Catalytic stereodivergent and simultaneous construction of axial and point chirality

Wen-Tao Wang a, Sen Zhang a, Wenxuan Lin b, Zhang-Hong Luo a, Dan Hu a, Fen Huang a, Ruopeng Bai *b, Yu Lan b, Linghui Qian *a and Jia-Yu Liao *ac
aNational Key Laboratory of Advanced Drug Delivery and Release Systems, College of Pharmaceutical Sciences, and Hangzhou Institute of Innovative Medicine, Zhejiang University, Hangzhou 310058, China. E-mail: lhqian@zju.edu.cn; jyliao@zju.edu.cn
bSchool of Chemistry and Chemical Engineering, Chongqing Key Laboratory of Theoretical and Computational Chemistry, Chongqing University, Chongqing 401331, China. E-mail: ruopeng@cqu.edu.cn
cInnovation Institute for Artificial Intelligence in Medicine of Zhejiang University, Hangzhou 310018, China

Received 13th February 2024 , Accepted 30th April 2024

First published on 1st May 2024


Abstract

Atropisomers possessing additional stereogenic centres have found wide applications in various fields; however, the development of catalytic asymmetric reactions for their preparation through a single transformation has been much less studied. In particular, no example of catalytic diastereodivergent and enantioselective simultaneous construction of axial and point chirality has been disclosed, which stands for an unmet synthetic challenge. Herein, we report the realization of such a goal for the first time. An unprecedented silver-catalyzed desymmetric [3 + 2] cycloaddition reaction of N-quinazolinone maleimides with isocyanoacetates was developed, producing a series of bicyclic N-heterocycles bearing a remote chiral N–N axis and three contiguous stereogenic carbon centres with high efficiencies and high-to-excellent stereoselectivities. Most strikingly, an interesting diastereodivergent synthesis was also achieved allowing facile access to both endo- and exo-cycloadducts. DFT calculations revealed that the stereoselectivity is mainly controlled by the coordination type of the ligand, ligand–substrate non-bonding interaction, and distortion of the annulation transition state.


Introduction

Owing to the profound applications of atropisomers,1–3 developing effective approaches for their preparation concerning structural complexity and diversity has attracted great attention from the chemical science community over the past few decades.4 Along these lines, incorporating a second chiral element into atropisomers would certainly enrich their topological chemical space and thus open up opportunities to discover new chemical entities with distinct application potential.5 In particular, atropisomers featuring additional stereogenic centres have found widespread applications as bioactive molecules in drug discovery or as chiral organocatalysts/ligands in asymmetric catalysis (Scheme 1a).6 Therefore, the implementation of efficient strategies to access such scaffolds is in high demand and has become an emerging research topic.
image file: d4qo00294f-s1.tif
Scheme 1 Background and project synopsis.

However, compared to the extensive studies on the synthesis of conventional atropisomers bearing a sole stereogenic axis (scenario I, Scheme 1b), the development of catalytic asymmetric reactions towards the synchronous control of axial and point chirality in one step has been much less studied and stands for a daunting task (scenario II).7,8 In this context, both issues of diastereo- and enantioselectivity need to be addressed properly to prevent the generation of a complex mixture of stereoisomers. Furthermore, it would be ideal and extremely desirable to establish catalyst-controlled diastereodivergent processes9,10 that allow access to different diastereomers from identical starting materials (scenario III), since individual diastereomers may exhibit significantly distinct biological activities against the same target.11 For example, (S)-atropisomer A was found to be 200-fold more potent than its corresponding (R)-epimer A′ as APJ receptor agonists.6b Nevertheless, achieving such a goal is much more difficult. Generally, generating one of the possible diastereomers is inherently preferred through substrate control, while the selective formation of other diastereomers requires overriding the natural diastereoselective bias.12 In this regard, to the best of our knowledge, no example of the catalytic diastereodivergent and enantioselective simultaneous creation of both axial and point chirality in a single step has been disclosed so far and remains an unmet synthetic challenge.

In alignment with our continuous interest in isocyanide chemistry,8d,g,13,14 especially inspired by our previous work on the simultaneous construction of C–N axial and point chirality (Scheme 1c),8d we report herein the development of an unprecedented silver-catalyzed desymmetrization reaction15 of newly designed prochiral N-quinazolinone maleimides16 with isocyanoacetates. This reaction delivers a wide range of highly functionalized 5,5-fused bicyclic N-heterocycles possessing a remote chiral N–N axis and three contiguous stereogenic carbon centres in high yields with high-to-excellent stereoselectivities (Scheme 1d). Most importantly, intriguing ligand-induced diastereodivergence was also achieved: while the phosphine-squaramide bifunctional ligand L3 resulted in the formation of endo-selective [3 + 2] cycloadducts, employing Trost ligand L9 led to a complete reversal to the exo-cycloadducts, realizing the first example of the diastereodivergent synthesis of atropisomers bearing both axial and point chirality via a single transformation. In addition, the use of either ent-L3 or ent-L9 allowed facile access to the corresponding opposite enantiomers, respectively. Moreover, DFT calculations were performed to elucidate the origin of such stereoselectivity.

