.6, 149.9, 144.3, 138.7, 135.2, 135.1, 134.4, 132.5, 129.8, 129.5, 128.6, 128.5, 128.2, 128.1, 127.7, 126.1, 125.9, 123.9, 66.5, 65.5, 50.0, 29.1, 28.1, 21.6 ppm; HRMS (ESI) calculated for C29H27ClN3O5S [M + H]+: 564.1354, found 564.1342; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 10.36 min, tR (minor) = 11.31 min. 3da: White solid, yield: 40 mg (74%); mp 176−177 °C; dr > 99:1; ee = 80%; [α]18 D: 116 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.87 (d, J = 8.0 Hz, 1H), 7.52 (d, J = 8.0 Hz, 2H), 7.30− 7.20 (m, 5H), 7.19−7.16 (m, 3H), 6.68 (s, 1H), 6.28 (s, 1H), 5.44− 5.35 (m, 1H), 5.21 (d, J = 10.0 Hz, 1H), 4.78 (d, J = 16.8 Hz, 1H), 3.32 (s, 3H), 2.76 (d, J = 10.0 Hz, 1H), 2.64 (s, 3H), 2.42 (s, 3H), 2.36 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 170.2, 165.5, 150.1, 143.9, 139.2, 136.6, 134.7, 133.9, 132.8, 130.7, 129.2, 128.8, 128.5, 127.9, 127.8, 127.2, 126.3, 126.1, 122.9, 66.3, 66.0, 50.6, 29.0, 28.0, 21.5, 21.4 ppm; HRMS (ESI) calculated for C30H30N3O5S [M + H]+: 544.1901, found 544.1905; HPLC separation (Chiralpak AD-H column, solvent: hexane/ethanol = 95/5, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 11.68 min, tR (major) = 13.13 min. 3ea: White solid, yield: 48 mg (85%); mp 218−219 °C; dr > 99:1; ee = 92%; [α]18 D: 110 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 8.02 (s, 1H), 7.55 (d, J = 8.0 Hz, 2H), 7.28−7.21 (m, 6H), 7.15 (d, J = 7.6 Hz, 2H), 6.81 (d, J = 8.0 Hz, 1H), 6.26 (s, 1H), 5.41− 5.32 (m, 1H), 5.24 (dd, J = 10.0, 1.2 Hz, 1H), 4.83 (d, J = 16.8 Hz, 1H), 3.32 (s, 3H), 2.80 (d, J = 9.6 Hz, 1H), 2.64 (s, 3H), 2.42 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.9, 165.6, 149.9, 144.4, 138.7, 137.7, 134.5, 133.6, 131.7, 130.0, 129.4, 128.6, 128.1, 127.7, 127.3, 126.8, 126.5, 126.1, 123.6, 66.4, 65.6, 50.1, 29.1, 28.0, 21.5 ppm; HRMS (ESI) calculated for C29H27ClN3O5S [M + H]+: 564.1354, found 564.1355; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 8.77 min, tR (major) = 9.61 min. 9297

