Next Article in Journal
Ultrafast Photochemical Dynamics of Dinitrosyl Iron Complexes Investigated by Femtosecond Time-Resolved Infrared Spectroscopy
Previous Article in Journal
Cinnarizine, a Calcium Channel Blocker, Partially Prevents the Striatal Dopamine Decline and Loss of Nigral Dopamine Neurons in the Lactacystin-Induced Rat Model of Parkinson’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalyst-Free Assembly of δ-Lactam-Based Hydrazide–Hydrazone Compounds from 3-Arylglutaconic Anhydrides and Aldazines

Department of Medicinal Chemistry, Institute of Chemistry, Saint Petersburg State University, 199034 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(18), 8834; https://doi.org/10.3390/ijms26188834
Submission received: 24 July 2025 / Revised: 30 August 2025 / Accepted: 2 September 2025 / Published: 10 September 2025
(This article belongs to the Special Issue Synthesis and Application of Natural and Inspired-Natural Products)

Abstract

A novel general approach to cyclic hydrazide–hydrazone compounds with a dihydropyridine-2-one core has been developed, involving annulation of symmetrical aldazines with 3-arylglutaconic anhydrides. This approach provides the benefits of straightforward and catalyst-free procedures, diastereoselectivity, and the ability to switch between two isomeric dihydropyridine-2-one cores based on the reaction temperature. Several post-modifications were performed on the side functional groups and the core to demonstrate the synthetic potential of the resulting products. This approach significantly expands the chemical diversity of medicinally relevant N-functionalized δ-lactams.

Graphical Abstract

1. Introduction

Acyl hydrazides [1] (R1CONR2NR3R4) and hydrazones [2,3] (R1R2C=NR3NR4R5) play a crucial role in modern drug development due to their success, with more than 50 years of extensive clinical use (Figure 1) as antibacterial and antifungal agents, anticonvulsants, and antidepressants. Recent developments have expanded their applications to include the treatment of cancer [4,5,6], chronic inflammatory diseases [7,8], neurodegenerative diseases, and viral infections [9,10]. These two structural motifs are often combined into a single functional group, and such compounds are then referred to as hydrazide–hydrazone [11,12,13,14] derivatives. Their structure often enhances drug delivery (e.g., to tumor sites) and activity by affecting solubility, stability, and toxicity. The hydrazone moiety undergoes hydrolysis under physiological conditions to produce less toxic metabolites, making the hydrazide–hydrazone functional group a promising platform for prodrug design [14]. Their sensitivity to acid-promoted hydrolysis specifically provides hydrazones with the unique ability to release the attached pharmacophore at the tumor cells having acidic pH, enabling pH-responsive targeted drug delivery [15]. Furthermore, hydrazide–hydrazone derivatives have been used in molecular switches, sensing, metal complexation [12], and the synthesis of heterocyclic compounds (e.g., β-lactams [16]). These practical applications make improving the structural diversity and synthetic accessibility of hydrazide–hydrazone derivatives highly desirable in medicinal chemistry and organic synthesis.
Although acyclic derivatives are readily available, there are relatively few known methods for preparing the corresponding cyclic compounds. One of the most interesting of these is the δ-lactam scaffold, which is considered privileged [17] in drug discovery. All three possible degrees of saturation for such a scaffold, namely piperidine-2-one, pyridine-2-one and dihydropyridine-2-one, are widely spread among the structures of bioactive compounds. The methodologies for constructing dihydropyridones offer greater versatility and efficiency as they provide access to all three types of lactams mentioned above through a single annulation strategy combined with post-condensation oxidation and reduction.
Published approaches to assembling such molecules functionalized with hydrazide–hydrazone fragments include a group of similar [4 + 2] annulation substrates of the croton series (e.g., aldehydes, acyl halides, acids, and esters) with acyclic hydrazide–hydrazone derivatives from alkyl glyoxylate (Figure 2a). These reactions are often promoted by an N-heterocyclic carbene (NHC) catalyst [18,19,20,21,22] and can be performed in an electrocatalytic [23] format. Another approach [24] is based on a multicomponent reaction involving isocyanides, N-arylidene-2-cyanoacetohydrazides, and dimethylpropiolate (Figure 2b). A significant limitation of these methods is the lack of variability of the substituent at position 6—only an ester group can be present. Additionally, only one type of regioisomer can be obtained, mostly 5,6-dihydropyridin-2(1H)-ones.
In 2017, our group reported the synthesis of cyclic hydrazide–hydrazone compounds involving selective mono-annulation of aldazines with homophthalic anhydride (Castagnoli–Cushman reaction, CCR [25]), which was used to obtain compounds with the same functional group but based on a different scaffold, namely tetrahydroisoquinolone [26].
In this study, we present a novel approach to rare cyclic hydrazide–hydrazones with a dihydropyridine-2-one scaffold that employs the same type of aldazine reactivity but extends it to a new type of anhydride: 3-arylglutaconic anhydrides (Figure 2c). This methodology significantly increases the structural diversity of available unsaturated δ-lactam-derived hydrazide–hydrazones, enabling variability in the substituent at position 6 and the introduction of an additional substituent at position 5. It also allows for the switchable, selective formation of 3,6-dihydropyridin-2(1H)-ones and 5,6-dihydropyridin-2(1H)-ones.

2. Results and Discussion

Our investigation began with conducting a reaction of anhydride 1a (Ar = p-FC6H4) with aldazine 2a (R = p-OMeC6H4) in DMSO at room temperature (Scheme 1). NMR monitoring revealed that at first acid A was formed via the Castagnoli–Cushman reaction [25,27] pathway. The key steps of CCR include Mannich-type addition of a C-nucleophilic enolized anhydride (Scheme 1, compound 1, right structure) to the electrophilic carbon atom of the imine and subsequent intramolecular N-acylation. Notably, compound A was formed as a single regio- and diastereomer (according to 1H NMR), in contrast to the similar reactions of anhydride 1a with N-alkyl imines [28] and oximes [29]. Similar acids have been shown to be thermally unstable [28], which are difficult to isolate in pure form due to decarboxylation. Therefore, in this work, we performed in situ esterification using a methyl iodide/potassium carbonate system. This protocol enabled us to isolate six esters 3a–f with a 5,6-dihydropyridin-2(1H)-one core in moderate yields (13–49% for two steps). The relative trans-configuration for compound 3 was assigned based on the 3JHH values (<1 Hz) of the vicinal methine protons from the lactam ring, according to our previous studies [28] on a similar reaction involving N-alkyl imines and 3-arylglutaconic anhydrides (the cis-isomers demonstrated 3JHH values of >5 Hz).
Considering that the initial reaction products A are prone to decarboxylation, we investigated the possibility of a targeted synthesis of the corresponding products 4 (Scheme 2). We screened the reaction conditions in order to identify the optimal parameters for the isolation of the desired product 4 (ESI, Table S1). The best results were achieved using 2 equivalents of anhydride in dry DMSO at 80 °C for 16 h without a catalyst. The protocol allowed the isolation of ten novel cyclic hydrazide–hydrazone compounds 4a–j in up to 81% yield. It should be noted that the expected [28] double-bond migration occurred alongside anionic decarboxylation, leading to the formation of another regioisomer, namely 3,6-dihydropyridin-2(1H)-one. This isomerization is presumably the result of tautomerization of the allyl anion due to stabilization of the negative charge by the electron-withdrawing carbonyl group.
Both electron-rich and electron-poor aryl groups were well tolerated in position 3 of the anhydride, including the medicinally relevant sulfonamide group (compound 4f). The scope of the imine substrate included symmetrical aldazines with aryl or alkyl groups. Substrates with strong electron-donating groups, such as alkoxy and dialkylamino groups, provided higher yields than substrates with weaker electron-donating alkyl or electron-withdrawing halogen groups. Interestingly, an aliphatic enolizable aldazine with an isopropyl moiety produced the corresponding product, 4c, in a 66% yield. Previously, similar substrates (imines and oximes) were found to be unreactive towards 3-arylglutaconic anhydrides. The structure of 3,6-dihydropyridin-2(1H)-one and the E-configuration of compound 4h were determined using X-ray crystallographic data (CCDC 2448075).
While screening reaction conditions for the synthesis of compound 4, we found that conducting the reaction at a higher temperature (150 °C) in the presence of scandium triflate leads to the formation of a second isomer, 4aa (5,6-dihydropyridin-2(1H)-one). It can be isolated in a pure form after preparative HPLC. However, our attempts to synthesize product 4aa by heating pure, isolated compound 4a and/or treating it with a base (DBU) only resulted in the decomposition of the starting material. Although for similar compounds with N-alkyl [30], NOH [29] or NH [31] functional groups the same approach was successful.
We also investigated switching from symmetrical aldazines to different types of structurally related, unsymmetrical substrates. Specifically, we examined N-substituted hydrazones with the general formula ArCH=N-N(H)PG, where PG is either a C(S)NH2, Boc, phthalimide, formyl, benzoyl, or benzhydrylidene moiety (Figure 3). However, no reaction occurred with 3-arylglutaconic anhydrides under various conditions. Therefore, we conclude that the presence of a second N-arylidene or N-alkylidene moiety in the substrate’s structure is crucial for the reactivity of the first C=N bond and for the developed dihydropyridone-based hydrazide–hydrazone approach. Presumably, moieties other than the N-arylidene/N-alkylidene serve as electron acceptors, reducing the nucleophilicity of the second nitrogen atom and/or creating steric hindrance. Both of these factors slow down the N-acylation step required to form the ring.
The synthetic potential of compounds 3a and 4a was demonstrated through a series of post-condensational modifications (see Scheme 3). 5,6-Dihydropyridin-2(1H)-one 3a was found to be selectively reduced with various hydride reagents. Reduction of the ester moiety to an alcohol 5 occurred when sodium borohydride was used. In contrast, treatment of the same substrate with sodium cyanoborohydride reduced the C=N bond only, yielding N-alkyl hydrazide 6. Another important transformation involving the hydrazide–hydrazone moiety was the exchange of the N-alkylidene residue with an N-arylidene using p-nitrobenzaldehyde. This allowed for the preparation of unsymmetrical 1-N-arylidene/6-alkyl product 7 from 3,6-dihydropyridin-2(1H)-one 4c. Notably, compound 7 could not be obtained selectively from the reaction of the anhydride and the unsymmetrical aldazine iPrCH=N-N=CHp-C6H4, as both [26] azine moieties would react to give isomeric products. Additionally, the dihydropyridine core was shown to be oxidizable with DDQ under mild conditions to produce N-arylidene pyridine-2-one 8. The latter was then treated with hydrazine in trifluoroethanol, yielding deprotected NH2-hydrazide 9.

