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

Isothiourea catalysed enantioselective generation of point and axially chiral iminothia- and iminoselenazinanones

Alastair J. Nimmo, Alister S. Goodfellow, Jacob T. Guntley, Aidan P. McKay, David B. Cordes, Michael Bühl* and Andrew D. Smith*
EaStCHEM, School of Chemistry, University of St Andrews, Fife, KY16 9ST, UK. E-mail: buehl@st-andrews.ac.uk; ads10@st-andrews.ac.uk

Received 31st March 2025 , Accepted 29th April 2025

First published on 30th April 2025


Abstract

Symmetrical and unsymmetrical thioureas, as well as unsymmetrical selenoureas, are used in an isothiourea-catalysed Michael addition-lactamisation protocol using α,β-unsaturated pentafluorophenyl esters to generate iminothia- and iminoselenazinanone heterocycles with high enantioselectivity (up to 99[thin space (1/6-em)]:[thin space (1/6-em)]1 er). The scope and limitations of this process have been widely investigated (40 examples in total) with unsymmetrical thio- and selenoureas containing ortho-substituted N-aryl substituents giving atropisomeric products, leading to an effective process for iminothia- and iminoselenazinanones heterocyclic products containing both point and axially chiral stereogenic elements with excellent stereocontrol (up to >95[thin space (1/6-em)]:[thin space (1/6-em)]5 dr and 98[thin space (1/6-em)]:[thin space (1/6-em)]2 er). Mechanistic investigation showed that (i) the catalytically liberated aryloxide could deprotonate an electron-deficient thiourea; (ii) in the absence of an isothiourea catalyst, this leads to formation of racemic product; (iii) a crossover experiment indicates the reversibility of the thia-Michael addition. Computational analysis has identified the factors leading to enantioselectivity within this process, with stereocontrol arising from the lactamisation step within the catalytic cycle.


1. Introduction

Since the introduction of isothioureas as catalysts for the acylative kinetic resolution of alcohols by Birman,1 these versatile Lewis bases have been widely developed and applied in enantioselective processes. Their simple and scalable synthesis, combined with their ability to generate multiple reactive intermediates 1–5 (acyl ammonium,2 α,β-unsaturated acyl ammonium,3 C(1)-ammonium enolate,4 betaine5 and silyl ammonium6 species) from simple starting materials has led to their widespread popularity (Scheme 1A). Isothiourea-catalysed Michael additions to in situ generated α,β-unsaturated acyl ammonium species have been used extensively to achieve enantioselective C–C bond formation.7 However, their use to selectively generate C-heteroatom bonds through conjugate addition of non-carbon centred nucleophiles is limited to relatively few C–S and C–N bond-forming processes. Within this area, Matsubara has reported a thia-Michael addition-lactamisation strategy for the synthesis of 1,5-benzothiazepines 9 using N-tosylated aminothiophenols 6 and α,β-unsaturated mixed anhydrides 7 in the presence of (R)-BTM 8 (Scheme 1B).8 The products were isolated in good to excellent yields (70 to 99%) and with excellent enantioselectivity (94[thin space (1/6-em)]:[thin space (1/6-em)]6 to 99[thin space (1/6-em)]:[thin space (1/6-em)]1 er). Mechanistic investigations revealed that the high enantioselectivity is a result of a dynamic kinetic process determined by the relative lactamisation rates of the diastereomeric thia-Michael addition products rather than the thia-Michael addition, which is readily reversible. This methodology was extended to 3-substituted and 2,3-disubstituted benzothiazepines with excellent stereoselectivity (79[thin space (1/6-em)]:[thin space (1/6-em)]21 to 94[thin space (1/6-em)]:[thin space (1/6-em)]6 er).9 Birman subsequently reported a route to enantioenriched thiochromenes 13 using thioesters 10 as α,β-unsaturated acyl ammonium precursors (Scheme 1C).10
image file: d5sc02435h-s1.tif
Scheme 1 (A) Isothiourea derived catalytic intermediate. (B) & (C) Enantioselective C–S bond forming processes using α,β-unsaturated acyl ammonium species.

Enantioselective C–S bond formation via thia-Michael addition is followed by an aldol-lactonisation cascade forming β-lactones 12, which readily decarboxylate, giving the corresponding thiochromene 13 in good to excellent yields (53 to 99%) and with excellent enantiocontrol (97[thin space (1/6-em)]:[thin space (1/6-em)]3 to >99[thin space (1/6-em)]:[thin space (1/6-em)]1 er). Alongside these advances, in previous work we developed the aza-Michael addition of 2-hydroxybenzophenone imines to α,β-unsaturated esters allowing access to β-amino amides in excellent yield (95%) and enantioselectivity (96[thin space (1/6-em)]:[thin space (1/6-em)]4 er).11