DOI: 10.1021/acs.joc.8b01268 J. Org. Chem. 2018, 83, 9291−9299

The Journal of Organic Chemistry



3ef: White solid, yield: 48 mg (83%); mp 208−209 °C; dr > 99:1; ee = 89%; [α]18 D: 180 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 8.01 (d, J = 1.6 Hz, 1H), 7.54 (d, J = 8.0 Hz, 2H), 7.27 (d, J = 7.6 Hz, 2H), 7.22(dd, J = 8.0, 1.6 Hz, 1H), 7.05 (q, J = 8.0 Hz, 4H), 6.81 (d, J = 8.0 Hz, 1H), 6.19 (s, 1H), 5.41−5.32 (m, 1H), 5.24 (d, J = 10.0 Hz, 1H), 4.83 (d, J = 16.4 Hz, 1H), 3.31 (s, 3H), 2.79 (d, J = 10.0 Hz, 1H), 2.66 (s, 3H), 2.43 (s, 3H), 2.28 (s, 3H) ppm; 13 C NMR (100 MHz, CDCl3): δ 169.9, 165.6, 150.0, 144.3, 137.9, 137.8, 135.6, 134.4, 133.5, 131.7, 130.1, 129.4, 129.3, 127.7, 127.3, 126.7, 126.5, 126.1, 123.5, 66.4, 65.5, 50.0, 29.1, 28.0, 21.6, 21.1 ppm; HRMS (ESI) calculated for C30H29ClN3O5S [M + H]+: 578.1511, found 578.1504; HPLC separation (Chiralpak IC column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 8.76 min, tR (major) = 9.85 min. 3eg: White solid, yield: 40 mg (67%); mp 233−234 °C; dr > 99:1; ee = 94%; [α]18 D: 162 (c = 0.10, CHCl3). 1H NMR (400 MHz, CDCl3): δ 8.01 (d, J = 2.0 Hz 1H), 7.53 (d, J = 8.4 Hz 2H), 7.29− 7.21 (m, 5H), 7.13 (d, J = 8.0 Hz, 2H), 6.79 (d, J = 8.0 Hz, 1H), 6.27 (s, 1H), 5.40−5.31 (m, 1H), 5.25 (dd, J = 10.0, 1.6 Hz, 1H), 4.82 (d, J = 16.4 Hz, 1H), 3.30 (s, 3H), 2.74−2.69 (m, 4H), 2.43 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.6, 165.2, 149.8, 144.5, 137.8, 137.4, 134.2, 133.9, 133.8, 130.9, 129.6, 129.5, 128.8, 127.8, 127.7, 127.2, 126.9, 126.5, 123.8, 65.4, 65.3, 50.7, 29.2, 28.1, 21.6 ppm; HRMS (ESI) calculated for C29H26Cl2N3O5S [M + H]+: 598.0965, found 598.0947; HPLC separation (Chiralpak AD column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 6.28 min, tR (minor) = 8.30 min. 3fg: White solid, yield: 48 mg (83%); mp 127−128 °C; dr > 99:1; ee = 92%; [α]18 D: 146 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.78−7.75 (m, 1H), 7.53 (d, J = 8.0 Hz, 2H), 7.28−7.23 (m, 4H), 7.15 (d, J = 8.4 Hz, 2H), 6.97−6.93 (m, 1H), 6.82−6.79 (m, 1H), 6.30 (s, 1H), 5.41−5.32 (m, 1H), 5.26−5.24 (m, 1H), 4.82 (d, J = 16.4 Hz, 1H), 3.30 (s, 3H), 2.75−2.70 (m, 4H), 2.42 (s, 3H) ppm; 13C NMR (100 MHz, CDCl3): δ 169.6, 165.2, 162.2 (d, J = 245.0 Hz), 149.9, 144.5, 138.0 (d, J = 11.0 Hz), 137.5, 134.2, 133.9, 129.8, 129.5, 128.8, 128.0 (d, J = 3.0 Hz), 127.8, 127.7, 126.6 (d, J = 9.0 Hz), 123.6, 114.9 (d, J = 24.0 Hz), 113.6 (d, J = 21.0 Hz), 65.5, 65.3, 50.8, 29.1, 28.1, 21.6 ppm; HRMS (ESI) calculated for C29H26ClFN3O5S [M + H]+: 582.1260, found 582.1252; HPLC separation (Chiralpak AD column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (major) = 7.31 min, tR (minor) = 8.91 min. 3 fc: White solid, yield: 48 mg (77%); mp 180−181 °C; dr > 99:1; ee = 97%; [α]18 D: 160 (c = 0.10, CHCl3); 1H NMR (400 MHz, CDCl3): δ 7.72 (dd, J = 9.6, 2.4 Hz, 1H), 7.63 (d, J = 8.4 Hz, 2H), 7.45 (d, J = 8.0 Hz, 1H), 7.30−7.21 (m, 4H), 7.11−7.07 (m, 1H), 7.05−7.00 (m, 1H), 6.96−6.92 (m, 1H), 6.55 (s, 1H), 5.45−5.38 (m, 1H), 5.25 (dd, J = 10.0, 1.2 Hz 1H), 4.86 (d, J = 16.8 Hz, 1H), 3.28 (s, 3H), 3.01 (d, J = 10.0 Hz, 1H), 2.60 (s, 3H), 2.43 (s, 3H) ppm; 13 C NMR (100 MHz, CDCl3): δ 169.3, 166.4, 161.9 (d, J = 245.0 Hz), 149.9, 144.4, 138.6, 136.9 (d, J = 11.0 Hz), 134.8, 132.5, 131.3, 130.5 (d, J = 4.0 Hz), 130.4, 129.8, 129.4, 127.8, 127.6, 126.7 (d, J = 8.0 Hz), 123.9, 120.8, 115.0 (d, J = 24.0 Hz), 114.0 (d, J = 21.0 Hz), 65.2, 65.0, 49.5, 29.2, 28.0, 21.6 ppm; HRMS (ESI) calculated for C29H26BrFN3O5S [M + H]+: 626.0755, found 626.0760; HPLC separation (Chiralpak AD column, solvent: hexane/ethanol = 90/10, flow rate = 1.0 mL/min, λ = 254 nm): retention times tR (minor) = 6.22 min, tR (major) = 7.27 min.



Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] ORCID

Hong-Wu Zhao: 0000-0001-7933-4801 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank Beijing Municipal Commission of Education (No. JC015001200902), Beijing Natural Science Foundation (Nos. 7102010, 2122008, and 2172003), Basic Research Foundation of Beijing University of Technology (X4015001201101), Funding Project for Academic Human Resources Development in Institutions of Higher Learning Under the Jurisdiction of Beijing Municipality (No. PHR201008025), and Doctoral Scientific Research Start-up Foundation of Beijing University of Technology (No. 52015001200701) for financial supports.



REFERENCES

(1) For selected examples, see: (a) Guo, J.; Joubran, C.; Luzietti, R. A., Jr.; Basarab, G. S.; Grimm, S. W.; Vishwanathan, K. Absorption, distribution, metabolism and elimination of 14C-ETX0914, a novel inhibitor of bacterial type-II topoisomerases in rodents. Xenobiotica 2017, 47, 31−49. (b) Bhaskarachar, R. K.; Revanasiddappa, V. G.; Hegde, S.; Balakrishna, J. P.; Reddy, S. Y. Design, synthesis and anticancer activity of functionalized spiro-quinolines with barbituric and thiobarbituric acids. Med. Chem. Res. 2015, 24, 3516−3528. (c) Guo, J.; Joubran, C.; Luzietti, R. A., Jr.; Zhou, F.; Basarab, G. S.; Vishwanathan, K. In vitro and in vivo metabolism of 14C-AZ11, a novel inhibitor of bacterial DNA gyrase/type II topoisomerase. Xenobiotica 2015, 45, 158−170. (d) Scott, G. K.; Benz, C. C. WO 2016077632 A2, 2016. (2) For selected examples, see: (a) Alm, R. A.; Lahiri, S. D.; Kutschke, A.; Otterson, L. G.; McLaughlin, R. E.; Whiteaker, J. D.; Lewis, L. A.; Su, X.; Huband, M. D.; Gardner, H.; Mueller, J. P. Characterization of the novel DNA Gyrase Inhibitor AZD0914: Low Resistance Potential and Lack of Cross-Resistance in. Antimicrob. Agents Chemother. 2015, 59, 1478−1486. (b) Basarab, G. S.; Doig, P.; Galullo, V.; Kern, G.; Kimzey, A.; Kutschke, A.; Newman, J. P.; Morningstar, M.; Mueller, J.; Otterson, L.; Vishwanathan, K.; Zhou, F.; Gowravaram, M. Discovery of Novel DNA Gyrase Inhibiting Spiropyrimidinetriones Benzisoxazole Fusion with N-Linked Oxazolidinone Substituents Leading to a Clinical Candidate (ETX0914). J. Med. Chem. 2015, 58, 6264−6282. (c) Jacobsson, S.; Golparian, D.; Alm, R. A.; Huband, M.; Mueller, J.; Jensen, J. S.; Ohnishi, M.; Unemo, M. High In Vitro Activity of the Novel Spiropyrimidinetrione AZD0914, a DNA Gyrase Inhibitor, against Multidrug-Resistant Neisseria gonorrhoeae Isolates Suggests a New Effective Option for Oral Treatment of Gonorrhea. Antimicrob. Agents Chemother. 2014, 58, 5585−5588. (d) Waites, K. B.; Crabb, D. M.; Duffy, L. B.; Huband, M. D. In vitro antibacterial activity of AZD0914 against human Mycoplasmas and Ureaplasmas. Antimicrob. Agents Chemother. 2015, 59, 3627−3629. (e) Huband, M. D.; Bradford, P. A.; Otterson, L. G.; Basarab, G. S.; Kutschke, A. C.; Giacobbe, R. A.; Patey, S. A.; Alm, R. A.; Johnstone, M. R.; Potter, M. E.; Miller, P. F.; Mueller, J. P. In vitro Antibacterial Activity of AZD0914, a New Spiropyrimidinetrione DNA Gyrase/Topoisomerase Inhibitor with Potent Activity against Gram-Positive, Fastidious Gram-Negative, and Atypical Bacteria. Antimicrob. Agents Chemother. 2015, 59, 467−474. (f) Papp, J. R.; Lawrence, K.; Sharpe, S.; Mueller, J.; Kirkcaldy, R. D. In vitro growth of multidrug-resistant Neisseria gonorrhoeae isolates is inhibited by ETX0914, a novel spiropyrimidinetrione. Int. J. Antimicrob. Agents 2016, 48, 328−330. (3) For selected examples, see: (a) Chan, P. F.; Srikannathasan, V.; Huang, J.; Cui, H.; Fosberry, A. P.; Gu, M.; Hann, M. M.; Hibbs, M.;

