Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Stereoselective synthesis of 1,3-disubstituted dihydroisoquinolines viaL-phenylalanine-derived dihydroisoquinoline N-oxides

Jesús Flores-Ferrándiz , Nicholas Carter , Maria José González-Soria , Malgorzata Wasinska , Daniel Gill , Beatriz Maciá * and Vittorio Caprio *
School of Science and the Environment, Manchester Metropolitan University, All Saints Campus, Oxford Road, Manchester, M15 6BH, UK. E-mail: b.macia-ruiz@mmu.ac.uk; v.caprio@mmu.ac.uk

Received 16th August 2018 , Accepted 10th September 2018

First published on 12th September 2018


Abstract

The preparation of chiral pool-derived nitrone 3 and its use in the protecting-group free, stereoselective synthesis of a range of 1,3-disubstituted tetrahydroisoquinolines is described. Grignard reagent additions to nitrone 3 yielded trans-1,3-disubstituted N-hydroxytetrahydroisoquinolines 6 with good levels of selectivity, while 1,3-dipolar cycloadditions to this nitrone provided access to 3-(2-hydroxyalkyl)isoquinolines 12 as single diastereomers.


Introduction

Tetrahydroisoquinoline (THIQ) systems occur in a wide range of natural products and therapeutic drugs1 and have been highlighted both as privileged scaffolds2 and as useful moieties in fragment-based drug discovery.3 Moreover, there is interest in the use of tetrahydroisoquinolines as ligands in asymmetric catalysis4 and as chiral bases.5 The methods for accessing 1,3-disubstituted THIQ's, can be broadly grouped into two strategies, based on either cyclisations of suitably substituted aromatic precursors6 or functionalisation of isoquinoline-based scaffolds.7 In an effort to develop a strategy towards enantiodefined tetrahydroisoquinolines, of general use, we became interested in an approach of the latter type utilizing a chiral pool-derived isoquinoline unit with a potentially broad reactivity profile. The only previous approaches with these criteria utilize either 1-lithiated derivative 18 or imine 25 – with reactivity restricted to additions to electrophiles or nucleophiles respectively (Fig. 1).
image file: c8ob02007h-f1.tif
Fig. 1 Previous chiral pool-derived THIQ precursors.

We envisaged that nitrone 3 would exhibit much utility in this regard. The core scaffold is readily accessible, in enantioenriched form from L-phenylalanine, and the reactivity of the nitrone group, functionalised via 1,3-dipolar cycloaddition,9 nucleophilic addition10 and reductive coupling with electrophiles,11 confers wide scope to this system (Scheme 1). Furthermore, the hydroxymethyl group functions as a readily removable stereocontrol element.12


image file: c8ob02007h-s1.tif
Scheme 1 Synthetic approaches to 1,3-disubstituted tetrahydroisoquinolines utilising nitrone 3.

Results and discussion

Preparation of nitrone 3

Our approach to nitrone 3 centred on the oxidation of the known amine 5 (Scheme 2), accessed by Pictet–Spengler reaction of L-phenylalanine,13 followed by borane-mediated reduction of the resulting acid.5 Trialing a range of methods for the direct oxidative conversion of amines to nitrones, revealed that the corresponding isoquinoline N-oxide was a significant byproduct. Best results were achieved using Oxone as oxidant, providing dihydroisoquinoline derived nitrone 3 cleanly, in moderate yield (58%), with no evidence of overoxidation to the corresponding isoquinoline N-oxide.14
image file: c8ob02007h-s2.tif
Scheme 2 Synthesis of nitrone 3.

Addition of organometallic reagents to nitrone 3

With nitrone 3 in hand, the nucleophilic addition of organometallic reagents was initially investigated by screening of a range of solvents and temperatures, using methylmagnesium bromide and methyllithium as test reagents (Table 1). In all cases, full conversion was achieved using 3 equivalents of nucleophile, with the 1,3-trans isomer 6a as the major product. THF provided the best diastereomeric ratios and no advantage was conferred by performing the reaction under cryogenic conditions see (entries 7–9). It is noteworthy that, while the addition of the more reactive methyllithium proceeded with full conversion, stereoselectivity was lower than that obtained with the corresponding Grignard reagent (entry 10 vs. entry 7). Consequently, subsequent studies were performed using Grignard species under the conditions outlined in entry 7.
Table 1 Optimization of the addition of organometallic reagents to nitrone 3a

image file: c8ob02007h-u1.tif

Entry RM Solvent T (°C) 6a/7ab Conv.c (%)
a Reaction conditions: 3 (0.1 mmol), RM (3 eq.), T, solvent. b Ratio determined by analysis of 1H-NMR spectra of crude products. Relative stereochemistry determined by 2D-NOESY studies; see ESI for further details. c Determined by 1H-NMR analysis of crude products.
1 MeMgBr CH2Cl2 0 4.5/1 >99
2 MeMgBr PhMe 0 7.1/1 >99
3 MeMgBr Dioxane 0 3.0/1 >99
4 MeMgBr t BuOMe 0 3.5/1 >99
5 MeMgBr Et2O 0 5.5/1 >99
6 MeMgBr THF 20 6.0/1 >99
7 MeMgBr THF 0 7.5/1 >99
8 MeMgBr THF −20 7.5/1 >99
9 MeMgBr THF −40 7.5/1 >99
10 MeLi THF 0 3.8/1 >99


The addition of different Grignard reagents to nitrone 3 provided, in all cases, the trans-isomer 6 as the major product (Table 2). No discernable trends are evident in Table 2, although the best levels of selectivity were obtained with the larger phenylmagnesium bromide (entry 5). Cis/trans stereochemistry of products 6 and 7 was assigned by 2D NOESY analysis (see Fig. 2 for an example). Strong correlations between the protons of the C1-substituent and H3 were observed in the trans-adduct, while the trans-diaxial correlations between H1 and H3 revealed the cis-diequatorial substitution pattern of the minor adduct, as illustrated on the ethyl-adduct 6b/7b.


image file: c8ob02007h-f2.tif
Fig. 2 Selected correlations observed in the 2D-NOESY spectra of tetrahydroisoquinolines 6b and 7b.
Table 2 Addition of organometallic reagents to nitrone 3a

image file: c8ob02007h-u2.tif

Entry RM Products 6[thin space (1/6-em)]:[thin space (1/6-em)]7a,b Yieldc (%)
a Reaction conditions: 3 (0.3 mmol), RMgBr (3 eq.), THF, 0 °C to r.t. b Ratio determined by analysis of crude 1H-NMR spectra. Relative stereochemistry determined by 2D-NOESY studies, see Fig. 2 and ESI for further details. c Isolated combined yield after column chromatography.
1 MeMgBr 6a/7a 7.5/1 75
2 EtMgBr 6b/7b 6.5/1 87
3 i PrMgBr 6c/7c 5.9/1 85
4 AllylMgBr 6d/7d 2.8/1 90
5 PhMgBr 6e/7e 10/1 96


The effect of protection of the 3-hydroxymethyl group, in combination with increasing size of the 3-substituent, was investigated by studying the addition of different Grignard reagents to the TBS-protected nitrone 8 (Table 3), prepared by sodium tungstate-catalysed-oxidation of the known TBS-derivative of hydroxymethyl-THIQ 5.5,15 The addition of both ethyl- and phenylmagnesium bromide to 8 proceeded in favour of the trans-adducts 9, but with much lower selectivities than those obtained with the unprotected nitrone 3 (compare Table 3, entries 1 and 2 with Table 2, entries 2 and 5, respectively).