In addition, in recent years there has been an explosion of interest concerning the development of selective synthetic methods for the preparation of atropisomeric species.12 While axially chiral biaryl (and heterobiaryl) species containing a C–C axis have been most widely studied, recent work has shown that the generation of axially chiral and configurationally stable C–N atropisomers is possible.13 Given these precedents, we considered alternative ways of developing enantioselective C–S and C–Se bond formation via an α,β-unsaturated acyl ammonium intermediate that could lead to heterocycles bearing both point and axially chiral stereogenic elements. In recent work Jin, Chi and coworkers used thioureas as dinucleophiles in an atropselective synthesis of iminothiazinones 18 using the NHC derived from precatalyst 17 (20 mol%, Scheme 2A).14 Thia-Michael addition-lactamisation of the thiourea 14 with an in situ generated acyl azolium intermediate gives iminothiazinones 18 in moderate to high yields (29 to 71%) with excellent enantioselectivity (91[thin space (1/6-em)]:[thin space (1/6-em)]9 to >99[thin space (1/6-em)]:[thin space (1/6-em)]1 er) with the formation of a C–N stereogenic axis. While effective, stoichiometric quantities (3 equiv.) of oxidant 16 was required to achieve oxidation of the Breslow intermediate in this protocol and furan was required as a solvent. This manuscript describes an alternative methodology to access related heterocyclic products that utilises α,β-unsaturated ester substrates and so avoids the need for in situ oxidation. Reaction via an in situ generated α,β-unsaturated acyl ammonium species derived from an α,β-unsaturated aryl ester and an isothiourea was postulated.15 In such processes, the aryloxide liberated upon isothiourea acylation can fulfil multiple roles, including that of Brønsted base.10,15,16 The hydrogen bond donor ability of electron-deficient thioureas, commonly exploited in catalytic applications, is concomitant with a pKa that facilitates deprotonation by the catalytically liberated aryloxide,17 activating this towards thia-Michael addition with an α,β-unsaturated acyl ammonium intermediate 23 (Scheme 2B). This simple methodology would require no additional reagents and allow access to iminothiazinanone heterocycles 25 that possess a stereogenic centre. Furthermore, the incorporation of a 2-substituted aryl substituent on the thiourea would generate atropisomeric products 24 that also contain a stereogenic centre, while extension to selenoureas would give access to chiral iminoselenazinanones for the first time.


image file: d5sc02435h-s2.tif
Scheme 2 (A) NHC-catalysed synthesis of atropisomeric iminothiazinones. (B) This work: isothiourea catalysed preparation of point and axially chiral heterocycles.

2. Results and discussion

2.1 Development of model process with symmetrical thioureas

Optimisation began with the reaction of α,β-unsaturated para-nitrophenyl (PNP) ester 27 with Schreiner's thiourea 26 and (2S,3R)-HyperBTM 21 (20 mol%) in MeCN (0.1 M) (Table 1, entry 1) giving product 33 in quantitative yield with 78[thin space (1/6-em)]:[thin space (1/6-em)]22 er. Variation of the reaction solvent (see ESI for full range of solvents trialled) indicated that most common solvents gave >95% conversion to product. CH2Cl2 (entry 2) gave reduced enantioselectivity (66[thin space (1/6-em)]:[thin space (1/6-em)]34 er), while EtOAc, DMF, and THF (entries 3–5) all gave improved enantioselectivity. The use of 2-MeTHF and i-PrOAc also led to high product enantioselectivity (entries 6 and 7) with further optimisation using THF. Catalyst variation and loading were next tested. While HyperBTM 20 (5 mol%) gave 33 in quantitative yield and 91[thin space (1/6-em)]:[thin space (1/6-em)]9 er, the use of alternative isothioureas 8 and 34 led to only moderate selectivity (entries 8–10). Variation of the nucleofuge within the α,β-unsaturated ester or anhydride reaction component was next trialled. Among the range of aryl esters tested (entries 11–14) the observed product enantioselectivity correlated with the pKa of the corresponding phenol.18 Pentafluorophenyl (PFP) ester 28 (entry 11) with the most acidic phenol (pKa 5.53 in H2O) gave the greatest enantioselectivity (93[thin space (1/6-em)]:[thin space (1/6-em)]7 er) at 90% conversion, while 3,5-bis(trifluoromethyl)phenyl ester 31 (entry 14) with the least acidic phenol (pKa 8.26 in H2O) gave the lowest observed product enantioselectivity (90[thin space (1/6-em)]:[thin space (1/6-em)]10 er). The use of in situ generated pivalic anhydride 32 did not lead to improved selectivity (entry 15, 83%, 90[thin space (1/6-em)]:[thin space (1/6-em)]10 er). The use of PFP ester 28 at 0 °C led to optimal reaction selectivity, giving product 33 in 95% isolated yield and 95[thin space (1/6-em)]:[thin space (1/6-em)]5 er (entry 16).
Table 1 Initial optimisation

image file: d5sc02435h-u1.tif

Entry Acyl donor Catalyst (mol%) Solvent Yielda erb
a Yield determined by 1H NMR analysis relative to internal standard 1,3,5-trimethoxybenzene (isolated yield in parentheses).b er measured by HPLC analysis on a chiral stationary phase.c Performed at 0 °C.
1 27 21 (20) MeCN Quant. 78[thin space (1/6-em)]:[thin space (1/6-em)]22
2 27 21 (20) CH2Cl2 Quant. 66[thin space (1/6-em)]:[thin space (1/6-em)]34
3 27 21 (20) EtOAc Quant. 90[thin space (1/6-em)]:[thin space (1/6-em)]10
4 27 21 (20) DMF Quant. 84[thin space (1/6-em)]:[thin space (1/6-em)]16
5 27 21 (20) THF Quant. 91[thin space (1/6-em)]:[thin space (1/6-em)]9
6 27 21 (20) 2-MeTHF Quant. 91[thin space (1/6-em)]:[thin space (1/6-em)]9
7 27 21 (20) i-PrOAc 98 90[thin space (1/6-em)]:[thin space (1/6-em)]10
8 27 21 (5) THF Quant. 91[thin space (1/6-em)]:[thin space (1/6-em)]9
9 27 8 (5) THF 96 84[thin space (1/6-em)]:[thin space (1/6-em)]16
10 27 34 (5) THF 98 33[thin space (1/6-em)]:[thin space (1/6-em)]67
11 28 21 (5) THF 90 93[thin space (1/6-em)]:[thin space (1/6-em)]7
12 29 21 (5) THF Quant. 92[thin space (1/6-em)]:[thin space (1/6-em)]8
13 30 21 (5) THF 96 92[thin space (1/6-em)]:[thin space (1/6-em)]8
14 31 21 (5) THF 95 90[thin space (1/6-em)]:[thin space (1/6-em)]10
15 32 21 (5) THF 83 90[thin space (1/6-em)]:[thin space (1/6-em)]10
16c 28 21 (5) THF (95) 95[thin space (1/6-em)]:[thin space (1/6-em)]5