Results and discussion

To initiate our study, the easily accessible prochiral N-quinazolinone maleimide 1a was prepared and applied in the reaction with isocyanoacetate 2a for the optimization of reaction conditions (Table 1). The structure of 1a was unambiguously determined by X-ray diffraction analysis. A panel of chiral phosphine ligands was tested in combination with 5 mol% Ag2CO3 as the catalyst for this model reaction. In general, 1,2-diamine-derived bifunctional ligands bearing a hydrogen-bond donor moiety17 such as urea (L1 and L5), thiourea (L2), and squaramide (L3 and L4) all generated the endo-selective [3 + 2] cycloadduct 3aa as the major product (entries 1–5). Among them, L3 proved superior to all others leading to a highly efficient synthesis of 3aa (entry 3, >20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr, 91% yield, and 98% ee). In comparison, the use of both Dixon-type ligands (L6 and L7)18 and Trost ligands (L8 and L9) led to a reversal in the diastereoselectivity, affording the exo-selective [3 + 2] cycloadduct 4aa as the major product (entries 6–9). In particular, L9 resulted in the formation of 4aa with >20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr, 80% yield, and 95% ee (entry 9). This observation adds a new entry to the intriguing ligand-induced diastereodivergent reactions.9 Notably, this is the first time to report that Trost ligands could be successfully employed in catalytic asymmetric reactions of isocyanoacetates,13c thus expanding the choice of chiral ligands in this field. With the optimal ligands being identified, further evaluation of other parameters such as metal salts and solvents was carried out; however, it did not lead to better results (see the ESI for details). The use of Ag2CO3 alone led to an obvious decrease in both efficiency and diastereoselectivity (entry 10, 2[thin space (1/6-em)]:[thin space (1/6-em)]1 dr, 22% yield), while no reaction took place without Ag2CO3 (entries 11 and 12). These results indicated that: (1) Ag2CO3 is indispensable for the reaction; (2) L3/L9 is crucial for tuning diastereoselectivity as well as enhancing reactivity. It is noteworthy that all reactions were conducted under an ambient atmosphere without the requirement to exclude air or moisture.
Table 1 Identification of the ligand-induced diastereodivergent reaction profilea

image file: d4qo00294f-u1.tif

Entry Ligand drb (3aa[thin space (1/6-em)]:[thin space (1/6-em)]4aa) Yieldc (%) ee of 3aae (%) ee of 4aae (%)
a Reaction conditions: 1a (0.1 mmol), 2a (0.1 mmol), Ag2CO3 (5 mol%), and ligand (10 mol%) in 1.0 mL DCE at 25 °C for 12 h. b Determined by crude 1H NMR. c Isolated yield in pure form. d Combined yield. e Determined by chiral HPLC. f Without Ag2CO3.
1 L1 9[thin space (1/6-em)]:[thin space (1/6-em)]1 43 94
2 L2 6[thin space (1/6-em)]:[thin space (1/6-em)]1 38d 82
3 L3 >20[thin space (1/6-em)]:[thin space (1/6-em)]1 91 98
4 L4 4[thin space (1/6-em)]:[thin space (1/6-em)]1 82d 92
5 L5 >20[thin space (1/6-em)]:[thin space (1/6-em)]1 75 84
6 L6 1[thin space (1/6-em)]:[thin space (1/6-em)]13 52 36
7 L7 1[thin space (1/6-em)]:[thin space (1/6-em)]17 49 39
8 L8 1 : >20 50 80
9 L9 1 : >20 80 95
10 2[thin space (1/6-em)]:[thin space (1/6-em)]1 22d
11f L3 <5
12f L9 <5


After establishing the optimal conditions, we first explored the scope of the Ag/L3-catalyzed endo-selective [3 + 2] cycloaddition reaction (Scheme 2). Of note, high-to-excellent diastereoselectivities (10[thin space (1/6-em)]:[thin space (1/6-em)]1 to >20[thin space (1/6-em)]:[thin space (1/6-em)]1) and uniformly excellent enantioselectivities (≥94%) were obtained. The generality of N-quinazolinone maleimides was evaluated with 2a as the model isocyanoacetate substrate. ortho-tert-Butyl-substituted N-quinazolinone maleimides possessing an electron-donating (Me and OMe) or withdrawing group (F, Cl, Br, I, and CF3) at different positions of the quinazolinone skeleton proved to be well tolerated, affording the desired products 3ba–3ma with uniformly excellent stereoselectivities (16[thin space (1/6-em)]:[thin space (1/6-em)]1–>20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr and 97%–>99% ee). In addition, the tert-butyl substituent at the ortho position could be replaced by other alkyl groups, such as isopropyl, cyclohexyl, and benzyl, giving the corresponding 3na–3pa in good-to-high yields with excellent diastereo- and enantioselectivities. Moreover, N-quinazolinone maleimides bearing an ortho-aryl (3qa–3va) or a heteroaryl (3wa) substituent turned out to be compatible substrates as well, undergoing smooth reactions with 2a under the standard conditions, further expanding the substrate scope as well as highlighting the versatility of this catalytic system.


image file: d4qo00294f-s2.tif
Scheme 2 Endo-selective [3 + 2] cycloaddition of N-quinazolinone maleimides with isocyanoacetates. Reaction conditions: 1 (0.1 mmol), 2 (0.1 mmol), Ag2CO3 (5 mol%), and L3 (10 mol%) in 1.0 mL DCE at 25 °C for 12 h.

The scope of isocyanoacetates was also examined under the standard conditions with 1a as the model maleimide substrate. To our delight, the introduction of different substituents (OMe, Cl, and Me) into the phenyl ring did not affect the stereoselectivity too much (3ab–3aevs.3aa). Besides, naphthyl-substituted isocyanoacetate was also applicable, delivering the desired product 3af with >20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr, 84% yield, and 99% ee. The relative and absolute configurations of 3pa were unambiguously determined by X-ray diffraction analysis, and those of other products were assigned by analogy.