Table 3 Addition of organometallic reagents to nitrone 8a

image file: c8ob02007h-u3.tif

Entry RM Product 9/10b Yieldc (%)
a Reaction conditions: 8 (0.3 mmol), RMgBr (3 eq.), THF, 0 °C to r.t. b Ratio determined by 1H-NMR on the reaction crude. Relative stereochemistry determined by 2D-NOESY studies. c Isolated combined yield after column chromatography.
1 EtMgBr 9a/10a 2.5/1 56
2 PhMgBr 9b/10b 3/1 53


These results indicate that, in addition to the metal, (see Table 1, entries 7 and 10, for the addition of MeMgBr and MeLi, respectively), the nitrone oxygen and the free hydroxyl group play an important role in determining the mode and level of diastereoselectivity in the addition to 3. This hypothesis is supported by the results obtained with imine 2, where the selectivities of the addition of organolithium species are generally lower than those reported in Table 2 for all but the largest nucleophiles.5 We propose the model depicted in Fig. 3 to account for this selectivity, which involves a cyclic magnesium chelate similar to that hypothesized during additions to N-glycosyl nitrones.16 In our model, the syn addition is hindered by the pseudo-axial proton at C4.


image file: c8ob02007h-f3.tif
Fig. 3 Proposed mode of addition of Grignard reagents to nitrone 3.

Cycloaddition reactions of nitrone 3

Next, the preparation of more functionalized, stereodefined 1,3-THIQ systems was investigated via 1,3-dipolar cycloaddition of 3 with a range of alkenes, followed by reduction of the resulting isoxazolidines 11 (Table 4). Styrene and hex-1-ene were chosen as the prototypical model dipolarophiles (entries 1 and 2), alongside functionalized butenes, selected to access products of subsequent synthetic utility (entries 3 and 4). Optimum yields/selectivity were obtained in toluene under reflux conditions. The stereochemistry of the cycloaddition was readily deduced by analysis of 2D-NOESY spectra of cycloadducts 11 (see an example in Fig. 4), which revealed strong correlations between H2/H6 and H10b/CH2(OH). In all cases, the exo-1,3-trans-isoxazolidine was exclusively obtained in moderate to good yield and no traces of any other isomers were detected in the 1H NMR spectra of crude products. The cycloaddition reactions of nitrone 3 with a range of electron deficient systems, including acrylates, were trialed but, unfortunately, failed to provide the corresponding cycloadducts.
image file: c8ob02007h-f4.tif
Fig. 4 Selected correlations observed in the 2D-NOESY spectrum of 11a.
Table 4 Cycloaddition of nitrone 3 and reduction of cycloadducts 11a

image file: c8ob02007h-u4.tif

Entry R Product 11 (% yield) 12 (% yield)
a Reaction conditions: (i) CH2[double bond, length as m-dash]CH2R (5 eq.), toluene, 110 °C; (ii) AcOH, Zn, reflux. b Isolated yield after flash chromatography. Relative stereochemistry determined by 2D-NOESY studies, see Fig. 4 and ESI for further details.
1 Ph 11a/12a 86 33
2 (CH2)3CH3 11b/12b 74 90
3 (CH2)2OBn 11c/12c 61 51
4 (CH2)2SPh 11d/12d 64 47


The subsequent reductive N–O bond cleavage of isoxazolidine 11, with zinc powder in AcOH, proceeded with no stereochemical degradation,17 providing the corresponding 1,3-THIQ 12a–d in good to moderate yields (Table 2).

Conclusions

In conclusion, we have developed an approach to the synthesis of a range of optically active 1,3-disubstituted tetrahydroisoquinolines from (S)-3-(hydroxymethyl)-3,4-dihydroisoquinoline 2-oxide 3. The addition of Grignard reagents to 3 proceeded with good yields and stereoselectivities, and unprecedented levels of stereoselection were found for the dipolar cycloaddition reaction of alkenes to nitrone 3. This strategy allows the direct assembly of highly functionalized tetrahydroisoquinoline units with up to three stereogenic centres in a simple and effective manner, starting from a readily available chiral substrate.

Experimental

General

1H NMR and 13C NMR have been recorded on a JEOL® ECS-400 (400 and 100.6 MHz, respectively) using CDCl3 as solvent. Chemical shift values are reported in ppm with TMS as internal standard (CDCl3: δ 7.26 for 1H-NMR, δ 77.0 for 13C-NMR). Data are reported as follows: chemical shifts, multiplicity (s = singlet, quint = quintuplet, m = multiplet), coupling constants (Hz), and integration. IR spectra were recorded on Nicolet® 380 FT/IR – Fourier Transform Infrared Spectrometer. Only the most significant frequencies have been considered during the characterization, and have been reported in cm−1. High resolution mass spectra were measured on an Agilent Technologies® 6540 Ultra-High-Definition (UHD) Accurate-Mass equipped with a time of flight (Q-TOF) analyzer and the samples were ionized by ESI techniques and introduced through a high pressure liquid chromatography (HPLC) model Agilent Technologies® 1260 Infinity Quaternary LC system. Optical rotations were measured on a Bellingham + Stanley® ADP 440 + Polarimeter with a 0.5 cm cell (c given in g per 100 mL).

All reactions were monitored by thin-layer chromatography using precoated sheets of silica gel 60, 0.25 mm thick (F254 Merck KGaA®). The components were visualized by UV light (254 nm) and phosphomolybdic acid. Flash column chromatography was done using Geduran® Silica gel 60, 40–63 microns RE. The eluent used is mentioned in each particular case. All glassware employed during inert atmosphere experiments was flame-dried under a stream of dry argon. Alkenes reagents were purchased from Sigma-Aldrich or Fisher Scientific and used without further purification. Grignard and organolithium reagents were purchased from Sigma-Aldrich. Anhydrous THF was obtained from a Pure Solv™ Solvent Purification Systems.

(S)-3-(Hydroxymethyl)-3,4-dihydroisoquinoline-2-oxide (3). NaHCO3 (6.72 g, 80.0 mmol) was added to a solution of tetrahydroisoquinoline 5 (2.6 g, 16.0 mmol) in THF–CH3CN 1[thin space (1/6-em)]:[thin space (1/6-em)]4 (28 mL) and 0.01 M aqueous EDTA solution (22.4 mL) at 5 °C. Oxone (10.3 g, 16.8 mmol) was then added portionwise over 2 h, and the mixture stirred for 20 min, at 5 °C. Ethyl acetate (80 mL) was added and the aqueous layer further extracted with ethyl acetate (2 × 80 mL). The combined organic extracts were dried over anhydrous MgSO4 and concentrated under reduced pressure to give nitrone 3 (1.66 g, 59%) as a pale yellow solid; m.p. decomposes T > 145 °C; [α]25D 174 (c = 3.3, CHCl3); IR (ATR): ν = 3242, 3054, 2945, 2897 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.80 (s, 1H), 7.35–7.25 (m, 2H), 7.25–7.20 (m, 1H), 7.18–7.14 (m, 1H), 4.51 (br s, 1H), 4.26–4.16 (m, 1H), 4.04 (dd, J = 12.0, 6.9 Hz, 1H), 3.94 (dd, J = 12.0, 3.6 Hz, 1H), 3.22–3.07 (m, 2H) ppm; 13C NMR (101 MHz, CDCl3): δC = 135.4, 130.0, 129.8, 127.6, 127.5, 127.2, 125.6, 67.2, 62.7, 29.8 ppm; MS (EI, 70 eV): m/z (%) = 177 (M+, 8), 130 (100), 115 (34), 103 (22), 77 (26); HRMS (ESI-MeOH): m/z calcd for C10H12NO2 [M + H]+: 178.0863 found 178.0862.