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.joc.8b01268. Copies of NMR for products 3; X-ray single crystal structure analysis data for 3ac; copies of HPLC spectra (PDF) Supplementary crystallographic data for compound 3ac (CIF) 9298

DOI: 10.1021/acs.joc.8b01268 J. Org. Chem. 2018, 83, 9291−9299

Article

The Journal of Organic Chemistry

Commun. 2017, 53, 10030−10033. (c) Mei, G. J.; Bian, C. Y.; Li, G. H.; Xu, S. L.; Zheng, W. Q.; Shi, F. Catalytic Asymmetric Construction of the Tryptanthrin Skeleton via an Enantioselective Decarboxylative [4 + 2] Cyclization. Org. Lett. 2017, 19, 3219−3222. (d) Leth, L. A.; Glaus, F.; Meazza, M.; Fu, L.; Thogersen, M. K.; Bitsch, E. A.; Jorgensen, K. A. Decarboxylative [4 + 2] Cycloaddition by Synergistic Palladium and Organocatalysis. Angew. Chem., Int. Ed. 2016, 55, 15272−15276. (10) For selected examples, see: (a) Guo, C.; Fleige, M.; JanssenMuller, D.; Daniliuc, C. G.; Glorius, F. Cooperative N-Heterocyclic Carbene/Palladium-Catalyzed Enantioselective Umpolung Annulations. J. Am. Chem. Soc. 2016, 138, 7840−7843. (b) Guo, C.; JanssenMuller, D.; Fleige, M.; Lerchen, A.; Daniliuc, C. G.; Glorius, F. Mechanistic Studies on a Cooperative NHC Organocatalysis/ Palladium Catalysis System: Uncovering Significant Lessons for Mixed Chiral Pd(NHC)(PR3) Catalyst Design. J. Am. Chem. Soc. 2017, 139, 4443−4451. (11) For selected examples, see: (a) Han, B.; Huang, W.; Ren, W.; He, G.; Wang, J.-h.; Peng, C. Asymmetric Synthesis of CyclohexaneFused Drug-Like Spirocyclic Scaffolds Containing Six Contiguous Stereogenic Centers via Organocatalytic Cascade Reactions. Adv. Synth. Catal. 2015, 357, 561−568. (b) Zhao, H. W.; Tian, T.; Li, B.; Yang, Z.; Pang, H. L.; Meng, W.; Song, X. Q.; Chen, X. Q. Diastereoselective Synthesis of Dispirobarbiturates through Et3NCatalyzed [3 + 2] Cycloaddition of Barbiturate-Based Olefins with 3Isothiocyanato Oxindoles. J. Org. Chem. 2015, 80, 10380−10385. (c) Dieskau, A. P.; Holzwarth, M. S.; Plietker, B. Fe-catalyzed Allylic C-C-Bond Activation: Vinylcyclopropanes As Versatile a1,a3,d5Synthons in Traceless Allylic Substitutions and [3 + 2]-Cycloadditions. J. Am. Chem. Soc. 2012, 134, 5048−5051. (d) Zhao, H.-W.; Tian, T.; Pang, H.-L.; Li, B.; Chen, X.-Q.; Yang, Z.; Meng, W.; Song, X.-Q.; Zhao, Y.-D.; Liu, Y.-Y. Organocatalytic [3 + 2] Cycloadditions of Barbiturate-Based Olefins with 3-Isothiocyanato Oxindoles: Highly Diastereoselective and Enantioselective Synthesis of Dispirobarbiturates. Adv. Synth. Catal. 2016, 358, 2619−2630. (e) Soleimani, E.; Yazdani, H.; Saei, P. Synthesis of spiro 3-bromo-4,5-dihydroisoxazoles via [1,3]dipolar cycloaddition reactions. Tetrahedron Lett. 2015, 56, 1635−1637. (f) Girgis, A. S.; Farag, H.; Ismail, N. S.; George, R. F. Synthesis, hypnotic properties and molecular modeling studies of 1,2,7,9-tetraaza-spiro[4.5]dec-2-ene-6,8,10-triones. Eur. J. Med. Chem. 2011, 46, 4964−4969. (12) CCDC 1828461 contains the supplementary crystallographic data for compound 3ac. These data can be obtained free of charge from The Cambridge Crystallographic Data Center via www. ccdc.cam.ac.uk/data_request/cif.