To further expand the scope of the Ag/L3 catalytic system, the reaction of N-indole maleimide 5a with 2a was tested under the same set of conditions (Scheme 3). Gratifyingly, the desired endo-cycloadduct 6a was obtained with >20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr, 93% yield, and 99% ee. Access to the opposite enantiomer ent-6a was also achieved by simply replacing L3 with ent-L3. Encouraged by these exciting results, we moved forward to examine the scope of N-indole maleimides. As shown, the incorporation of various substituents at different positions on the indole ring did not affect the result obviously, affording 6b–6e with uniformly excellent stereoselectivities (>20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr and >99% ee for all cases). Replacing the ortho ester group with amide was tolerated as well to generate 6f with >20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr, 72% yield, and 90% ee. In addition, enantiopure N-indole maleimides derived from complex bioactive molecules (L-Menthol, 5g; L-Borneol, 5h) underwent reactions smoothly with 2a under the standard conditions, delivering 6g and 6i in 91% and 95% yields, respectively, with uniformly excellent diastereoselectivities (>20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr for both cases). On the other hand, the reaction of 5g and 5h with 2a conducted with ent-L3 instead of L3 resulted in the synthesis of opposite diastereomers 6h and 6j with comparable yields and diastereoselectivities. These experimental results suggested a mechanism involving catalyst control rather than substrate control when using enantiomerically pure substrates. The relative and absolute configurations of 6a were unambiguously determined by X-ray diffraction analysis, and those of other products were assigned by analogy.


image file: d4qo00294f-s3.tif
Scheme 3 Endo-selective [3 + 2] cycloaddition of N-indole maleimides with 2a. Reaction conditions: 5 (0.1 mmol), 2a (0.1 mmol), Ag2CO3 (5 mol%), and L3 (10 mol%) in 1.0 mL DCE at 25 °C for 12 h. a With ent-L3.

Next, the scope of the Ag/L9-catalyzed exo-selective [3 + 2] cycloaddition reaction was evaluated. As shown in Scheme 4, a wide range of ortho-tert-butyl-substituted N-quinazolinone maleimides with diverse substitution patterns underwent the reactions smoothly with 2a to give the corresponding exo-cycloadducts 4ba–4na in good yields (56%–78%) with uniformly excellent diastereo- and enantioselectivities (>20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr for all cases, 90%–98% ee). However, different from the Ag/L3 catalytic system, an undesired ortho-substituent effect was discovered. Changing the tert-butyl group to less steric substituents, such as isopropyl or phenyl, all led to complex mixtures. In comparison, the scope of isocyanoacetates proved to be broad. The presence of substituents with different electronic properties (OMe, Cl, and Me) at diverse positions (para, meta, and ortho) on the isocyanoacetate structure had little influence on the stereocontrol, delivering 4ab–4ae with >20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr and 90%–95% ee. Other than benzyl-substituted isocyanoacetates, the use of 2g bearing a methyl substitute afforded the desired product 4af with >20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr and 51% yield, albeit with a moderate 71% ee.


image file: d4qo00294f-s4.tif
Scheme 4 Exo-selective [3 + 2] cycloaddition of N-quinazolinone maleimides with isocyanoacetates. Reaction conditions: 1 (0.1 mmol), 2 (0.1 mmol), Ag2CO3 (5 mol%), and L9 (10 mol%) in 1.0 mL DCE at 25 °C for 12 h.

Considering that our products represent a new member of N–N atropisomers,19 the stereochemical stability regarding the chiral N–N axis was evaluated by performing thermal epimerization experiments. Specifically, structurally different compounds 3aa, 3pa, 3qa, 4aa, and 6a were selected as representative compounds and were heated in DMSO-d6 at 100 or 80 °C for 12 h, respectively, and epimerization was monitored by 1H NMR (see the ESI for details). Gratifyingly, all these compounds belong/near to class 3 atropisomers according to the LaPlante classification system,1b which is beneficial for further synthetic transformations and applications. As shown, no corresponding epimer generated by the rotation of the N–N bond was detected for 3aa, 3pa, and 4aa (Fig. 1). In comparison, relatively lower stability was observed for 3qa and 6a, with rotation barriers determined to be 30.7 and 28.7 kcal mol−1, respectively.20


image file: d4qo00294f-f1.tif
Fig. 1 Evaluating the stability of N–N axial chirality.

To illustrate the stereodivergence of this study, we performed the reaction between 1a and 2a in the presence of catalytic Ag2CO3 with four different chiral ligands: L3, ent-L3, L9, and ent-L9 (Scheme 5a). In this way, we were able to obtain four stereoisomers 3aa, ent-3aa, 4aa, and ent-4aa, respectively, in uniformly high yields with excellent stereoselectivities. In addition, gram-scale syntheses of 3aa, 4aa, and 6g were achieved with comparable efficiencies and stereoselectivities without any variation of the standard conditions (Scheme 5b), demonstrating the practicality and robustness of our method. Moreover, the existence of an imine functionality in our products laid the roots for further synthetic transformations. For example, upon treatment with NaBH3CN, 4aa could be facilely reduced to the corresponding pyrrolidine 7 in 84% yield (Scheme 5c). The relative and absolute configurations of 7 were unambiguously determined by X-ray diffraction analysis, and those of other products were assigned by analogy. Alternatively, the nucleophilic addition of indole to 4aa took place smoothly with the use of AgBF4, generating 8 in an almost quantitative yield with exclusive diastereoselectivity. The absolute configuration of the newly generated stereogenic centre in 8 was determined by NOE analysis (see the ESI for details).


image file: d4qo00294f-s5.tif
Scheme 5 Synthetic applications.