General procedure for the diastereoselective addition of Grignard reagents

A solution of RMgBr (0.9 mmol, 3 eq.) was added to a solution of nitrone 3 (53.1 mg, 0.3 mmol) in THF (1.5 mL). The mixture was stirred at 0 °C for 1 h and then warmed to room temperature and quenched with water. The mixture was extracted with EtOAc (3 × 10 mL) and the combined organic phases dried over MgSO4, and concentrated under reduced pressure. The crude product, was purified by flash column chromatography eluting with n-hexane/EtOAc (gradient from 10 to 60%) to give the corresponding 1,3-disubstituted tetrahydroisoquinoline-2-ol as a yellow oil.
(1R,3S)-3-(Hydroxymethyl)-1-methyl-3,4-dihydroisoquinolin-2(1H)-ol (6a) and (1S,3S)-3-(hydroxymethyl)-1-methyl-3,4-dihydroisoquinolin-2(1H)-ol (7a). Following the general procedure, the reaction of methylmagnesium bromide (3.0 M in Et2O, 300 μL) with nitrone 3 gave the 1,3-disubstituted tetrahydroisoquinoline-2-ol 6a (38 mg, 66% yield) as a yellow oil. [α]25D −15 (c = 1, CHCl3); IR (ATR): ν = 3240, 3019, 2973, 2932 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.17–7.10 (m, 2H), 7.10–7.01 (m, 2H), 4.32 (q, J = 7.1, 1H), 3.82 (dd, J = 11.5, 7.6 Hz, 1H), 3.72 (br s, 1H), 3.41–3.34 (m, 1H), 2.93 (dd, J = 16.7, 11.5 Hz, 1H), 2.53 (d, J = 14.8 Hz, 1H), 1.41 (d, J = 7.1 Hz, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 137.6, 132.2, 128.7, 127.3, 126.4, 126.2, 63.2, 62.4, 56.5, 25.8, 21.0 ppm; HRMS (ESI-MeOH): m/z calcd for C11H16NO2 [M + H]+: 194.1176 found 194.1778. 7a (5 mg, 9% yield) as a yellow oil. [α]25D −27 (c = 1, CHCl3); IR (ATR): ν = 3257, 2981, 2930, 2880 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.22–7.11 (m, 3H), 7.10–7.04 (m, 1H), 4.06 (q, J = 6.6 Hz, 1H), 3.82–3.78 (m, 2H), 3.16 (brs, 1H), 3.03 (dd, J = 16.2, 12.1 Hz, 1H), 2.65 (d, J = 14.8 Hz, 1H), 1.61 (d, J = 6.6 Hz, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 137.2, 133.0, 128.2, 126.6, 126.4, 126.0, 64.3, 64.0, 62.6, 29.7, 18.4 ppm; HRMS (ESI-MeOH): m/z calcd for C11H16NO2 [M + H]+: 194.1176 found 194.1177.
(1R,3S)-1-Ethyl-3-(hydroxymethyl)-3,4-dihydroisoquinolin-2(1H)-ol (6b) and (1S,3S)-1-ethyl-3-(hydroxymethyl)-3,4-dihydroisoquinolin-2(1H)-ol (7b). Following the general procedure, the reaction of ethylmagnesium bromide (3.0 M in Et2O, 300 μL) with nitrone 3 gave the 1,3-disubstituted tetrahydroisoquinoline-2-ol 6b (47 mg, 76% yield) as a yellow oil. [α]25D +3 (c = 1, CHCl3); IR (ATR): ν = 3285, 3062, 2931, 2873 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.20–7.12 (m, 2H), 7.10–7.07 (m, 2H), 4.07 (t, J = 7.5 Hz, 1H), 3.84 (dd, J = 10.2, 9.0 Hz, 1H), 3.68 (dd, J = 11.0, 3.5 Hz, 1H), 3.40–3.33 (m, 1H), 2.90 (dd, J = 16.7, 11.7 Hz, 1H), 2.52 (dd, J = 16.7, 5.0 Hz, 1H), 1.76–1.63 (m, 2H), 1.10 (t, J = 7.5 Hz, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 136.7, 132.4, 128.8, 127.8, 126.4, 126.0, 68.9, 63.1, 56.2, 29.4, 24.6, 11.9 ppm; HRMS (ESI-MeOH): m/z calcd for C12H18NO2 [M + H]+: 208.1332 found 208.1334. 7b (7 mg, 11% yield) as a yellow oil. [α]25D −29 (c = 1, CHCl3); IR (ATR): ν = 3279, 2961, 2931, 2873 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.22–7.11 (m, 3H), 7.08 (d, J = 7.1 Hz, 1H), 4.00 (t, J = 4.6 Hz, 1H), 3.81 (d, J = 4.9 Hz, 2H), 3.17–3.09 (m, 1H), 2.96 (dd, J = 16.1, 12.0 Hz, 1H), 2.64 (dd, J = 16.1, 4.0 Hz, 1H), 2.29–2.17 (m, 1H), 2.01–1.89 (m, 1H), 0.92 (t, J = 7.3 Hz, 1H) ppm; 13C NMR (101 MHz, CDCl3): δC = 136.0, 133.9, 128.2, 126.3, 126.3, 125.9, 67.7, 64.9, 63.4, 29.7, 24.1, 9.2 ppm; HRMS (ESI-MeOH): m/z calcd for C12H18NO2 [M + H]+: 208.1332 found 208.1334.
(1R,3S)-3-(Hydroxymethyl)-1-isopropyl-3,4-dihydroisoquinolin-2(1H)-ol (6c) and (1S,3S)-3-(hydroxymethyl)-1-isopropyl-3,4-dihydroisoquinolin-2(1H)-ol (7c). Following the general procedure, the reaction of isopropylmagnesium bromide (0.75 M in THF, 1.2 mL) with nitrone 3 gave the 1,3-disubstituted tetrahydroisoquinoline-2-ol 6c (48 mg, 73% yield) as a yellow oil. [α]25D +15 (c = 1, CHCl3); IR (ATR): ν = 3308, 3020, 2956, 2871 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.20–7.12 (m, 2H), 7.12–7.05 (m, 2H), 3.85 (dd, J = 11.3, 8.6 Hz, 1H), 3.79 (d, J = 8.7 Hz, 1H), 3.64 (dd, J = 11.3, 3.6 Hz, 1H), 3.46–3.35 (m, 1H), 2.81 (dd, J = 17.1, 11.3 Hz, 1H), 2.57 (dd, J = 17.1, 6.1 Hz, 1H), 1.85–1.70 (m, 1H), 1.09 (d, J = 6.7 Hz, 3H), 0.97 (d, J = 6.7 Hz, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 135.0, 132.8, 129.3, 128.8, 126.5, 125.2, 73.3, 63.2, 56.2, 33.2, 24.5, 21.2, 20.8 ppm; HRMS (ESI-MeOH): m/z calcd for C13H20NO2 [M + H]+: 222.1489 found 222.1483. 