2.2 Scope of the developed process with symmetrical thioureas

With optimised conditions for the generation of model heterocycle 33 developed, the scope and limitations of this process were investigated through variation of the N-aryl substituents within the thiourea and the β-substituent within a range of α,β-unsaturated esters (Scheme 3). Using β-trifluoromethyl α,β-unsaturated PFP ester 28 as standard, the incorporation of 4-F3CC6H4 substituents within the bis-N-arylthiourea led to reduced product conversion at 0 °C, but at RT gave 35 in 79% yield with excellent enantioselectivity (95[thin space (1/6-em)]:[thin space (1/6-em)]5 er). The incorporation of strongly electron-withdrawing 4-O2NC6H4 and 4-NCC6H4 N-aryl substituents within the thiourea showed similar reactivity to that of the model substrate, giving products 36 and 37 in excellent yields (95% and 83%) and enantioselectivity (98[thin space (1/6-em)]:[thin space (1/6-em)]2 er and 97[thin space (1/6-em)]:[thin space (1/6-em)]3 er respectively). The absolute configuration within (6R,Z)-37 was confirmed by single crystal X-ray analysis, with the configuration within all other products assigned by analogy.19 Further work probed the scope of the α,β-unsaturated ester Michael acceptors using bis-4-O2NC6H4 substituted thiourea. The incorporation of perhalogenated β-CF2Cl- and perfluoroethyl β-substituents gave products 38 and 39 in good yield (78% and 59%) with excellent enantioselectivity (both 98[thin space (1/6-em)]:[thin space (1/6-em)]2 er).20 While β-CF2H and β-CH2F substituents within the ester were tolerated giving products 40 and 41 respectively, reduced enantioselectivity was observed in both cases (80[thin space (1/6-em)]:[thin space (1/6-em)]20 er and 83[thin space (1/6-em)]:[thin space (1/6-em)]17 er respectively). To further test this methodology the use of β-alkyl substituted esters was probed as these generally show moderate reactivity in reactions involving the corresponding α,β-unsaturated acyl ammonium species. In this context, 42–45 were generated in excellent yields and promising enantioselectivity (86[thin space (1/6-em)]:[thin space (1/6-em)]14 er to 94[thin space (1/6-em)]:[thin space (1/6-em)]6 er) although these reactions were carried out at room temperature and required higher catalyst loadings (20–30 mol%). Unfortunately the use of β-aryl substituted PFP esters (β-aryl = Ph, 4-NO2C6H4) were unreactive in this process using bis-4-O2NC6H4 substituted thiourea and represent a limitation of this methodology.
image file: d5sc02435h-s3.tif
Scheme 3 All yields are isolated; all er ratios determined by HPLC analysis on a chiral stationary phase; [a] 10 mol% 21 used at 0 °C; [b] 20 mol% 21 used at RT; [c] 30 mol% 21 used at RT.

2.3 Application to unsymmetrical thio- and selenoureas

Further work extended this process to the use of unsymmetrical thio- and selenoureas (Scheme 4). In principle these starting materials could lead to the formation of regioisomeric products, and so to bias regioselectivity variation in the electronic properties of the nitrogen substituents was used to dictate reactivity. Initial studies considered the utility of a range of thioureas bearing electronically distinct N-substituents with PFP ester 28 (Scheme 4A). In each case a single regioisomeric product was observed but with varying levels of enantioselectivity. Reaction to generate 46 required 20 mol% catalyst and 48 h at RT to reach completion, giving 46 in 77% yield but in racemic form. Replacing the N-4-nitrophenyl substituent with an N-tosyl substituent led to improved reactivity, giving 47 in high yield (88%) using 5 mol% catalyst at 0 °C but with moderate enantioselectivity (67[thin space (1/6-em)]:[thin space (1/6-em)]33 er). Using an N-benzyl substituted derivative gave 48 in 98% yield with improved enantioselectivity (78[thin space (1/6-em)]:[thin space (1/6-em)]22 er), while N-aryl variants gave 49–51 in high yields (79 to 91%) and good enantioselectivity (88[thin space (1/6-em)]:[thin space (1/6-em)]12 to 91[thin space (1/6-em)]:[thin space (1/6-em)]9 er). Variation within the electron-withdrawing substituents was also tolerated, with N-4-methoxybenzenesulfonyl and N-trifluoroacetyl variants giving 52 and 53 with high enantioselectivity (91[thin space (1/6-em)]:[thin space (1/6-em)]9 er and 96[thin space (1/6-em)]:[thin space (1/6-em)]4 er). As a further test, it was postulated that regioselectivity could be achieved through a steric bias within two electron-withdrawing N-aryl substituents. Using the sterically demanding N-2,4,6-trichlorophenyl substituent to disfavour cyclisation led to a single regioisomer of 54 in excellent yield (92%) with excellent enantioselectivity (98[thin space (1/6-em)]:[thin space (1/6-em)]2 er). The constitution of 46 and 54 were confirmed by 1H–15N HMBC (see ESI for further information).
image file: d5sc02435h-s4.tif
Scheme 4 All yields are isolated; all er ratios determined by HPLC analysis on a chiral stationary phase; [a] 20 mol% 21 used at RT; [b] 10 mol% 21 used at 0 °C; [c] 10 mol% 21 used at RT; [d] 30 mol% 21 used at RT.

