Diana Hurtado-Rodrígueza,
Diana Becerraa,
Hugo Rojasa,
Jovanny A. Gómez Castaño*b,
Mario A. Macías*c and
Juan-Carlos Castillo*a
aGrupo de Catálisis de la UPTC, Escuela de Ciencias Químicas, Universidad Pedagógica y Tecnológica de Colombia, Avenida Central del Norte 39-115, Tunja 150003, Colombia. E-mail: juan.castillo06@uptc.edu.co
bGrupo Química-Física Molecular y Modelamiento Computacional (QUIMOL), Escuela de Ciencias Químicas, Universidad Pedagógica y Tecnológica de Colombia, Avenida Central del Norte 39-115, Tunja 150003, Colombia. E-mail: jovanny.gomez@uptc.edu.co
cCristalografía y Química de Materiales CrisQuimMat, Departamento de Química, Universidad de los Andes, Carrera 1 No. 18A-10, Bogotá 111711, Colombia. E-mail: ma.maciasl@uniandes.edu.co
First published on 12th August 2024
A series of 3-cyano-2(1H)-pyridones 4a–c were efficiently synthesized using an expeditious microwave-assisted multicomponent approach. Single-crystal XRD analysis revealed the presence of six independent molecules in the asymmetric unit cell for all compounds, with the crystal packing stabilized by a network of cyclic dimers formed by N–H⋯OC and C–H⋯OC intermolecular interactions. Additional supramolecular interactions, including C–H⋯π, C–N⋯π, and π⋯π, and C–H⋯X (for halogenated derivatives, i.e., 4b and 4c), appear crucial for crystal stabilization. Density Functional Theory (DFT) calculations were employed to understand the electronic structures and potential binding affinities. Comprehensive spectroscopic characterization by FT-IR, UV-Vis, NMR, and HMRS techniques confirmed the structures of all synthesized compounds. Differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA) were employed to evaluate the thermal stability of these compounds. The in vitro anticancer activity was evaluated against a panel of 60 human cancer cell lines, demonstrating promising activity against non-small-cell lung and breast cancer cell lines. Notably, compounds 4a and 4c exhibited the highest anticancer activity against the HOP-92 and MCF7 cell lines, with growth inhibition percentages (GI%) of 54.35 and 40.25, respectively.
The properties of 2(1H)-pyridones have fueled their exploration in various pharmacological applications (Fig. 1). Naturally occurring examples, such as huperzine A employed for Alzheimer's treatment,6 and fredericamycin A a promising anticancer lead,7 showcase their inherent biological potential. The recent emergence of 2(1H)-pyridone derivatives in FDA-approved drugs further underscores their translational value.8–10 Tazemetostat, a 2020-approved oral inhibitor of enhancer of zeste homolog 2 (EZH2), exemplifies this class.11 Interestingly, its active pharmacophore is the 3-(aminomethyl)-pyridin-2(1H)-one moiety, strategically obtained via chemoselective reduction of a 3-cyano-2(1H)-pyridone precursor. It highlights the continued need for efficient synthetic strategies to access diverse 3-cyano-2(1H)-pyridone derivatives, paving the way for further exploration of their therapeutic potential in various diseases.