7c (8 mg, 12% yield) as a yellow oil. [α]25D +88 (c = 1, CHCl3); IR (ATR): ν = 3311, 3020, 2956, 2871 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.22–7.07 (m, 4H), 4.12 (d, J = 3.7 Hz, 1H), 3.93 (dd, J = 11.1, 3.0 Hz, 1H), 3.72 (dd, J = 11.1, 4.5 Hz, 1H), 3.11–3.01 (m, 1H), 2.88 (dd, J = 15.6, 12.4 Hz, 1H), 2.59 (dd, J = 15.6, 3.2 Hz, 1H), 2.29–2.12 (m, 1H), 1.06 (d, J = 6.8 Hz, 3H), 0.87 (d, J = 6.8 Hz, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 136.8, 134.9, 127.5, 126.8, 126.1, 126.0, 73.8, 64.9, 63.7, 34.6, 30.9, 20.6, 17.9 ppm; HRMS (ESI-MeOH): m/z calcd for C13H20NO2 [M + H]+: 222.1489 found 222.1482.
(1R,3S)-1-Allyl-3-(hydroxymethyl)-3,4-dihydroisoquinolin-2(1H)-ol (6d) and (1S,3S)-1-allyl-3-(hydroxymethyl)-3,4-dihydroisoquinolin-2(1H)-ol (7d). Following the general procedure, the reaction of allylmagnesium bromide (1.0 M in Et2O, 900 μL) with nitrone 3 gave the 1,3-disubstituted tetrahydroisoquinoline-2-ol 6d (44 mg, 67% yield) as a yellow oil. [α]25D −51 (c = 1, CHCl3); IR (ATR): ν = 3302, 3076, 3021, 2928, 2044 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.21–7.13 (m, 2H), 7.13–7.04 (m, 2H), 6.01–5.85 (m, 1H), 5.09–5.05 (m, 2H), 4.26 (t, J = 7.4 Hz, 1H), 3.83 (dd, J = 11.2, 8.4 Hz, 1H), 3.68 (dd, J = 11.2, 3.4 Hz, 1H), 3.44–3.35 (m, 1H), 2.92 (dd, J = 16.7, 11.7 Hz, 1H), 2.53 (dd, J = 16.7, 4.7 Hz, 1H), 2.49–2.39 (m, 2H) ppm; 13C NMR (101 MHz, CDCl3): δC = 135.8, 135.5, 132.5, 128.8, 127.7, 126.6, 126.0, 117.1, 66.8, 63.1, 56.3, 40.6, 24.7 ppm; HRMS (ESI-MeOH): m/z calcd for C13H18NO2 [M + H]+: 220.1332 found 220.1325. 7d (15 mg, 23% yield) as a yellow oil. [α]25D −41 (c = 1, CHCl3); IR (ATR): ν = 3280, 3075, 3021, 2927, 2042 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.26–7.14 (m, 3H), 7.14–7.06 (m, 1H), 6.02–5.87 (m, 1H), 5.16 (dd, J = 17.2, 1.6 Hz, 1H), 5.06 (d, J = 10.3 Hz, 1H), 4.15 (t, J = 5.7 Hz, 1H), 3.85–3.74 (m, 2H), 3.21–3.12 (m, 1H), 3.02–2.95 (m, 2H), 2.78–2.69 (m, 1H), 2.63 (dd, J = 16.3, 3.8 Hz, 1H) ppm; 13C NMR (101 MHz, CDCl3): δC = 136.3, 135.8, 133.8, 128.4, 126.5, 126.4, 126.1, 116.4, 66.1, 64.6, 63.3, 36.8, 28.9 ppm; HRMS (ESI-MeOH): m/z calcd for C13H18NO2 [M + H]+: 220.1332 found 220.1326.
(1R,3S)-3-(Hydroxymethyl)-1-phenyl-3,4-dihydroisoquinolin-2(1H)-ol (6e) and (1S,3S)-3-(hydroxymethyl)-1-phenyl-3,4-dihydroisoquinolin-2(1H)-ol (7e). Following the general procedure, reaction of phenylmagnesium bromide (3.0 M in Et2O, 300 μL) with nitrone 3 gave the 1,3-disubstituted tetrahydroisoquinoline-2-ol 6e (67 mg, 87% yield) as a yellow oil. [α]25D −84 (c = 1, CHCl3); IR (ATR): ν = 3240, 3061, 3026, 2889 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.29–7.24 (m, 3H), 7.23–7.18 (m, 2H), 7.18–7.12 (m, 1H), 7.12–7.06 (m, 2H), 6.91 (d, J = 7.6 Hz, 1H), 5.33 (s, 1H), 3.77 (dd, J = 11.4, 7.6 Hz, 1H), 3.55 (d, J = 9.2 Hz, 1H), 3.36–3.26 (m, 1H), 2.96 (dd, J = 16.9, 10.1 Hz, 1H), 2.70 (dd, J = 16.9, 5.1 Hz, 1H) ppm; 13C NMR (101 MHz, CDCl3): δC = 140.6, 134.0, 133.3, 130.2, 129.0, 128.6, 128.2, 127.7, 127.0, 126.2, 70.3, 62.8, 56.7, 26.2 ppm; HRMS (ESI-MeOH): m/z calcd for C16H18NO2 [M + H]+: 256.1332 found 256.1336. 7e (6 mg, 9% yield) as a yellow oil. [α]25D +4.5 (c = 1, CHCl3); IR (ATR): ν = 3238, 3061, 3025, 2890 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.35–7.30 (m, 2H), 7.29–7.23 (m, 2H), 7.15–7.08 (m, 2H), 7.03–6.95 (m, 1H), 6.54 (d, J = 7.8 Hz, 1H), 5.35 (br s, 1H), 4.88 (s, 1H), 3.95–3.82 (m, 1H), 3.77 (dd, J = 11.8, 5.3 Hz, 1H), 3.21 (d, J = 11.8 Hz, 1H), 3.14 (dd, J = 27.5, 11.8 Hz, 2H), 2.82 (d, J = 15.6 Hz, 1H) ppm; 13C NMR (101 MHz, CDCl3): δC = 141.9, 136.6, 133.2, 130.3, 128.3, 128.2, 127.8, 127.7, 126.7, 125.9, 75.1, 64.8, 64.6, 31.6 ppm; HRMS (CI-CH4): m/z calcd for C16H18NO2 [M + H]+: 256.1332 found 256.1337.