3. Materials and Methods

3.1. General Information

NMR spectra were recorded with a 400 MHz Bruker Avance III spectrometer (Bruker, Billerica, MA, USA) (400.13 MHz for 1H and 100.61 MHz for 13C) in CDCl3 or DMSO-d6 and were referenced to residual solvent proton signals (δH = 7.26 and 2.50, respectively) and solvent carbon signals (δC = 77.16 and 39.52, respectively). Mass spectra were recorded with an HRMS-ESI-qTOF spectrometer Nexera LCMS-9030 (Shimadzu, Kyoto, Japan) or MaXis II Bruker Daltonic GmbH (Bruker, Billerica, MA, USA) (electrospray ionization mode, positive ions detection). Flash column chromatography on silica (Merck, 230–400 mesh) was carried out using the Biotage Isolera Prime instrument (Biotage, Uppsala, Sweden). TLC was performed on aluminum-backed pre-coated plates (0.25 mm) with silica gel 60 F254 with a suitable solvent system and was visualized using UV fluorescence. Preparative HPLC was carried out on a compact preparative system ECOM ECS28P00 (ECOM, Chrastany u Prahy, Czech Republic), equipped with a spectrophotometric detector or Shimadzu LC-20AP (Shimadzu, Kyoto, Japan). Column: YMC-Pack SIL-06 (YMC, Kyoto, Japan), 5 µm, 250 × 20 mm or Agilent Zorbax prepHT XDB-C18 (Agilent, Santa Clara, California), 5 μm, 21.2 × 150 mm. Dimethylsulfoxide, DMSO, was dried by distillation from calcium hydride and was stored over activated molecular sieves 4Å.

3.2. General Procedure for the Preparation of Lactams 3a-f and Their Analytical Data

In a screw-cap vial equipped with a magnetic stir bar, the corresponding arylglutaconic anhydride [28] 1 (2 equivalents, 0.06–0.54 mmol) and the corresponding aldazine 2 (1 equivalents, 0.03–0.27 mmol) were mixed in dry DMSO (0.06–0.54 M 0.5 mL). The resulting mixture was placed in a pre-heated to 30 °C metal heating block. DMSO was removed from the reaction mixture after 16 h using freeze-drying at 10−2 mbar. Next, the reaction mixture was dissolved in acetonitrile, ACN (1 mL), followed by the addition of CH3I (5 equivalents, 0.3–1.35 mmol) and K2CO3 (1.5 equivalents, 0.09–0.41 mmol). After 16 h, the mixture was concentrated under reduced pressure and diluted with EA (20 mL) and water (20 mL). The aqueous layer was extracted with EA (3 × 20 mL), then the combined organic phase was dried over Na2SO4 and evaporated. Then, the resulting mixture was purified by column chromatography on silica gel with a linear gradient (5−50%) of acetone in hexane (total volume of eluent, 400 mL) to provide pure compounds (3a–f). Prepared compound 3 should not be stored as chloroform solutions due to possible hydrolysis.

3.2.1. Methyl (2RS,3RS)-4-(4-Fluorophenyl)-1-(((E)-4-methoxybenzylidene)amino)-2-(4-methoxyphenyl)-6-oxo-1,2,3,6-tetrahydropyridine-3-carboxylate (3a)

Yield 35 mg, 49%, dr > 20:1, yellow solid. 1H NMR (400 MHz, CDCl3) δ 8.62 (s, 1H), 7.74–7.60 (m, 2H), 7.40–7.29 (m, 2H), 7.24–7.19 (m, 2H), 7.04 (t, J = 8.62 Hz, 2H), 6.90–6.79 (m, 4H), 6.49 (s, 1H), 5.80 (d, J = 1.69 Hz, 1H), 4.07 (d, J = 1.72 Hz, 1H), 3.82 (s, 3H), 3.76 (s, 3H), 3.72 (s, 3H). 13C NMR (101 MHz, CDCl3) δ = 170.3, 163.7 (d, J = 251.0), 161.7, 161.6, 159.5, 151.1, 142.8, 133.0 (d, J = 3.4), 129.7, 129.6, 128.3 (d, J = 8.5), 127.3, 122.6 (d), 116.1 (d, J = 21.7), 114.6, 114.1, 62.9, 55.5, 55.4, 53.4, 51.4. 19F NMR (376 MHz, CDCl3) δ −110.72. HRMS (ESI) m/z: [M+H]+ Calcd for C28H26FN2O5+ 489.1820; Found 489.1821.

3.2.2. Methyl (2RS,3RS)-1-(((E)-4-Methoxybenzylidene)amino)-2-(4-methoxyphenyl)-6-oxo-4-(4-(pyrrolidin-1-ylsulfonyl)phenyl)-1,2,3,6-tetrahydropyridine-3-carboxylate (3b)

Yield 19 mg, 40%, dr > 20:1, yellow oil. 1H NMR (400 MHz, CDCl3) δ 8.66 (s, 1H), 7.84–7.75 (m, 2H), 7.68–7.61 (m, 2H), 7.51–7.42 (m, 2H), 7.24–7.16 (m, 2H), 6.92–6.80 (m, 4H), 6.58 (s, 1H), 5.86–5.79 (m, 1H), 4.10 (d, J = 1.67 Hz, 1H), 3.82 (s, 3H), 3.77 (s, 3H), 3.74 (s, 3H), 3.27–3.18 (m, 4H), 1.82–1.72 (m, 4H). 13C NMR (126 MHz, CDCl3) δ 13C NMR (126 MHz, CDCl3) δ = 169.8, 161.7, 161.0, 159.4, 151.9, 142.3, 140.8, 138.1, 132.0, 129.7, 128.0, 127.2, 126.9, 126.8, 124.6, 114.5, 114.0, 62.8, 55.4, 55.3, 53.4, 51.2, 47.9, 25.3. HRMS (ESI) m/z: [M+H]+ Calcd for 604.2112 C32H34N3O7S+; Found 604.2112.