To elucidate the mechanism and stereoselectivity of the silver-catalyzed [3 + 2] cycloaddition reaction, DFT calculations were performed. Fig. 2a illustrates the proposed mechanism for the asymmetric [3 + 2] cycloaddition between N-quinazolinone maleimide 1a and isocyanoacetate 2a catalyzed by silver complexes with ligands L3 and L9. Deprotonation of 2a coordinated to a ligand-Ag2CO3 complex forms the metal-substrate complex L3/L9-Int1, which is set as a relative zero point in our calculated free energy profiles. A concerted [3 + 2] cycloaddition then occurs between metal-bound 2a and 1a through the transition state L3/L9-TS2, giving intermediate L3-Int3-A and L9-Int3-B exothermically by 13.1/10.1 kcal mol−1. Finally, the protonation of L3/L9-Int3 yields products 3aa/4aa. This cycloaddition step is critical as it determines axial and point chirality as well as diastereoselectivity.


image file: d4qo00294f-f2.tif
Fig. 2 Mechanistic study based on DFT calculations. (a) Proposed divergent reaction pathways for the cycloaddition of 1a with 2a. (b) Geometrical structures of cycloaddition transition states. (c) Dihedral information of L3-int1, L3-TS2-A, and L3-TS2-B. (d) Front molecular orbital analysis and bond angle information of L9-TS2-A and L9-TS2-B.

Thus, we further analysed the cycloaddition transition states L3/L9-TS2 (Fig. 2b). When squaramide L3 is employed as the chiral ligand, 1a can approach 2a by either the upper (L3-TS2-A) or lower (L3-TS2-B) side in the cycloaddition step. In the favoured transition state L3-TS2-A, both N–H bonds in the squaramide moiety can form intermolecular hydrogen bonds with carbonyls in 1a. However, in the disfavoured transition state L3-TS2-B, only one intermolecular hydrogen bond can be formed. In addition, a larger distortion is detected in L3-TS2-B: the dihedral angle of N–C–C–N in the diaminocyclohexane moiety is 51.0°, considerably deviating from the value in L3-Int1 (61.2°), indicating a large twist of the squaramide ligand (Fig. 2c). In contrast, the dihedral angle value in the favoured transition state L3-TS2-A is 58.5°, close to the value in L3-Int1, indicating a definitely smaller twist. The extra hydrogen bond and smaller twist in L3-TS2-A provide a relative free energy of 4.2 kcal mol−1 lower than L3-TS2-B, leading to the major product 3aa. On the other hand, the Trost ligand L9 has two phosphine atoms, and therefore it is capable of serving as a bidentate ligand and binds silver and 1a in a T-shaped configuration. We found that stereoselectivity is mainly attributed to the distortion: the bond angle of N–C–Ag in L9-TS2-A is 143.0°, considerably deviating from the linear structure, indicating a large distortion (Fig. 2d). In addition, the molecular orbital analysis showed an insufficient orbital overlap between 1a and the Ag centre in HOMO−3. In contrast, the bond angle of N–C–Ag in L9-TS2-B is 165.7°, indicating a smaller distortion; meanwhile, the molecular orbital analysis showed a better overlap in HOMO−3. As a result, the calculated relative free energy of L9-TS2-B is 2.2 kcal mol−1 lower than L9-TS2-A, favouring the formation of 4aa.

Conclusions

In conclusion, we have developed an unprecedented desymmetric [3 + 2] cycloaddition reaction of prochiral N-quinazolinone maleimides with isocyanoacetates. Under silver catalysis, a broad range of structurally novel and complex N–N atropisomers containing additional three contiguous stereogenic carbon centres were obtained in good yields with high-to-excellent diastereo- and enantioselectivities. Of particular note, by simply switching the chiral ligand (L3/L9), both endo- and exo-cycloadducts were generated. DFT calculations revealed that L3 and L9 act in different ways. Squaramide L3 acts as a monodentate ligand, and the stronger ligand–substrate non-bonding interaction and smaller distortion of the ligand in the annulation step lead to the generation of endo-selective cycloadducts. Trost ligand L9 acts as a bidentate ligand, and the smaller distortion of isocyanoacetate–Ag coordination together with better Ag–C σ-orbital overlap contributes to the generation of exo-selective [3 + 2] cycloadducts. The success of this study not only represents the first example of the catalytic stereodivergent and simultaneous construction of axial and point chirality but also inspires the design of other useful stereodivergent processes.

Author contributions

L. Q. and J.-Y. L. designed the project. W.-T. W., S. Z., Z.-H. L., D. H., and F. H. carried out the experiments and analysed the data. W. L., R. B., and Y. L. performed the computational study. W.-T. W., S. Z., and W. L. contributed equally to this work. J.-Y. L. directed the project and wrote the manuscript. All authors discussed the results and commented on the manuscript.

Data availability

Experimental details and characterization data for all new compounds in this article are available in the ESI. Crystallographic data for compounds 1a, 3pa, 6a, and 7 have been deposited in the Cambridge Crystallographic Data Centre (CCDC) under accession numbers CCDC 2261237, 2261239, 2261241, and 2261240, respectively.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We acknowledge financial support from the National Natural Science Foundation of China (82273887 and 22271034), the Joint Funds of the Zhejiang Provincial Natural Science Foundation of China (LTZ22B020001), and Zhejiang University. We also thank Mr Jiyong Liu (Chemistry Instrumentation Center, Zhejiang University) for X-ray crystallographic analysis and Mr Jiadong Xue (Innovation Institute for Artificial Intelligence in Medicine of Zhejiang University) for performing NMR spectroscopy for structure elucidation.