(S)-3-(((tert-Butyldimethylsilyl)oxy)methyl)-3,4-dihydroisoquinoline 2-oxide (8). A solution of (S)-3-(((tert-butyldimethylsilyl)oxy)methyl)-3,4-dihydroisoquinoline5,15 (0.76 g, 2.77 mmol) was added to a solution of sodium tungstate (0.14 g, 0.046 mmol) in methanol (1 mL). A 30% aqueous solution of H2O2 (1.0 mL, 8.31 mmol) was added at 0 °C over a period of 30 min and the reaction mixture was warmed to room temperature and stirred overnight. The mixture was concentrated under reduced pressure, diluted with brine and extracted with CH2Cl2 (2 × 15 mL). The combined organic layers were dried over anhydrous MgSO4, and concentrated under reduced pressure. The residue was purified by column chromatography, using CHCl3–CH3OH (10[thin space (1/6-em)]:[thin space (1/6-em)]1) as eluent, to give nitrone 10 (0.46 g, 57%) as a viscous oil. [α]25D −36 (c = 1, CHCl3); IR (ATR): ν = 2952, 2927, 2883, 2855 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.73 (s, 1H), 7.28–7.18 (m, 3H), 7.08 (d, J = 6.9 Hz, 1H), 4.16–4.08 (m, 1H), 3.96 (dd, J = 9.9, 3.8 Hz, 1H), 3.84 (dd, J = 9.9, 7.7 Hz, 1H), 3.42 (dd, J = 16.7, 7.3 Hz, 1H), 3.25 (dd, J = 16.8, 3.2 Hz, 1H), 0.79 (s, 9H), −0.01 (s, 3H), −0.03 (s, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 134.8, 129.6, 129.2, 127.7, 127.6, 127.2, 125.1, 68.8, 61.8, 29.2, 25.5, 18.0, −5.7, −5.78 ppm; HRMS (ESI-MeOH): m/z calcd for C16H26NO2Si [M + H]+: 292.1727 found 292.1734.
(1R,3S)-3-(((tert-Butyldimethylsilyl)oxy)methyl)-1-ethyl-3,4-dihydroisoquinolin-2(1H)-ol (9a) and (1S,3S)-3-(((tert-butyldimethylsilyl)oxy)methyl)-1-ethyl-3,4-dihydroisoquinolin-2(1H)-ol (10a). Following the general procedure, the reaction of ethylmagnesium bromide (3.0 M in Et2O, 300 μL) with nitrone 8 gave the 1,3-disubstituted tetrahydroisoquinoline-2-ol 9a (38 mg, 40% yield) as a yellow solid; m.p. 50–51 °C; [α]25D −12 (c = 1.2, CHCl3); IR (ATR): ν = 3229, 3063, 2954, 2928, 2856, 1462, 1252, 1100, 834, 774, 748 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.18–7.09 (m, 4H), 6.42 (br s, 1H), 4.05 (dd, J = 8.3, 6.3 Hz, 1H), 3.99 (dd, J = 10.0, 5.6 Hz, 1H), 3.79 (dd, J = 10.0, 6.1 Hz, 1H), 3.31 (td, J = 11.1, 5.6 Hz, 1H), 2.98 (dd, J = 16.7, 11.0 Hz, 1H), 2.66 (dd, J = 16.7, 5.2 Hz, 1H), 1.76–1.67 (m, 2H), 1.09 (t, J = 7.4Hz, 3H), 0.92 (s, 9H), 0.11 (s, 3H), 0.10 (s, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 137.4, 133.3, 128.9, 127.8, 126.3, 126.0, 68.7, 65.5, 56.6, 29.7, 26.1, 25.6, 18.5, 11.9, −5.1, −5.2 ppm; HRMS (ESI-EtOAc): m/z calcd for C18H31NO2Si [M + H]+ 322.2202 found 322.2197. 10a (16 mg, 16% yield) as a yellow oil. [α]25D −7 (c = 1.5, CHCl3); IR (ATR): ν = 3263, 3066, 2954, 2928, 2855, 1462, 1252, 1103, 834, 775, 739 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.19–7.11 (m, 4H),5.05 (br s, 1H), 4.03 (dd, J = 10.0, 4.9 Hz, 1H), 3.96 (t, J = 4.6 Hz, 1H), 3.77 (dd, J = 10.0, 6.2 Hz, 1H), 3.09 (m, 1H), 2.93–2.86 (m 1H), 2.79 (dd, J = 16.2, 4.2 Hz, 1H), 2.28–2.17 (m, 1H), 1.98–187 (m, 1H), 0.92 (s, 9H), 0.89 (t, J = 7.2 Hz, 1H), 0.11 (s, 3H), 0.10 (s, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 136.6, 134.6, 128.3, 126.3, 126.2, 126.0, 68.1, 66.1, 63.6, 29.9, 26.1, 24.6, 18.5, 9.2, −5.1, −5.2 ppm; HRMS (ESI-EtOAc): m/z calcd for C18H31NO2Si [M + H]+ 322.2202 found 322.2197.
(1R,3S)-3-(((tert-Butyldimethylsilyl)oxy)methyl)-1-phenyl-3,4-dihydroisoquinolin-2(1H)-ol (9b) and (1S,3S)-3-(((tert-butyldimethylsilyl)oxy)methyl)-1-phenyl-3,4-dihydroisoquinolin-2(1H)- (10b). Following the general procedure, the reaction of phenylmagnesium bromide (3.0 M in Et2O, 300 μL) with nitrone 8 gave the 1,3-disubstituted tetrahydroisoquinoline-2-ol 9b (44 mg, 40% yield) as a yellow oil. [α]25D −4 (c = 1, CHCl3); IR (ATR): ν = 3213, 3062, 2927, 2855 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.39–7.31 (m, 3H), 7.29–7.28 (m, 2H), 7.25–7.17 (m, 3H), 6.96 (d, J = 7.7 Hz, 1H), 6.92 (s, 1H), 5.28 (s, 1H), 4.06 (dd, J = 10.0, 5.2 Hz, 1H), 3.82 (dd, J = 10.0, 6.7 Hz, 1H), 3.48–3.40 (m, 1H), 3.14 (dd, J = 16.8, 8.3 Hz, 1H), 2.98 (dd, J = 16.8, 5.3 Hz, 1H), 0.92 (s, 9H), 0.08 (s, 3H), 0.06 (s, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 141.8, 134.9, 133.7, 130.0, 128.9, 128.7, 128.1, 127.4, 126.7, 125.9, 70.0, 63.4, 57.5, 27.2, 25.8, 18.2, −5.4, −5.5 ppm; HRMS (ESI-MeOH): m/z calcd for C22H32NO2Si [M + H]+ 370.2197 found 370.2205. 10b (15 mg, 13% yield) as a yellow oil. [α]25D −55 (c = 1, CHCl3); IR (ATR): ν = 3306, 3028, 2927, 2854 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.39–7.34 (m, 5H), 7.16–7.11 (m, 2H), 7.04–6.97 (m, 1H), 6.59 (d, J = 7.8 Hz, 1H), 4.89 (s, 2H), 4.16 (dd, J = 9.9, 4.4 Hz, 1H), 3.79 (dd, J = 9.9, 6.9 Hz, 1H), 3.29–3.18 (m, 1H), 3.09 (s, 1H), 3.07 (s, 1H), 0.93 (s, 9H), 0.10 (s, 3H), 0.10 (s, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 142.9, 137.3, 133.5, 129.9, 128.4, 128.1, 127.8, 127.5, 126.5, 125.8, 74.7, 65.3, 64.1, 25.9, 22.7, 18.3, −5.2, −5.3 ppm; HRMS (ESI-MeOH): m/z calcd for C22H32NO2Si [M + H]+ 370.2197 found 370.2203.