Traditional synthetic approaches for the preparation of 4,6-diaryl-3-cyano-2(1H)-pyridones often rely on a two-step sequence: (i) Michael-type addition involving enolate and α,β-unsaturated carbonyl acceptors, and (ii) subsequent oxidative intramolecular cyclization.12–15 A diverse range of α,β-unsaturated carbonyl substrates have been effectively employed in this strategy, encompassing ethyl 2-cyano-3-arylacrylates,12,13 2-cyano-3-arylacrylamides,14 and 1,3-diaryl-2-propen-1-ones.15
Multicomponent reactions (MCRs) have emerged as powerful tools for efficiently synthesizing diverse heterocyclic scaffolds, including aza-heterocycles.16,17 However, the application of MCRs to prepare 4,6-diaryl-3-cyano-2(1H)-pyridones 4 remains relatively unexplored.18–21 Pioneering work by Rong et al. described a three-component reaction for the synthesis of compounds 4 with good yields using malononitrile as the active methylene component and sodium hydroxide as the base under both grinding and conventional heating conditions (routes i and ii, Scheme 2a).18,19 This approach was further extended by the same group, replacing malononitrile with cyanoacetamide as the active methylene compound for synthesizing analogous products with good yields (route iii, Scheme 2a).20 El-Sayed et al. subsequently reported a solvent-free four-component reaction utilizing ethyl cyanoacetate as the active methylene compound and ammonium acetate (only 1 equivalent) acting dually as an ammonia surrogate and base, resulting in 4,6-diaryl-3-cyano-2(1H)-pyridones 4 with good yields (route iv, Scheme 2a).21
Despite reported synthetic procedures and spectroscopic characterization of 4,6-diaryl-3-cyano-2(1H)-pyridones 4a–c (Scheme 2),18–21 critical knowledge gap persists regarding their comprehensive physicochemical and biological properties. Notably, prior studies have not elucidated their crystal structure, performed detailed computational analysis using DFT, investigated their thermal behavior, or explored their potential as anticancer agents. This study meticulously addresses these shortcomings, offering a multifaceted exploration of compounds 4a–c. Through a combination of crystallographic analysis, DFT calculations, thermal studies, and in vitro anticancer activity assays against a panel of 60 human cancer cell lines, we unveil valuable insights into their potential for future drug development endeavors.
The UV-Vis spectra of compounds 4a–c displayed three distinct absorption bands within the range of 217–275 nm (Table 2). These bands were attributed to various electronic transitions involving both n → π* and π → π* states. In addition, the spectra revealed two broad bands between 275 and 410 nm, which were assigned to the electronic transitions of the HOMO−1 (πPh(C)) and HOMO (nO) to LUMO orbitals (Table 2). These specific assignments were achieved by a combined approach using vertical electronic and natural bond orbital (NBO) calculations at the B3LYP/6-311++G(2df,2dp) level of theory, in conjunction with previously reported spectra for structurally similar 4,6-diaryl-3-cyano-2(1H)-pyridones.12
Compound | Solvent | Experimental | Calculated | Δλ (nm) | Transition | Assignmenta | ||||
---|---|---|---|---|---|---|---|---|---|---|
λ (nm) | ε × 104 (L mol−1 cm−1) | λ (nm) | f | Normalized coefficient | Transition percentage | |||||
a For ring labeling (i.e., A, B or C), see Fig. 5. | ||||||||||
4a | EtOH | 204 | 2.80 | 241.6 | 0.148 | 0.67332 | 90.7 | −37.6 | HOMO−1–LUMO+1 | |
EtOH | 221 | 1.55 | 247.7 | 0.027 | 0.65845 | 86.7 | −26.7 | HOMO–LUMO+2 | ||
EtOH | 257 | 1.60 | 263.0 | 0.212 | 0.68038 | 92.6 | −6.0 | HOMO–LUMO+1 | ||
EtOH | 301 | 0.85 | 328.9 | 0.238 | 0.69747 | 97.3 | −27.9 | HOMO−1–LUMO | ||
EtOH | 365 | 1.20 | 357.4 | 0.493 | 0.69727 | 97.2 | 7.6 | HOMO–LUMO | ||
ACN | 369 | 1.40 | 357.1 | 0.493 | 0.69717 | 97.2 | 11.9 | HOMO–LUMO | ||
DMSO | 376 | 1.25 | 358.1 | 0.509 | 0.69725 | 97.2 | 17.9 | HOMO–LUMO | ||
4b | EtOH | 203 | 3.40 | 242.6 | 0.001 | 0.69800 | 97.4 | −39.6 | HOMO−6–LUMO | |
EtOH | 217 | 2.05 | 246.7 | 0.012 | 0.65984 | 87.1 | −29.7 | HOMO–LUMO+3 | ||
EtOH | 260 | 1.95 | 264.6 | 0.194 | 0.68574 | 94.0 | −4.6 | HOMO–LUMO+1 | ||