References

  1. (a) J. Clayden, W. J. Moran, P. J. Edwards and S. R. LaPlante, The challenge of atropisomerism in drug discovery, Angew. Chem., Int. Ed., 2009, 48, 6398–6401 CrossRef CAS PubMed; (b) S. R. LaPlante, P. J. Edwards, L. D. Fader, A. Jakalian and O. Hucke, Revealing atropisomer axial chirality in drug discovery, ChemMedChem, 2011, 6, 505–513 CrossRef CAS PubMed; (c) S. R. LaPlante, L. D. Fader, K. R. Fandrick, D. R. Fandrick, O. Hucke, R. Kemper, S. P. F. Miller and P. J. Edwards, Assessing atropisomer axial chirality in drug discovery and development, J. Med. Chem., 2011, 54, 7005–7022 CrossRef CAS PubMed; (d) S. T. Toenjes and J. L. Gustafson, Atropisomerism in medicinal chemistry: challenges and opportunities, Future Med. Chem., 2018, 10, 409–422 CrossRef CAS PubMed; (e) Z. Wang, L. Meng, X. Liu, L. Zhang, Z. Yu and G. Wu, Recent progress toward developing axial chirality bioactive compounds, Eur. J. Med. Chem., 2022, 243, 114700 CrossRef CAS PubMed.
  2. (a) R. Noyori and H. Takaya, BINAP: an efficient chiral element for asymmetric catalysis, Acc. Chem. Res., 1990, 23, 345–350 CrossRef CAS; (b) Y. Chen, S. Yekta and A. K. Yudin, Modified BINOL ligands in asymmetric catalysis, Chem. Rev., 2003, 103, 3155–3212 CrossRef CAS PubMed; (c) S. Shirakawa and K. Maruoka, Recent developments in asymmetric phase-transfer reactions, Angew. Chem., Int. Ed., 2013, 52, 4312–4348 CrossRef CAS PubMed; (d) D. Parmar, E. Sugiono, S. Raja and M. Rueping, Complete field guide to asymmetric BINOL-phosphate derived Brønsted acid and metal catalysis: history and classification by mode of activation; Brønsted acidity, hydrogen bonding, ion pairing, and metal phosphates, Chem. Rev., 2014, 114, 9047–9153 CrossRef CAS PubMed.
  3. L. Pu, Enantioselective fluorescent recognition of free amino acids: challenges and opportunities, Angew. Chem., Int. Ed., 2020, 59, 21814–21828 CrossRef CAS PubMed.
  4. For selected general reviews on the synthesis of atropisomers, see: (a) G. Bringmann, A. J. P. Mortimer, P. A. Keller, M. J. Gresser, J. Garner and M. Breuning, Atroposelective synthesis of axially chiral biaryl compounds, Angew. Chem., Int. Ed., 2005, 44, 5384–5427 CrossRef CAS PubMed; (b) J. Wencel-Delord, A. Panossian, F. R. Leroux and F. Colobert, Recent advances and new concepts for the synthesis of axially stereoenriched biaryls, Chem. Soc. Rev., 2015, 44, 3418–3430 RSC; (c) E. Kumarasamy, R. Raghunathan, M. P. Sibi and J. Sivaguru, Nonbiaryl and heterobiaryl atropisomers: molecular templates with promise for atropselective chemical transformations, Chem. Rev., 2015, 115, 11239–11300 CrossRef CAS PubMed; (d) R. Song, Y. Xie, Z. Jin and Y. R. Chi, Carbene-catalyzed asymmetric construction of atropisomers, Angew. Chem., Int. Ed., 2021, 60, 26026–26037 CrossRef CAS PubMed; (e) J. A. Carmona, C. Rodríguez-Franco, R. Fernández, V. Hornillos and J. M. Lassaletta, Atroposelective transformation of axially chiral (hetero)biaryls. From desymmetrization to modern resolution strategies, Chem. Soc. Rev., 2021, 50, 2968–2983 RSC; (f) J. K. Cheng, S.-H. Xiang, S. Li, L. Ye and B. Tan, Recent advances in catalytic asymmetric construction of atropisomers, Chem. Rev., 2021, 121, 4805–4902 CrossRef CAS PubMed; (g) G.-J. Mei, W. L. Koay, C.-Y. Guan and Y. Lu, Atropisomers beyond the C-C axial chirality: advances in catalytic asymmetric synthesis, Chem, 2022, 8, 1855–1893 CrossRef CAS; (h) M. C. Kozlowski, S. J. Miller and S. Perreault, Atropisomers: synthesis, analysis, and applications, Acc. Chem. Res., 2023, 56, 187–188 CrossRef CAS PubMed; (i) C. B. Roos, C.-H. Chiang, L. A. M. Murray, D. Yang, L. Schulert and A. R. H. Narayan, Stereodynamic strategies to induce and enrich chirality of atropisomers at a late stage, Chem. Rev., 2023, 123, 10641–10727 CrossRef CAS PubMed; (j) Y. Zhang, H. Cai, X. Gan and Z. Jin, N-Heterocyclic carbene-catalyzed enantioselective (dynamic) kinetic resolutions and desymmetrizations, Sci. China: Chem., 2024, 67, 482–511 CrossRef CAS.
  5. (a) X. Bao, J. Rodriguez and D. Bonne, Enantioselective synthesis of atropisomers with multiple stereogenic axes, Angew. Chem., Int. Ed., 2020, 59, 12623–12634 CrossRef CAS PubMed; (b) F. Ye, Z. Xu and L.-W. Xu, The discovery of multifunctional chiral P ligands for the catalytic construction of quaternary carbon/silicon and multiple stereogenic centers, Acc. Chem. Res., 2021, 54, 452–470 CrossRef CAS PubMed; (c) T. A. Schmidt and C. Sparr, Catalyst control over twofold and higher-order stereogenicity by atroposelective arene formation, Acc. Chem. Res., 2021, 54, 2764–2774 CrossRef CAS PubMed; (d) X.-F. Bai, Y.-M. Cui, J. Cao and L.-W. Xu, Atropisomers with axial and point chirality: synthesis and applications, Acc. Chem. Res., 2022, 55, 2545–2561 CrossRef CAS PubMed.