General procedure for the cycloaddition reactions of nitrone 3

A solution of nitrone 3 (0.1 g, 0.56 mmol) in dry toluene (5 mL) was heated to 80 °C. A solution of the corresponding alkene (5 eq.) in dry toluene (5 mL) was added dropwise and the reaction was heated at 110 °C overnight. Removal of the solvent under reduced pressure yielded the crude product as a dark brown oil. Purification by flash chromatography eluting with EtOAc/cyclohexane (1[thin space (1/6-em)]:[thin space (1/6-em)]1) gave the corresponding products 11.
((2R,5S,10bR)-2-Phenyl-1,5,6,10b-tetrahydro-2H-isoxazolo[3,2-a]isoquinolin-5-yl)methanol (11a). Following the general procedure, reaction of styrene (2.8 mmol, 322 μL) with nitrone 3 gave the cycloadduct 11a (26 mg, 86% yield) as a yellow oil; [α]25D −29 (c = 3.6, CHCl3); IR (ATR): ν = 3372, 3027, 2937, 2886, 1733, 1241, 1047, 927, 794, 699 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.46–7.43 (m, 2H), 7.42–7.37 (m, 2H), 7.36–7.31 (m, 1H), 7.21–7.14 (m, 3H), 7.09–7.07 (m, 1H), 5.44 (dd, J = 9.5, 4.5 Hz, 1H), 4.88 (dd, J = 10.3, 8.2 Hz, 1H), 3.94–3.93 (m, 2H), 3.35–3.30 (m, 1H), 2.97 (dd, J = 16.4, 11.5 Hz, 1H), 2.82–2.72 (m, 2H), 2.68–2.62 (m, 1H) ppm; 13C NMR (101 MHz, CDCl3): δC = 141.1, 134.7, 132.2, 128.6, 128.2, 128.0, 127.0, 126.9, 126.4, 79.1, 66.2, 64.1, 56.9, 45.5, 31.5 ppm; HRMS (ESI-EtOAc): m/z calcd for C18H19NO2 [M + H]+: 282.1494 found 282.1489.
((2S,5S,10bR)-2-Butyl-1,5,6,10b-tetrahydro-2H-isoxazolo[3,2-a]isoquinolin-5-yl)methanol (11b). Following the general procedure reaction of 1-hexene (2.8 mmol, 350 μL) with nitrone 3 gave the cycloadduct 11b (17 mg, 74% yield) as a yellow oil; [α]25D −6 (c = 1, CHCl3); IR (ATR): ν = 3405, 2953, 2858, 1454, 1050, 944, 748, 713 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.20–7.14 (m, 2H), 7.10–7.09 (m, 2H), 4.60 (t, J = 9.2, 1H), 4.41 (dq, J = 12.6, 6.1 Hz, 1H), 3.86–3.85 (m, 2H), 3.29 (br s, 1H), 3.20–3.14 (m, 1H), 2.85 (dd, J = 16.3, 11.6 Hz, 1H), 2.70 (dd, J = 16.4, 3.6 Hz, 1H), 2.40–2.31 (m, 2H), 1.72–1.66 (m, 1H), 1.64 (m, 1H), 1.44–1.30 (m, 4H), 0.92 (t, J = 6.9 Hz, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 135.3, 132.2, 128.3, 127.1, 126.8, 126.4, 77.6, 66.7, 63.9, 56.5, 42.4, 35.0, 31.6, 28.2, 22.8, 14.1 ppm; HRMS (ESI-EtOAc): m/z calcd for C16H23NO2 [M + H]+: 262.1807 found 262.1802.
((2R,5S,10bR)-2-(2-(Benzyloxy)ethyl)-1,5,6,10b-tetrahydro-2H-isoxazolo[3,2-a]isoquinolin-5-yl)methanol (11c). Following the general procedure, reaction of 4-benzyloxybut-1-ene18 (2.8 mmol, 454 mg) with nitrone 3 gave the cycloadduct 11c (42 mg, 61% yield) as a yellow oil. [α]25D −1.21 (c = 9.9, CHCl3); IR (ATR): ν = 3421, 3027, 2916, 2882, 1454, 1361, 1096, 910, 733, 697 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.38–7.27 (m, 5H), 7.21–7.15 (m, 2H), 7.12–7.07 (m, 2H), 4.64–4.55 (m, 2H), 4.52 (s, 2H), 3.86 (d, 4.2 Hz, 2H), 3.65–3.60 (m, 2H), 3.24 (br s, 1H), 3.17 (ddd, J = 11.7, 7.9, 4.0 Hz, 1H), 2.88 (dd, J = 16.3, 11.6 Hz, 1H), 2.72 (dd, J = 16.4, 3.6 Hz, 1H), 2.48–2.36 (m, 2H), 2.05–1.92 (m, 2H) ppm; 13C NMR (101 MHz, CDCl3): δC = 138.1, 135.0, 132.1, 128.3, 128.1, 127.6, 127.5, 126.9, 126.7, 126.2, 77.2, 74.8, 73.1, 66.9, 66.1, 63.6, 56.5, 42.2, 35.2, 31.4 ppm; HRMS (ESI-EtOAc): m/z calcd for C21H25NO3 [M + H]+: 340.1913 found 340.1907.
((2R,5S,10bR)-2-(2-(Phenylthio)ethyl)-1,5,6,10b-tetrahydro-2H-isoxazolo[3,2-a]isoquinolin-5-yl)methanol (11d). Following the general procedure, reaction of but-3-en-1-yl(phenyl)sulfane19 (2.8 mmol, 459 mg) with nitrone 3 gave the cycloadduct 11d (50 mg, 64% yield) as a yellow oil. [α]25D −91 (c = 1, CHCl3); IR (ATR): ν = 3408, 3058, 2930, 2884, 1582, 1438, 1051, 1024, 944, 736, 690 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.37–7.35 (m, 2H), 7.32–7.28 (m, 2H), 7.21–7.06 (m, 5H), 4.63–4.54 (m, 2H), 3.86 (s, 2H), 3.19–3.07 (m, 3H), 2.99 (ddd, J = 13.1, 8.3, 7.2 Hz, 1H), 2.88 (dd, J = 16.2, 11.6 Hz, 1H), 2.72 (dd, J = 16.4, 3.6 Hz, 1H), 2.45–2.30 (m, 2H), 2.08–1.91 (m, 2H) ppm; 13C NMR (101 MHz, CDCl3): δC = 135.8, 134.8, 132.1, 129.2, 128.9, 128.1, 126.9, 126.8, 126.3, 126.1, 75.7, 66.1, 63.7, 56.7, 42.1, 34.7, 31.4, 29.9 ppm; HRMS (ESI-EtOAc): m/z calcd for C20H23NO2S [M + H]+: 342.1528 found 342.1522.