3.2.3. Methyl (2RS,3RS)-1-(((E)-4-Methoxybenzylidene)amino)-2,4-bis(4-methoxyphenyl)-6-oxo-1,2,3,6-tetrahydropyridine-3-carboxylate (3c)

Yield 43 mg, 37%, dr > 20:1, yellow solid. 1H NMR (400 MHz, CDCl3) δ 8.59 (s, 1H), 7.69–7.57 (m, 2H), 7.38–7.32 (m, 2H), 7.24–7.19 (m, 2H), 6.89–6.85 (m, 4H), 6.85–6.82 (m, 2H), 6.50 (s, 1H), 5.79–5.77 (m, 1H), 4.12 (d, J = 1.63 Hz, 1H), 3.81 (s, 3H), 3.80 (s, 3H), 3.75 (s, 3H), 3.71 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 170.6, 162.0, 161.5, 161.1, 159.4, 150.4, 143.2, 129.9, 129.5, 128.9, 127.8, 127.5, 127.4, 120.5, 114.5, 114.4, 114.1, 62.8, 55.5, 55.5, 55.4, 53.3, 51.1. HRMS (ESI) m/z: [M+H]+ Calcd for C29H29N2O6+ 501.2020; Found 501.2014.

3.2.4. Methyl (2RS,3RS)-1-(((E)-4-Methoxybenzylidene)amino)-2-(4-methoxyphenyl)-6-oxo-4-phenyl-1,2,3,6-tetrahydropyridine-3-carboxylate (3d)

Yield 26 mg, 21%, dr > 20:1, yellow solid. 1H NMR (400 MHz, CDCl3) δ 8.61 (s, 1H), 7.68–7.63 (m, 2H), 7.25–7.19 (m, 2H), 6.92–6.80 (m, 5H), 6.56 (s, 1H), 5.86–5.66 (m, 1H), 4.14 (d, J = 1.65 Hz, 1H), 3.81 (s, 3H), 3.75 (s, 3H), 3.71 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 170.4, 161.7, 161.6, 159.4, 150.9, 143.9, 136.8, 129.9, 129.7, 129.6, 129.0, 127.4, 126.3, 125.2, 122.6, 114.5, 114.1, 62.9, 55.5, 55.4, 53.3, 51.3. HRMS (ESI) m/z: [M+H]+ Calcd for C28H27N2O5+ 471.1914; Found 471.1919.

3.2.5. Methyl (2RS,3RS)-4-(4-Fluorophenyl)-2-isopropyl-1-(((E)-2-methylpropylidene)amino)-6-oxo-1,2,3,6-tetrahydropyridine-3-carboxylate (3e)

Yield 28 mg, 16%, dr > 20:1, yellow solid. 1H NMR (400 MHz, CDCl3) δ 8.24 (d, J = 5.80 Hz, 1H), 7.52–7.44 (m, 2H), 7.15–7.02 (m, 2H), 6.35 (s, 1H), 4.22 (dd, J = 7.63, 1.47 Hz, 1H), 3.85 (d, J = 1.49 Hz, 1H), 3.66 (s, 3H), 2.68–2.56 (m, 1H), 2.26–2.14 (m, 1H), 1.14 (dd, J = 6.85, 2.34 Hz, 6H), 1.01 (d, J = 6.88 Hz, 3H), 0.95 (d, J = 6.76 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ 171.1, 163.8, 163.7 (d, J = 243.00 Hz), 143.2, 133.0 (d, J = 3.39 Hz), 128.2 (d, J = 8.36 Hz), 122.2, 116.2 (d, J = 21.96 Hz), 66.5, 53.1, 45.0, 32.7, 31.6, 20.3, 20.1, 19.9. 19F NMR (376 MHz, CDCl3) δ −110.9. HRMS (ESI) m/z: [M+H]+ Calcd for C20H26FN2O3+ 361.1922; Found 361.1931.

3.2.6. Methyl (2RS,3RS)-4-(4-Fluorophenyl)-1-(((E)-4-methylbenzylidene)amino)-6-oxo-2-(p-tolyl)-1,2,3,6-tetrahydropyridine-3-carboxylate (3f)

Yield 14 mg, 13%, dr > 20:1, white solid. 1H NMR (400 MHz, CDCl3) δ 8.62 (s, 1H), 7.59 (d, J = 8.12 Hz, 2H), 7.37–7.31 (m, 2H), 7.21–7.09 (m, 6H), 7.09–6.97 (m, 2H), 6.49 (s, 1H), 5.86–5.80 (m, 1H), 4.11 (d, J = 1.63 Hz, 1H), 3.72 (s, 3H), 2.35 (s, 3H), 2.30 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 170.2, 163.8 (d, J = 250.99 Hz), 161.7, 150.9, 143.0, 140.8, 138.1, 134.7, 133.0 (d, J = 3.36 Hz), 131.9, 129.9, 129.4, 128.3 (d, J = 8.46 Hz), 128.0, 126.0, 122.6, 116.1 (d, J = 21.83 Hz), 63.1, 53.4, 51.4, 21.6, 21.1. 19F NMR (376 MHz, CDCl3) δ −110.7. HRMS (ESI) m/z: [M+H]+ Calcd for C28H26FN2O3+ 457.1922; Found 457.1920.

3.3. Synthesis of Lactams 4a–j and Their Analytical Data

The corresponding arylglutaconic anhydride [28] 1 (2 equivalents, 0.06–0.54 mmol) and the aldazine 2 (1 equivalents, 0.03–0.27 mmol) were mixed in dry DMSO (0.5 mL; 0.06–0.54 M) and placed in a screw-cap vial equipped with a magnetic stir bar. The resulting mixture was heated to 80 °C using a pre-heated metal heating block and left stirring for 16 h. After cooling to room temperature, the mixture was purified by column chromatography on silica gel with a linear gradient (5−50%) of EA in Et2O (total volume of eluent, 400 mL) to provide pure compounds 4a–j *.
* For 4d, the reaction time was 2 days;
For 4g scandium (III) trifluoromethanesulfonate (10% mol, 0.005 mmol) was added to the reaction mixture, the reaction time was 2 days;
For 4c, the corresponding aldazine was used in excess (5 equivalents, 1.2 mmol) and the reaction temperature was 60 °C; all prepared compounds 4 should not be stored as chloroform solutions due to possible hydrolysis.

3.3.1. (E)-4-(4-Fluorophenyl)-1-((4-methoxybenzylidene)amino)-6-(4-methoxyphenyl)-3,6-dihydropyridin-2(1H)-one (4a)

Yield 18 mg, 81%, dr > 20:1, white solid. 1H NMR (400 MHz, CDCl3) δ = 8.61 (s, 1H), 7.61 (d, J = 8.8, 2H), 7.47–7.32 (m, 2H), 7.25–7.15 (m, 2H), 7.05 (t, J = 8.7, 2H), 6.94–6.76 (m, 4H), 6.19 (dd, J = 4.5, 2.1, 1H), 5.63 (d, J = 3.9, 1H), 3.82 (s, 3H), 3.77 (s, 3H), 3.75–3.55 (m). 13C NMR (101 MHz, CDCl3) δ 13C NMR (101 MHz, CDCl3) δ = 164.1, 162.8 (d, J = 248.0), 162.0, 159.5, 158.0, 133.9 (d, J = 3.4), 131.9, 129.8, 129.7, 128.4, 127.0, 126.9, 121.7 (d, J = 1.3), 115.7 (d, J = 21.5), 114.5, 114.1, 65.1, 55.5, 55.4, 36.0. 19F NMR (376 MHz, CDCl3) δ –110.83. HRMS (ESI) m/z: [M+H]+ Calcd for C26H24FN2O3+ 431.1756; Found 431.1773.

3.3.2. (E)-1-((4-Methoxybenzylidene)amino)-4,6-bis(4-methoxyphenyl)-3,6-dihydropyridin-2(1H)-one (4b)

Yield 15 mg, 74%, dr > 20:1, yellow oil. 1H NMR (400 MHz, CDCl3) δ 8.61 (s, 1H), 7.71–7.54 (m, 2H), 7.45–7.34 (m, 2H), 7.26–7.20 (m, 2H), 6.98–6.61 (m, 6H), 6.16 (dd, J = 4.54, 1.90 Hz, 1H), 5.62 (q, J = 3.68 Hz, 1H), 3.82 (s, 3H), 3.81 (s, 3H), 3.77 (s, 3H), 3.69 (dd, J = 3.79, 2.17 Hz, 1H), 3.62 (dd, J = 21.04, 2.96 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ 164.4, 161.9, 159.8, 159.4, 157.6, 132.3, 130.2, 129.9, 129.7, 128.4, 127.1, 126.3, 120.0, 114.4, 114.1, 114.1, 65.1, 55.5, 55.4, 35.9. HRMS (ESI) m/z: [M+H]+ Calcd for C27H27N2O4+ 443.1965; Found 443.1973.