EtOH | 290 | 1.00 | 325.8 | 0.252 | 0.69615 | 96.9 | −35.8 | HOMO−1–LUMO | ||
EtOH | 368 | 1.35 | 360.9 | 0.515 | 0.69817 | 97.5 | 7.1 | HOMO–LUMO | ||
ACN | 372 | 1.40 | 360.6 | 0.514 | 0.69814 | 97.5 | 11.4 | HOMO–LUMO | ||
DMSO | 378 | 1.25 | 361.6 | 0.530 | 0.69830 | 97.5 | 16.4 | HOMO–LUMO | ||
4c | EtOH | 204 | 1.80 | 239.5 | 0.031 | 0.51929 | 53.9 | −35.5 | HOMO−1–LUMO+2 | |
EtOH | 222 | 1.48 | 246.9 | 0.011 | 0.65532 | 85.9 | −24.9 | HOMO–LUMO+3 | ||
EtOH | 259 | 1.46 | 267.8 | 0.221 | 0.68072 | 92.7 | −8.8 | HOMO–LUMO+1 | ||
EtOH | 295 | 0.60 | 320.3 | 0.246 | 0.69833 | 97.5 | −25.3 | HOMO−1–LUMO | ||
EtOH | 368 | 1.21 | 360.4 | 0.489 | 0.69941 | 97.8 | 7.6 | HOMO–LUMO | ||
ACN | 372 | 1.33 | 360.2 | 0.489 | 0.69938 | 97.8 | 11.8 | HOMO–LUMO | ||
DMSO | 380 | 1.06 | 361.1 | 0.504 | 0.69954 | 97.9 | 18.9 | HOMO–LUMO |
The electronic absorption assignments observed, alongside FT-IR analysis, support the predominance of the lactam tautomeric form in both solution and solid states. This conclusion is further strengthened by the presence of CO stretching vibrations in the FT-IR spectra (1636–1639 cm−1) and the corresponding CO carbon signals in the 13C-NMR spectra (161.7–162.5 ppm) of compounds 4a–c. The UV-Vis spectroscopy results, coupled with established insights into the lactam–lactim equilibrium, underscore the influence of solvent polarity in determining the predominant tautomeric form. The 2-hydroxypyridine tautomer 5 exhibits lower polarity than the lactam tautomer 4, which can be attributed to the presence of a charge-separated mesomeric form 4′ for the lactam tautomer (Scheme 3).12
Scheme 3 Influence of mesomeric form 4′ on the lactam–lactim tautomeric equilibrium in 4,6-diaryl-3-cyano-2(1H)-pyridones. |
Consequently, electron-withdrawing substituents at positions 4 and 6 of the pyridine ring and polar protic solvents significantly stabilize the charge-separated resonance structure 4′, favoring the more polar lactam form 4. The lactam tautomer 4 achieves stability when exposed to polar protic solvents due to its ability to act as a hydrogen-bond donor to carbonyl oxygen. Conversely, the 2-hydroxypyridine tautomer 5 may gain some stabilization in solvents that primarily function as hydrogen-bond acceptors.12 To gain further insight into the n → π* electronic transition of the CO group associated with the lactam–lactim equilibrium in compounds 4a–c, a solvatochromic study was undertaken using a range of polar solvents with varying dielectric constants (ε): ethanol (ε = 24.5), acetonitrile (ε = 37.5), and dimethylsulfoxide (DMSO, ε = 46.7). As illustrated in Fig. 3 and Table 2, all compounds displayed a moderate bathochromic shift (red shift) in the n → π* electronic transition of the CO group with increasing solvent polarity. As mentioned earlier, this phenomenon can be attributed to the enhanced stabilization of the charge-separated mesomeric form 4′ with increasing solvent polarity (Scheme 3).
Fig. 3 UV-Vis spectra and solvatochromic study of compounds 4a (R1 = Me), 4b (R1 = Br), and 4c (R1 = Cl). |
Furthermore, we investigated the influence of substituents at the 4-position of the aromatic ring on the electronic transitions. The presence of a more electronegative substituent (R1 = Br and Cl) resulted in a bathochromic shift in the n → π* electronic transition from the HOMO to LUMO orbitals. This observation reveals a correlation between the bathochromic shift and a reduction in the band gap for compounds containing electron-withdrawing substituents (R1 = Br and Cl), which contrasts with the behavior observed for the compound containing an electron-donating group (R1 = Me). While Fig. 3 reveals that compounds 4a–c exhibited their highest absorbance in acetonitrile, interestingly, the n → π* electronic transition displayed a hypochromic effect in DMSO. This observation indicates that the electron-withdrawing substituent (Cl in compound 4c) reduces the absorbance (1.06 × 104 L mol−1 cm−1), which contrasts with the slightly higher absorbance observed for the compound containing an electron-donating substituent (CH3 in compound 4a, 1.25 × 104 L mol−1 cm−1).