General procedure for the reduction of cycloadducts 1116

Zinc dust (5.0 eq.) was added to a solution of cycloadduct 11 (1 eq.) in 50% acetic acid[thin space (1/6-em)]:[thin space (1/6-em)]water at room temperature. The mixture was stirred under reflux for 3 h. After cooling, the mixture was neutralized with saturated aqueous NaHCO3 and the aqueous phase was extracted with CH2Cl2 (3 × 10 mL). The combined organic phases were dried over MgSO4, concentrated under reduced and the residue purified by column chromatography.
(R)-2-((1R,3S)-3-(Hydroxymethyl)-1,2,3,4-tetrahydroisoquinolin-1-yl)-1-phenylethanol (12a). Following the general procedure, reduction of cycloadduct 11a (50 mg, 0.18 mmol) gave 1,3-disubstituted THIQ 12a (17 mg, 33% yield) as a yellow oil. [α]25D −0.42 (c = 2.4, CHCl3); IR (ATR): ν = 3300, 3060, 3024, 2918, 1655, 1450, 1057, 1028, 743, 700 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.45 (d, J = 7.4 Hz, 2H), 7.35 (t, J = 7.6 Hz, 2H), 7.27–7.23 (m, 1H), 7.12–7.03 (m, 3H), 6.82–6.80 (m, 1H), 5.09 (t, J = 4.3 Hz, 1H), 4.4 (br s, 2H), 4.17 (dd, J = 11.2, 2.2 Hz, 1H), 3.76 (dd, J = 11.1, 3.8 Hz, 1H), 3.59 (dd, J = 11.2, 6.3 Hz, 1H), 3.30–3.24 (m, 1H), 2.69 (dd, J = 16.5, 4.2 Hz, 1H), 2.56 (dd, J = 16.4, 10.7 Hz, 1H), 2.44 (ddd, J = 14.8, 11.4, 3.6 Hz, 1H), 1.98 (ddd, J = 14.7, 5.2, 2.8 Hz, 1H), ppm; 13C NMR (101 MHz, CDCl3): δC = 144.7, 138.0, 134.0, 129.4, 128.2, 126.7, 126.2, 125.9, 125.4, 72.3, 65.1, 51.5, 48.6, 41.8, 30.4 ppm; HRMS (ESI-EtOAc): m/z calcd for C18H21NO2 [M + H]+: 284.1651 found 284.1645.
(S)-1-((1R,3S)-3-(Hydroxymethyl)-1,2,3,4-tetrahydroisoquinolin-1-yl)hexan-2-ol (12b). Following the general procedure, reduction of cycloadduct 11b (17 mg, 0.07 mmol) gave 1,3-disubstituted THIQ 12b (16 mg, 90% yield) as a yellow oil. [α]25D−0.21 (c = 4.7, CHCl3); IR (ATR): ν = 3310, 3062, 2926, 2857, 1452, 1123, 1035, 743 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.14–7.01 (m, 4H), 4.43 (dd, J = 10.8, 2.4 Hz, 1H), 3.82 (br s, 3H), 3.70 (dd, J = 11.1, 3.7 Hz, 1H), 3.54 (dd, J = 11.0, 7.3 Hz, 1H), 3.26–3.20 (m, 1H), 2.73 (dd, J = 16.4, 4.1 Hz, 1H), 2.56 (dd, J = 16.4, 10.2 Hz, 1H), 2.10 (ddd, J = 14.2, 11.3, 3.0 Hz, 1H), 1.73–1.61 (m, 2H), 1.56–1.1.25 (m, 6H), 0.91 (t, J = 6.9 Hz, 3H) ppm; 13C NMR (101 MHz, CDCl3): δC = 138.3, 133.8, 129.4, 126.3, 126.2, 126.0, 69.4, 64.8, 51.3, 48.9, 40.3, 36.8, 30.5, 28.3, 22.8, 14.1 ppm; HRMS (ESI-EtOAc): m/z calcd for C16H25NO2 [M + H]+: 264.1964 found 264.1958.
(R)-4-(Benzyloxy)-1-((1R,3S)-3-(hydroxymethyl)-1,2,3,4-tetrahydroisoquinolin-1-yl)butan-2-ol (12c). Following the general procedure, reduction of cycloadduct 11c (27 mg, 0.08 mmol) gave 1,3-disubstituted THIQ 12c (14 mg, 51% yield) as a yellow oil. [α]25D −0.31 (c = 3.2, CHCl3); IR (ATR): ν = 3298, 3061, 3028, 2916, 2857, 1452, 1362, 1090, 1027, 740, 697 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.28–7.18 (m, 5H), 7.06–6.89 (m, 4H), 4.43 (s, 2H), 4.36 (dd, J = 10.9, 2.5 Hz, 1H), 4.01–3.92 (m, 4H), 3.64 (dd, J = 11.2, 3.8 Hz, 1H), 3.53 (t, J = 6.1 Hz, 2H), 3.45 (dd, J = 11.1, 7.4 Hz, 1H), 3.20–3.14 (m, 1H), 2.64 (dd, J = 16.5, 4.3 Hz, 1H), 2.49 (dd, J = 16.5, 10.4 Hz, 1H), 2.02–1.95 (m, 1H), 1.88–1.81 (m, 1H), 1.78–1.64 (m, 2H) ppm; 13C NMR (101 MHz, CDCl3): δC = 138.4, 138.2, 133.7, 129.5, 128.5, 127.9, 127.8, 126.6, 126.5, 126.2, 73.4, 68.5, 67.8, 64.9, 51.4, 49.1, 41.1, 36.8, 30.6 ppm; HRMS (ESI-EtOAc): m/z calcd for C21H27NO3 [M + H]+: 342.2069 found 342.2064.
(R)-1-((1R,3S)-3-(Hydroxymethyl)-1,2,3,4-tetrahydroisoquinolin-1-yl)-4-(phenylthio)butan-2-ol (12d). Following the general procedure, reduction of cycloadduct 11d (47 mg, 0.13 mmol) gave 1,3-disubstituted THIQ 12d (22 mg, 47% yield) as a white solid; m.p. 60–62 °C; [α]25D −6 (c = 2, CHCl3); IR (ATR): ν = 3369, 3070, 2919, 1581, 1479, 1437, 1090, 1027, 735, 689 cm−1; 1H NMR (400 MHz, CDCl3): δH = 7.25 (d, J = 7.4 Hz, 2H), 7.18 (t, J = 7.6 Hz, 2H), 7.09–7.02 (m, 31H), 7.00–6.97 (m, 1H), 6.92–6.90 (m, 1H), 4.31 (dd, J = 10.8, 2.6 Hz, 1H), 3.94–3.88 (m, 1H), 3.71 (br s, 3H), 3.61 (dd, J = 11.1, 3.8 Hz, 1H), 3.44 (dd, J = 11.1, 7.2 Hz, 1H), 3.16–3.03 (m, 2H), 2.95–2.88 (m, 1H), 2.64 (dd, J = 16.5, 4.3 Hz, 1H), 2.46 (d, J = 16.4, 101 Hz, 1H), 2.02 (ddd, J = 14.2, 11.2, 3.1), 1.96–1.87 (m, 1H), 1.75–1.67 (m, 1H), 1.62 (ddd, J = 14.7, 6.6, 3.1 Hz, 1H) ppm; 13C NMR (101 MHz, CDCl3): δC = 128.3, 136.6, 133.9, 129.7, 129.0, 128.9, 126.5, 126.4, 126.2, 125.9, 68.6, 65.1, 51.6, 49.0, 40.4, 36.5, 30.7, 30.3 ppm; HRMS (ESI-EtOAc): m/z calcd for C20H25NO2S [M + H]+: 344.1684 found 344.1679.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

J. F.-F. thanks the University of Alicante for a placement studentship. M. W. thanks the Erasmus+programme for funding her internship. B. M. thanks the European Commission for a Marie Curie Career Integration Grant, the EPSRC for a First Grant and the RS for a travel grant.