3.3.3. (E)-4-(4-Fluorophenyl)-6-isopropyl-1-((2-methylpropylidene)amino)-3,6-dihydropyridin-2(1H)-one (4c)

Yield 48 mg, 66%, dr > 20:1, orange oil. 1H NMR (400 MHz, CDCl3) δ 8.20 (d, J = 5.57 Hz, 1H), 7.43–7.29 (m, 2H), 7.11–6.95 (m, 2H), 6.07 (dd, J = 4.51, 2.39 Hz, 1H), 4.40 (p, J = 4.03 Hz, 1H), 3.49 (ddd, J = 21.02, 4.05, 2.45 Hz, 1H), 3.36 (dd, J = 20.94, 2.84 Hz, 1H), 2.67 (pd, J = 6.88, 5.53 Hz, 1H), 2.31 (dtt, J = 10.82, 6.94, 3.91 Hz, 1H), 1.24–1.12 (m, 6H), 1.04 (d, J = 7.07 Hz, 3H), 0.81 (d, J = 6.77 Hz, 3H). 13C NMR (101 MHz, Chloroform-d) δ 169.6, 161.7 (d, J = 248.5 Hz), 159.5, 134.4 (d, J = 3.2 Hz), 128.8, 127.4 (d, J = 7.9 Hz), 126.8, 116.1 (d, J = 21.9 Hz), 66.4, 49.4, 30.8, 21.8, 19.0. HRMS (ESI) m/z: [M+H]+ Calcd for C18H24FN2O+ 303.1867; Found 303.1857.

3.3.4. (E)-1-((4-Methoxybenzylidene)amino)-6-(4-methoxyphenyl)-4-(4-nitrophenyl)-3,6-dihydropyridin-2(1H)-one (4d)

Yield 12 mg, 62%, dr > 20:1, yellow solid. 1H NMR (400 MHz, CDCl3) δ 8.66 (s, 1H), 8.25–8.21 (m, 2H), 7.66–7.54 (m, 4H), 7.25–7.21 (m, 2H), 6.90–6.85 (m, 4H), 6.43 (dd, J = 4.53, 1.99 Hz, 1H), 5.68 (q, J = 3.76 Hz, 1H), 3.74 (dd, J = 3.87, 2.21 Hz, 3H), 3.65 (dd, J = 21.07, 2.98 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ 163.3, 162.1, 159.6, 158.6, 147.5, 144.0, 131.2, 129.8, 129.0, 128.4, 126.8, 126.0, 125.7, 124.1, 114.6, 114.2, 65.2, 55.5, 55.4, 35.5. HRMS (ESI) m/z: [M+H]+ Calcd for C26H24N3O5+ 458.1710; Found 458.1720.

3.3.5. (E)-1-((4-Methoxybenzylidene)amino)-6-(4-methoxyphenyl)-4-phenyl-3,6-dihydropyridin-2(1H)-one (4e)

Yield 13 mg, 59%, dr > 20:1, yellow oil. 1H NMR (400 MHz, CDCl3) δ 8.63 (s, 1H), 7.65–7.54 (m, 2H), 7.48–7.41 (m, 2H), 7.41–7.34 (m, 2H), 7.34–7.30 (m, 1H), 7.24 (d, J = 8.71 Hz, 2H), 6.93–6.82 (m, 4H), 6.26 (dd, J = 4.53, 2.01 Hz, 1H), 5.64 (q, J = 3.73 Hz, 1H), 3.82 (s, 3H), 3.77 (s, 3H), 3.73 (dd, J = 3.81, 2.19 Hz, 1H), 3.65 (dd, J = 21.20, 3.00 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ 164.3, 161.9, 159.4, 157.8, 137.8, 132.0, 130.6, 129.7, 128.8, 128.4, 128.3, 127.0, 125.2, 121.8, 114.5, 114.1, 65.1, 55.5, 55.4, 35.8. HRMS (ESI) m/z: [M+H]+ Calcd for C26H25N2O3+ 413.1860; Found 413.1875.

3.3.6. (E)-1-((4-Methoxybenzylidene)amino)-6-(4-methoxyphenyl)-4-(4-(pyrrolidin-1-ylsulfonyl)phenyl)-3,6-dihydropyridin-2(1H)-one (4f)

Yield 45 mg, 53%, dr > 20:1, white solid. 1H NMR (400 MHz, CDCl3) δ 8.65 (s, 1H), 7.83 (d, J = 8.26 Hz, 2H), 7.61 (d, J = 8.54 Hz, 2H), 7.57 (d, J = 8.14 Hz, 2H), 7.23 (d, J = 8.51 Hz, 2H), 6.87 (dd, J = 8.60, 3.83 Hz, 4H), 6.38 (dd, J = 4.70, 2.12 Hz, 1H), 5.67 (q, J = 3.78 Hz, 1H), 3.82 (s, 3H), 3.78 (s, 3H), 3.76–3.72 (m, 1H), 3.65 (dd, J = 21.01, 2.94 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ 163.6, 162.1, 159.6, 158.5, 141.9, 136.7, 131.5, 129.8, 129.4, 128.4, 128.1, 126.9, 125.7, 124.6, 114.6, 114.2, 65.3, 55.5, 55.4, 48.1, 35.6, 25.4. HRMS (ESI) m/z: [M+H]+ Calcd for C30H32N3O5S+ 546.2057; Found 546.2076.

3.3.7. (E)-4-(4-Chlorophenyl)-1-((4-methoxybenzylidene)amino)-6-(4-methoxyphenyl)-3,6-dihydropyridin-2(1H)-one (4g)

Yield 11 mg, 52%, dr > 20:1, white solid. 1H NMR (400 MHz, CDCl3) δ 8.61 (s, 1H), 7.66–7.56 (m, 2H), 7.35 (d, J = 3.72 Hz, 4H), 7.26–7.18 (m, 2H), 6.90–6.80 (m, 4H), 6.23 (dd, J = 4.54, 1.95 Hz, 1H), 5.63 (q, J = 3.76 Hz, 1H), 3.82 (s, 3H), 3.77 (s, 3H), 3.68 (dd, J = 3.86, 2.23 Hz, 1H), 3.60 (dd, J = 21.11, 2.98 Hz, 1H). 13C NMR (101 MHz, Toluene-d8) δ 164.0, 162.4, 160.2, 157.6, 138.1, 137.6, 137.0, 134.4, 133.5, 130.1, 130.0, 129.1, 127.0, 123.1, 114.9, 114.6, 66.9, 55.0, 55.0, 36.5. HRMS (ESI) m/z: [M+H]+ Calcd for C26H24ClN2O3+ 447.1470; Found 447.1477.

3.3.8. (E)-1-((4-(Dimethylamino)benzylidene)amino)-6-(4-(dimethylamino)phenyl)-4-(4-fluorophenyl)-3,6-dihydropyridin-2(1H)-one (4h)

Yield 11 mg, 48%, dr > 20:1, red solid. 1H NMR (400 MHz, Toluene-d8) δ 9.42 (s, 1H), 7.73–7.67 (m, 2H), 7.33–7.27 (m, 2H), 7.00–6.92 (m, 2H), 6.85–6.77 (m, 2H), 6.61–6.55 (m, 2H), 6.46–6.37 (m, 2H), 5.97 (dd, J = 4.60, 2.05 Hz, 1H), 5.59 (q, J = 3.67 Hz, 1H), 3.56 (ddd, J = 20.96, 3.76, 2.21 Hz, 1H), 3.46 (dd, J = 20.91, 2.92 Hz, 1H), 2.56 (s, 6H), 2.48 (s, 6H). 13C NMR (101 MHz, Toluene-d8) δ 163.6, 163.2 (d, J = 246.73 Hz), 159.1, 152.5, 150.8, 135.1 (d, J = 3.32 Hz), 129.9, 128.9, 127.4 (d, J = 7.94 Hz), 124.1, 123.2, 123.2, 115.8 (d, J = 21.28 Hz), 113.4, 112.4, 66.8, 40.5, 40.0, 36.8. 19F NMR (376 MHz, Toluene-d8) δ −114.52. HRMS (ESI) m/z: [M+H]+ Calcd for C28H30FN4O+ 457.2398; Found 457.2401.