Fig. 4 Calculated frontier molecular orbitals (HOMO and LUMO) and band gap of compounds 4a–c using the B3LYP/6-311++G(2df,2pd) approximation level. |
Table 3 presents the calculated global reactivity descriptors for compounds 4a–c, including ionization potential (IP), electron affinity (EA), electrophilicity index (ω), chemical potential (μ), electronegativity (χ), and hardness (η). Koopmans' theorem establishes a correlation between the HOMO and LUMO energy levels of a molecule and its IP and EA, respectively.22 Electronegativity (χ) is calculated as the average of the HOMO and LUMO energies using the equation χ = (IP + EA)/2.23 Hardness (η), a valuable indicator of chemical stability, is inversely proportional to the HOMO–LUMO energy gap.24,25 Parr et al. introduced the electrophilicity index (ω) defined as ω = μ2/2η, where μ is the chemical potential calculated as μ = −(IP + EA)/2.26
Parameters | 4a | 4b | 4c |
---|---|---|---|
HOMO energy | −6.440 | −6.612 | −6.604 |
LUMO energy | −2.596 | −2.786 | −2.776 |
HOMO–LUMO energy gap | 3.844 | 3.826 | 3.828 |
Ionization potential (IP) | 6.440 | 6.612 | 6.604 |
Electron affinity (EA) | 2.596 | 2.786 | 2.776 |
Electrophilicity index (ω) | 2.655 | 2.886 | 2.873 |
Chemical potential (μ) | −4.518 | −4.699 | −4.690 |
Electronegativity (χ) | 4.518 | 4.699 | 4.690 |
Hardness (η) | 3.844 | 3.826 | 3.828 |
The calculated electrophilicity index values for 4a (2.66 eV), 4b (2.89 eV), and 4c (2.87 eV) denote moderate electrophilicity toward electron-rich species (Table 3). The chemical potential values (μ) for 4a (−4.52 eV), 4b (−4.70 eV), and 4c (−4.69 eV) imply a moderate ability to attract electrons. The electronegativity values (χ) follow the same trend, ranging from 4.52 to 4.70 eV. The high hardness values (η) calculated from the HOMO–LUMO energy gap ranging from 3.83 to 3.84 eV denote significant resistance to electron transfer for all compounds, resulting in reduced chemical reactivity.
Fig. 5 Molecular structures of compounds (a) 4a, (b) 4b, and (c) 4c show anisotropic thermal vibrations. The ellipsoids represent a 30% probability level, and hydrogen atoms are depicted as spheres with arbitrary radii. Blue arrows denote the free rotation between C7A–C1A and C3A–C13A atoms, resulting in six independent molecules in the asymmetric unit for 4a–c. The dihedral angles between the planes containing rings A–B and A–C for compounds 4a–c were measured using Mercury software (Fig. 4 and Table 4).27 |
Dihedral angles (°) | 4a | 4b | 4c |
---|---|---|---|
A–B | 26.8(19) | 23.0(3) | 27.6(13) |
32.2(18) | 24.1(3) | 23.1(14) | |
30.1(18) | 27.7(3) | 28.2(14) | |
32.0(19) | 28.5(2) | 23.9(13) | |
27.2(19) | 27.4(3) | 27.9(13) | |
29.7(17) | 28.4(3) | 26.8(13) | |
A–C | 53.0(2) | 49.7(3) | 59.0(14) |
57.2(2) | 48.3(3) | 52.1(15) | |
48.3(2) | 46.2(3) | 58.9(14) | |
57.8(2) | 47.1(3) | 49.3(15) | |
53.9(2) | 58.5(3) | 46.1(13) | |
49.3(2) | 57.3(3) | 45.6(15) |
Further analysis of the crystal structures of compounds 4a–c revealed the presence of various intermolecular interactions that govern their solid-state packing arrangements (Fig. 6 and Table 5). Independent of the conformational orientation of individual molecules, inversion symmetry-related molecules are linked through a combination of N–H⋯O and C–H⋯O hydrogen bonds. Notably, some of these hydrogen bonds exhibit a split character, where a single oxygen atom acts as a hydrogen bond acceptor for two (N,C)–H donors. As expected, N–H⋯O hydrogen bonds display shorter distances compared to C–H⋯O interactions due to the enhanced acidity of the N–H proton.