Notes and references

  1. (a) A. H. Jackson, in Chemistry and Biology of Isoquinoline Alkaloids, ed. J. D. Phillipson, M. F. Margaret and M. H. Zenk, Springer, Berlin, Germany, 1985, pp. 62–78 Search PubMed; (b) J. D. Scott and R. M. Williams, Chem. Rev., 2002, 102, 1669–1730 CrossRef PubMed; (c) K. W. Bentley, Nat. Prod. Rep., 2006, 23, 444–463 RSC; (d) S. Kotha, D. Deodhar and P. Khedkar, Org. Biomol. Chem., 2014, 12, 9054–9091 RSC; (e) V. H. Le, R. M. Williams and T. Kan, Nat. Prod. Rep., 2015, 32, 328–347 RSC.
  2. (a) M. J. Fisher, R. Backer, S. Husain, H. M. Hsiung, J. T. Mullaney, T. P. O'Brian, P. L. Ornstein, R. R. Rothhaar, J. M. Zgombick and K. Briner, Bioorg. Med. Chem. Lett., 2005, 15, 4459–4462 CrossRef PubMed; (b) M. E. Welsch, S. A. Snyder and B. R. Stockwell, Curr. Opin. Chem. Biol., 2010, 14, 347–361 CrossRef PubMed; (c) B. A. Granger, K. Kaneda and S. F. Martin, Org. Lett., 2011, 13, 4542–4545 CrossRef PubMed.
  3. C. W. Murray and D. C. Rees, Angew. Chem., Int. Ed., 2016, 55, 488–492 CrossRef PubMed.
  4. (a) J. Feng, S. Dastgir and C.-J. Li, Tetrahedron Lett., 2008, 49, 668–671 CrossRef; (b) B. K. Peters, S. K. Chakka, T. Naicker, G. E. M. Maguire, H. G. Kruger, P. G. Andersson and T. Govender, Tetrahedron: Asymmetry, 2010, 21, 679–687 CrossRef; (c) T. Naicker, K. Petzold, T. Singh, P. I. Arvidsson, H. G. Kruger, G. E. M. Maguire and T. Govender, Tetrahedron: Asymmetry, 2010, 21, 2859–2867 CrossRef; (d) S. K. Chakka, Z. E. D. Cele, S. C. Sosibo, V. Francis, P. I. Arvidsson, H. G. Kruger, G. E. M. Maguire and T. Govender, Tetrahedron: Asymmetry, 2010, 21, 846–852 CrossRef; (e) P. D. McCloud, A. M. Reckling and C.-J. Li, Heterocycles, 2010, 80, 1319–1337 CrossRef; (f) S. K. Chakka, P. G. Andersson, G. E. M. Maguire, H. G. Kruger and T. Govender, Eur. J. Org. Chem., 2010, 972–980 CrossRef; (g) T. Naicker, P. I. Arvidsson, H. G. Kruger, G. E. M. Maguire and T. Govender, Eur. J. Org. Chem., 2011, 6923–6932 CrossRef; (h) Z. E. D. Cele, S. C. Sosibo, P. G. Andersson, H. G. Kruger, G. E. M. Maguire and T. Govender, Tetrahedron: Asymmetry, 2013, 24, 191–195 CrossRef; (i) D. T. Kong and Z. M. A. Judeh, Synthesis, 2016, 48, 2271–2279 CrossRef.
  5. V. K. Aggarwal, P. S. Humphries and A. Fenwick, J. Chem. Soc., Perkin Trans. 1, 1999, 2883–2889 RSC.
  6. For selected recent examples see: (a) D. Enders, J. X. Liebich and G. Raabe, Chem. – Eur. J., 2010, 16, 9763–9766 CrossRef PubMed; (b) D. L. Priebbenow, S. G. Stewart and F. M. Pfeffer, Tetrahedron Lett., 2012, 53, 1468–1471 CrossRef; (c) S. Fustero, I. Ibanez, P. Barrio, M. A. Maestro and S. Catalan, Org. Lett., 2013, 15, 832–835 CrossRef PubMed; (d) B. Dulla, N. D. Tangellamudi, S. Balasubramanian, S. Yellanki, R. Medishetti, R. K. Banote, G. H. Chaudhari, P. Kulkarni, J. Iqbal, O. Reiser and M. Pal, Org. Biomol. Chem., 2014, 12, 2552–2558 RSC; (e) L. R. Peacock, R. S. L. Chapman, A. C. Sedgwick, M. F. Mahon, D. Amans and S. D. Bull, Org. Lett., 2015, 17, 994–997 CrossRef PubMed; (f) K. Mori, N. Umehara and T. Akiyama, Adv. Synth. Catal., 2015, 357, 901–906 CrossRef; (g) S. Raghavan and P. Senapati, J. Org. Chem., 2016, 81, 6201–6210 CrossRef PubMed.
  7. For selected recent examples see: (a) N. Chatani, T. Asaumi, S. Yorimitsu, T. Ikeda, F. Kakiuchi and S. Murai, J. Am. Chem. Soc., 2001, 123, 10935–10941 CrossRef PubMed; (b) N. Y. Kuznetsov, V. N. Khrustalev, I. A. Godovikov and Y. N. Bubnov, Eur. J. Org. Chem., 2006, 113–120 CrossRef; (c) A. Peschiulli, V. Smout, T. E. Storr, E. A. Mitchell, Z. Elis, W. Herrebout, D. Berthelot, L. Meerpoel and B. U. Maes, Chem. – Eur. J., 2013, 19, 10378–10387 CrossRef PubMed; (d) G. Lahm and T. Opatz, Org. Lett., 2014, 16, 4201–4203 CrossRef PubMed; (e) Y. Kita, K. Yamaji, K. Higashida, K. Sathaiah, A. Iimuro and K. Mashima, Chem. – Eur. J., 2015, 21, 1915–1927 CrossRef PubMed.
  8. A. Monsees, S. Laschat and I. Dix, J. Org. Chem., 1998, 63, 10018–10021 CrossRef.
  9. For reviews covering cycloadditions to nitrones see: (a) M. Fredrickson, Tetrahedron, 1997, 53, 403–425 CrossRef; (b) P. de March, M. Figueredo and J. Font, Heterocycles, 1999, 50, 1213–1226 CrossRef; (c) H. M. I. Osborn, N. Gemmell and L. M. Harwood, J. Chem. Soc., Perkin Trans. 1, 2002, 2419–2438 RSC; (d) A. E. Koumbis and J. K. Gallos, Curr. Org. Chem., 2003, 7, 585–628 CrossRef; (e) L. M. Stanley and M. P. Sibi, Chem. Rev., 2008, 108, 2887–2902 CrossRef PubMed; (f) S. Kanemasa, Heterocycles, 2010, 82, 87–200 CrossRef; (g) T. M. V. D. Pinho e Melo, Eur. J. Org. Chem., 2010, 3363–3376 CrossRef; (h) G. Molteni, Heterocycles, 2016, 92, 2115–2140 CrossRef; Also, see: (i) W. S. Jen, J. J. M. Weiner and D. W. C. MacMillan, J. Am. Chem. Soc., 2000, 122, 9874–9875 CrossRef; (j) P. H. Poulsen, S. Vergura, A. Monleon, D. K. B. Jørgensen and K. A. Jørgensen, J. Am. Chem. Soc., 2016, 138, 6412–6415 CrossRef PubMed; (k) J. Vesely, R. Rios, I. Ibrahem, G.-L. Zhao, L. Eriksson and A. Cordova, Chem. – Eur. J., 2008, 14, 2693–2698 CrossRef PubMed.
  10. For reviews covering nucleophilic additions to nitrones see: (a) R. Matute, S. Garcia-Viñuales, H. Hayes, M. Ghirardello, A. Darù, T. Tejero, I. Delso and P. Merino, Curr. Org. Synth., 2016, 13, 669–686 CrossRef; (b) M. Lombardo and C. Trombini, Curr. Org. Chem., 2002, 6, 695–713 CrossRef; (c) M. Lombardo and C. Trombini, Synthesis, 1999, 905–917 Search PubMed.
  11. (a) G. Masson, S. Py and Y. Vallee, Angew. Chem., Int. Ed., 2002, 41, 1772–1775 CrossRef; (b) D. Riber and T. Skrydstrup, Org. Lett., 2003, 5, 229–231 CrossRef PubMed; (c) M. Szostak, N. J. Fazakerley, D. Parmar and D. Procter, Chem. Rev., 2014, 114, 5959–6039 CrossRef PubMed.
  12. (a) I. M. P. Huber and H. Seebach, Helv. Chim. Acta, 1987, 70, 1944–1954 CrossRef; (b) M. Gurram, B. Gyimothy, R. Wang, S. Q. Lam, F. Ahmed and R. J. Herr, J. Org. Chem., 2011, 76, 1605–1613 CrossRef PubMed.
  13. M. P. Chelopo, S. A. Pawar, M. K. Sokhela, T. Govender, H. G. Kruger and G. E. M. Maguire, Eur. J. Med. Chem., 2013, 66, 407–414 CrossRef PubMed.
  14. C. Gella, E. Ferrer, R. Alibes, F. Busque, P. de March, M. Figueredo and J. A. Font, J. Org. Chem., 2009, 74, 6365–6367 CrossRef PubMed.
  15. D. Page, A. Naismith, R. Schmidt, R. Coupal, E. Labarre, M. Gosselin, D. Bellemare, K. Payza and W. Brown, J. Med. Chem., 2001, 44, 2387–2390 CrossRef PubMed.
  16. M. Bonanni, M. Marradi, S. Cicchi, C. Faggi and A. Goti, Org. Lett., 2005, 7, 319–322 CrossRef PubMed.
  17. P. Aschwanden, L. Kværnø, R. W. Geisser, F. Kleinbeck and E. M. Carreira, Org. Lett., 2005, 7, 5741–5742 CrossRef PubMed.
  18. S. Bag, R. Jayarajan, R. Mondal and D. Maiti, Angew. Chem., Int. Ed., 2017, 56, 3182–3186 CrossRef PubMed.
  19. H. Leicht, I. Göttker-Schnetmann and S. Mecking, J. Am. Chem. Soc., 2017, 139, 6823–6826 CrossRef PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c8ob02007h

This journal is © The Royal Society of Chemistry 2018