Homes, P.; Ingraham, K.; Pizzollo, J.; Shen, C.; Shillings, A. J.; Spitzfaden, C. E.; Tanner, R.; Theobald, A. J.; Stavenger, R. A.; Bax, B. D.; Gwynn, M. N. Structural basis of DNA gyrase inhibition by antibacterial QPT-1, anticancer drug etoposide and moxifloxacin. Nat. Commun. 2015, 6, 10048−10060. (b) Ruble, J. C.; Hurd, A. R.; Johnson, T. A.; Sherry, D. A.; Barbachyn, M. R.; Toogood, P. L.; Bundy, G. L.; Graber, D. R.; Kamilar, G. M. Synthesis of (−)-PNU286607 by Asymmetric Cyclization of Alkylidene Barbiturates. J. Am. Chem. Soc. 2009, 131, 3991−3997. (4) For selected examples, see: (a) Zhu, S.; Chen, C.; Xiao, M.; Yu, L.; Wang, L.; Xiao, J. Construction of the tetrahydroquinoline spiro skeleton via cascade [1,5]-hydride transfer-involved C(sp3)−H functionalization “on water. Green Chem. 2017, 19, 5653−5658. (b) Jiang, B.; Cao, L.-J.; Tu, S.-J.; Zheng, W.-R.; Yu, H.-Z. Highly Diastereoselective Domino Synthesis of 6-Spirosubstituted Pyrido[2,3-d]pyrimidine Derivatives in Water. J. Comb. Chem. 2009, 11, 612−616. (c) Bhuyan, D.; Sarma, R.; Prajapati, D. Microwave-assisted efficient synthesis of spiroquinoline derivatives via a catalyst- and solvent-free aza-Diels−Alder reaction. Tetrahedron Lett. 2012, 53, 6460−6463. (d) Krasnov, K.; Kartsev, V. Synthesis of Spiroheterocyclic Systems from Barbituric Acids and N,N-Disubstituted oAminobenzaldehydes. Russ. J. Org. Chem. 2005, 41, 901−906. (e) Sarmah, M. M.; Borthakur, S.; Bhuyan, D.; Prajapati, D. Ultrasound mediated efficient synthesis of spironaphthoquinolines. RSC Adv. 2015, 5, 68839−68842. (f) Wang, X.-S.; Zhang, M.-M.; Jiang, H.; Yao, C.-S.; Tu, S.-J. Unexpected Spiro-benzoquinolines in the Reaction of N-(Arylidene)naphthalen-2-amine, Arylaldehyde, and 1,3-Dimethylbarbituric Acid in Water. Chem. Lett. 2007, 36, 450−451. (5) Basarab, G. S.; Brassil, P.; Doig, P.; Galullo, V.; Haimes, H. B.; Kern, G.; Kutschke, A.; McNulty, J.; Schuck, V. J.; Stone, G.; Gowravaram, M. Novel DNA Gyrase Inhibiting Spiropyrimidinetriones with a Benzisoxazole Scaffold: SAR and in Vivo Characterization. J. Med. Chem. 2014, 57, 9078−9095. (6) For a review, see: (a) Wang, C.; Pahadi, N.; Tunge, J. A. Decarboxylative Cyclizations and Cycloadditions of Palladiumpolarized Aza-ortho-Xylylenes. Tetrahedron 