D–H⋯A | D–H | H⋯A | D⋯A | D–H⋯A | Symmetry code |
---|---|---|---|---|---|
Compound 4a | |||||
N1A–H1A⋯O1D | 0.86 | 1.99 | 2.817(4) | 161 | 1 − x,1 − y,−z |
N1B–H1B⋯O1F | 0.86 | 1.99 | 2.800(4) | 165 | x,y,z |
N1C–H1C⋯O1C | 0.86 | 1.99 | 2.788(4) | 161 | 1 − x,1 − y,−z |
N1D–H1D⋯O1A | 0.86 | 1.99 | 2.827(4) | 164 | 1 − x,1 − y,−z |
N1E–H1E⋯O1E | 0.86 | 2.00 | 2.819(4) | 158 | −x,1 − y,1 − z |
N1F–H1F⋯O1B | 0.86 | 1.96 | 2.803(4) | 165 | x,y,z |
C8C–H8C⋯O1C | 0.93 | 2.57 | 3.189(4) | 125 | 1 − x,1 − y,−z |
C8D–H8D⋯O1A | 0.93 | 2.57 | 3.202(4) | 126 | 1 − x,1 − y,−z |
C8F–H8F⋯O1B | 0.93 | 2.55 | 3.205(4) | 128 | x,y,z |
Compound 4b | |||||
N1A–H1A⋯O1A | 0.86 | 1.98 | 2.797(6) | 158 | 1 − x,−y,1 − z |
N1B–H1B⋯O1C | 0.86 | 1.97 | 2.812(6) | 166 | 1 − x,1 − y,1 − z |
N1C–H1C⋯O1B | 0.86 | 1.98 | 2.815(5) | 164 | 1 − x,1 − y,1 − z |
N1D–H1D⋯O1E | 0.86 | 1.97 | 2.806(6) | 165 | −x,1 − y,1 − z |
N1E–H1E⋯O1D | 0.86 | 2.01 | 2.834(6) | 161 | −x,1 − y,1 − z |
N1F–H1F⋯O1F | 0.86 | 1.95 | 2.781(5) | 163 | −x,1 − y,1 − z |
C8A–H8A⋯O1A | 0.93 | 2.53 | 3.215(7) | 130 | 1 − x,−y,1 − z |
C8B–H8B⋯O1C | 0.93 | 2.52 | 3.214(6) | 132 | 1 − x,1 − y,1 − z |
C8C–H8C⋯O1B | 0.93 | 2.45 | 3.148(7) | 132 | 1 − x,1 − y,1 − z |
C8D–H8D⋯O1E | 0.93 | 2.49 | 3.167(7) | 130 | −x,1 − y,1 − z |
C8F–H8F⋯O1F | 0.93 | 2.49 | 3.159(7) | 129 | −x,1 − y,1 − z |
Compound 4c | |||||
N1A–H1A⋯O1C | 0.86 | 1.97 | 2.816(3) | 166 | x,1 + y,z |
N1B–H1B⋯O1B | 0.86 | 1.99 | 2.809(3) | 158 | 1 − x,2 − y,1 − z |
N1C–H1C⋯O1A | 0.86 | 1.97 | 2.815(3) | 165 | x,−1 + y,z |
N1D–H1D⋯O1F | 0.86 | 1.99 | 2.827(3) | 165 | 1 − x,−y,−z |
N1E–H1E⋯O1E | 0.86 | 1.94 | 2.777(3) | 163 | −x,−y,−z |
N1F–H1F⋯O1D | 0.86 | 2.00 | 2.830(3) | 162 | 1 − x,−y,−z |
C8E–H8E⋯O1E | 0.93 | 2.48 | 3.159(3) | 130 | −x,−y,−z |
The observed free rotation around the C7A–C1A and C3A–C13A bonds, leading to six independent molecules within the asymmetric units of 4a–c, is influenced by packing constraints. Interestingly, the formation of precisely six distinct conformers in each crystal structure is associated with the development of directional molecular chains extending along the [100], [010], and [100] crystallographic directions for 4a, 4b, and 4c, respectively. Within these chains, the 4-substituted aromatic rings (4-MeC6H4, 4-BrC6H4, and 4-ClC6H4) adopt conformations that favor the formation of C–H⋯π interactions, with H⋯π distances of approximately 2.9 Å, contributing to the overall structural stability (Fig. 6). Furthermore, unconventional C–N⋯π contacts (N⋯π distances ranging from 3.4 to 3.6 Å) participate in the assembling of supramolecular chains by facilitating the appropriate molecular orientations necessary for the formation of weak π⋯π interactions.
