Introduction

α-Hydroxyketones (also known as acyloins) are structural units ubiquitously found in natural products1,2,3,4,5 and pharmaceuticals6,7. They are also oft-employed synthetic precursors in a panel of high-value transformations (Fig. 1a)8,9,10,11,12,13. The construction of these important molecules is therefore the subject of substantial synthetic efforts14. Traditional benzoin condensation method assembles α-hydroxyketones via condensation of different aldehydes, thus limits its applicability within this substrate class15,16,17. The alternate oxidative pathways that encompass α-hydroxylation of ketones18,19,20,21,22 and ketohydroxylation of olefins23,24,25,26,27 are certainly enabling, but continues to be challenged in terms of substrate diversity and poor selectivity. Hence, devising complementary routes towards these useful entities from readily available starting materials is highly relevant and desirable.

Fig. 1: Precedent works and proposed Passerini-type coupling reaction with boronic acids as nucleophilic agents.
figure 1

a α-Hydroxyketones in bioactive moleculars or as synthetic precursors. b Classic Passerini or Ugi reaction. c Petasis boronic acid-Mannich reaction. d Passerini-type coupling reaction of boronic acids (this work).

Multicomponent reactions (MCRs) are often prized for their concise and modular features in forging complex molecules with synthetic and biological interest28,29,30,31,32,33,34,35,36. The representative Passerini reaction37,38,39,40,41,42,43,44,45,46 or Ugi reaction47,48,49,50,51,52,53,54 efficiently assembles α-acyloxyamides or α-acylaminoamides from several reactant components via the intermediacy of nitrilium species in single-pot operation (Fig. 1b)55,56,57,58. Interception of this electrophilic intermediate in Passerini reaction pathway by carbon nucleophiles (in place of conventionally used carboxylic acids) would offer an intriguing access to α-hydroxyketone products; yet such synthetic maneuver remains underexplored59,60. Central to the successful establishment of this chemistry would lie in choosing suitable carbon nucleophiles that would not interfere with the formation of the nitrilium intermediate while possess sufficient nucleophilicity to capture this electrophile.

On the other hand, boronic acids are easily available, benign and common building blocks for C-C bond cross-coupling reactions, in both transition-metal catalysis61,62 and metal-free catalysis regimes63,64,65,66,67,68,69,70,71,72,73,74,75. In boronic acid-Mannich reaction (or Petasis reaction), for instance, the nucleophilic feature of boronic acids effects the formation of boron “ate” complex, leading to functionalized amines following 1,3-metallate migration (Fig. 1c)76,77,78,79,80. To this end, a recent endeavor of our group has unraveled a 1,4-metallate shift of boron “ate” nitrilium species generated from nitrile oxide and arylboronic acid, thus mediating stereospecific formation of C-C bond between oxime chlorides and arylboronic acids under metal-free conditions81. Grounded in these knowledges, we envisioned that a boron “ate” nitrilium intermediate could be released from co-treatment of aldehyde, isocyanide, and boronic acid; 1,4-metallate shift of which will invoke C-C bond coupling and α-hydroxyketones could be revealed on hydrolysis (Fig. 1d). Here, we disclose the development of a Passerini-type coupling reaction, which afforded α-hydroxyketones from the combination of readily available aldehydes, isocyanides, and boronic acids (aryl, alkenylboronic acids, and alkynyl trifluoroborate salts) under transition-metal-free conditions. Mild reaction conditions, ease of execution, high functional group tolerance, broad substrate scope, and utility are practical features of this methodology.

Results

Investigation of reaction conditions

Exploratory investigations towards our envisioned Passerini-type reaction involving boronic acids were conducted with phenylpropyl aldehyde (1a), tertbutyl isocyanide (2a) and 4-methoxyphenyl boronic acid (3a) as test substrates (Table 1). To our delights, simple mixing of the three reactants (1a, 2a, and 3a) without any other additive in DCM furnished the desired α-hydroxyketone product 4a in 60% isolated yield (entry 1). A solvent screen of DCE, MeCN, toluene, MeOH, and THF revealed that the best reaction efficiency was endowed by CHCl3, whereas using MeOH caused a complete reaction inhibition (entries 1−7). As reaction temperature was decreased to 10 °C, the yield of 4a improved to 68% (entry 8). Binary mixture of CHCl3 and water in a ratio of 7:3 (entries 9−11) minimally but meaningfully enhanced the delivery of 4a to 72% yield (entry 11). This has guided our subsequent study of mixed solvent system with CHCl3 against various buffer solutions (entries 12−15) where the combination with pH = 8.0 buffer delightfully provided 81% yield of target product (entry 14). We reasoned that a basic reaction medium could sequester the byproduct B(OH)3 generated during reaction, thus promoting this boronic acid-involved Passerini-type reaction. It was further established that on replacement of tertbutyl isocyanide (2a) with cyclohexyl isocyanide (2b), benzyl isocyanide (2c), or ethyl 2-isocyanoacetate (2d), formation efficiency of α-hydroxyketone product 4a was diminished (entry 16). None of the other ratios of the three reagents resulted in higher yields (entries 17−18).