  6. (a) Q. Zhang, A. Mándi, S. Li, Y. Chen, W. Zhang, X. Tian, H. Zhang, H. Li, W. Zhang, S. Zhang, J. Ju, T. Kurtán and C. Zhang, N-N-Coupled indolo-sesquiterpene atropo-diastereomers from a marine-derived actinomycete, Eur. J. Org. Chem., 2012, 5256–5262 CrossRef CAS; (b) S. Su, A. Clarke, Y. Han, H. J. Chao, J. Bostwick, W. Schumacher, T. Wang, M. Yan, M.-Y. Hsu, E. Simmons, C. Luk, C. Xu, M. Dabros, M. Galella, J. Onorato, D. Gordon, R. Wexler, P. S. Gargalovic and R. M. Lawrence, Biphenyl acid derivatives as APJ receptor agonists, J. Med. Chem., 2019, 62, 10456–10465 CrossRef CAS PubMed; (c) J. Chen, X. Gong, J. Li, Y. Li, J. Ma, C. Hou, G. Zhao, W. Yuan and B. Zhao, Carbonyl catalysis enables a biomimetic asymmetric Mannich reaction, Science, 2018, 360, 1438–1442 CrossRef CAS PubMed; (d) J. F. Teichert and B. L. Feringa, Phosphoramidites: privileged ligands in asymmetric catalysis, Angew. Chem., Int. Ed., 2010, 49, 2486–2528 CrossRef CAS PubMed.
  7. For reviews, see: (a) H.-H. Zhang, T.-Z. Li, S.-J. Liu and F. Shi, Catalytic asymmetric synthesis of atropisomers bearing multiple chiral elements: an emerging field, Angew. Chem., Int. Ed., 2023, 63, e202311053 CrossRef PubMed; (b) A. Luc and J. Wencel-Delord, One reaction - double stereoinduction: C-H activation as a privileged route towards complex atropisomeric molecules, Chem. Commun., 2023, 59, 8159–8167 RSC; (c) H.-H. Zhang and F. Shi, Organocatalytic atroposelective synthesis of indole derivatives bearing axial chirality: strategies and applications, Acc. Chem. Res., 2022, 55, 2562–2580 CrossRef CAS PubMed; (d) C. Portolani, G. Centonze, P. Righi and G. Bencivenni, Role of cinchona alkaloids in the enantio- and diastereoselective synthesis of axially chiral compounds, Acc. Chem. Res., 2022, 55, 3551–3571 CrossRef CAS PubMed.
  8. For selected recent examples, see: (a) B.-B. Gou, Y. Tang, Y.-H. Lin, L. Yu, Q.-S. Jian, H.-R. Sun, J. Chen and L. Zhou, Modular construction of heterobiaryl atropisomers and axially chiral styrenes via all-carbon tetrasubstituted VQMs, Angew. Chem., Int. Ed., 2022, 61, e202208174 CrossRef CAS PubMed; (b) Y. Li, Y.-C. Liou, J. C. A. Oliveira and L. Ackermann, Ruthenium(II)/imidazolidine carboxylic acid-catalyzed C-H alkylation for central and axial double enantio-induction, Angew. Chem., Int. Ed., 2022, 61, e202212595 CrossRef CAS PubMed; (c) J.-Y. Du, T. Balan, T. D. W. Claridge and M. D. Smith, Counterion-mediated enantioconvergent synthesis of axially chiral medium rings, J. Am. Chem. Soc., 2022, 144, 14790–14797 CrossRef CAS PubMed; (d) S. Zhang, Z.-H. Luo, W.-T. Wang, L. Qian and J.-Y. Liao, Simultaneous construction of C-N axial and central chirality via silver-catalyzed desymmetrizative [3 + 2] cycloaddition of prochiral N-aryl maleimides with activated isocyanides, Org. Lett., 2022, 24, 4645–4649 CrossRef CAS PubMed; (e) Y.-B. Chen, L.-G. Liu, C.-M. Chen, Y.-X. Liu, B. Zhou, X. Lu, Z. Xu and L.-W. Ye, Construction of axially chiral arylpyrroles via atroposelective diyne cyclization, Angew. Chem., Int. Ed., 2023, 62, e202303670 CrossRef CAS PubMed; (f) J. A. Carmona, P. Rodríguez-Salamanca, R. Fernández, J. M. Lassaletta and V. Hornillos, Dynamic kinetic resolution of 2-(quinolin-8-yl)benzaldehydes: atroposelective iridium-catalyzed transfer hydrogenative allylation, Angew. Chem., Int. Ed., 2023, 62, e202306981 CrossRef CAS PubMed; (g) F. Huang, L.-F. Tao, J. Liu, L. Qian and J.-Y. Liao, Diastereo- and enantioselective synthesis of biaryl aldehydes bearing both axial and central chirality, Chem. Commun., 2023, 59, 4487–4490 RSC; (h) H. Jiang, X.-K. He, X. Jiang, W. Zhao, L.-Q. Lu, Y. Cheng and W.-J. Xiao, Photoinduced cobalt-catalyzed desymmetrization of dialdehydes to access axial chirality, J. Am. Chem. Soc., 2023, 145, 6944–6952 CrossRef CAS PubMed; (i) R. Mi, Z. Ding, S. Yu, R. H. Crabtree and X. Li, O-Allylhydroxyamine: a bifunctional olefin for construction of axially and centrally chiral amino alcohols via asymmetric carboamidation, J. Am. Chem. Soc., 2023, 145, 8150–8162 CrossRef CAS PubMed; (j) C. Rodríguez-Franco, A. Ros, P. Merino, R. Fernández, J. M. Lassaletta and V. Hornillos, Dynamic kinetic resolution of indole-based sulfenylated heterobiaryls by rhodium-catalyzed atroposelective reductive aldol reaction, ACS Catal., 2023, 13, 12134–12141 CrossRef PubMed; (k) L. Dai, X. Zhou, J. Guo, X. Dai, Q. Huang and Y. Lu, Diastereo- and atroposelective synthesis of N-arylpyrroles enabled by light-induced phosphoric acid catalysis, Nat. Commun., 2023, 14, 4813 CrossRef CAS PubMed.