Subsequent work extended this process to the use of unsymmetrical selenoureas to give iminoselenazinanones (Scheme 4B). Pleasingly, N-Bz,N-Ph-substituted selenourea led to 55 in excellent yield and enantioselectivity (97%, 99[thin space (1/6-em)]:[thin space (1/6-em)]1 er). In this series, alternative N-aryl substitution was investigated, with the incorporation of the antibiotic sulfamethoxazole within a selenourea giving 56 in good yield (61%) with excellent 98[thin space (1/6-em)]:[thin space (1/6-em)]2 er. 4-MeOC6H4 substitution gave 57 with excellent yield (97%) and enantioselectivity (98[thin space (1/6-em)]:[thin space (1/6-em)]2 er). Variation within the N-benzoyl substitution was also well tolerated, with 4-MeOC6H4, 4-BrC6H4, and 4-O2NC6H4 substituents giving products 58, 59, and 60 respectively in high yields (80 to 92%) and excellent enantioselectivity (97[thin space (1/6-em)]:[thin space (1/6-em)]3 to 99[thin space (1/6-em)]:[thin space (1/6-em)]1 er). To demonstrate that the selenoureas were reactive with alternative Michael acceptors, two representative examples were chosen. Reaction with ethyl ester substituted Michael acceptor at RT gave product 61 in 71% yield with high enantioselectivity (93[thin space (1/6-em)]:[thin space (1/6-em)]7 er). The reactivity of selenoureas was pushed to a limit with the β-methyl substituted Michael acceptor with the use of 30 mol% 21 at RT for 48 h giving 62 in low 29% yield (87[thin space (1/6-em)]:[thin space (1/6-em)]13 er) and so represents a limit of this methodology. Further work investigated the use of thioureas and a selenourea bearing ortho-substituted aryl substituents to give atropisomeric products (Scheme 4C). Initial studies showed that chalcogenoureas bearing an ortho-substituted aryl substituent give atropisomeric products 63–66 upon treatment with pentafluorophenyl acrylate and HyperBTM. Although the isolated yields of products 63–66 were good to excellent (71 to 98%), the enantioselectivities were only moderate (73[thin space (1/6-em)]:[thin space (1/6-em)]27 er to 87[thin space (1/6-em)]:[thin space (1/6-em)]13 er). The configurational stability of 63–66 was determined as described by Armstrong,21 with increasing barriers to rotation observed with increased steric hindrance of the 2-aryl substituent (Ph < i-Pr < t-Bu), and with Se-containing 66 possessing a lower barrier than its sulfur analogue. Intrigued by these observations, application to products containing both a stereogenic centre and a stereogenic axis were investigated. Using the previously developed conditions, treatment of a thiourea bearing an ortho-iso-propyl substituent with PFP ester 28 and HyperBTM gave 67 in excellent yield (94%) and stereoselectivity (92[thin space (1/6-em)]:[thin space (1/6-em)]8 dr, 98[thin space (1/6-em)]:[thin space (1/6-em)]2 er). Similarly, thioureas bearing ortho-phenyl and ortho-tert-butyl substituents gave the corresponding products 68 and 69 with excellent stereoselectivity. Using a selenourea bearing an ortho-iso-propyl substituent gave 70 in 86% yield and excellent stereoselectivity (>95[thin space (1/6-em)]:[thin space (1/6-em)]5 dr, 98[thin space (1/6-em)]:[thin space (1/6-em)]2 er). The generality of the enantio- and atropselective methodology was further investigated by reaction of an ortho-iso-propyl substituted thiourea with three alternative β-substituted α,β-unsaturated esters. CF2Cl substitution gave 71 in excellent yield (89%) and stereoselectivity (>95[thin space (1/6-em)]:[thin space (1/6-em)]5 dr, 98[thin space (1/6-em)]:[thin space (1/6-em)]2 er), while pleasingly β-alkyl substituted Michael acceptors were also tolerated, giving 72 and 73 with high yields (82% and 86%) and stereoselectivity (88[thin space (1/6-em)]:[thin space (1/6-em)]12 dr, 92[thin space (1/6-em)]:[thin space (1/6-em)]8 er and 92[thin space (1/6-em)]:[thin space (1/6-em)]8 dr, 95[thin space (1/6-em)]:[thin space (1/6-em)]5 er). The absolute configuration of (6R,Ra,Z)-67 was confirmed by single crystal X-ray crystallography, with the relative configuration within 72 confirmed by 1H NOESY NMR analysis (see ESI Section 13 for further information). The configuration within all other products was assigned by analogy.