3.3.9. Crystallographic Data for 4h

Crystals of C28H29FN4O (M = 456.55) are monoclinic, space group P21/n, at 100(2) K: a = 13.6438(3) Å, b = 6.3904(2) Å and c = 26.9347(7) Å, α = 90°, β = 100.653(3)°, γ = 90°, V = 2307.94(11) Å3, Z = 4, dcalc = 1.314 g/cm3, μ = 0.698 mm−1, F(000) = 968.0. 16,247 reflections were measured, and 4256 independent reflections (Rint = 0.0468) were used in a further refinement. The final R1 was 0.0994 (I ≥ 2σ (I)), and wR2 was 0.2876 (all data). X-ray diffraction study of 4h was performed at 100(2) K on the Rigaku XtaLAB Synergy-S (Rigaku, Tokyo, Japan) diffractometer (HyPix-6000HE type detector (Rigaku, Tokyo, Japan)) using Cu Kα (λ = 1.54184 Å) radiation. The structure was solved with the ShelXT [32] structure solution program using Intrinsic Phasing and refined with the ShelXL [33] refinement package incorporated in the OLEX2 program package [34] using Least Squares minimization. Empirical absorption correction was applied in the CrysAlisPro program complex using spherical harmonics, implemented in the SCALE3 ABSPACK scaling algorithm. The hydrogen atom positions were fixed geometrically at calculated distances and allowed to ride on the parent atoms.
CCDC 2448075 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via http://www.ccdc.cam.ac.uk (accessed on 23 July 2025).

3.3.10. (E)-4-(4-Fluorophenyl)-1-((4-methylbenzylidene)amino)-6-(p-tolyl)-3,6-dihydropyridin-2(1H)-one (4i)

Yield 32 mg, 34%, dr > 20:1, yellow solid. 1H NMR (400 MHz, CDCl3) δ 8.65 (s, 1H), 7.55 (d, J = 8.03 Hz, 2H), 7.43–7.33 (m, 2H), 7.23–7.09 (m, 6H), 7.08–7.02 (m, 2H), 6.20 (dd, J = 4.66, 2.15 Hz, 1H), 5.67 (q, J = 3.40 Hz, 1H), 3.73 (ddd, J = 21.49, 3.62, 2.23 Hz, 1H), 3.62 (dd, J = 21.21, 2.81 Hz, 1H), 2.35 (s, 3H), 2.31 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 164.4, 162.8 (d, J = 247.84 Hz), 156.8, 141.2, 138.0, 136.8, 133.9 (d, J = 3.33 Hz), 131.6, 129.9, 129.6, 129.4, 128.1, 126.9 (d, J = 8.10 Hz), 126.8, 121.7, 115.7 (d, J = 21.59 Hz), 65.1, 36.0, 21.6, 21.2. 19F NMR (376 MHz, CDCl3) δ -113.54. HRMS (ESI) m/z: [M+H]+ Calcd for C26H24FN2O+ 399.1867; Found 399.1883.

3.3.11. (E)-1-((4-Chlorobenzylidene)amino)-6-(4-chlorophenyl)-4-(4-fluorophenyl)-3,6-dihydropyridin-2(1H)-one (4j)

Yield 22 mg, 21%, dr > 20:1, yellow solid. 1H NMR (400 MHz, CDCl3). δ 8.89 (s, 1H), 7.57 (d, J = 8.39 Hz, 2H), 7.44–7.37 (m, 3H), 7.33 (d, J = 8.40 Hz, 4H), 7.07 (t, J = 8.53 Hz, 3H), 6.16 (dd, J = 4.62, 2.01 Hz, 1H), 5.67 (q, J = 3.75 Hz, 1H), 3.72 (dt, J = 21.85, 2.67 Hz, 1H), 3.63 (dd, J = 21.31, 2.98 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ = 164.7, 163.0 (d, J = 248.6), 155.8, 138.5, 137.0, 134.1, 133.5 (d, J = 3.3), 132.9, 130.4, 129.4, 129.2, 129.1, 128.4, 127.0 (d, J = 8.2), 120.9, 115.9 (d, J = 21.6), 65.6, 36.1.19F NMR (376 MHz, CDCl3) δ -113.0. HRMS (ESI) m/z: [M+H]+ Calcd for C24H18Cl2FN2O+ 439.0775; Found 439.0785.

3.3.12. (E)-4-(4-Fluorophenyl)-1-((4-methoxybenzylidene)amino)-6-(4-methoxyphenyl)-5,6-dihydropyridin-2(1H)-one (4aa)

In a screw-cap vial equipped with a magnetic stir bar, the corresponding arylglutaconic anhydride 1a (2 equivalents, 0.5 mmol) and the corresponding aldazine 2a (1 equivalents, 0.24 mmol) were mixed in dry DMSO (0.5 mL, 0.5 M). The reaction mixture was heated to 150 °C for 16h using a pre-heated metal heating block. After cooling to room temperature, the mixture was separated by column chromatography on silica gel with a linear gradient (5−50%) of ethyl acetate in Et2O (total volume of eluent, 400 mL) to afford pure compounds 4a (28 mg, 26%, dr > 20:1) and 4aa.
Yield 23 mg, 22%, dr > 20:1, white solid. 1H NMR (400 MHz, CDCl3) δ 8.59 (s, 1H), 7.62 (d, J = 8.8 Hz, 2H), 7.48–7.35 (m, 2H), 7.20 (d, J = 8.7 Hz, 2H), 7.05 (t, J = 8.6 Hz, 2H), 6.85 (dd, J = 10.9, 8.7 Hz, 3H), 6.40 (d, J = 2.3 Hz, 1H), 5.38 (dd, J = 6.8, 3.3 Hz, 1H), 3.82 (s, 3H), 3.76 (s, 3H), 3.48 (ddd, J = 17.1, 6.8, 2.4 Hz, 1H), 3.11 (dd, J = 17.1, 3.3 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ = 164.0, 162.8 (d, J = 248.2), 162.0, 159.5, 157.9, 133.9 (d, J = 3.4), 132.0, 129.7, 129.7, 128.4, 127.0, 126.9 (d, J = 8.1), 121.7 (d, J = 1.1), 115.7 (d, J = 21.6), 114.5, 114.1, 65.1, 55.5, 55.4, 36.0. 19F NMR (376 MHz, CDCl3) δ -113.58. HRMS (ESI) m/z: [M+H]+ Calcd for C26H24FN2O3+ 431.1756; Found 431.1774.

3.4. Preparation of Compounds 5–9

3.4.1. (E)-4-(4-fluorophenyl)-5-(hydroxymethyl)-1-((4-methoxybenzylidene)amino)-6-(4-methoxyphenyl)-5,6-dihydropyridin-2(1H)-one (5)

Compound 3a (45 mg, 0.09 mmol) was dissolved in MeOH (2 mL) and cooled with an ice-water bath under stirring, followed by the addition of solid NaBH4 (18 mg, 0.48 mmol) in one portion. After stirring for 16 h at room temperature, 5 drops of acetic acid were added to the reaction mixture. The resulting solution was diluted with ether (25 mL) and water (10 mL). The organic layer was removed, and the aqueous layer was extracted twice with 10 mL of ether. The combined organics were washed with NaHCO3 sat., water, and brine, dried over sodium sulfate, filtered, and concentrated. The residue was purified using HPLC on silica (mobile phase: linear gradient of acetone in hexane 2–50%, total volume 250 mL). Yield 24 mg, 59%, dr > 20:1, yellow powder.
1H NMR (400 MHz, CDCl3) δ 8.22 (s, 1H), 7.63 (d, J = 8.8 Hz, 1H), 7.41–7.27 (m, 2H), 7.18 (d, J = 8.7 Hz, 1H), 7.01 (t, J = 8.6 Hz, 2H), 6.83 (dd, J = 8.6, 5.2 Hz, 4H), 6.38 (s, 1H), 5.68 (s, 1H), 3.80 (s, 3H), 3.77–3.71 (m, 5H), 3.42 (t, J = 6.8 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ 161.9, 160.5 (d, J = 242.0 Hz), 148.4, 146.9, 132.3 (d, J = 3.3 Hz), 132.1, 130.4, 129.8, 128.3 (d, J = 8.4 Hz), 127.3, 127.1, 120.9, 116.1 (d, J = 21.8 Hz), 114.6, 114.2, 62.6, 59.5, 55.5, 55.4, 49.6. HRMS (ESI) m/z: [M+Na]+ Calcd for C26H26FN2NaO4+ 483.1691; Found 483.1697.