From a supramolecular perspective, the crystal structures of compounds 4a–c can be envisioned as assemblies of structurally similar molecular chains, resembling supramolecular bricks (Fig. 6). In compound 4a, these bricks are connected primarily through weak C–H⋯π interactions and van der Waals forces. In contrast, the packing arrangement in compounds 4b and 4c involves additional intermolecular C–H⋯Cl and C–H⋯Br halogen bonding interactions, with H⋯Cl and H⋯Br distances ranging from 2.8 to 3.1 Å. Furthermore, C–H⋯N–C interactions involving the cyano group (CN), with H⋯N distances of approximately 2.9 Å, contribute to the interconnection of the chains (Fig. 6). It is noteworthy that despite the variation in the 4-substituent on the aromatic ring (4-MeC6H4, 4-BrC6H4, and 4-ClC6H4), the intermolecular interactions governing the assembly between the molecular chains exhibit a high degree of similarity across compounds 4a–c.
As previously mentioned, the carbonyl oxygen atom in compounds 4a–c appears to be crucial for the stability of these crystalline systems. It forms two simultaneous hydrogen bonds: one with the amine's hydrogen and another with aromatic hydrogen from a symmetrically related molecular unit (Fig. 6 and 7-top). To test this hypothesis, we performed QTAIM-C analysis on the unmodified structure of the corresponding dimer extracted from these crystal systems. As shown in Fig. 7-bottom, the topological analysis revealed the presence of two bond critical points (BCPs) and two corresponding bond paths (BPs) connecting each carbonyl oxygen atom with the hydrogen atoms involved in these hydrogen bonding interactions. The topological characterization of these BCPs, such as ρ(r) and ∇2ρ(r), confirms the presence of closed-shell hydrogen bonding interactions, which is consistent with the structural analysis observations. The electron density (ρ) at the BCPs indicates the strength of these interactions, with a higher electron density corresponding to stronger hydrogen bonding. Electron density values at BCPs of the type CO⋯H–Ph HB ranged from 0.007 to 0.009 au in the three systems, indicating weak-to-moderate strength hydrogen bonding interactions (0.002 ≤ ρ ≤ 0.4 au, for neutral systems).28 As shown in Fig. 7, dimer 4a exhibited slightly lower electronic density values for these types of BCPs compared to 4b and 4c. On the other hand, the calculated electron density at the CO⋯H–N BCPs attained values of 0.025 au for system 4a and 0.024 au for systems 4b and 4c, revealing stronger bidirectional dimer interactions between the carbonyl oxygen atom and the amine hydrogen in compounds 4a–c. These findings suggest that a carbonyl oxygen atom in compounds 4a–c is critical for stable crystalline assemblies.
Our topological analysis unexpectedly revealed a long intermolecular CN⋯H–C bond path (approximately 7.97 au) in halogenated dimers 4b and 4c, connecting the cyano group with an aromatic hydrogen atom from a neighboring molecule (Fig. 7-bottom). The electron densities at these BCPs reached low values of approximately 3.7 × 10−4 au, suggesting weak but potentially significant dispersive interactions contributing to the overall supramolecular organization observed in these molecular crystals. In addition, relatively high ellipticities (ε ≈ 2.2) were observed on these BCPs, indicating that these interactions exhibit partial π-character. The absence of intermolecular CN⋯H–C interactions in compound 4a suggests that the introduction of halogen atoms in compounds 4b and 4c promotes the formation of weak non-covalent interactions, which may play an auxiliary role in stabilizing the molecular packing arrangements.