2.4 Mechanistic analysis and control studies

2.4.1 Role of aryloxide. Subsequent studies considered the role of aryloxide liberated upon catalyst acylation and its ability to promote thia-Michael addition through thiourea activation by deprotonation. Following the method described by Yatsimirsky, the feasibility of this deprotonation process was interrogated by NMR titration.22 Schreiner's thiourea 26 was titrated with NBu4OPNP in d8-THF (0.1 M) (Scheme 5A). Upon addition of only 20 mol% of NBu4OPNP the NH proton resonance (H, δH = 9.7 ppm) disappeared which is indicative of deprotonation.22 This is consistent with the reported pKa of 26 (pKa 8.5 in DMSO) and 4-nitrophenol (pKa 10.8 in DMSO).17 As the concentration of NBu4OPNP was incrementally increased (to 100 mol%), the C(4)-ortho proton signal (H) experienced a significant upfield shift (from δH = 7.8 ppm to δH = 7.4 ppm) consistent with the equilibrium shifting towards the deprotonated thiourea. This titration suggests that under the reaction conditions the thiourea is likely to be deprotonated by a catalytically liberated aryloxide. Further work considered if this deprotonation could promote a racemic background reaction to generate 33. In the absence of any isothiourea catalyst or base, no conversion to 33 was observed, but on addition of 10 mol% NBu4OPNP the product 33 was formed in 99% yield by 1H NMR analysis (Scheme 5B). While this experiment does not give quantitative data regarding the rate of the background reaction compared to the isothiourea catalysed reaction, it highlights the benefit of the aryloxide base being generated catalytically and in proximity to the proposed α,β-unsaturated acyl ammonium intermediate. Further titration studies (see ESI Section 15 for further information) indicate that the thiourea does not hydrogen bond and activate the ester carbonyl group.
image file: d5sc02435h-s5.tif
Scheme 5 (A) Determining role of aryloxide by 1H NMR titration. (B) Aryloxide catalysed racemic background reaction. (C) Probing reversibility of the thia-Michael addition by crossover experiment.
2.4.2 Reversibility of thia-Michael addition. The potential reversibility of the proposed S- or Se-conjugate addition was next probed. The reaction of thiourea 74 and pentafluorophenyl acrylate 75 was stopped after 5 h, giving pre-cyclised thia-Michael addition product 76 in 55% yield. In this case, the tert-butyl substituent presumably hinders cyclisation, allowing aryloxide catalyst turnover (Scheme 5C). To investigate the potential reversibility of the thia-Michael addition, 76 was reacted with (±)-HyperBTM 21 and thiourea 77, giving crossover product 78 in 51% isolated yield. The formation of 78 indicates that thia-Michael addition is reversible under the reaction conditions, similar to the mechanism proposed by Matsubara (Scheme 1B).9

2.5 Computational analysis and proposed mechanism

Based upon these observations a catalytic cycle for this process can be outlined, with the origins of stereocontrol probed by DFT calculations performed at the M06- 2XSMD(THF)/def2-TZVP//M06-2XSMD(THF)/def2-SVP level of theory using Gaussian16 (Scheme 6). Using the symmetric bis-4-O2NC6H4 substituted thiourea and PFP ester 28 as a model, N-acylation of the PFP ester 28 by the Lewis base (2R,3S)-HyperBTM 21 generates the corresponding acyl ammonium ion pair I with a stabilising 1,5-O⋯S chalcogen bonding interaction (nOσ*S–C, E(2) = 6.0 kcal mol−1, described by NBO second-order perturbation analysis).23–32 This ensures coplanarity between the 1,5-O- and S-atoms and provides a conformational bias.26 Deprotonation of the thiourea promotes thia-Michael addition via TS1 to give II. Subsequent proton transfer gives III, with intramolecular cyclisation via TS2 generating IV. Catalyst release is promoted by collapse of the tetrahedral intermediate IV via TS3 generating the heterocyclic product 36 in high enantioselectivity (Scheme 6A).
image file: d5sc02435h-s6.tif
Scheme 6 (A) Proposed catalytic cycle. (B) DFT analysis of the stereodetermining transition state. (C) Computed reaction profile leading to enantiomeric products. M06-2XSMD(THF)/def2-TZVP//M06-2XSMD(THF)/def2-SVP Gibbs free energies (ΔG273) shown in kcal mol−1.

Stereoselectivity in the initial Michael addition considered the anionic thiourea nucleophile approaching either the Re- or Si-face of acyl ammonium intermediate I (Scheme 6B and C). Approach to the Si-face is hindered by the stereodirecting phenyl group within (2S,3R)-HyperBTM leading to favoured addition to the Re-face (TS1, ΔΔG273 = 1.3 kcal mol; 92[thin space (1/6-em)]:[thin space (1/6-em)]8 er (R)). After proton transfer, the pendant nitrogen nucleophile can initiate lactamisation (TS2) to generate the tetrahedral cyclised product adduct, before catalyst turnover (TS3) to generate the product. These diastereomeric lactamisation transition states (TS2) were computed to be higher in energy than the thia-Michael addition transition states, TS1 (2.5 kcal mol−1 higher for the (R)-pathway), due to the geometric constraints of ring-closure. While the stereogenic centre is formed in the Michael addition (TS1), computation suggested that the observed enantioselectivity arises from the energetic difference between the higher energy lactamisation transition states (TS2). Turnover of the catalyst (TS3) is strongly exergonic and proceeds through a lower energy transition state compared to TS2. Every step of the reaction (Scheme 6C) was computed to be reversible up until catalyst turnover (TS3), in agreement with the experimentally observed crossover and retro-Michael addition in Scheme 5C. Lactamisation (TS2) was computed to be enantiodetermining with the favoured (R)-pathway proceeding through nucleophilic addition from the less hindered Si-face of the acyl ammonium. The disfavoured (S)-pathway is initially positioned to lactamise from the hindered Re-face of the acyl ammonium (TS2-(S)′, ΔΔG273 = +5.3 kcal mol−1). However, rotation allows lactamisation from the less hindered Si-face (TS2-(S), ΔΔG273 = +2.7 kcal mol−1).