3.4.2. Methyl 4-(4-fluorophenyl)-1-((4-methoxybenzyl)amino)-2-(4-methoxyphenyl)-6-oxo-1,2,3,6-tetrahydropyridine-3-carboxylate (6)

Compound 3a (20 mg, 0.04 mmol) was dissolved in a mixture of methanol and acetic acid (2 + 0.2 mL) and cooled to 0 °C with an ice-water bath, followed by the addition of solid NaBH3CN (8 mg, 0.13 mmol) in one portion. After stirring for 16 h at room temperature, the reaction mixture was diluted with water (15 mL) and ether (25 mL). The organic layer was separated, washed with water, and aq. NaHCO3 sat., water and brine, dried over Na2SO4, filtered and concentrated in vacuo to provide pure title compound as a yellow solid. Yield 15 mg, 75%, dr > 20:1.
1H NMR (400 MHz, CDCl3) δ 7.34–7.27 (m, 4H), 7.15–7.08 (m, 2H), 7.05–6.97 (m, 2H), 6.84 (dd, J = 8.6, 5.7 Hz, 4H), 6.43 (s, 1H), 5.34 (d, J = 1.5 Hz, 1H), 3.97 (s, 2H), 3.89 (d, J = 1.5 Hz, 1H), 3.79 (s, 3H), 3.76 (s, 3H), 3.65 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 170.3, 163.7, 163.6 (d, J = 250.7 Hz), 159.5, 159.2, 143.5, 133.3 (d, J = 3.5 Hz), 130.8, 130.4, 129.4, 128.2 (d, J = 8.5 Hz), 127.5, 121.1, 116.0 (d, J = 21.8 Hz), 114.4, 113.9, 63.5, 55.4, 55.4, 54.2, 53.1, 51.4. HRMS (ESI) m/z: [M+H]+ Calcd for C28H28FN2O5+ 491.1977; Found 491.1979.

3.4.3. (E)-4-(4-Fluorophenyl)-6-isopropyl-1-((4-nitrobenzylidene)amino)-5,6-dihydropyridin-2(1H)-one (7)

In a screw-cap vial equipped with a septum cap and a magnetic stir bar, compound 4c (40 mg, 0.13 mmol) and p-nitrobenzaldehyde (98 mg, 0.65 mmol) were mixed in dry toluene (0.5 mL). The resulting mixture was placed in a pre-heated to 110 °C metal heating block, and a needle was inserted into the septum. After 16 h, the mixture was cooled to room temperature, concentrated under reduced pressure, and purified by column chromatography on silica gel with a linear gradient (5−50%) of acetone in hexane (total volume of eluent, 400 mL) to provide the pure title compound. Yield 11 mg, 21%, dr > 20:1, yellow solid.
1H NMR (400 MHz, CDCl3) δ 9.64 (s, 1H), 8.28–8.21 (m, 2H), 7.89–7.82 (m, 2H), 7.59–7.48 (m, 2H), 7.18–7.10 (m, 2H), 6.26 (d, J = 2.42 Hz, 1H), 4.09 (td, J = 6.89, 2.81 Hz, 1H), 3.15 (ddd, J = 17.65, 6.87, 2.49 Hz, 1H), 2.94 (dd, J = 17.61, 2.84 Hz, 1H), 2.36 (h, J = 6.82 Hz, 1H), δ 1.0 (d, J = 6.9 Hz, 3H), 1.0 (d, J = 6.7 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ 164.0 (d, J = 251.18 Hz), 163.4, 148.7, 148.5, 147.9, 142.1, 133.4 (d, J = 3.37 Hz), 128.0 (d, J = 8.43 Hz), 127.9, 124.1, 116.2 (d, J = 21.81 Hz), 66.3, 31.2, 20.1, 19.5. 19F NMR (376 MHz, DMSO-d6) δ −110.3. HRMS (ESI) m/z: [M+H]+ Calcd for C19H19FN3O4+ 382.1561; Found 382.1571.

3.4.4. (E)-4-(4-Fluorophenyl)-1-((4-methoxybenzylidene)amino)-6-(4-methoxyphenyl)pyridin-2(1H)-one (8)

To a stirred solution of compound 4a (30 mg, 0.07 mmol) in dry DCM (1 mL), 4,5-dichloro-3,6-dioxocyclohexa-1,4-diene-1,2-dicarbonitrile (DDQ) (1.1 equivalents, 0.08 mmol) was added in one portion. After stirring 16 h at room temperature, the precipitate formed was centrifuged, washed with DCM (5 mL), 10% NaOH solution (5 mL), and water (5 mL), then dried in a vacuum to obtain the target product.
Yield 22 mg, 72%, dr > 20:1, white solid. 1H NMR (400 MHz, DMSO-d6) δ 8.76 (s, 1H), 7.61–7.55 (m, 4H), 7.38–7.32 (m, 2H), 7.10 (t, J = 8.62 Hz, 2H), 6.89–6.77 (m, 4H), 6.69 (d, J = 2.21 Hz, 1H), 6.38 (d, J = 2.29 Hz, 1H), 3.78 (s, 3H), 3.75 (s, 3H). 13C NMR (101 MHz, CDCl3) δ = 166.7, 161.6 (d, J = 283.4), 160.1, 148.9, 148.3, 133.9 (d, J = 3.3), 131.2, 130.8, 129.7, 128.8 (d, J = 8.5), 127.4, 125.9, 125.7, 116.1 (d, J = 21.7), 115.5, 114.3, 113.5, 106.5, 55.6, 55.5. 19F NMR (376 MHz, CDCl3) δ −111.93. HRMS (ESI) m/z: [M+H]+ Calcd for 429.1609 C26H22FN2O3+; Found 429.1600.

3.4.5. 1-Amino-4-(4-fluorophenyl)-6-(4-methoxyphenyl)pyridin-2(1H)-one (9)

To the mixture of compound 7 (30 mg, 0.07 mmol) in dry trifluoroethanol (0.3 mL), hydrazine hydrate (0.2 mL) was added. After stirring at room temperature for 16 h, the precipitate formed was centrifuged, washed with Et2O (5 mL), and then dried in a vacuum to obtain the title product.
Yield 18 mg, 82%, dr >20:1, white solid. 1H NMR (400 MHz, DMSO-d6) δ 7.91–7.71 (m, 2H), 7.65 (d, J = 8.70 Hz, 2H), 7.30 (t, J = 8.83 Hz, 2H), 7.03 (d, J = 8.84 Hz, 2H), 6.76 (d, J = 2.25 Hz, 1H), 6.50 (d, J = 2.32 Hz, 1H), 3.82 (s, 3H). 13C NMR (101 MHz, DMSO-d6) δ 162.8 (d, J = 246.80 Hz), 160.3, 159.8, 148.2, 147.3, 133.5 (d, J = 3.09 Hz), 130.8, 129.1 (d, J = 8.50 Hz), 126.5, 115.8 (d, J = 21.52 Hz), 113.3, 111.3, 105.1, 55.3. 19F NMR (376 MHz, DMSO-d6) δ -112.6. HRMS (ESI) m/z: [M+H]+ Calcd for 311.1190 C18H16FN2O2+; Found 311.1191.

4. Conclusions

In conclusion, we have developed a novel, general approach to synthesizing δ-lactam-based hydrazide–hydrazone compounds from 3-arylglutaconic anhydrides and aldazines. Two simple, catalyst-free protocols have been developed to selectively prepare products with 3,6-dihydropyridin-2(1H)-one or 5,6-dihydropyridin-2(1H)-one cores, depending on the reaction temperature. Furthermore, all compounds were isolated as single diastereomers with respect to the hydrazone moiety and the relative configuration of the substituents at the 5,6-dihydropyridin-2(1H)-one core. We carried out a series of post-condensation modifications to demonstrate useful synthetic transformations of the side functional groups and the core. These transformations include the selective reduction of the ester or hydrazone moiety, hydrazone moiety exchange, the oxidation of the lactam core to 2-pyridone, and the deprotection of the hydrazone moiety to the NH2-hydrazide. Thus, the developed strategy enables the preparation of several medicinally relevant and poorly available N-functionalized δ-lactam scaffolds from the same set of substrates and a single annulation strategy. Our future studies will continue to evaluate the antibacterial and antitubercular activity profiles of the obtained compounds to identify their medical potential.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms26188834/s1.