The two-dimensional (2D) fingerprint plots in Fig. 8 depict the contributions of various interatomic contacts to the overall HS for each compound. A qualitative comparison of these plots reveals similar patterns for compounds 4b and 4c, which differ from those observed for 4a. These differences can be attributed to the varying 4-substituents on the aromatic ring C (4-MeC6H4, 4-BrC6H4, and 4-ClC6H4), which influence the formation of intermolecular contacts. In all compounds, H⋯H interactions represent the most significant contribution of the HS maps, with the highest percentage observed for compound 4a, as expected. The presence of 4-BrC6H4 and 4-ClC6H4 fragments in compounds 4b and 4c leads to a notable contribution from C–H⋯(Br, Cl) interactions (14.1% and 14.0% for 4b and 4c, respectively), highlighting the importance of these interactions in stabilizing the crystal packing.
Beyond the conventional N–H⋯O, C–H⋯O hydrogen bonds, C–H⋯π and π⋯π interactions, HS analysis revealed the presence of additional, potentially weak C–N⋯π contacts relevant to the supramolecular assembly in all compounds. These C–N⋯π contacts contribute approximately 4.3–5.5% to the total HS map for each compound. Fig. 9 presents the shape index maps, emphasizing the distribution of these C–N⋯π contacts. The presence of red hollows on the shape index maps indicates areas where the surfaces around C–N fragments come into proximity with the surfaces of the aromatic rings in each compound. These hollows, centered on the centroids of the π-systems, further support the existence of these weak C–N⋯π contacts (N⋯π distances: 3.4–3.6 Å).
Fig. 11 Band gap values assessed by a correlated curve of (F(R)hν)2 set against photon energy plots of compounds 4a–c. |
Analysis of the combined TGA and DSC thermograms for compound 4a (R1 = Me) revealed a melting process within the temperature range of 292–331 °C, characterized by a distinct endothermic peak at 315 °C (ΔHfus = 117 J g−1). Similarly, the TGA and DSC curves for compounds 4b (R1 = Br) and 4c (R1 = Cl) displayed analogous melting transitions within 321–348 °C and 303–337 °C, respectively. These melting processes were accompanied by prominent endothermic peaks observed at 347 °C (ΔHfus = 125 J g−1) for compound 4b and 320 °C (ΔHfus = 132 J g−1) for compound 4c.
The introduction of halogen substituents (Br and Cl) into compounds 4b and 4c appears to result in a slight increase in the enthalpy of fusion and the melting temperature range compared to compound 4a. This observed trend can be attributed to stronger intermolecular interactions facilitated by halogen incorporation. As evidenced by single-crystal X-ray diffraction analysis (Section 2.5.1), these stronger interactions likely involve C–H⋯Cl and C–H⋯Br hydrogen bonding.
Compound | Most sensitive cell lines | GIa (%) |
---|---|---|
a GI% = 100 − G%. | ||
4a | HOP-92 (non-small cell lung cancer) | 54.35 |
HCT-15 (colon cancer) | 28.96 | |
UO-31 (renal cancer) | 26.78 | |
SR (leukemia) | 26.72 | |
T-47D (breast cancer) | 26.54 | |
4b | MCF7 (breast cancer) | 24.72 |
HOP-92 (non-small cell lung cancer) | 23.42 | |
RPMI-8226 (leukemia) | 23.33 | |
MOLT-4 (leukemia) | 22.34 | |
NCI-H522 (non-small cell lung cancer) | 21.22 | |
4c | MCF7 (breast cancer) | 40.25 |
HOP-92 (non-small cell lung cancer) | 31.79 | |
T-47D (breast cancer) | 29.86 | |
UO-31 (renal cancer) | 28.23 | |
UACC-62 (melanoma) | 28.18 |
The 3-cyano-2(1H)-pyridones 4a and 4c exhibited the most promising in vitro anticancer activity. These compounds displayed significant growth inhibition percentages (GI%) against the HOP-92 non-small-cell lung (54.35 for 4a) and MCF7 (40.25 for 4c) breast cancer cell lines. In addition, compounds 4b and 4c demonstrated moderate activity against the HOP-92 cell line with GI% values of 23.42 and 31.79, respectively. Conversely, compounds 4a and 4b showed low activity against the T-47D and MCF7 breast cancer cell lines, with GI% values of 26.54 and 24.72, respectively. Likewise, compounds 4a and 4c demonstrated low activity against the UO-31 renal cancer cell line, with GI% values of 26.78 and 28.23, respectively.