3. Conclusion

In conclusion, a range of symmetrical and unsymmetrical thioureas and selenoureas can be utilised in an isothiourea-catalysed protocol to generate iminothia- and iminoselenazinanone heterocycles with high enantioselectivity (up to 99[thin space (1/6-em)]:[thin space (1/6-em)]1 er). Notably, unsymmetrical thio- and selenoureas containing ortho-substituted N-aryl substituents generate atropisomeric products. This allows the generation of iminothia- and iminoselenazinanone heterocyclic products containing both point and axially chiral stereogenic elements with excellent stereocontrol (up to >95[thin space (1/6-em)]:[thin space (1/6-em)]5 dr and 98[thin space (1/6-em)]:[thin space (1/6-em)]2 er) for the first time. Mechanistic investigations indicate that catalytically liberated aryloxide can deprotonate an electron-deficient thiourea, while a crossover experiment indicates the reversibility of the thia-Michael addition. Extensive computational analysis has identified the factors leading to enantioselectivity within this process, with stereocontrol shown to arise from the lactamisation step within the catalytic cycle.

Data availability

All data (experimental procedures, characterisation data and cartesian coordinates for all DFT calculations) that support the findings of this study are available within the article and its ESI. The research data supporting this publication can be accessed from “Isothiourea Catalysed Enantioselective Generation of Point and Axially Chiral Iminothia- and Iminoselenazinanones”. Pure ID: 314763684. University of St Andrews Research Portal “PURE” [https://doi.org/10.17630/3ffd33a9-50ae-441d-bd88-6a3343351c63].

Author contributions

Alastair J. Nimmo – conceptualization, investigation, writing. Alister S. Goodfellow – formal analysis, investigation writing-review and editing. Jacob T. Guntley – investigation. Aidan P. McKay, David. B. Cordes, investigation (X-ray analysis). Michael Bühl – investigation, funding acquisition, writing-review and editing. Andrew D. Smith – conceptualization, funding acquisition, project administration writing-review and editing.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

We acknowledge the EaSICAT Centre for Doctoral Training (AJN, ASG) for funding. ADS thanks the EPSRC Programme Grant “Boron: Beyond the Reagent” (EP/W007517) for support. MB thanks EaStCHEM and the School of Chemistry for support. Computations were performed on a local HPC cluster maintained by Dr H. Früchtl. To meet institutional and research funder open access requirements, any accepted manuscript arising shall be open access under a Creative Commons Attribution (CC BY) reuse licence with zero embargo.