Table 1 Optimization of the reaction conditionsa.

Scope of aldehydes

Having optimized the model coupling of this Passerini-type reaction, we examined the generality of these conditions with respect to a range of aldehyde components (Fig. 2). Delightfully, diverse aliphatic aldehydes were aptly transformed in moderate to high yields. Phenylpropyl aldehydes with strong electron-withdrawing groups and 3-(furan-2-yl)propanal furnished the α-hydroxyketone products 4b4d in 66% to 90% yields. The chain length of aldehydes posed no effect on the effectiveness of this reaction, providing respective α-hydroxyketones (4e4g) in moderate yields. Primary aldehydes bearing ester, adamantyl, and benzyloxy moieties were tolerated well to yield 4h4j in moderate efficiencies. Secondary aldehydes comprised of acyclic and cyclic analogs (cyclopropyl, cyclohexyl, piperidinyl) were incorporated in 4k4q with moderate to good yields as well. The diastereomeric ratios (dr) of compounds 4l and 4n are 1.13:1 and 1.38:1. Comparable outcome was observed for a tertiary 1-phenylcyclobutane-1-carbaldehyde substrate, which afforded 4r in 54% yield. It merits mention that transformation of paraformaldehyde has given rise to 4s, which serves as versatile synthetic intermediate for a variety of bioactive molecules. More importantly, this reaction was well suited to diverse aromatic aldehydes when treated in concert with cyclohexyl isocyanide (2b). The electronic property and the position of substituents on the benzene ring had minimal bearing on the efficiency of this transformation. Neutral (4t), electron-rich (4u4y), or electron-deficient (4z4aa) functionalities found good compatibility and were left unscathed in respective molecular outputs. The accommodation of halogen substituents (4ab4ae) signified potential structural elaborations from these handles. Fused ring reactants including 2-naphthaldehyde (4af) and 1-naphthaldehyde (4ag) were also suitable candidates for this MCR.

Fig. 2: Scope of aldehydesa.
figure 2

Reaction conditions: aaldehyde 1 (0.2 mmol, 1 equiv), 2a (0.3 mmol), 3a (0.36 mmol), and CHCl3/pH = 8 buffer (7:3, 1 mL) under an argon atmosphere for 24 hours unless otherwise specified; bThe dr was determined by 1H NMR analysis. caldehyde 1 (0.2 mmol, 1 equiv), 2b (0.3 mmol), 3a (0.36 mmol), and DCM/pH = 8 buffer (7:3, 1 mL) under an argon atmosphere for 24 hours.

Scope of boronic acids

This protocol featured an admirable scope with respect to arylboronic acid substrates (Fig. 3). For electron-rich congeners, good reactivities were exhibited. Arylboronic acids with electronically neutral meta-para-dimethyl, para-methyl, and para-tertbutyl substituents produced α-hydroxyketones 5a5c in moderate yields. Analogs with electron-rich substituents such as acetal, alkoxy, and diphenylamino groups reacted smoothly towards products 5d–5l in 51–85% yields. Inclusion of alkenyl or alkynyl group was noteworthy; from which products 5j and 5k were acquired in 80% and 77% yield. This study was auspiciously and effortlessly extendable to a series of heteroarylboronic acids containing furan (5m), thiophene (5n–5p), benzofuran (5q), benzothiophen (5r), protected or unprotected indoles (5s, 5u), 7-azaindole (5t), dibenzothiophene (5w) and carbazole (5x) cores. Remarkably, both aryl and alkyl substituted alkenylboronic acids could rendered the corresponding α-hydroxy enones 5y and 5z in 82% and 63% yields, which broadly expand the scope of the products. For electron-deficient substituted boronic acids, such as the halobenzene boronic acids, only trace amounts of products could be obtained, which probably is due to their low nucleophilicity that cannot capture the nitrilium intermediates. Aliphatic boronic acids, such as phenethylboronic acid and cyclopentylboronic acid, do not react under our standard conditions, perhaps owing to the lack of π electrons which makes 1,4-alkyl shift difficult68.