Interestingly, some studies have documented the anticancer activity of 2(1H)-pyridone derivatives against the MCF7 breast cancer cell line,33–36 highlighting the 4-aryl-6-[benzo[f]coumarin-3-yl]-3-cyano-2(1H)-pyridones with GI50 values ranging from 0.003 to >50 μM, as well as 3,4,6-triaryl-2(1H)-pyridones exhibiting higher anticancer activity than the standard drug Nolvadex.34,35 Conversely, limited data exists regarding their efficacy against HOP-92 and UO-31 cancer cell lines. Overall, the promising in vitro anticancer activity of compounds 4a–c against the HOP-92 non-small-cell lung and breast cancer cell lines can be attributed to the presence of the 2(1H)-pyridone scaffold, aligning with findings on the antiproliferative properties of 2-pyridone derivatives.33–36 However, further investigation is warranted to elucidate the underlying mechanisms of action of these compounds and to explore their potential for development as potential anticancer agents.
Density functional theory (DFT) calculations at the B3LYP/6-311++G(2df,2pd) level of theory were employed to optimize the molecular geometries of 4a–c in the gas phase and various in silico solvent environments (ethanol, acetonitrile, and DMSO). Time-dependent DFT (TD-DFT) calculations provided insights into the electronic transitions responsible for the UV-Vis spectral bands. Natural bond orbital (NBO) population analysis yielded the frontier molecular orbital surfaces.
The topological analysis of the electron density, conducted using the QTAIM-C approach, highlighted the crucial role of the carbonyl oxygen atom in forming strong bidirectional hydrogen bonding interactions, primarily with the amine hydrogen and secondarily with aromatic hydrogens from neighboring molecules. Furthermore, the analysis revealed the presence of weak intermolecular CN⋯H–C contacts in the halogenated compounds 4b and 4c, which may contribute to the overall stability of the supramolecular assemblies. Hirshfeld surface analysis offered a quantitative and visual representation of these close contacts in real space, providing valuable insights into the nature of the noncovalent interactions that govern the crystalline packing of the 3-cyano-2(1H)-pyridones. These findings can be further utilized to design and synthesize new functional materials with specific properties.
Initial assessment of anticancer activity revealed that all compounds exhibited moderate cytotoxicity against human lung carcinoma (A549) and breast adenocarcinoma (MCF7) cell lines. Compounds 4a and 4c exhibited the most significant anticancer activity against HOP-92 non-small-cell lung and MCF7 breast cancer cell lines, with growth inhibition percentages (GI%) of 54.35 and 40.25, respectively. Further studies are needed to elucidate their mechanisms of action and enhance the anticancer potential of other 3-cyano-2(1H)-pyridones.
Topological analysis of the electron density (ρ) within the crystalline packing of compounds 4a–c was performed using the quantum theory of atoms in molecules and crystals (QTAIM-C) approach developed by Bader.44 To represent the crystalline environment, a dimeric model was considered for each compound by extracting two symmetrically related molecules from the corresponding unit cell. The wavefunctions of these dimers were calculated using Gaussian 09 software at the B3LYP/6-311++G(2df,p) level of theory. These wavefunctions were then used as input for the AIM2000 program45 to perform the QTAIM-C analysis. Critical points on the gradient vector field of the electron density (∇ρ) were located using Newton's method with a maximum of 120 iterations and a step size factor of 0.5. The nature of each CP was determined by analyzing its Laplacian (∇2ρ) and eigenvalue properties. The molecular graphs were constructed following the established QTAIM-C protocol. This involved tracing the paths of steepest descent (downhill paths) from bond critical points (BCPs, (3,+1) CPs) and paths of steepest ascent (uphill paths) from ring critical points (RCPs, (3,−1) CPs). In addition, bond paths connecting neighboring BCPs were identified.
Intermolecular interactions in the crystal packing of compounds 4a–c were investigated using Hirshfeld surface (HS) analysis.29 CrystalExplorer21.5 software30 was employed to generate the HS and associated two-dimensional fingerprint plots. The three-dimensional dnorm (normalized contact distance) HS map was generated for each compound. A standard surface resolution was used for HS generation, and a fixed color scale of −0.2023 to +1.3588 au was applied to represent the range of normalized contact distances.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2335843–2335845. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4ra04563g |
This journal is © The Royal Society of Chemistry 2024 |