Notes and references

  1. V. B. Birman and X. Li, Org. Lett., 2006, 8, 1351–1354 Search PubMed.
  2. For reviews see: (a) J. Merad, J.-M. Pons, O. Chuzel and C. Bressy, Eur. J. Org Chem., 2016, 2016, 5589–5610 Search PubMed; (b) J. E. Taylor, S. D. Bull and J. M. J. Williams, Chem. Soc. Rev., 2012, 41, 2109–2121 Search PubMed; (c) V. B. Birman, Aldrichimica Acta, 2016, 49, 23–41 Search PubMed.
  3. For reviews see: (a) S. Vellalath and D. Romo, Angew. Chem., Int. Ed., 2016, 55, 13934–13943 Search PubMed; (b) J. Bitai, M. T. Westwood and A. D. Smith, Org. Biomol. Chem., 2021, 19, 2366–2384 Search PubMed.
  4. For reviews see (a) L. C. Morrill and A. D. Smith, Chem. Soc. Rev., 2014, 43, 6214–6226 Search PubMed; (b) C. McLaughlin and A. D. Smith, Chem.–Eur. J., 2021, 27, 1533–1555 Search PubMed.
  5. (a) L. S. Vogl, P. Mayer, R. Robiette and M. Waser, Angew. Chem., Int. Ed., 2024, 63, e202315345 Search PubMed; (b) M. Piringer, M. Hofer, L. S. Vogl, P. Mayer and M. Waser, Adv. Synth. Catal., 2024, 366, 2115–2122 Search PubMed.
  6. (a) C. I. Sheppard, J. L. Taylor and S. L. Wiskur, Org. Lett., 2011, 13, 3794–3797 Search PubMed; (b) R. W. Cark, T. M. Deaton, Y. Zhang, M. I. Moore and S. L. Wiskur, Org. Lett., 2013, 15, 6132–6135 Search PubMed; (c) T. Zhang, B. K. Redden and S. L. Wiskur, Eur. J. Org Chem., 2019, 30, 4827–4831 Search PubMed; (d) L. Wang, R. K. Akhani and S. L. Wiskur, Org. Lett., 2015, 17, 2408–2411 Search PubMed; (e) T. Hu, Y. Zhang, W. Wang, Q. Li, L. Huang, J. Gao, Y. Kuang, C. Zhao, S. Zhou, Z. Su and Z. Song, J. Am. Chem. Soc., 2024, 146, 23092–23102 Search PubMed.
  7. (a) A. Matviitsuk, M. D. Greenhalgh, D.-J. B. Antúnez, A. M. Z. Slawin and A. D. Smith, Angew. Chem., Int. Ed., 2017, 56, 12282–12287 Search PubMed; (b) E. R. T. Robinson, D. M. Walden, C. Fallan, M. D. Greenhalgh, P. H.-Y. Cheong and A. D. Smith, Chem. Sci., 2016, 7, 6919–6927 Search PubMed; (c) G. Liu, M. E. Shirley, K. N. Van, R. L. McFarlin and D. Romo, Nat. Chem., 2013, 5, 1049–1057 Search PubMed; (d) E. R. T. Robinson, C. Fallan, C. Simal, A. M. Z. Slawin and A. D. Smith, Chem. Sci., 2013, 4, 2193–2200 Search PubMed; (e) A. Matviitsuk, J. E. Taylor, D. B. Cordes, A. M. Z. Slawin and A. D. Smith, Chem.–Eur. J., 2016, 22, 17748–17757 Search PubMed; (f) E. R. T. Robinson, A. B. Frost, P. Elías-Rodríguez and A. D. Smith, Synthesis, 2017, 49, 409–423 Search PubMed.
  8. Y. Fukata, K. Asano and S. Matsubara, J. Am. Chem. Soc., 2015, 137, 5320–5323 Search PubMed.
  9. Y. Fukata, K. Yao, R. Miyaji, K. Asano and S. Matsubara, J. Org. Chem., 2017, 82, 12655–12668 Search PubMed.
  10. N. A. Ahlemeyer and V. B. Birman, Org. Lett., 2016, 18, 3454–3457 Search PubMed.
  11. J. E. Lapetaje, C. M. Young, C. Shu and A. D. Smith, Chem. Commun., 2022, 58, 6886–6889 Search PubMed.
  12. For selected review articles, see: (a) E. Kumarasamy, R. Raghunathan, M. P. Sibi and J. Sivaguru, Chem. Rev., 2015, 115, 11239–11300 Search PubMed; (b) B. Zilate, A. Castrogiovanni and C. Sparr, ACS Catal., 2018, 8, 2981–2988 Search PubMed; (c) V. Corti and G. Bertuzzi, Synthesis, 2020, 52, 2450–2468 Search PubMed; (d) T.-Z. Li, S.-J. Liu, W. Tan and F. Shi, Chem.–Eur. J., 2020, 26, 15779–15792 Search PubMed; (e) J. K. Cheng, S.-H. Xiang, S. Li, L. Ye and B. Tan, Chem. Rev., 2021, 121, 4805–4902 Search PubMed; (f) R. Song, Y. Xie, Z. Jin and Y. R. Chi, Angew. Chem., Int. Ed., 2021, 60, 26026–26037 Search PubMed; (g) G. Hedouin, S. Hazra, F. Gallou and S. Handa, ACS Catal., 2022, 12, 4918–4937 Search PubMed; (h) X. Zhang, K. Zhao and Z. Gu, Acc. Chem. Res., 2022, 55, 1620–1633 Search PubMed; (i) O. F. B. Watts, J. Berreur, B. S. L. Collins and J. Clayden, Acc. Chem. Res., 2022, 55, 3362–3375 Search PubMed.
  13. For selected reviews on C–N atropisomerism, see: (a) I. Takahashi, Y. Suzuki and O. Kitagawa, Org. Prep. Proced. Int., 2014, 46, 1–23 Search PubMed; (b) J. Frey, S. Choppin, F. Colobert and J. Wencel-Delord, Chimia, 2020, 74, 883–889 Search PubMed; (c) O. Kitagawa, Acc. Chem. Res., 2021, 54, 719–730 Search PubMed; (d) Y.-J. Wu, G. Liao and B.-F. Shi, Green Synth. Catal., 2022, 3, 117–136 Search PubMed; (e) P. Rodríguez-Salamanca, R. Fernández, V. Hornillos and J. M. Lassaletta, Chem.–Eur. J., 2022, 28, e202104442 Search PubMed; (f) J. S. Sweet and P. C. Knipe, Synthesis, 2022, 54, 2119–2132 Search PubMed; (g) G.-J. Mei, W. L. Koay, C.-Y. Guan and Y. Lu, Chem, 2022, 8, 1855–1893 Search PubMed; (h) X. Xiao, B. Chen, Y.-P. Yao, H.-J. Zhou, X. Wang, N.-Z. Wang and F.-E. Chen, Molecules, 2022, 27, 6583 Search PubMed; (i) A. D. G. Campbell and R. J. Armstrong, Synthesis, 2023, 55, 2427–2438 Search PubMed.
  14. T. Li, C. Mou, P. Qi, X. Peng, S. Jiang, G. Hao, W. Xue, S. Yang, L. Hao, Y. R. Chi and Z. Jin, Angew. Chem., Int. Ed., 2021, 60, 9362–9367 Search PubMed.