2009, 65, 5102−5109. (b) Wang, Q.; Li, T. R.; Lu, L. Q.; Li, M. M.; Zhang, K.; Xiao, W. J. Catalytic Asymmetric [4 + 1] Annulation of Sulfur Ylides with Copper-Allenylidene Intermediates. J. Am. Chem. Soc. 2016, 138, 8360−8363. (c) Wang, Y. N.; Wang, B. C.; Zhang, M. M.; Gao, X. W.; Li, T. R.; Lu, L. Q.; Xiao, W. J. Hydrogen Bond Direction Enables Palladium-Catalyzed Branch- and Enantioselective Allylic Aminations and Beyond. Org. Lett. 2017, 19, 4094−4097. (d) Li, M.-M.; Wei, Y.; Liu, J.; Chen, H.-W.; Lu, L.-Q.; Xiao, W.-J. Sequential Visible-Light Photoactivation and Palladium Catalysis Enabling Enantioselective [4 + 2] Cycloadditions. J. Am. Chem. Soc. 2017, 139, 14707−14713. (7) For selected examples, see: (a) Jin, J. H.; Wang, H.; Yang, Z. T.; Yang, W. L.; Tang, W.; Deng, W. P. Asymmetric Synthesis of 3,4Dihydroquinolin-2-ones via a Stereoselective Palladium-Catalyzed Decarboxylative [4 + 2] Cycloaddition. Org. Lett. 2018, 20, 104−107. (b) Wei, Y.; Lu, L. Q.; Li, T. R.; Feng, B.; Wang, Q.; Xiao, W. J.; Alper, H. P, S Ligands for the Asymmetric Construction of Quaternary Stereocenters in Palladium-Catalyzed Decarboxylative [4 + 2] Cycloadditions. Angew. Chem., Int. Ed. 2016, 55, 2200−2204. (8) For selected examples, see: (a) Li, T. R.; Tan, F.; Lu, L. Q.; Wei, Y.; Wang, Y. N.; Liu, Y. Y.; Yang, Q. Q.; Chen, J. R.; Shi, D. Q.; Xiao, W. J. Asymmetric trapping of zwitterionic intermediates by sulphur ylides in a palladium-catalysed decarboxylation-cycloaddition sequence. Nat. Commun. 2014, 5, 5500−5509. (b) Wang, Q.; Qi, X.; Lu, L. Q.; Li, T. R.; Yuan, Z. G.; Zhang, K.; Li, B. J.; Lan, Y.; Xiao, W. J. Iron-Catalyzed Decarboxylative (4 + 1) Cycloadditions: Exploiting the Reactivity of Ambident Iron-Stabilized Intermediates. Angew. Chem., Int. Ed. 2016, 55, 2840−2844. (9) For selected examples, see: (a) Wang, C.; Tunge, J. A. Asymmetric Cycloadditions of Palladium-Polarized Aza-o-xylylenes. J. Am. Chem. Soc. 2008, 130, 8118−8119. (b) Mei, G. J.; Li, D.; Zhou, G. X.; Shi, Q.; Cao, Z.; Shi, F. A catalytic asymmetric construction of a tetrahydroquinoline-based spirooxindole framework via a diastereoand enantioselective decarboxylative [4 + 2] cycloaddition. Chem. 9299

DOI: 10.1021/acs.joc.8b01268 J. Org. Chem. 2018, 83, 9291−9299