Fig. 3: Scope of boronic acidsa.
figure 3

Reaction conditions: a1a (0.2 mmol, 1 equiv), 2a (0.3 mmol), 3 (0.36 mmol), and CHCl3/pH = 8 buffer (7:3, 1 mL) under an argon atmosphere for 24 hours unless otherwise specified. b1 (0.2 mmol, 1 equiv), 2a (0.3 mmol), 3 (0.36 mmol), and DCM/pH = 8 buffer (7:3, 1 mL) under an argon atmosphere for 24 hours.

Passerini-type reaction of alkynylboron compounds

α-Hydroxy alkynylketones are important intermediates for the synthesis of natural products and drug molecules82,83. However, the synthesis of such α-hydroxyketones has faced significant challenges and usually multiple steps are required82,83. We sought to explore the Passerini-type reaction on alkynylboron compounds, if successful, a straightforward and efficient method could be disclosed for the synthesis of α-hydroxy alkynylketones, which further demonstrates the strengths and capability of our protocol (Fig. 4). Alkynyl trifluoroborate salt was employed as the source of alkyne in our transformation owing to the instability of alkynylboronic acid. To our delight, the Passerini-type reaction of alkynyl trifluoroborate salt could proceed smoothly under the action of Lewis acid (Sc(OTf)3), and the target products (7a7d) could be obtained in a moderate yield. This reaction could not occur without Lewis acid (Sc(OTf)3), probably because a four-coordinated boron “ate” nitrilium intermediate could not be generated from potassium phenyltrifluoroborate, aldehyde, and isocyanide. Lewis acid may promote the conversion of potassium phenyltrifluoroborate (6) into phenyldifluoroborane, which could form a four-coordinated boron intermediate84.

Fig. 4: Passerini-type reaction of alkynyl trifluoroborate salt.
figure 4

Reaction conditions: 1 (0.2 mmol, 1 equiv), 2a (0.5 mmol), 6 (0.6 mmol), and THF (1.5 mL) under an argon atmosphere for 12 hours unless otherwise specified.

Late-stage modifications of complex molecules

The excellent functional group compatibility prompted our endeavors to extrapolate this synthesis scheme to late-stage modification of bioactive or therapeutic agents (Fig. 5). A series of bioactive or drug molecules (Ibuprofen, Naproxen, Ketoprofen, Gemfibrozil, Indometacin, L-Menthol and L-Borneol, and Cholesterol) were derivatized into corresponding aldehydes which, upon treatment with 4-methoxyphenyl boronic acid under established Passerini-type coupling conditions, were smoothly incorporated in eventual α-hydroxyketone derivatives 8a8h. Futhermore, conversions of arylboronic acids that were derived from drug molecules such as Epiandrosterone and Clofibrate had brought forth drug analogs 8i and 8j in moderate yields. It was thus envisioned that this method would simplify access to discover other bioactive molecules. The previous MCRs involving isocyanide exhibit poor stereoselectivity. This Passerini-type reaction of boronic acids showed similar results in terms of stereochemical control. In most cases (4l, 4n, 8a8c, and 8f8i), the dr values remained between 1:1 and 2.5:1 (see the Supplementary Information for details).

Fig. 5: Late-stage modifications of bioactive or drug molecules.
figure 5

Reaction conditions: a1a (0.2 mmol, 1 equiv), 2a (0.3 mmol), 3 (0.36 mmol), and CHCl3/pH = 8 buffer (7:3, 1 mL) under an argon atmosphere for 24 hours unless otherwise specified. bThe dr was determined by 1H NMR analysis. c1 (0.2 mmol, 1 equiv), 2a (0.3 mmol), 3 (0.36 mmol) and DCM/pH = 8 buffer (7:3, 1 mL) under an argon atmosphere for 24 hours. dThe dr was determined by HPLC analysis.

Gram-scale synthesis and synthetic applications

The practical constraint of this Passerini-type MCR with boronic acids was next evaluated through translation to gram-scale synthesis. As shown in Fig. 6a, reaction efficiencies were preserved on 2 gram-scale (10 mmol, 50 times), thus implying the application potential for industrial production of the α-hydroxyketones.

Fig. 6: Gram-scale synthesis and synthetic applications.
figure 6

a Gram-scale synthesis. b Transformations of α-hydroxyketones. c Synthesis of Harmandianone. d Syntheis of α-acyloxy lactone. e Synthesis of α, β-unsaturated lactone.

The readiness of α-hydroxyketone products for chemical manipulations was pronounced in production of 1,2-diol (9), 1,2-dione (10), quinoxaline (11), cyclic sulfamate imine (12), and poly-substituted oxazole (13) (Fig. 6b). The innate step economy of MCRs has also presented an abbreviated route towards (±) Harmandianone, a phenylpropanoid derivative isolated from Clausena Harmandiana fruits85, from simple building blocks (Fig. 6c). Of further significance, products could be precursors for entities that constitute the structural core of bioactive compounds such as α-acyloxy lactone and α,β-unsaturated lactone. The former (19) was fabricated upon lactonization of α-hydroxyketone 18 formed from methyl 4-oxobutanoate (17) (Fig. 6d). A two-step olefination and ring-closing olefin metathesis of 4a afforded α,β-unsaturated lactone product (±) 22 (Fig. 6e). The high diastereoselectivity of compound 20 may originate from the chelation of hydroxyl, carbonyl oxygen, and magnesium86.

In conclusion, we have realized the application of boronic acid as carbon nucleophiles in the manifold of Passerini reaction. Accordingly, this protocol provided simplified modular access of α-hydroxyketones from aldehydes, isocyanide, and boronic acids. The functional group tolerance of this chemistry has supported late-stage diversifications of bioactive products and pharmaceuticals through this three-component coupling reaction. The wealth of follow-up chemical conversions that could be performed on procured α-hydroxyketones has additionally illustrated the utility of this method.

Methods

General procedure A for the synthesis of α-hydroxyketones from alkylaldehydes

In air, a 10 mL schlenk tube was charged with arylboronic acids (0.36 mmol, 1.8 equiv). The tube was evacuated and filled with argon for three cycles. Then, chloroform (0.7 mL), pH = 8 buffer (0.3 mL), alkylaldehydes (0.20 mmol, 1 equiv), tertbutyl isocyanide (34 μl, 0.30 mmol, 1.5 equiv) were added under argon. The reaction was allowed to stir at corresponding temperature for 24 hours. Upon completion, proper amount of silica gel was added to the reaction mixture. After removal of the solvent, the crude reaction mixture was purified on silica gel (petroleum ether and ethyl acetate) to afford the desired products.

General procedure B for the synthesis of α-hydroxyketones from arylaldehydes

In air, a 10 mL schlenk tube was charged with arylboronic acids (0.36 mmol, 1.8 equiv). The tube was evacuated and filled with argon for three cycles. Then, dichloromethane (0.7 mL), pH = 8 buffer (0.3 mL), arylaldehydes (0.20 mmol, 1 equiv), cyclohexyl isocyanide (37 μl, 0.30 mmol, 1.5 equiv) were added under argon. The reaction was allowed to stir at room temperature for 24 hours. Upon completion, proper amount of silica gel was added to the reaction mixture. After removal of the solvent, the crude reaction mixture was purified on silica gel (petroleum ether and ethyl acetate) to afford the desired products.

General procedure C for the synthesis of α-hydroxyketones from alkynyl trifluoroborate salt

In air, a 10 mL schlenk tube was charged with alkynyl trifluoroborate salt (0.60 mmol, 3 equiv) and Sc(OTf)3 (30.0 mg, 0.06 mmol, 0.3 equiv). The tube was evacuated and filled with argon for three cycles. Then, THF (1.5 mL), aldehydes (0.20 mmol, 1 equiv), and tertbutyl isocyanide (57 μl, 0.50 mmol, 2.5 equiv) were added under argon. The reaction was allowed to stir at room temperature for 12 hours. Upon completion, proper amount of silica gel was added to the reaction mixture. After removal of the solvent, the crude reaction mixture was purified on silica gel (petroleum ether and ethyl acetate) to afford the desired products.