  9. For reviews on diastereodivergent catalysis, see: (a) M. Bihani and J. C.-G. Zhao, Advances in asymmetric diastereodivergent catalysis, Adv. Synth. Catal., 2017, 359, 534–575 CrossRef CAS; (b) S. Krautwald and E. M. Carreira, Stereodivergence in asymmetric catalysis, J. Am. Chem. Soc., 2017, 139, 5627–5639 CrossRef CAS PubMed; (c) I. P. Beletskaya, C. Nájera and M. Yus, Stereodivergent catalysis, Chem. Rev., 2018, 118, 5080–5200 CrossRef CAS PubMed; (d) D. Moser, T. A. Schmidt and C. Sparr, Diastereodivergent catalysis, JACS Au, 2023, 3, 2612–2630 CrossRef CAS PubMed.
  10. For rare examples of stereodivergent synthesis of atropisomers bearing both axial and point chirality in a stepwise manner with two independent catalysts, see: (a) Y. Kwon, A. J. Chinn, B. Kim and S. J. Miller, Divergent control of point and axial stereogenicity: catalytic enantioselective C-N bond-forming cross-coupling and catalyst-controlled atroposelective cyclodehydration, Angew. Chem., Int. Ed., 2018, 57, 6251–6255 CrossRef CAS PubMed; (b) C. Portolani, G. Centonze, S. Luciani, A. Pellegrini, P. Righi, A. Mazzanti, A. Ciogli, A. Sorato and G. Bencivenni, Synthesis of atropisomeric hydrazides by one-pot sequential enantio- and diastereoselective catalysis, Angew. Chem., Int. Ed., 2022, 61, e202209895 CrossRef CAS PubMed.
  11. Drug Stereochemistry: Analytical Methods and Pharmacology, ed. K. Jozwiak, W. J. Lough and I. W. Wainer, Informa, 2012 Search PubMed.
  12. Comprehensive Asymmetric Catalysis, ed. E. N. Jacobsen, A. Pfaltz and H. Yamamoto, Springer Berlin, Heidelberg, 1999 Search PubMed.
  13. For comprehensive reviews on isocyanide chemistry, see: (a) A. V. Gulevich, A. G. Zhdanko, R. V. A. Orru and V. G. Nenajdenko, Isocyanoacetate derivatives: synthesis, reactivity, and application, Chem. Rev., 2010, 110, 5235–5331 CrossRef CAS PubMed; (b) M. Giustiniano, A. Basso, V. Mercalli, A. Massarotti, E. Novellino, G. C. Tron and J. Zhu, To each his own: isonitriles for all flavors. Functionalized isocyanides as valuable tools in organic synthesis, Chem. Soc. Rev., 2017, 46, 1295–1357 RSC; (c) J. Luo, G.-S. Chen, S.-J. Chen, Z.-D. Li and Y.-L. Liu, Catalytic enantioselective isocyanide-based reactions: beyond Passerini and Ugi multicomponent reactions, Chem. – Eur. J., 2021, 27, 6598–6619 CrossRef CAS PubMed.
  14. (a) L. Qian, L.-F. Tao, W.-T. Wang, E. Jameel, Z.-H. Luo, T. Zhang, Y. Zhao and J.-Y. Liao, Catalytic atroposelective dynamic kinetic resolution of biaryl lactones with activated isocyanides, Org. Lett., 2021, 23, 5086–5091 CrossRef CAS PubMed; (b) Z.-H. Luo, W.-T. Wang, T.-Y. Tang, S. Zhang, F. Huang, D. Hu, L.-F. Tao, L. Qian and J.-Y. Liao, Torsional strain-independent catalytic enantioselective synthesis of biaryl atropisomers, Angew. Chem., Int. Ed., 2022, 61, e202211303 CrossRef CAS PubMed; (c) L.-F. Tao, S. Zhang, F. Huang, W.-T. Wang, Z.-H. Luo, L. Qian and J.-Y. Liao, Diastereo- and enantioselective silver-catalyzed [3 + 3] cycloaddition and kinetic resolution of azomethine imines with activated isocyanides, Angew. Chem., Int. Ed., 2022, 61, e202202679 CrossRef CAS PubMed; (d) W.-T. Wang, S. Zhang, L.-F. Tao, Z.-Q. Pan, L. Qian and J.-Y. Liao, Cooperative catalysis-enabled C-N bond cleavage of biaryl lactams with activated isocyanides, Chem. Commun., 2022, 58, 6292–6295 RSC; (e) L.-F. Tao, F. Huang, X. Zhao, L. Qian and J.-Y. Liao, Atroposelective synthesis of eight-membered lactam-bridged N-arylindoles via stepwise cut-and-sew strategy, Cell Rep. Phys. Sci., 2023, 4, 101697 CrossRef CAS.
  15. For selected reviews on catalytic asymmetric desymmetrization reactions, see: (a) X.-P. Zeng, Z.-Y. Cao, Y.-H. Wang, F. Zhou and J. Zhou, Catalytic enantioselective desymmetrization reactions to all-carbon quaternary stereocenters, Chem. Rev., 2016, 116, 7330–7396 CrossRef CAS PubMed; (b) A. J. Metrano and S. J. Miller, Peptide-based catalysts reach the outer sphere through remote desymmetrization and atroposelectivity, Acc. Chem. Res., 2019, 52, 199–215 CrossRef CAS PubMed; (c) C. Nájera, F. Foubelo, J. M. Sansano and M. Yus, Enantioselective desymmetrization reactions in asymmetric catalysis, Tetrahedron, 2022, 106–107, 132629 CrossRef; (d) Y. Xu, T.-Y. Zhai, Z. Xu and L.-W. Ye, Recent advances towards organocatalytic enantioselective desymmetrizing reactions, Trends Chem., 2022, 4, 191–205 CrossRef CAS.