Author Contributions

Conceptualization, O.B.; methodology, O.B.; investigation, A.A., E.K., O.B. and D.S.; data curation, A.A., O.B. and D.S.; writing—original draft preparation, O.B.; writing—review and editing, A.A. and G.K.; supervision, O.B.; funding acquisition, G.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 22-13-00005-P.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We are grateful to the Research Centre for Magnetic Resonance, the Centre for Chemical Analysis and Materials Research, the Research Centre for X-ray Diffraction Studies and the Cryogenic department of Saint Petersburg State University Research Park for the analytical data and supplies.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Teixeira, S.; Castanheira, E.M.S.; Carvalho, M.A. Hydrazides as Powerful Tools in Medicinal Chemistry: Synthesis, Reactivity, and Biological Applications. Molecules 2025, 30, 2852. [Google Scholar] [CrossRef]
  2. Wahbeh, J.; Milkowski, S. The Use of Hydrazones for Biomedical Applications. SLAS Technol. Transl. Life Sci. Innov. 2019, 24, 161–168. [Google Scholar] [CrossRef]
  3. Tafere, D.A.; Gebrezgiabher, M.; Elemo, F.; Sani, T.; Atisme, T.B.; Ashebr, T.G.; Ahmed, I.N. Hydrazones, hydrazones-based coinage metal complexes, and their biological applications. RSC Adv. 2025, 15, 6191–6207. [Google Scholar] [CrossRef]
  4. Du, Y.; Chen, W.; Fu, X.; Deng, H.; Deng, J. Synthesis and biological evaluation of heterocyclic hydrazone transition metal complexes as potential anticancer agents. RSC Adv. 2016, 6, 109718–109725. [Google Scholar] [CrossRef]
  5. Yancheva, D.; Argirova, M.; Georgieva, I.; Milanova, V.; Guncheva, M.; Rangelov, M.; Todorova, N.; Tzoneva, R. Antiproliferative and Pro-Apoptotic Activity and Tubulin Dynamics Modulation of 1H-Benzimidazol-2-yl Hydrazones in Human Breast Cancer Cell Line MDA-MB-231. Molecules 2024, 29, 2400. [Google Scholar] [CrossRef]
  6. Raucci, A.; Zwergel, C.; Valente, S.; Mai, A. Advancements in Hydrazide-Based HDAC Inhibitors: A Review of Recent Developments and Therapeutic Potential. J. Med. Chem. 2025, 68, 14171–14194. [Google Scholar] [CrossRef]
  7. Kajal, A.; Bala, S.; Sharma, N.; Kamboj, S.; Saini, V. Therapeutic potential of hydrazones as anti-inflammatory agents. Int. J. Med. Chem. 2014, 2014, 761030. [Google Scholar] [CrossRef] [PubMed]
  8. Schuster, D.; Zederbauer, M.; Langer, T.; Kubin, A.; Furtmuller, P.G. Pharmacophore-based discovery of 2-(phenylamino)aceto-hydrazides as potent eosinophil peroxidase (EPO) inhibitors. J. Enzym. Inhib. Med. Chem. 2018, 33, 1529–1536. [Google Scholar] [CrossRef] [PubMed]
  9. Sharma, P.C.; Sharma, D.; Sharma, A.; Saini, N.; Goyal, R.; Ola, M.; Chawla, R.; Thakur, V.K. Hydrazone comprising compounds as promising anti-infective agents: Chemistry and structure-property relationship. Mater. Today Chem. 2020, 18, 100349. [Google Scholar] [CrossRef]
  10. Yang, Z.; Li, P.; Gan, X. Novel Pyrazole-Hydrazone Derivatives Containing an Isoxazole Moiety: Design, Synthesis, and Antiviral Activity. Molecules 2018, 23, 1798. [Google Scholar] [CrossRef] [PubMed]
  11. Popiolek, L. Hydrazide-hydrazones as potential antimicrobial agents: Overview of the literature since 2010. Med. Chem. Res. 2017, 26, 287–301. [Google Scholar] [CrossRef] [PubMed]
  12. Ribeiro, N.; Correia, I. A review of hydrazide-hydrazone metal complexes’ antitumor potential. Front. Chem. Biol. 2024, 3, 1398873. [Google Scholar] [CrossRef]
  13. Angelova, V.; Karabeliov, V.; Andreeva-Gateva, P.A.; Tchekalarova, J. Recent Developments of Hydrazide/Hydrazone Derivatives and Their Analogs as Anticonvulsant Agents in Animal Models. Drug Dev. Res. 2016, 77, 379–392. [Google Scholar] [CrossRef]
  14. Murugappan, S.; Dastari, S.; Jungare, K.; Barve, N.M.; Shankaraiah, N. Hydrazide-hydrazone/hydrazone as enabling linkers in anti-cancer drug discovery: A comprehensive review. J. Mol. Struct. 2024, 1307, 138012. [Google Scholar] [CrossRef]
  15. Sonawane, S.J.; Kalhapure, R.S.; Govender, T. Hydrazone linkages in pH responsive drug delivery systems. Eur. J. Pharm. Sci. 2017, 99, 45–65. [Google Scholar] [CrossRef]
  16. Popiolek, L. The application of hydrazones and hydrazide-hydrazones in the synthesis of bioactive azetidin-2-one derivatives: A mini review. Biomed. Pharmacother. 2023, 163, 114853. [Google Scholar] [CrossRef]
  17. Welsch, M.E.; Snyder, S.A.; Stockwell, B.R. Privileged scaffolds for library design and drug discovery. Curr. Opin. Chem. Biol. 2010, 14, 347–361. [Google Scholar] [CrossRef]
  18. Cao, J.; Gillard, R.; Jahanbakhsh, A.; Breugst, M.; Lupton, D.W. Enantioselective N-Heterocyclic Carbene Catalysis via Acyl Azolium without Exogenous Oxidants. ACS Catal. 2020, 10, 11791–11796. [Google Scholar] [CrossRef]
  19. Sun, B.; Gao, L.; Shen, S.; Yu, C.; Li, T.; Xie, Y.; Yao, C. NHC-catalyzed [4 + 2] annulation of 2-bromo-2-enals with acylhydrazones: Enantioselective synthesis of delta-lactams. Org. Biomol. Chem. 2017, 15, 991–997. [Google Scholar] [CrossRef]
  20. Wu, X.; Zhang, Y.; Wang, Y.; Ke, J.; Jeret, M.; Reddi, R.N.; Yang, S.; Song, B.A.; Chi, Y.R. Polyhalides as Efficient and Mild Oxidants for Oxidative Carbene Organocatalysis by Radical Processes. Angew. Chem. 2017, 56, 2942–2946. [Google Scholar] [CrossRef] [PubMed]
  21. Jia, W.-Q.; Zhang, H.-M.; Zhang, C.-L.; Gao, Z.-H.; Ye, S. N-Heterocyclic carbene-catalyzed [4 + 2] annulation of α,β-unsaturated carboxylic acids: Enantioselective synthesis of dihydropyridinones and spirocyclic oxindolodihydropyridinones. Org. Chem. Front. 2016, 3, 77–81. [Google Scholar] [CrossRef]
  22. Xu, J.; Jin, Z.; Chi, Y.R. Organocatalytic enantioselective gamma-aminoalkylation of unsaturated ester: Access to pipecolic acid derivatives. Org. Lett. 2013, 15, 5028–5031. [Google Scholar] [CrossRef] [PubMed]
  23. Zhou, P.; Li, W.; Lan, J.; Zhu, T. Electroredox carbene organocatalysis with iodide as promoter. Nat. Commun. 2022, 13, 3827. [Google Scholar] [CrossRef]
  24. Li, M.; Kong, W.; Wen, L.-R.; Liu, F.-H. Facile isocyanide-based one-pot three-component regioselective synthesis of highly substituted pyridin-2(1H)-one derivatives at ambient temperature. Tetrahedron 2012, 68, 4838–4845. [Google Scholar] [CrossRef]
  25. Ramiro, J.L.; Martinez-Caballero, S.; Neo, A.G.; Diaz, J.; Marcos, C.F. The Castagnoli-Cushman Reaction. Molecules 2023, 28, 2654. [Google Scholar] [CrossRef]
  26. Mikheyev, A.; Kantin, G.; Krasavin, M. Aldazines in the Castagnoli–Cushman Reaction. Synthesis 2018, 50, 2076–2086. [Google Scholar] [CrossRef]
  27. Krasavin, M.; Dar’in, D. Current diversity of cyclic anhydrides for the Castagnoli–Cushman-type formal cycloaddition reactions: Prospects and challenges. Tetrahedron Lett. 2016, 57, 1635–1640. [Google Scholar] [CrossRef]
  28. Firsov, A.; Bakulina, O.; Dar’in, D.; Guranova, N.; Krasavin, M. Further Insight into the Castagnoli-Cushman-type Synthesis of 1,4,6-Trisubstituted 1,6-Dihydropyridin-2-(3H)-ones from 3-Arylglutaconic Acid Anhydrides. J. Org. Chem. 2020, 85, 6822–6829. [Google Scholar] [CrossRef]