  15. (a) J. Wu, C. M. Young, A. A. Watts, A. M. Z. Slawin, G. R. Boyce, M. Bühl and A. D. Smith, Org. Lett., 2022, 24, 4040–4045 Search PubMed; (b) J. Bitai, A. J. Nimmo, A. M. Z. Slawin and A. D. Smith, Angew. Chem., Int. Ed., 2022, 61, e202202621 Search PubMed; (c) M. D. Greenhalgh, S. Qu, A. M. Z. Slawin and A. D. Smith, Chem. Sci., 2018, 9, 4909–4918 Search PubMed; (d) D. Yuan, A. S. Goodfellow, K. Kasten, Z. Duan, T. Kang, D. B. Cordes, A. P. McKay, M. Bühl, G. R. Boyce and A. D. Smith, Chem. Sci., 2023, 14, 14146–14156 Search PubMed.
  16. C. McLaughlin, A. M. Z. Slawin and A. D. Smith, Angew. Chem., Int. Ed., 2019, 58, 15111–15119 Search PubMed.
  17. G. Jakab, C. Tancon, Z. Zhang, K. M. Lippert and P. R. Schreiner, Org. Lett., 2012, 14, 1724–1727 Search PubMed.
  18. (a) A. Kütt, V. Movchun, T. Rodima, T. Dansauer, E. B. Rusanov, I. Leito, I. Kaljurand, J. Koppel, V. Pihl, I. Koppel, G. Ovsjannikov, L. Toom, M. Mishima, M. Medebielle, E. Lork, G.-V. Röschenthaler, I. A. Koppel and A. A. Kolomeitsev, J. Org. Chem., 2008, 73, 2607–2620 Search PubMed; (b) S. Espinosa, E. Bosch and M. Roses, J. Chromatogr. A, 2002, 964, 55–66 Search PubMed; (c) J. Han and F. M. Tao, J. Phys. Chem. A, 2006, 110, 257–263 Search PubMed.
  19. Crystallographic data for compounds 37, 55 and 67 have been deposited with the Cambridge Crystallographic Data Centre under deposition numbers 2410987, 2410988 and 2410989 respectively..
  20. In the perfluoroethyl case a regioisomeric product was isolated in 10% yield as a racemate. For further information and structural assignment see ESI Section 10..
  21. J.-P. Heeb, J. Clayden, M. D. Smith and R. J. Armstrong, Nat. Protoc., 2023, 18, 2745–2771 Search PubMed.
  22. C. Pérez-Casas and A. K. Yatsimirsky, J. Org. Chem., 2008, 73, 2275–2284 Search PubMed.
  23. V. B. Birman, X. Li and Z. Han, Org. Lett., 2007, 9, 37–40 Search PubMed.
  24. P. Liu, X. Yang, V. B. Birman and K. N. Houk, Org. Lett., 2012, 14, 3288–3291 Search PubMed.
  25. M. E. Abbasov, B. M. Hudson, D. J. Tantillo and D. Romo, J. Am. Chem. Soc., 2014, 136, 4492–4495 Search PubMed.
  26. E. R. T. Robinson, D. M. Walden, C. Fallan, M. D. Greenhalgh, P. H.-Y. Cheong and A. D. Smith, Chem. Sci., 2016, 7, 6919–6927 Search PubMed.
  27. M. D. Greenhalgh, S. M. Smith, D. M. Walden, J. E. Taylor, Z. Brice, E. R. T. Robinson, C. Fallan, D. B. Cordes, A. M. Z. Slawin, H. C. Richardson, M. A. Grove, P. H.-Y. Cheong and A. D. Smith, Angew. Chem., Int. Ed., 2018, 57, 3200–3206 Search PubMed.
  28. C. M. Young, A. Elmi, D. J. Pascoe, R. K. Morris, C. McLaughlin, A. M. Woods, A. B. Frost, A. de la Houpliere, K. B. Ling, T. K. Smith, A. M. Z. Slawin, P. H. Willoughby, S. L. Cockroft and A. D. Smith, Angew. Chem., Int. Ed., 2020, 59, 3705–3710 Search PubMed.
  29. Y. Nagao, S. Miyamoto, M. Miyamoto, H. Takeshige, K. Hayashi, S. Sano, M. Shiro, K. Yamaguchi and Y. Sei, J. Am. Chem. Soc., 2006, 128, 9722–9729 Search PubMed.
  30. B. R. Beno, K.-S. Yeung, M. D. Bartberger, L. D. Pennington and N. A. Meanwell, J. Med. Chem., 2015, 58, 4383–4438 Search PubMed.
  31. D. J. Pascoe, K. B. Ling and S. L. Cockroft, J. Am. Chem. Soc., 2017, 139, 15160–15167 Search PubMed.
  32. For a selection of other enantioselective processes that involve chalcogen bonding interactions see: (a) M. Breugst and J. J. Koenig, Eur. J. Org Chem., 2020, 2020, 5473–5487 Search PubMed; (b) A. J. Mukherjee, S. S. Zade, H. B. Singh and R. B. Sunoj, Chem. Rev., 2010, 110, 4357–4416 Search PubMed; (c) K.-i. Fujita, M. Iwaoka and S. Tomoda, Chem. Lett., 2006, 23, 923–926 Search PubMed; (d) K.-i. Fujita, K. Murata, M. Iwaoka and S. Tomoda, J. Chem. Soc., Chem. Commun., 1995, 1995, 1641–1642 Search PubMed; (e) K.-i. Fujita, K. Murata, M. Iwaoka and S. Tomoda, Tetrahedron Lett., 1995, 36, 5219–5222 Search PubMed; (f) T. Wirth, Angew. Chem., Int. Ed., 1995, 34, 1726–1728 Search PubMed; (g) C. Bleiholder, R. Gleiter, D. B. Werz and H. Köppel, Inorg. Chem., 2007, 46, 2249–2260 Search PubMed; (h) S. Kolb, G. A. Oliver and D. B. Werz, Angew. Chem., Int. Ed., 2020, 59, 22306–22310 Search PubMed; (i) S. Benz, J. López-Andarias, J. Mareda, N. Sakai and S. Matile, Angew. Chem., Int. Ed., 2017, 56, 812–815 Search PubMed; (j) P. Wonner, L. Vogel, M. Düser, L. Gomes, F. Kniep, B. Mallick, D. B. Werz and S. M. Huber, Angew. Chem., Int. Ed., 2017, 56, 12009–12012 Search PubMed; (k) W. Wang, H. Zhu, S. Liu, Z. Zhao, L. Zhang, J. Hao and Y. Wang, J. Am. Chem. Soc., 2019, 141, 9175–9179 Search PubMed.

Footnote

Electronic supplementary information (ESI) available. CCDC 2410987–2410989. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5sc02435h

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.