  16. For reviews on maleimide chemistry, see: (a) P. Chauhan, J. Kaur and S. S. Chimni, Asymmetric organocatalytic addition reactions of maleimides: a promising approach towards the synthesis of chiral succinimide derivatives, Chem. – Asian J., 2013, 8, 328–346 CrossRef CAS PubMed; (b) R. Manoharan and M. Jeganmohan, Alkylation, annulation, and alkenylation of organic molecules with maleimides by transition-metal-catalyzed C-H bond activation, Asian J. Org. Chem., 2019, 8, 1949–1969 CrossRef CAS; (c) S.-L. Liu, C. Ye and X. Wang, Recent advances in transition-metal-catalyzed directed C-H alkenylation with maleimides, Org. Biomol. Chem., 2022, 20, 4837–4845 RSC.
  17. Y. Xiong, Z. Du, H. Chen, Z. Yang, Q. Tan, C. Zhang, L. Zhu, Y. Lan and M. Zhang, Well-designed phosphine-urea ligand for highly diastereo- and enantioselective 1,3-dipolar cycloaddition of methacrylonitrile: a combined experimental and theoretical study, J. Am. Chem. Soc., 2019, 141, 961–971 CrossRef CAS PubMed.
  18. (a) F. Sladojevich, A. Trabocchi, A. Guarna and D. J. Dixon, A new family of cinchona-derived amino phosphine precatalysts: application to the highly enantio- and diastereoselective silver-catalyzed isocyanoacetate aldol reaction, J. Am. Chem. Soc., 2011, 133, 1710–1713 CrossRef CAS PubMed; (b) P.-L. Shao, J.-Y. Liao, Y. A. Ho and Y. Zhao, Highly diastereo- and enantioselective silver-catalyzed double [3 + 2] cyclization of α-imino esters with isocyanoacetate, Angew. Chem., Int. Ed., 2014, 53, 5435–5439 CrossRef CAS PubMed.
  19. For reviews, see: (a) G. Centonze, C. Portolani, P. Righi and G. Bencivenni, Enantioselective strategies for the synthesis of N-N atropisomers, Angew. Chem., Int. Ed., 2023, 62, e202303966 CrossRef PubMed; (b) J. Feng and R.-R. Liu, Catalytic asymmetric synthesis of N-N biaryl atropisomers, Chem. – Eur. J., 2023, 30, e202303165 CrossRef PubMed. For selected examples, see: (c) G.-J. Mei, J. J. Wong, W. Zheng, A. A. Nangia, K. N. Houk and Y. Lu, Rational design and atroposelective synthesis of N-N axially chiral compounds, Chem, 2021, 7, 2743–2757 CrossRef CAS; (d) X.-M. Wang, P. Zhang, Q. Xu, C.-Q. Guo, D.-B. Zhang, C.-J. Lu and R.-R. Liu, Enantioselective synthesis of nitrogen-nitrogen biaryl atropisomers via copper-catalyzed Friedel-Crafts alkylation reaction, J. Am. Chem. Soc., 2021, 143, 15005–15010 CrossRef CAS PubMed; (e) W. Lin, Q. Zhao, Y. Li, M. Pan, C. Yang, G.-H. Yang and X. Li, Asymmetric synthesis of N-N axially chiral compounds via organocatalytic atroposelective N-acylation, Chem. Sci., 2022, 13, 141–148 RSC; (f) K.-W. Chen, Z.-H. Chen, S. Yang, S.-F. Wu, Y.-C. Zhang and F. Shi, Organocatalytic atroposelective synthesis of N-N axially chiral indoles and pyrroles by de novo ring formation, Angew. Chem., Int. Ed., 2022, 61, e202116829 CrossRef CAS PubMed; (g) Y. Gao, L.-Y. Wang, T. Zhang, B.-M. Yang and Y. Zhao, Atroposelective synthesis of 1,1′-bipyrroles bearing a chiral N-N axis: chiral phosphoric acid catalysis with Lewis acid induced enantiodivergence, Angew. Chem., Int. Ed., 2022, 61, e202200371 CrossRef CAS PubMed; (h) S.-Y. Yin, Q. Zhou, C.-X. Liu, Q. Gu and S.-L. You, Enantioselective synthesis of N-N biaryl atropisomers through iridium(I)-catalyzed C-H alkylation with acrylates, Angew. Chem., Int. Ed., 2023, 62, e202305067 CrossRef CAS PubMed; (i) Y. Wang, X. Zhu, D. Pan, J. Jing, F. Wang, R. Mi, G. Huang and X. Li, Rhodium-catalyzed enantioselective and diastereodivergent access to diaxially chiral heterocycles, Nat. Commun., 2023, 14, 4661 CrossRef CAS PubMed.
  20. (a) D. B. Guthrie and D. P. Curran, Asymmetric radical and anionic cyclizations of axially chiral carbamates, Org. Lett., 2009, 11, 249–251 CrossRef CAS PubMed; (b) S. Lin, D. Leow, K.-W. Huang and C.-H. Tan, Enantioselective protonation of itaconimides with thiols and the rotational kinetics of the axially chiral C-N bond, Chem. – Asian J., 2009, 4, 1741–1744 CrossRef CAS PubMed.

Footnotes

Electronic supplementary information (ESI) available: Experimental procedures, characterization of all new compounds, and crystallographic data of 1a, 3pa, 6a, and 7. CCDC 2261237 and 2261239–2261241. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4qo00294f
These authors contributed equally.

This journal is © the Partner Organisations 2024