  29. Bannykh, A.; Levashova, E.; Bakulina, O.; Krasavin, M. New reagent space and new scope for the Castagnoli-Cushman reaction of oximes and 3-arylglutaconic anhydrides. Org. Biomol. Chem. 2022, 20, 8643–8648. [Google Scholar] [CrossRef]
  30. Firsov, A.; Chupakhin, E.; Dar’in, D.; Bakulina, O.; Krasavin, M. Three-Component Castagnoli-Cushman Reaction of 3-Arylglutaconic Acids with Aromatic Aldehydes and Amines Delivers Rare 4,6-Diaryl-1,6-dihydropyridin-2(3H)-ones. Org. Lett. 2019, 21, 1637–1640. [Google Scholar] [CrossRef] [PubMed]
  31. Peshkov, A.A.; Bakulina, O.; Dar’in, D.; Kantin, G.; Bannykh, A.; Peshkov, V.A.; Krasavin, M. Three-Component Castagnoli-Cushman Reaction of 3-Arylglutaconic Acid Anhydrides, Carbonyl Compounds, and Ammonium Acetate: A Quick and Flexible Way to Assemble Polysubstituted NH-delta-lactams. Eur. J. Org. Chem. 2021, 2021, 1726–1731. [Google Scholar] [CrossRef]
  32. Sheldrick, G.M. SHELXT-integrated space-group and crystal-structure determination. Acta Crystallogr. A Found. Adv. 2015, 71 Pt 1, 3–8. [Google Scholar] [CrossRef]
  33. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C Struct. Chem. 2015, 71 Pt 1, 3–8. [Google Scholar] [CrossRef] [PubMed]
  34. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
Figure 1. Structures of commercial drugs containing hydrazide, hydrazone, or hydrazide–hydrazone moieties.
Figure 1. Structures of commercial drugs containing hydrazide, hydrazone, or hydrazide–hydrazone moieties.
Ijms 26 08834 g001
Figure 2. Published approaches to cyclic hydrazide–hydrazones with dihydropyridine-2-one core and the aim of the present study: (a) based on [4 + 2] annulation of acrolein derivatives with hydrazide–hydrazones derivatived from alkyl glyoxylate [18,19,20,21,22,23] (b) based on multicomponent reaction involving isocyanides, N-arylidene-2-cyanoacetohydrazides, and dimethylpropiolate [24] (c) based on reaction of cyclic anhydrides and aldazines.
Figure 2. Published approaches to cyclic hydrazide–hydrazones with dihydropyridine-2-one core and the aim of the present study: (a) based on [4 + 2] annulation of acrolein derivatives with hydrazide–hydrazones derivatived from alkyl glyoxylate [18,19,20,21,22,23] (b) based on multicomponent reaction involving isocyanides, N-arylidene-2-cyanoacetohydrazides, and dimethylpropiolate [24] (c) based on reaction of cyclic anhydrides and aldazines.
Ijms 26 08834 g002
Scheme 1. Preparation of hydrazide–hydrazones with 5,6-dihydropyridin-2(1H)-one core from 3-arylglutaconic anhydrides and aldazines. 1: C-nucleophilic enolized anhydride; 2: Imine; A: reaction product A; 3a–f: Six esters 3a–f with a 5,6-dihydropyridin-2(1H)-one core in moderate yields.
Scheme 1. Preparation of hydrazide–hydrazones with 5,6-dihydropyridin-2(1H)-one core from 3-arylglutaconic anhydrides and aldazines. 1: C-nucleophilic enolized anhydride; 2: Imine; A: reaction product A; 3a–f: Six esters 3a–f with a 5,6-dihydropyridin-2(1H)-one core in moderate yields.
Ijms 26 08834 sch001
Scheme 2. Preparation of hydrazide–hydrazones with 3,6-dihydropyridin-2(1H)-one core from 3-arylglutaconic anhydrides and aldazines. a Isolated from the reaction conducted at 150 °C in the presence of Sc(OTF)3 as a catalyst. b PMP = p-methoxyphenyl. c Conducted at 60 °C. d Reaction was conducted in the presence of Sc(OTf)3 (10%mol). A: reaction product A from Scheme 1; 4a–j: ten novel cyclic hydrazide–hydrazone compounds; 4aa: 5,6-dihydropyridin-2(1H)-one.
Scheme 2. Preparation of hydrazide–hydrazones with 3,6-dihydropyridin-2(1H)-one core from 3-arylglutaconic anhydrides and aldazines. a Isolated from the reaction conducted at 150 °C in the presence of Sc(OTF)3 as a catalyst. b PMP = p-methoxyphenyl. c Conducted at 60 °C. d Reaction was conducted in the presence of Sc(OTf)3 (10%mol). A: reaction product A from Scheme 1; 4a–j: ten novel cyclic hydrazide–hydrazone compounds; 4aa: 5,6-dihydropyridin-2(1H)-one.
Ijms 26 08834 sch002
Figure 3. Structures of N-protected hydrazones, which were found to be unsuccessful reaction partners for 3-arylglutaconic anhydrides.
Figure 3. Structures of N-protected hydrazones, which were found to be unsuccessful reaction partners for 3-arylglutaconic anhydrides.
Ijms 26 08834 g003
Scheme 3. Post-condensational modifications of compounds 3 and 4: reduction of side groups, oxidation of the core, exchange and deprotection of the hydrazide moiety. a A thermally induced isomerization to 5,6-dihydropyridin-2(1H)-one occurred in addition to the target reaction. 3a: Methyl (2RS,3RS)-4-(4-Fluorophenyl)-1-(((E)-4-methoxybenzylidene)amino)-2-(4-methoxyphenyl)-6-oxo-1,2,3,6-tetrahydropyridine-3-carboxylate; 4a: (E)-4-(4-Fluorophenyl)-1-((4-methoxybenzylidene)amino)-6-(4-methoxyphenyl)-3,6-dihydropyridin-2(1H)-one; 4c: (E)-4-(4-Fluorophenyl)-6-isopropyl-1-((2-methylpropylidene)amino)-3,6-dihydropyridin-2(1H)-one. 5: Alcohol; 6: N-alkyl hydrazide; 7: Unsymmetrical 1-N-arylidene/6-alkyl product; 8: N-arylidene pyridine-2-one; 9: NH2-hydrazide.
Scheme 3. Post-condensational modifications of compounds 3 and 4: reduction of side groups, oxidation of the core, exchange and deprotection of the hydrazide moiety. a A thermally induced isomerization to 5,6-dihydropyridin-2(1H)-one occurred in addition to the target reaction. 3a: Methyl (2RS,3RS)-4-(4-Fluorophenyl)-1-(((E)-4-methoxybenzylidene)amino)-2-(4-methoxyphenyl)-6-oxo-1,2,3,6-tetrahydropyridine-3-carboxylate; 4a: (E)-4-(4-Fluorophenyl)-1-((4-methoxybenzylidene)amino)-6-(4-methoxyphenyl)-3,6-dihydropyridin-2(1H)-one; 4c: (E)-4-(4-Fluorophenyl)-6-isopropyl-1-((2-methylpropylidene)amino)-3,6-dihydropyridin-2(1H)-one. 5: Alcohol; 6: N-alkyl hydrazide; 7: Unsymmetrical 1-N-arylidene/6-alkyl product; 8: N-arylidene pyridine-2-one; 9: NH2-hydrazide.
Ijms 26 08834 sch003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ananeva, A.; Karchuganova, E.; Spiridonova, D.; Kantin, G.; Bakulina, O. Catalyst-Free Assembly of δ-Lactam-Based Hydrazide–Hydrazone Compounds from 3-Arylglutaconic Anhydrides and Aldazines. Int. J. Mol. Sci. 2025, 26, 8834. https://doi.org/10.3390/ijms26188834

AMA Style

Ananeva A, Karchuganova E, Spiridonova D, Kantin G, Bakulina O. Catalyst-Free Assembly of δ-Lactam-Based Hydrazide–Hydrazone Compounds from 3-Arylglutaconic Anhydrides and Aldazines. International Journal of Molecular Sciences. 2025; 26(18):8834. https://doi.org/10.3390/ijms26188834

Chicago/Turabian Style

Ananeva, Anna, Elizaveta Karchuganova, Dar’ya Spiridonova, Grigory Kantin, and Olga Bakulina. 2025. "Catalyst-Free Assembly of δ-Lactam-Based Hydrazide–Hydrazone Compounds from 3-Arylglutaconic Anhydrides and Aldazines" International Journal of Molecular Sciences 26, no. 18: 8834. https://doi.org/10.3390/ijms26188834

APA Style

Ananeva, A., Karchuganova, E., Spiridonova, D., Kantin, G., & Bakulina, O. (2025). Catalyst-Free Assembly of δ-Lactam-Based Hydrazide–Hydrazone Compounds from 3-Arylglutaconic Anhydrides and Aldazines. International Journal of Molecular Sciences, 26(18), 8834. https://doi.org/10.3390/ijms26188834

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop