Skip to main content
ACS Omega logoLink to ACS Omega
. 2023 Oct 31;8(45):42632–42646. doi: 10.1021/acsomega.3c05372

Crystal Structure, Hirshfeld Surface Analysis, and Biological Activities of Schiff-Base Derivatives of 4-Aminoantipyrine

Esteban Aguilar-Llanos , Saskya E Carrera-Pacheco , Rebeca González-Pastor , Johana Zúñiga-Miranda , Cristina Rodríguez-Pólit , Arianna Mayorga-Ramos , Oscar Carrillo-Naranjo , Linda P Guamán , Juan Carlos Romero-Benavides §, Carlos Cevallos-Morillo , Gustavo A Echeverría , Oscar E Piro , Christian D Alcívar-León , Jorge Heredia-Moya ‡,*
PMCID: PMC10652364  PMID: 38024734

Abstract

graphic file with name ao3c05372_0008.jpg

Eight Schiff bases, synthesized by the reaction of 4-aminoantipyrine with different cinnamaldehydes, were studied in the solid state by using vibrational spectroscopy (IR) and X-ray diffraction techniques. The analysis was extended to the solution phase through ultraviolet–vis, fluorescence spectroscopy, and cyclic voltammetry. Finally, the crystal structures of four compounds (3b, 3d, 3g, and 3h) were determined and studied. In addition to the experimental study, theoretical calculations using the semiempirical method PM6/ZDO were performed to understand better the compound’s molecular properties, UV–vis, and infrared spectra. The primary difference is the angular conformation of the terminal phenyl rings around the corresponding linking C–N and C–C σ-bonds. Furthermore, as a result of extended bonding, the > C=N– azomethine group-containing Cpyr–N=(CH)–(CR)=(CH)–Cbz chain (with R=H for 3b, 3d, and 3h, and R=CH3 for 3g) is planar, nearly coplanar, with the mean plane of the pyrazole ring. Hirshfeld surface (HS) analysis was used to investigate the crystal packing and intermolecular interactions, which revealed that intermolecular C–H···O and C–H···N hydrogen bonds, π···π stacking, and C–H···π and C=O···π interactions stabilize the compounds. The energy contributions to the lattice energies of potential hydrogen bonds were primarily dispersive and repulsive. All derivatives were tested in vitro on LPS-stimulated mouse macrophages to assess their ability to suppress the LPS-induced inflammatory responses. Only a slight reduction in the level of NO production was found in activated macrophages treated with 3h. Additionally, the derivatives were tested for antimicrobial activity against several clinical bacteria and fungi strains, including three biofilm-forming microorganisms. Nevertheless, only Schiff base 3f showed interesting antibacterial activities with minimum inhibitory concentration (MIC) values as low as 15.6 μM against Enterobacter gergoviae. On the other hand, Schiff base 3f and, to a lesser extent, 3b and 3h showed antifungal activity against clinical isolates of Candida. The lowest MIC value was for 3f against Candida albicans (15.6 μM). It is interesting to note that the same Schiff bases exhibit the highest activity in both biological evaluations.

Introduction

Schiff base derivatives of 4-aminoantipyrine (4-AAP) are an intriguing class of organic compounds with a five-membered ring containing two nitrogen atoms, a double bond, an imine, and a carbonyl group.1 They are easily synthesized by condensing the amine group in 4-AAP with the carbonyl group attached to different aldehydes or ketones, facilitating access to a broad range of compounds with high yields. Additionally, the electron-donating nitrogen in the imine bond generates L-type ligands, capable of forming complexes with nearly any metal;24 therefore, these compounds have been employed successfully in a variety of applications, including optical5 and nonlinear optical applications,6,7 analytical,8 sensor development,9 solar cells,10 catalysis,11 corrosion inhibitors,12 organic synthesis intermediates, new drug development,13,14 and more.

A wide range of biological activities, including anti-leishmanial, analgesic, anti-inflammatory, anticancer, and antiviral activities, has been reported for Schiff bases.15,16 They also have strong radical scavenging and AChE enzyme inhibition activity,1720 as well as antibacterial and antifungal activities against several pathogens.21,22 The mechanism of action of 4-AAP compounds is thought to involve their ability to interact with biological macromolecules such as proteins23,24 and DNA, causing changes in their structure and function;25,26 as a result, they have great potential as candidates for developing new drugs to treat a wide range of diseases.

The great majority of reports on Schiff bases derived from 4-AAP have concentrated primarily on the use of substituted benzaldehydes, with relatively few reports involving the use of cinnamaldehydes. The presence of a double-bound conjugated with the imino moiety is attractive not only for the synthesis of novel compounds with potential biological activity but also for potential uses in optical applications.27,28 We recently synthesized several Schiff base derivatives obtained from 4-AAP with various cinnamaldehydes and preliminary data show that these compounds present antibacterial activity and antiproliferative activity against several carcinoma cell lines.29 In order to continue our investigation of these compounds, we report here vibrational (IR), electronic (UV–vis) spectroscopy, electrochemical characterization of these Schiff base derivatives, and X-ray diffraction studies of four compounds. Likewise, the vibrational and electronic spectra were discussed and assigned using theoretical calculations at the PM6/ZDO level of theory. Furthermore, the Hirshfeld surface (HS) analysis was used to investigate their crystal packing and intermolecular interactions to extend and discuss the changes in crystal packing and supramolecular stability observed in related compounds.30 Lastly, the anti-inflammatory activity, antifungal activity, and biofilm inhibition results are reported.

Results and Discussion

Schiff bases 3ah (Scheme 1) were synthesized in good to excellent yield by the condensation of 4-amino-1,5-dimethyl-2-phenylpyrazol-3-one (1) with various cinnamaldehydes 2ah in ethanol, as previously reported.29 In all cases, the synthesized compounds correspond to trans isomers around the side chain olefinic double bond (Jα–β ≈ 16 Hz), with no evidence of cis isomer formation (see the Experimental Section and Supporting Information, Figures S1–S8).

Scheme 1. Synthesis of Schiff Base Derivatives 3ah.

Scheme 1

Crystallographic Structural Results

Molecular Structures

Recrystallization in ethanol produced single crystals of four compounds (3b, 3d, 3g, and 3h) that were suitable for crystallographic analysis. Figure 1 is an ORTEP32 drawing of the solid-state molecules, and their corresponding bond distances and angles are in Tables S1–S6. Although the compounds studied are derivatives of the same molecular framework, their solid-state structures are similar, with the main difference being the angular conformation of the terminal phenyl rings around the corresponding linking C–N and C–C σ-bonds.

Figure 1.

Figure 1

(ad) View of 3b, 3d, 3g, and 3h solid-state molecules showing the labeling of the non-H atoms and their displacement ellipsoids at the 30% probability level. Weak intramolecular CH···O bond is indicated by dashed lines.

Because of extended π-bonding delocalization, the Cpyr–N=(CH)–(CR)=(CH)–Cbz chain (where R=H for 3b, 3d, and 3h, and R=CH3 for 3g), which includes the > C=N– azomethine group of the Schiff base is the planar [rms deviation of atoms from the best least-squares plane of 0.055 Å (3b), 0.044 Å (3d), 0.034 Å (3g), and 0.013 Å (3h)] and favored by a weak intramolecular CH···O bond, nearly coplanar with the mean plane of the pyrazole ring [angled at 6.3(2)° (3b), 7.4(2)° (3d), 3.2(1)° (3g), and 6.9(2)° (3h)]. The phenyl ring attached to the pyrazole ring subtends angles of 61.03(6)° (3b), 49.33(6)° (3d), 46.75(6)° (3g), and 55.39(8)° (3h) with it and the phenyl ring at the other molecular end forms angles of 30.0(1)° (3b), 6.6(2)° (3d), 40.2(1)° (3g), and 3.0(2)° (3h) with the corresponding chain plane.

Observed bond distances and angles for all four compounds agree with the established organic chemistry rules. For convenience, we shall refer to the metric of the representative and better-refined 3g compound. In particular, the observed short C–N distance of 1.277(2) Å in the > C=N– azomethine group confirms this link’s formal double bond character. Phenyl rings C–C bond distances [from 1.357(3) to 1.394(2) Å] are as expected for a resonant-bond structure. Within the pyrazole ring, observed d(N–N) = 1.412(2) Å, N–C bond distances in the 1.357(2)–1.460(2) Å range and the sum of bond angles around the N-atoms less than 360° [346.6(2) and 352.4(3)°] confirm the single-bond character of the links involving both amine N-atoms. The observed OC-C and short H3CC=C bond lengths of 1.434(2) and 1.359(2) Å, respectively, are as expected for the single and double bond characters of these links.

The geometrical parameters, such as bond lengths and angles, were also determined theoretically from the most stable conformers of 3b, 3d, 3g, and 3h at the PM6/ZDO level of theory and were compared using the root-mean-square deviation (rmsd). Tables S23 and S24 show the experimental and calculated bond lengths (Å) and bond angles (deg), while Table S25 shows the calculated standard error values of the bond lengths and angles. The comparative analysis revealed that the theoretical bond lengths 3b, 3d, 3g, and 3h are in good agreement with the values observed by X-ray diffraction. However, the rmsd for angles evidenced divergence between the theoretical and experimental values due to the theoretical structures determined in the gas phase without evaluating crystal packing effects or intermolecular interactions that can stabilize higher energy conformations and the level of calculation used.

Supramolecular Features and HS Analysis

The crystal packing diagrams for 3b, 3d, 3g, and 3h are shown in Figure 2a–d. The crystal packing of the compounds is stabilized by intermolecular Csp3–H···O, Csp2–H···O, Csp3–H···N, and Csp2–H···N hydrogen bonds (Table 1). For 3b, the carbonyl group of the 1H-pyrazole moiety arises as bifurcated C10–H10A···O1 (d(H···O) = 2.490 Å, ∠(C–H···O) = 162°) and C19–H19···O1 (d(H···O) = 2.350 Å, ∠(C–H···O) = 142°) hydrogen bonds (Figure 2b). In this sense, the intermolecular contacts between the oxygen atom of the 1H-pyrazole with hydrogen atoms of methyl groups and aromatic protons generate the R33(24) graph set. On the other hand, in 3b and 3d compounds, the C2–H2···N (d(H···N) = 2.600 Å, ∠(C–H···N) = 172°) and C2–H2···N3 (d(H···N) = 2.630 Å, ∠(C–H···N) = 160°) hydrogen bonds generate structural dimers between nitrogen atoms of the imine group and a hydrogen atom of the benzene moiety that create R22(15) and R22(16) motifs, respectively (Figure 2a,b).

Figure 2.

Figure 2

(ad) Packing of 3b, 3d, 3g, and 3hvia Csp3–H···O, Csp2–H···O, Csp3–H···N, and Csp2–H···N intermolecular interactions.

Table 1. Potential Hydrogen Bond (Distances in Å; Angles in °) Interactions and Lattice Energies (kJ/mol) for 3b, 3d, 3g, and 3ha.
compound 3bb
D-H···A d(D-H) d(H···A) d(D···A) (D-H···A) Rc Eele Epol Edis Erep ETOT
C2–H2···Ni 0.930 2.600 3.5188(2) 172 6.63 –6.0 –3.6 –42.1 20.8 –29.5
C10–H10A···O1ii 0.960 2.490 3.4158(2) 162 6.91 –16.4 –7.6 –29.4 17.3 –34.2
C19–H19···O1iii 0.930 2.350 3.1411(2) 142 7.53 –29.3 –14.3 –30.3 18.6 –51.4
C12–H12···O1 (intra) 0.930 2.370 3.0376(2) 129            
C14–H14···O3 (intra) 0.930 2.280 2.8479(2) 119            
compound 3db
C10–H10A···O1i 0.960 2.410 3.3562(3) 167 7.05 –20.8 –7.9 –30.2 16.0 –40.6
C2–H2···N3ii 0,930 2.630 3,5180 160 6.98 –32.7 –10.6 –74.0 38.4 –75.7
C12–H12···O1 (intra) 0.930 2.380 3.0463(2) 128            
compound 3gb
C11–H11C···O1i 0.960 2.604 3.5300(2) 162 6.91 –16.4 –7.6 –29.4 17.3 –34.2
C10–H10B···N3 (intra) 0.960 2.610 2.9831(3) 103            
C12–H12···O1 (intra) 0.930 2.440 3.0998(3) 128            
compound 3hb
C10–H10A···O1i 0.960 2.380 3.3293(3) 172 6.91 –16.4 –7.6 –29.4 17.3 –34.2
C11–H11C···O3ii 0.960 2.580 3.4985(3) 160 4.92 –14.6 –6.2 –97.6 44.2 –70.9
C17–H17···O1iii 0.930 2.410 3.2295(3) 147 7.53 –29.3 –14.3 –30.3 18.6 –51.4
C19–H19···O2iv 0.930 2.460 3.2900(3) 149 15.66 –21.4 –4.9 –10.7 10.7 –25.9
C14–H14···O3 (intra) 0.930 2.370 3.0409(2) 129            
a

Symmetry code (3b): (i) 2 – x,– y,– z (ii) x,– 1 + y, z (iii) 3/2 – x,– 1/2 + y, 1/2 – z. (3d): (i) x, 1 + y, z (ii) 1 – x,– y,– z. (3g): (i) −1 + x, y, z. (3h): (i) x, 1 + y, z (ii) 1 – x, 2 – y,– z (iii) 1 – x, 1 – y,– z (iv) 2 – x, 2 – y,– z.

b

Numbering according to Figure 1.

c

Distance between molecular centroids (mean atomic position) in Å.

For 3h, the crystal packing is stabilized by R22(10) and R22(30) graph-set ring motifs, which form dimers in the supramolecular assembly (Figure 2d). These intermolecular interactions C19–H19···O2 (d(H···O) = 2.460 Å, ∠(C–H···O) = 149°) and C11–H11C···O3 (d(H···O) = 2.580 Å, ∠(C–H···O) = 160°) evidences the important participation of hydrogen atoms of methyl groups and a benzyl moiety with oxygen atoms of the –NO2 group. Likewise, for 3d, 3g, and 3h, the 1H-pyrazole moiety generates C–H···O intermolecular interactions that stabilized the crystalline packing by hydrogen bonds C10–H10A···O1, C11–H11C···O1, and C17–H17···O1 between hydrogen atoms of the –CH3 group and oxygen of the carbonyl group.

On the other hand, intermolecular interactions of type π···π stacking, C–H···π, and C=O···π were evidenced for 3b, 3d, 3g, and 3h (Figures S17–S20, Supporting Information). The C–H···π intermolecular interactions for 3b, 3g, and 3h evidenced a close distance of contact (Comp. 3b: 2.921 and 2.901 Å, Comp. 3g: 2.897 Å and 2.911 Å, and Comp. 3h: 2.988 Å) where the interaction arises principally between hydrogen atoms of the benzene moiety. For 3b, 3d, and 3g, the intermolecular π stacking contacts (Figures S17–S19, Supporting Information) are observed between the 1H-pyrazole rings (intercentroid distances: Comp. 3b: 4.162 Å, Comp. 3d: 4.127 Å, and Comp. 3g: 3.890 Å). Moreover, interactions π···π stacking arise between the 1H-pyrazole moiety and the benzene ring for 3b and 3h (intercentroid distances: Comp. 3b: 3.996 Å and Comp. 3h: 3.646 Å). Likewise, for 3d and 3g, we observed C=O···π intermolecular interactions between 1H-pyrazole rings (intercentroid distances: Comp. 3b: 3.123 Å and Comp. 3g: 3.905 Å).

The features associated with the energy of each intermolecular interaction shown in Tables S26–S29 were evaluated by the Crystal Explorer program, applying the CE-HF/3-21G energy model for the electron densities. Figures S21–S24 show the detailed crystal lattice analysis and 3D-energy frameworks. The analysis of 3b, 3d, 3g, and 3h evidenced that the C–H···O hydrogen bonds are dominated mainly by dispersive energies with values in a range of −97.6––29.4 kJ mol–1 and electrostatic energy contributions in a range of −32.7––14.6 kJ mol–1. Moreover, compounds 3b and 3d show C–H···N hydrogen bonds, where the nature of the interactions evidenced dispersive and repulsive contributions of energy (Comp. 3b: Edis: −42.1 kJ mol–1 and Erep: −20.8 kJ mol–1; Comp. 3d: Edis: −74.0 kJ mol–1 and Erep: −38.4 kJ mol–1).

The HS analysis was performed to explore the relative contribution of each intermolecular interaction in the crystal packing. Figures S25–S28 (row A) show small contacts or intermolecular interactions on the dnorm surface, as highlighted in red dots. The red circular regions on the dnorm surfaces are indicative of Csp3–H···O, Csp2–H···O, Csp3–H···N, and Csp2–H···N hydrogen bonds and other contacts, where the carbonyl group (C=O) and methyl group (−CH3) of the 1H-pyrazole ring; moreover, the benzyl, imine, and nitro moieties promote intermolecular interactions in the studied compounds. Likewise, the shape index [Figures S25–S28 (row B)] evidence π···π stacking by “bow-tie” patterns that are observed as red and blue triangles over the surface, while C–O···π and C=O···π interactions are also evident as large and flat green regions delineated by a blue outline over the curvedness surface [Figures S25–S28 (row C)]. On the other hand, Figure S29 shows 2D-fingerprint plots that allowed rationalization of the contribution of significant intermolecular interactions in the crystalline arrangements. In this sense, the H···H contacts have the most significant contribution to the total HS, with 41, 60, 61, and 43% for 3b, 3d, 3g, and 3h, respectively (Figure 3). Moreover, the H···H contacts also contribute to the HS with moderate contributions of 24, 26, 24, and 21%. This result reveals that the close planarity of the molecular structures contributes to the relatively high proportion of H···H and C···H contacts, related to π···π stacking and C–H···π and C=O···π weak intermolecular interactions that contribute to stabilizing the crystalline packing. Besides, the C–H··· O intermolecular hydrogen bonds show relative contributions for 3b and 3h with 23 and 22%, respectively, due to the –NO2 groups.

Figure 3.

Figure 3

Relative (%) contributions of the main intermolecular contacts to the HS areas for compounds 3b, 3d, 3g, and 3h.

Electrochemical Characterization

Electrochemistry is a powerful tool to probe reactions involving electron transfers.33 Cyclic voltammetry (CV) is an electroanalytical technique employing a triangular potential waveform to identify species, equilibria, binding properties, and oxidation estates of organic, inorganic, and organometallic compounds.34 The electrochemical behaviors of the Schiff bases were studied using the CV technique in a DMSO solution containing 0.1 M NMe4ClO4.Figure 4 shows voltammograms of all synthesized Schiff-base derivatives. First, in Figure S30, we showed the potential window and background for all derivatives, from −2.5 to 0.73 V vs ferrocene/ferrocenium, with a little signal at −1.079 V vs Fc/Fc+ due to oxygen reduction. The voltammogram of Figure 4a shows data from compound 3a, the base structure of the derivative family, exhibiting two irreversible peak signals. The peak at Ep,a = 0.526 V vs Fc/Fc+, can be attributed to the oxidation that takes place at the amino group of the dipyrone ring,35 while the reduction one at Ep,c = −2.030 V vs Fc/Fc+, can be assigned to imine reduction.36 The voltammogram of 3b (Figure 4b) has two extra irreversible reduction signals labeled as b.1 at Ep,c = 1.343 V and b.2 at Ep,c = −1.571. These signals correspond to the reduction of the nitro group; the first step in aprotic media includes one electron interchange to get the radical anion R-NO2, then the second electron transfer occurs to obtain the dianion species R-NO22–.37Figure 4c shows data from Schiff base 3d, which has the same shape as 3a, with an oxidation peak at 0.474 V and an irreversible reduction signal at −2.030 V. In the voltammogram of 3d (Figure 4d), one extra irreversible oxidation peak, labeled as d.1, is found at Ep,a = 0.147 V, and is attributed to amine group oxidation. Schiff bases 3e and 3g show voltammograms similar to 3a (Figure 4e,g), with one oxidation peak at 0.466 and 0.492 V and a reduction peak at −2.043 and −2.174 V vs Fc/Fc+, respectively. The voltammogram of Schiff base 3f (Figure 4f) shows an additional peak at −0.115 V vs Fc/Fc+, which can be attributable to the bromine atom, and a diminution of 210 mV in the oxidation signal of the amino group of the dipyrone ring. Finally, the voltammogram of 3g (Figure 4g) shows two irreversible signals, g.1 at −1.256 V and g.2 at −1.529 V vs ferrocene/ferrocenium; these peaks are due to the reduction of the nitro group, which behaves similarly to structure 3b.

Figure 4.

Figure 4

Voltammograms of Schiff-base derivatives 3ah in DMSO (1.0 mM) + 0.1 M tetramethylammonium perchlorate at a scan rate of 100 mV/s.

IR Spectroscopy

Figures S9–S16 show the experimental IR spectra of 3a3h in the solid state. Theoretical calculations for the most stable conformers at the PM6/ZDO level of theory assisted in the tentative assignments of the most relevant vibrational modes. Table 2 shows the unscaled theoretical vibrational frequency of the most representative bands discussed below.

Table 2. Experimental and Calculated Frequencies (cm–1) and Tentative Assignment of the Most Relevant Vibrational Modes for 3a3h.
compound 3a
compound 3b
assignment PED (%)a Exp. Calc.c
assignment PED (%)a Exp. Calc.c
  IRb Freq. Int.   IRb Freq. Int.
νs (C–H) (86) Ar 3045 (vw) 2768 250 νs (C–H) (88) Ar 3052 (vw) 2758 324
νas (C–H) (91) Ar 3035 (vw) 2759 267 νas (C–H) (96) Ar 3019 (vw) 2719 145
νas (CH3) (84) pyrazole 3001 (vw) 2678 175 νas (CH3) (97) pyrazole 2961 (vw) 2677 122
νas (CH3) (88) pyrazole 2960 (vw) 2659 118 νas (CH3) (88) pyrazole 2846 (vw) 2658 80
ν (C–H) (91) azomethine 2928 (vw) 2626 105 ν (C=O) (87) pyrazole 1643 (vs) 1859 422
ν (C=O) (89) pyrazole 1643 (vs) 1859 891 ν (C=C) (83) alkene 1607 (w) 1776 24
ν (C=C) (81) alkene 1594 (m) 1776 9 ν (C=N) (84) azomethine 1592 (w) 1721 21
ν (C=N) (77) azomethine 1587 (m) 1722 5 νas (NO2) (91) 1569 (s) 1706 839
ν (C=C) (69) pyrazole 1558 (m) 1657 368 ν (C=C) (63) pyrazole 1557 (m) 1649 319
δ (C–C–H) (31) ip Ar 1415 (m) 1509 282 δ (C–C–H) (32) ip Ar 1510 (m) 1509 293
ν (C–N) (53) pyrazole 1455 (w) 1413 307 νs (NO2) (74) 1444 (w) 1441 420
ν (N–N) (61) pyrazole 1447 (w) 1388 345 ν (C–N) (52) pyrazole 1428 (w) 1417 435
δ (CH3) (35) pyrazole 1307 (m) 1298 26 ν (N–N) (52) pyrazole 1413 (w) 1390 340
δ (C–C–N) (35) pyrazole 1133 (m) 1169 438 δ (CH3) (29) pyrazole 1077 (w) 1077 77
δ (N–C–H) (56) pyrazole 1045 (w) 1056 72 δ (C–C–H) (63) pyrazole 978 (m) 970 5
δ (CH3) (60) pyrazole 1117 (s) 1017 60        
δ (C–C–H) (78) pyrazole 858 (vw) 963 85        
compound 3c
compound 3d
assignment PED (%)a Exp. Calc.c
assignment PED (%)a Exp. Calc.c
  IRb Freq. Int.   IRb Freq. Int.
νs (C–H) (94) Ar 3035 (vw) 2769 245 νs (CH3) (34) pyrazole 3039 (vw) 2779 77
νas (C–H) (89) Ar 3007 (vw) 2763 259 νs (CH3) (19) pyrazole 2961 (vw) 2762 74
νs (C–H) (76) Alkene 2960 (vw) 2745 214 νas (C–H) (63) Ar 2926 (vw) 2758 235
νas (CH3) (93) pyrazole 2681 (vw) 2737 28 νas (C–H) (53) Ar 2892 (vw) 2752 233
νas (CH3) (89) pyrazole 2812 (vw) 2735 92 νas (C–H) (90) Ar 2853 (vw) 2735 93
ν (C=O) (86) pyrazole 1643 (vs) 1851 904 νs (CH3) (29) N,N-dimethyl 2810 (vw) 2673 94
ν (C=C) (78) alkene 1614 (w) 1772 102 νas (CH3) (28) N,N-dimethyl 2797 (vw) 2664 74
ν (C=N) (79) azomethine 1594 (w) 1737 274 ν (C=O) (86) pyrazole 1643 (s) 1850 418
ν (C=C) (74) pyrazole 1579 (w) 1686 606 ν (C=N) (47) azomethine 1600 (s) 1737 139
ν (C–N) (36) pyrazole 1435 (w) 1406 96 ν (C=C) (34) pyrazole 1564 (m) 1647 533
δ (CH3) (74) methoxy 1414 (w) 1385 631 ν (C–N) (30) N,N-dimethyl 1434 (m) 1440 250
ν (N–N) (19) pyrazole 1413 (w) 1361 320 ν (C–N) (22) pyrazole 1407 (w) 1404 146
δ (CH3) (80) pyrazole 1381 (w) 1331 72 ν (N–N) (27) pyrazole 1368 (s) 1385 375
δ (C=C–H) (74) azomethine 1346 (vw) 1294 139 δs (CH3) (19) N,N-dimethyl 1314 (w) 1353 37
δ (C–C–H) (48) ip Ar 1510 (m) 1245 76 δ (CH3) (34) pyrazole 1288 (m) 1332 79
δ (O–C–H) (51) methoxy 1074 (vw) 1069 1 δ (C=C–H) (22) azomethine 1239 (vw) 1294 58
δ (CH3) (52) pyrazole 1050 (w) 1062 46 δas (CH3) (27) N,N-dimethyl 1184 (w) 1252 61
δ (C–C–H)(46) oop alkene 991 (m) 967 43 δas (CH3) (18) N,N-dimethyl 1023 (m) 1024 74
        δ (C–C–H) (20) oop alkene 946 (m) 951 72
compound 3e
compound 3f
assignment PED (%)a Exp. Calc.c
assignment PED (%)a Exp. Calc.c
  IRb Freq. Int.   IRb Freq. Int.
νs (C–H) (26) Ar 3028 (vw) 2769 117 νas (CH3) (36) ketone 3103 (vw) 2778 83
νas (C–H) (57) Ar 3011 (vw) 2758 186 νs (C–H) (41) Ar 3067 (vw) 2768 229
νas (C–H) (62) Ar 2979 (vw) 2752 123 νas (C–H) (49) Ar 3054 (vw) 2759 242
νas (CH3) (52) acetoxy 2941 (vw) 2684 123 νas (C–H) (89) Ar 3042 (vw) 2735 93
νas (CH3) (51) pyrazole 2916 (vw) 2678 116 ν (C–H) (98) ketone 3029 (vw) 2728 134
νas (CH3) (77) acetoxy 2876 (vw) 2651 119 νas (CH3) (65) ketone 2995 (vw) 2678 106
νs (C–H) (48) azomethine 2828 (vw) 2624 74 νas (CH3) (56) pyrazole 2942 (vw) 2659 86
ν (C=O) (88) ester 1755 (vs) 1860 421 ν (C=O) (88) pyrazole 1645 (vs) 1857 393
ν (C=O) (88) pyrazole 1744sh (m) 1843 710 ν (C=N) (69) azomethine 1608 (w) 1728 13
ν (C=N) (77) azomethine 1655 (w) 1722 12 ν (C=C) (66) pyrazole 1591 (m) 1666 294
ν (C=C) (56) pyrazole 1640 (m) 1652 383 δ (C–C–H) (35) ip Ar 1567 (w) 1641 100
δ (C–C–H) (16) ip Ar 1594 (m) 1509 308 ν (C–N) (17) pyrazole 1492 (m) 1509 276
ν (C–N) (17) pyrazole 1509 (m) 1418 257 ν (C–N) (22) pyrazole 1509 (w) 1412 246
ν (N–N) (26) pyrazole 1483 (m) 1389 271 ν (N–N) (25) pyrazole 1417 (m) 1388 316
δ (CH3) (17) pyrazole 1467 (m) 1368 131 δ (CH3) (38) pyrazole 1337 (m) 1332 76
δ (CH3) (38) pyrazole 1368 (m) 1330 75 δ (C=C–H) (48) azomethine 1225 (vw) 1266 21
δ (CH3) (38) acetoxy 1344 (m) 1312 123 δ (CH3) (45) pyrazole 1178 (w) 1242 31
δ (CH3) (17) acetoxy 1289 (m) 1212 354 δ (CH3) (39) pyrazole 1137 (m) 1129 52
δ (CH3) (28) methoxy 1212 (m) 1123 68 δ (C=C–H) (18) alkene 1075 (w) 1194 41
δ (CH3) (36) methoxy 1125 (s) 1030 64 δ (CH3) (29) pyrazole 942 (m) 1057 52
δ (C–C–H)(28) oop alkene 1032 (s) 959 82 δ (CH3) (29) pyrazole 926 (w) 1018 57
ν (C–O) (37) ester 993 (vs) 940 115 ν (C–Br) (20) 627 (w) 614 60
compound 3g
compound 3h
assignment PED (%)a Exp. Calc.c
assignment PED (%)a Exp. Calc.c
  IRb Freq. Int.   IRb Freq. Int.
νs (CH3) (34) pyrazole 3145 (vw) 2779 82 νs (C–H) (27) Ar 3117 (vw) 2768 117
νs (C–H) (50) Ar 3099 (vw) 2768 156 νas (C–H) (56) Ar 3100 (vw) 2756 192
νas (C–H) (37) Ar 3067 (vw) 2761 239 νas (C–H) (55) Ar 3071 (vw) 2749 210
νas (C–H) (39) Ar 3052 (vw) 2753 111 νas (C–H) (80) Ar 3031 (vw) 2724 137
νas (C–H) (31) ketone 3040 (vw) 2748 76 νas (C–H) (76) Ar 3022 (vw) 2722 113
νas (C–H) (89) Ar 3026 (vw) 2735 92 νas (CH3) (42) pyrazole 2997 (vw) 2677 84
νas (CH3) (58) alkene 2984 (vw) 2691 84 νas (CH3) (47) pyrazole 2975 (vw) 2657 84
νas (CH3) (65) pyrazole 2959 (vw) 2679 77 ν (C–H) (49) alkene 2928 (vw) 2624 76
νas (CH3) (55) alkene 2943 (vw) 2675 97 ν (C=O) (81) pyrazole 1646 (vs) 1861 417
νas (CH3) pyrazole 2919 (vw) 2661 79 ν (C=N) (78) azomethine 1612sh (w) 1718 54
ν (C=O) (86) pyrazole 1641 (vs) 1850 427 ν (NO2) (48) 1594 (s) 1714 586
ν (C=N) (74) azomethine 1618sh (w) 1758 19 ν (C=C) (41) pyrazole 1568 (s) 1645 158
ν (C=C) (58) pyrazole 1588 (s) 1688 194 δ (C–C–H) (25) ip Ar 1549 (m) 1634 626
δ (C–C–H) (35) ip Ar 1571 (w) 1641 88 ν (C–N) (16) pyrazole 1511 (s) 1508 278
δ (C–C–H) (16) ip Ar 1494 (m) 1527 40 ν (NO2) (42) 1433 (m) 1443 725
δ (C–C–H) (17) ip Ar 1481 (m) 1510 272 ν (N–N) (23) pyrazole 1420 (w) 1392 286
ν (C–N) (24) pyrazole 1418 (m) 1409 103 ν (C–N) (14) azomethine 1410 (w) 1378 170
ν (N–N) (27) pyrazole 1387 (w) 1387 290 δ (CH3) (19) pyrazole 1377 (w) 1353 66
δ (C=C–H) (23) alkene 1371 (w) 1370 99 δ (CH3) (17) pyrazole 1335 (w) 1334 92
δ (CH3) (23) alkene 1313sh (w) 1342 90 δ (CH3) (37) pyrazole 1250 (w) 1217 75
δ (CH3) (18) alkene 1302 (m) 1336 85 δ (C=C–H) (44) alkene 972 (m) 958 92
δ (C=C–H) (76) azomethine 1191 (vw) 1282 13 δ (NO2) (31) 871 (m) 808 92
δ (CH3) (51) pyrazole 1176 (w) 1244 43 δ (NO2) (37) oop 826 (m) 764 109
δ (CH3) (39) pyrazole 1154 (w) 1229 52 δ (NO2) (39) ip 534 (w) 527 2
a

ν, δ, y, and γ represent stretching, in-plane deformation (ip), and out-of-plane deformation (op) vibrational modes, respectively. In parentheses, potential energy distribution (PED) analysis.

b

vs, very strong; s, strong; m, medium; w, weak; vw, very weak; and sh, shoulder.

c

Calculated infrared frequencies (cm–1) and intensities (km mol–1) in parentheses (B3LYP/6-311++G(d,p).

The IR spectra (3a3h) show a set of weak absorption bands in the 3145 to 2797 cm–1 region, which are attributed to stretching C–H vibrational modes, both symmetric and asymmetric due to the aromatic, methyl, and alkene groups. Strong absorption bands at 1643 cm–1 (3a3d), 1744 cm–1 (3e), 1645 cm–1 (3f), 1641 cm–1 (3g), and 1646 cm–1 (3h), are attributed to the carbonyl of the pyrazole ring, as previously reported for related molecules.38,39 The shift of the C=O band toward lower frequencies in the IR spectra could be explained by the short intermolecular interactions Csp2–O···H–Csp3 observed in the supramolecular features in the crystalline structures of 3b, 3d, 3g, and 3h. The stretching C=C of the pyrazole group appears as medium absorption bands at 1558, 1557, 1579, 1564, 1640, 1591, 1588, and 1568 cm–1, while the weak and medium absorption bands at 1455, 1428, 1435, 1407, 1509, 1492, 1418, and 1410 cm–1 were attributed to stretching C–N. Finally, stretching N–N is assigned to weak absorption bands at 1447, 1413, 1413, 1368, 1468, 1483, 1417, 1387, and 1420 cm–1. All these characteristic absorption bands for the pyrazole group are in good agreement with related compounds.3841

The vibration corresponding to the azomethine –CH=N group characterizes the Shiff base. The C=N stretching vibrational mode was attributed to the weak absorption bands at 1587, 1592, 1594, 1600, 1655, 1608, 1618, and 1612 cm–1, for 3a3h, analogous to the weak absorption bands reported for similar compounds.40,41

On the other hand, the IR spectra of 3b3e and 3h showed distinctive absorption bands for all different substituents of the benzyl moiety. For 3b and 3h, the –NO2 group exhibited significant absorption bands at 1569 cm–1 and 1594 m–1, respectively, attributed to a stretching vibrational mode. The methoxy group in 3b showed a weak absorption band at 1414 cm–1, which was assigned to the bending O–C–H vibrational mode. For 3d, the C–N stretching was assigned to a weak absorption band at 1407 cm–1, whereas bands at 1314, 1184, and 1023 cm–1 were assigned to symmetric and antisymmetric vibrational modes. The absorption bands at 1755, 1334, 1289, and 1125 cm–1 in the 3e IR spectra indicated vibrational modes attributable to stretching of C=O and deformations of the –CH3 moiety of the acetoxy and methoxy groups, respectively. In the case of Schiff bases with an α carbon substitution, 3f exhibits a weak absorption band at 627 cm–1 assigned to the stretching C–Br deformation mode, whereas 3g exhibits a medium absorption band at 1312 cm–1 attributable to the –CH3 deformation mode.

Optical Properties

The experimental electronic absorption spectra of 3a3h are shown in Figure 5A, while the analyses of the calculated and experimental electronic absorption spectra of all Schiff bases are shown in Table 3. Electronic spectra were calculated at the TD-PM6/ZDO level of theory, while the molecular orbitals involved in the electronic transitions 3a3h are shown in Figures S31–S38.

Figure 5.

Figure 5

(A) Electronic absorption spectra of Schiff bases 3a3h using acetonitrile as the solvent (3a: 60 μM, 3b: 35 μM, 3c: 25 μM, 3d: 12.5 μM, 3e: 25 μM, 3f: 45 μM, 3g: 30 μM, and 3h: 70 μM). (B) Emission spectra of Schiff bases 3a3h in acetonitrile at room temperature.

Table 3. Experimental and Calculated Electronic Spectra (in nm) of 3a3h and Tentative Assignment of the Absorption Bands.

compound 3a
compound 3b
Exp.a Calc.b assignmentc Exp. Calc. assignment
195 253 (0.239) H-3 → L+1 (26%) 194 263 (0.222) H-2 → L+5 (11%)
  257 (0.203) H-4 → L+2 (35%)   272 (0.019) H-3 → L+7 (9%)
  262 (0.176) H → L+8 (22%)   273 (0.191) H → L+9 (13%)
240 318 (0.419) H-1 → L+1 (15%) 243 318 (0.379) H-1 → L+2 (12%)
286 390 (0.060) H → L+5 (43%) 299 391 (0.076) H → L+8 (43%)
357 464 (0.518) H → L(50%) 381 466 (0.519) H → L+1 (27%)
compound 3c
compound 3d
Exp. Calc. assignment Exp. Calc. assignment
193 257 (0.204) H-4 → L+3 (31%) 193 257 (0.204) H-4 → L+3 (31%)
  263 (0.170) H → L+10 (20%)   263 (0.157) H-2 → L+10 (14%)
  276 (0.238) H-1 → L+2 (16%)   279 (0.149) H-5 → L+3 (27%)
242 327 (0.412) H-3 → L (14%) 245 331 (0.299) H-2 → L+1 (17%)
291 371 (0.095) H → L+1 (19%) 392 488 (0.568) H → L (36%)
365 465 (0.432) H → L (30%)      
compound 3e
compound 3f
Exp. Calc. assignment Exp. Calc. assignment
192 268 (0.179) H-4 → L+2 (13%) 194 284 (0.068) H-5 → L (10%)
  289 (0.204) H-7 → L+2 (7%)   290 (0.090) H → L+3 (13%)
243 394 (0.040) H-2 → L+2 (21%) 241 325 (0.343) H-1 → L+1 (23%)
291 436 (0.072) H → L (16%) 294 382 (0.108) H → L+7 (23%)
363 477 (0.508) H → L (34%) 360 462 (0.335) H → L (36%)
compound 3g
compound 3h
Exp. Calc. assignment Exp. Calc. assignment
193 315 (0.518) H-1 → L+1 (20%) 194 276 (0.076) H-8 → L+7 (67%)
344 430 (0.250) H → L (29%)   286 (0.099) H → L+11 (18%)
        295 (0.109) H-1 → L+6 (25%)
      243 322 (0.438) H-9 → L (30%)
      297 394 (0.053) H → L+6 (51%)
      402 474 (0.620) H → L (22%)
a

Absorption maxima and spectral positions are given in nm.

b

Oscillator strengths of calculated transitions, shown in parentheses, are in atomic units.

c

H: HOMO and L: LUMO.

The experimental electronic spectra 3a3h showed strong absorption bands from 192 to 195 nm. The electronic transitions in this range involve principally the participation of π → π* orbitals for 3a, 3b, 3f, and 3h due to H-3 → L+1, H-2 → L+5, H → L+3, and H-1 → L+6 excitations, respectively. These strong absorption bands arise by electronic transitions from the π-bonding orbitals of the benzene, 1H-pyrazole, azomethine, and double bond of the alkene moiety to π*-antibonding orbitals of the benzene ring. On the other hand, for the compounds 3c, 3d, 3e, and 3g that are present in the chemical structure –O–CH3, –N(CH3)3, –OCH3/–OCOCH3, and CH3 groups, the absorption electronic transitions arise by π → π* orbitals and n → π* orbitals, where the strong absorption bands show the participation of nonbonding contributions of the nitrogen of the 1H-pyrazole and azomethine groups.

The electronic spectra (3a3f and 3h) showed weak absorption bands in the 240–245 nm region. The electronic transitions of 3a3f involve the participation of π → π* orbitals, where the π-bonding orbitals along the system of Schiff base derivatives of 4-AAP show participation and transit to π*-antibonding orbitals of the benzene and 1H-pyrazole rings. The weak absorption band at 243 nm in the spectrum of 3h can be attributed to n → π*, where the transition arises from nonbonding contributions localized in the oxygen atoms of the –NO2 group.

On the other hand, for 3a3c, 3e3f, and 3h, the electronic spectra showed weak bands in a range of 286–299 nm. The electronic transitions arise from π → π* orbitals, where the π-bonding orbitals of the aromatic system benzene, 1H-pyrazole, and azomethine moiety transit to π*-antibonding orbitals localized principally in the benzene and 1H-pyrazole rings. Moreover, the electronic spectra showed a weak broadband range of 342–402 nm for all compounds 3a3h. These absorption bands arise mainly from HOMO → LUMO transitions, where the π-bonding orbitals of the aromatic system transit to π*-antibonding orbitals localized in the 1H-pyrazole, azomethine, and double bonds of the alkene moiety. Interestingly, none of the Schiff bases present emission when excited at their respective absorption maxima except 3d, which has a prominent emission peak at 480 nm (Figure 5B).

Biological Activity

Preliminary studies showed that some Schiff bases 3 had antitumoral activity with a better therapeutic index (TI) than cisplatin, a standard chemotherapeutic drug, as well as activity against Gram-positive bacteria;29 therefore, to expand the evaluation of the biological activity of synthesized compounds, anti-inflammatory and additional antimicrobial activities were evaluated.

Anti-inflammatory Activity

In order to assess the anti-inflammatory activity of Schiff base derivatives, RAW264.7 cells were pretreated with the compounds prior to stimulation with LPS, and the levels of NO in the culture media were determined by using the Griess method. The concentrations used were selected based on the IC50 values (Table S30), and cell viability at the end of the experiment was maintained at over 80% for all the controls and samples assayed (Figure 6). The activity of 3f could not be determined due to the high inhibition of RAW264.7 proliferation observed.

Figure 6.

Figure 6

Anti-inflammatory activity of compounds 3a3h. RAW264.7 cells were pretreated with the derivatives and then stimulated with LPS. The percentage of NO production was calculated using cells treated with LPS only (control + LPS) as 100% NO production. Dexamethasone (DEX) and aspirin (ASA) were used as positive controls, and cells were only incubated with cell media as negative controls (control). Bars represent the percentage of the NO production. Lines represent the % cell viability for each corresponding set of data (black, controls; purple, 1 μM; and blue, 25 μM).

As shown in Figure 6, LPS stimulation resulted in a marked induction of NO production compared to that of untreated cells (control). The values obtained after the pretreatment with the derivatives alone without LPS stimulation were the same as those obtained with the negative control, only incubated with cell media (not shown).

Although pretreatment with 25 μM of 3b, 3e, and 3h four h before LPS stimulation for 18 h reduced NO production compared to the control + LPS, the percentage of inhibition was lower when compared to the positive controls (dexamethasone and aspirin) (36.9 and 43.4% of NO production, respectively). The most effective of these three derivatives was 3h, which reduced the NO production by 87.9%, followed by 3b and 3e. These results are consistent with other studies where the introduction of nitro groups in Schiff bases, such as in 3h, presented higher anti-inflammatory activity compared to the presence of other groups.42

Antimicrobial Activity

On the other hand, to expand on the previously reported antibacterial activity of these compounds,29 their activity was tested against many Gram-negative bacteria, including Pseudomonas aeruginosa, Salmonella enterica, Klebsiella ozaenae, and Enterobacter gergoviae. The minimum inhibitory concentrations (MICs) are shown in Table 4. MIC values above 250 μM were considered ineffective and labeled “non-active” (NA).

Table 4. MICa and Type of Inhibitionb,c Determined by Growth Kinetics over 20 h (OD600) for Bacteria and over 72 h (OD530) for Fungi after Serial Microdilution in 96-Well Plates for Active Schiff Bases.
bacteria strain 3b 3f 3h
P. aeruginosa ATCC 27853 NA NA NA
S. enterica ATCC 14028 NA 31.25c NA
K. ozaenae (clinical isolate) NA 62.50c NA
E. gergoviae (clinical isolate) NA 15.60b NA
Fungal Strain
Candida albicans (clinical isolate) 250 15.60 125
Candida krusei (clinical isolate) NA 62.50 NA
Candida tropicalis (clinical isolate) NA 31.25 NA
Candida glabrata (clinical isolate) NA 62.50 NA
a

μM.

b

Bacteriostatic.

c

Bactericidal; NA: non-active.

We found that compounds 3a, 3b, 3c, 3d, 3e, 3g, and 3h did not show antibacterial activities against the tested bacteria; however, Schiff base 3f inhibited the bacterial growth of all strains, except P. aeruginosa, with the lowest MIC value (15.6 μM) against E. gergoviae. According to the observations, 3f had a bactericidal effect on S. enterica and K. ozaenae while having a bacteriostatic effect on E. gergoviae (Table 4). Interestingly, none of the Schiff bases inhibited P. aeruginosa growth, which is consistent with previous findings;16,43 however, it is unclear if the observed lack of activity is due to the Schiff base’s structure, the mechanism of action, or both.

In addition to the tested bacterial strains, the antifungal activities were determined by testing the eight Schiff bases against clinical isolates of Candida albicans, Candida krusei, Candida tropicalis, and Candida glabrata. Compounds 3a, 3b, 3c, 3d, 3e, and 3g did not show any activity against the tested Candida strains; however, Schiff bases 3b, 3f, and 3h showed fungal growth inhibition, and their MIC values are detailed in Table 4. MIC values above 250 μM were considered ineffective and labeled “NA”.

Of the three compounds that showed antifungal activities, 3f showed activity for all the tested strains (MIC < 100 μM), with the lowest MIC value of 15.6 μM for C. albicans. C. albicans was the only strain susceptible to three compounds (3b, 3f, and 3h). Interestingly, 3f showed higher antifungal activity than 3h and 3b within the same strain, with MIC values at least 8 times lower.

Forming biofilms is widely accepted as an important factor in developing long-term infections and resistance to conventional antimicrobial therapy.47 Studies have shown that biofilm communities are more resistant to antimicrobial treatments than planktonic cells,48 highlighting the need for new compounds with antibiofilm potential. In this context, Schiff bases 3 were tested for their biofilm inhibitory ability against three biofilm-forming microorganisms: Staphylococcus aureus, Enterococcus faecalis, and C. tropicalis, and their MIC are shown in Table 5. MIC values above 250 μM were considered ineffective and labeled “NA”.

Table 5. MICa of Biofilm Inhibition Activity of Schiff Bases 3ah against Biofilm-Forming Microorganisms at 24 h.
bacteria strain 3e 3f 3h
S. aureus ATCC 25923 NA 62.50 NA
E. faecalis ATCC 29212 250 125 250
C. tropicalis (clinical isolate) NA 62.50 250
a

μM; NA: non-active.

Compounds 3a, 3b, 3c, 3d, and 3g showed no biofilm inhibition activity against the tested microorganisms. On the other hand, compound 3f inhibited biofilm formation in S. aureus and C. tropicalis by 82.77 and 90.41%, respectively, at 62.5 μM, whereas E. faecalis inhibition was 75.69% at 125 μM. The other active compounds were 3h and 3e, which reduced biofilm formation in E. faecalis at a concentration of 250 μM by 76.63 and 69.05%, respectively. Finally, compound 3h inhibited biofilm formation in C. tropicalis by 22.11% at the highest concentration evaluated.

As in the antibacterial evaluation, 3f exhibited a remarkably high potential for biofilm inhibition. Bromide-containing compounds have been linked to this activity,49,50 and several bromine-derived compounds have been reported to interfere with biofilm development through Quorum-sensing inhibition since these compounds can inactivate receptors involved in biofilm formation and other virulence factors.51,52

Conclusions

The Schiff-base derivatives 3ah can be readily synthesized with high yields by the condensation reaction between 4-AAP (1) and various cinnamaldehydes. Vibrational (IR), electronic (UV–vis), and fluorescence spectroscopy were used to characterize the structure. The theoretical results support assigning the vibrational and electronic spectra absorption bands. The crystal packing of 3b, 3d, 3g, and 3h showed intermolecular C–H···O and C–H···N hydrogen bonds, where 1H-pyrazole, benzene, methyl, and NO2 moieties evidence important participation-generating intermolecular contacts that stabilize the supramolecular assembly by R33(24), R22(15), R22(16), R22(10), and R22(30) graph-set ring motifs. Moreover, the HS analysis revealed a dominant percentage of relative contributions C···H (range 21–26%) and H···H (range 41–61%) contacts due to the planarity of crystalline structures that generate π···π stacking, C–H···π, and C=O···π intermolecular interactions. Schiff-base derivative 3h showed minor anti-inflammatory activity. Furthermore, Schiff bases 3f and, to a lesser extent, 3h and 3b, have promising activity against a variety of clinically important bacteria and fungi and should be investigated further. The cytotoxicity observed in 3f could be attributed to the bromine atom’s oxidative properties, which favor microbial inhibition; nevertheless, more research is needed to understand this compound’s actual mode of action.

Experimental Section

General Information

Dimethyl sulfoxide (DMSO, Fisher-certified reagent), molecular sieve 4 Å (Merck), all solvents, and other reagents were from Aldrich (St. Louis, MO, USA) and were used as received without further purification. All melting points were uncorrected and determined on a Fisher-Johns analogue melting point apparatus. FTIR spectra were recorded by a PerkinElmer FTIR Spectrum One using an ATR system (4000–650 cm–1). The 1H and 13C NMR spectra were recorded at 298 K on a Bruker ADVANCE 500 MHz spectrometer equipped with a z-gradient, triple-resonance (1H, 13C, and 15N) cryoprobe using DMSO-d6 or CDCl3 as solvents. Chemical shifts are expressed in ppm with TMS as an internal reference (TMS, δ = 0 ppm) for protons. Reactions were monitored by TLC on silica gel using ethyl acetate/hexane mixtures as a solvent and compounds visualized by a UV lamp. The reported yields are for the purified materials and are not optimized. Electrochemical experiments were carried out using a PGSTAT128N Autolab potentiostat (Metrohm AG), with Nova 2.1 software into a three-electrode cell composed of a graphite counter electrode (ca. 0.35 cm2), a silver/silver chloride reference electrode (Ag/AgCl 3.0 M, CH Instruments), and a glassy carbon working electrode (3.0 mm diameter, CH Instruments). Voltammograms were obtained employing a scan rate of 100 mV/s into a 0.1 M solution of NMe4ClO4 with a concentration of ca. 1.0 mM of each target molecule. Before each measurement, argon was bubbled by ca. to 10 min to displace oxygen from the solution.

Synthesis of Schiff Bases 3a–h

All Schiff bases 3 were synthesized, as previously reported.16,29 Equimolar amounts of 4-amino-1,5-dimethyl-2-phenylpyrazol-3-one (1) (1.722 mmol) and substituted cinnamaldehydes 2ah (1.722 mmol) were dissolved in 5.0 mL of EtOH and refluxed for 1 to 24 h, except for the synthesis of 3f, which was carried out at room temperature. The progress of the reaction was monitored by TLC. After completion, the precipitates formed were filtered, washed, dried, and purified by recrystallization with ethanol. All compounds were characterized, and all of the data obtained agreed with the proposed structures (see Supporting Information).

X-ray Diffraction Data and Structural Solution and Refinement

The measurements were performed on an Oxford Xcalibur Gemini, Eos CCD diffractometer with graphite-monochromated MoKα (λ = 0.71073 Å) radiation. X-ray diffraction intensities were collected (ω scans with ϑ and κ-offsets), integrated, and scaled with the CrysAlisPro53 suite of programs. The unit cell parameters were obtained by least-squares refinement (based on the angular setting for all collected reflections with intensities larger than seven times the standard deviation of measurement errors) using CrysAlisPro. The data were corrected empirically for absorption employing the multiscan method implemented in CrysAlisPro.

The non-H crystal structures of 3b, 3d, and 3h were solved by the dual space procedure implemented in SHELXT54 and the molecular model refined by full-matrix least-squares with SHELXL of the SHELX suite of programs.55 At this stage, difference Fourier maps showed most hydrogen atoms. These, however, were positioned at their expected geometric locations and refined with the riding model. The methyl H-atoms were treated during the refinement as rigid groups allowed to rotate around the corresponding C–CH3 or N–CH3 bonds to maximize the sum of the residual electron density at the calculated H-positions.

Most of the diffraction pattern of the 3g crystal was interpreted in terms of a two-component monoclinic twin. The unit cell parameters for both crystal domains were equal to experimental accuracy. These domains were rotated 180° from each other around the reciprocal c* axis. The reflections were indexed in the reciprocal unit cell of the corresponding domains. It was resorted to the twin crystal data reduction facility implemented in CrysAlisPro to generate two datasets, namely, a regular one with the diffraction data indexed in the reciprocal unit cell of the largest domain (with about 57% diffracting power) and a second one including all reflections from both domains with the overlapping ones flagged for structure development and refinement. The first dataset was employed to solve and refine the non-H structure as described above for the other compounds. Despite the correct molecular model, however, the refinement showed evidence for the presence of overlap between reflections from both crystal twins. This evidence included (besides the visual rendering of the weighted reciprocal space implemented in CrysAlisPro) a relatively high agreement R1-factor (0.192) and the most disagreeable reflections list showing systematically larger F(obs) as compared with F(calc) values. We therefore refined the initial molecular model against the second dataset, which includes all collected reflections for both crystal domains, employing the untwining process implemented in SHELXL. Now that the R1-factor has dropped to 0.077, the sign of F(obs) – F(calc) difference for the most disagreeable reflections was more evenly distributed, and a difference Fourier map showed all H-atoms. These, however, were positioned and refined with the riding model as detailed above.

Crystal data and structure refinement results are summarized in Tables S1 and S2. Crystallographic structural data have been deposited at the Cambridge Crystallographic Data Centre (CCDC). Any request to the CCDC for this material should quote the full literature citation and the reference numbers CCDC 2226611 (3b), 2226612 (3d), 2226613 (3g), and 2226614 (3h).

HS Calculations

The CrystalExplorer 17.556 program was used for the development of lattice energy and HS calculations. In this sense, the HS was evaluated by descriptors of surface, dnorm (normalized contact distance), shape index, and curvedness that decode and quantify intermolecular interactions in the crystal lattice. The dnorm surfaces were mapped over a fixed color scale of −0.078 au (red) – 0.657 au (blue). In addition, the shape index was mapped in the color range of −1.0 au (concave) – 1.0 au (convex) Å and curvedness in the range of −4.0 au (flat) – 0.4 au (singular) Å. The 2D-fingerprint plots (2D-FP) were used to visualize and decode the relative contributions of the main intermolecular contacts to the HS. Moreover, the intermolecular interaction energies were calculated using the TONTO program with the energy model CE-HF···HF/3-21G electron densities, integrated into the CrystalExplorer 17.5 program. The interaction energy and intermolecular contacts between molecules are expressed in terms of four components: electrostatic (Eele), polarization (Epol), dispersion (Edis), and exchange-repulsion (Erep). The 2D-FP were generated by using the translated (1.0–2.8 Å) range, and reciprocal contacts were included.

Electrochemical Characterization

Tetramethylammonium perchlorate (NMe4ClO4, ACS reagent, supporting electrolyte) and ferrocene (>98%) were acquired from Sigma-Aldrich. All reagents were employed as received without further purification processes. Electrochemical experiments were carried out using a PGSTAT128N Autolab potentiostat (Metrohm AG) with Nova 2.1 software in a three-electrode cell. The cell was composed of a graphite bar as a counter electrode (ca. 0.35 cm2), a silver/silver chloride as a reference electrode (Ag/AgCl 3.0 M, CH Instruments, Inc.), and a glassy carbon as a working electrode (3.0 mm diameter, CH Instruments, Inc.). Voltammograms were obtained employing a scan rate of 100 mV/s into a 0.1 M solution of NMe4ClO4 in DMSO, with a concentration of ca. 1.0 mM of each target molecule. Before each measurement, argon was bubbled for approximately 10 min to displace oxygen from the solution.

Biological Evaluation

Evaluation of Anti-inflammatory Activity

RAW264.7 cells (1 × 105 cells/well) were seeded onto a 24–well culture plate and pretreated with 1, 25, and 50 μM of the compounds for 4 h. Dexamethasone (DEX) and aspirin (ASA) were used as controls. After, LPS (InvivoGen) at a final concentration of 1 μg/mL was added to the cells and incubated for an additional 18 h. NO production was monitored by adding 50 μL of Griess reagent (Sigma) to 50 μL of supernatants. After incubating for 10 min at room temperature in the dark, the absorbance was measured in a Cytation5 multimode detection system (BioTek) at 540 nm. The % reduction of NO was calculated using cells treated with LPS only as 100% NO production. Additionally, to assess cell viability after LPS stimulation, cells were washed with PBS and fixed with 4% paraformaldehyde for 20 min at room temperature before staining with 0.5% (w/v) crystal violet for 30 min at room temperature. Then, the plates were gently washed with water to eliminate the extra dye and left to dry at room temperature. Dried plates were scanned and quantified in a Cytation5 multimode detection system (BioTek) at 570 nm. Untreated cells were used as a 100% cell viability control.

Evaluation of Antibacterial Activity

The antibacterial activity of the eight Schiff bases was tested against the Gram-negative bacteria P. aeruginosa ATCC 27853, S. enterica ATCC 14028, K. ozaenae, and E. gergoviae, using the previously reported method.29

The bacterial inoculum, stock solutions, blanks, and controls were prepared as previously described.29 Several antibiotics were used as controls for growth inhibition at the recommended working concentrations for the tested strains. Specifically, chloramphenicol (20 μg/mL) was used for S. enterica ATCC 14028 and E. gergoviae, tetracycline (10 μg/mL) was used for K. ozaenae, and streptomycin (100 μg/mL) was used for P. aeruginosa ATCC 27853.

Concentrations (0.5–250 μM) were used for the Schiff bases. The MIC was determined after tracking bacterial growth over 20 h in samples exposed to the tested compound at different concentrations. The MIC was defined as the lowest concentration of the antibacterial agent that completely inhibited the growth of the microorganism, as determined by the optical density at 600 nm. These assays were performed at least in triplicates.

Evaluation of Antifungal Activity

Antifungal activity was tested against the clinical isolates of C. albicans 13932, C. krusei, C. glabrata, and C. tropicalis using the microdilution method.41,42,57

The fungal inoculum was prepared in yeast dextrose peptone (YPD) media to a final cell density of 5 × 102 cells/mL. Stock solutions were prepared as described before. The final concentration of DMSO was 2.5% v/v, which was shown to not affect the fungal growth. As a control, fungal cells were grown with 2.5% DMSO to rule out any potential growth inhibitory effect. Additionally, nourseothricin was used as a control for growth inhibition at 200 μg/mL. YPD alone and that supplemented with the compounds at different concentrations were used as blanks.

Drug susceptibility was assessed with concentrations ranging between 15.6 and 250 μM. The following modifications were included: first, compounds were serially diluted in DMSO, then 5 μL of each dilution was added to 195 μL of fungal suspension (5 × 102 cells/mL) up to a total volume of 200 μL. The plates were then incubated at 35 °C for 72 h without shaking, and the OD530 was monitored every 30 min on a Cytation5 multimode detection system (BioTek). The MIC was determined after fungal growth was monitored for 72 h in samples exposed to the tested compounds at different concentrations. These assays were performed at least in triplicate.

Evaluation of Biofilm Inhibition Activity

Bacterial strains: Gram-positive S. aureus ATCC 25923, E. faecalis ATCC 29212, and the fungal strain C. tropicalis (clinical isolate), were cultured in TSB + G (Tryptic Soy Broth medium supplemented with 1% Glucose) overnight.

1/100 dilution of the overnight culture was prepared, and aliquots of 150 μL were distributed into 96-well polystyrene microtiter plates along with a range of concentrations of Shiff bases 3ah. Plates were incubated under static conditions at 37 °C for 24 h.

The dilution stock solutions were prepared in 100% DMSO, and the final concentration of DMSO used in the assay was 2.5%. The minimum biofilm inhibitory concentration of the Shiff bases was determined by serial dilutions (250–15.62 μM). Concentrations were tested in technical duplicates during three independent biological replicates. All strains were also cultured with 2.5% DMSO (positive control) for the data analysis.

After a 24 h incubation period, the medium was carefully removed with a micropipette, and the plate was washed twice with PBS 1x (pH 7.2). The plate was dried inside a laboratory oven at 60 °C for 1 h. Subsequently, 150 μL of 0.1% crystal violet solution was added to each well for 20 min at room temperature. The staining solution was discarded, and wells were washed three times with 1x PBS (pH 7.2). Finally, 200 μL of 98% ethanol was added to each well, and the wells were incubated for 30 min at room temperature, and absorbance was measured at 590 nm. The biofilm inhibition percentage was calculated using the following formula

graphic file with name ao3c05372_m001.jpg

Computational Details

Quantum chemical calculations were performed for the ground state (gas phase) for 3ah with the program package Gaussian 09, Rev D.01. A conformational analysis around the free rotation bond C7–C8–N3–C12 (see Figures 1 and S39) was performed with the semiempirical method PM6/ZDO to explore minimum energy conformations. To confirm the stability of the optimized conformations, vibrational frequency calculations were performed at the same level of theory (PM6/ZDO). No scaling factor was applied to the computed frequencies. In all cases, the energy minima structures have no imaginary values of frequencies. The electronic transitions were calculated with the time-dependent density functional theory at the same theory level without the solvent’s implicit effect.

Acknowledgments

We thank CONICET (grant PIP 0651) and UNLP (grant 11/X857) of Argentina for their financial support. G.A.E. and O.E.P. are Research Fellows of CONICET. We would like to thank the Instituto de Microbiologia de la Universidad San Francisco de Quito for providing some ATCC bacteria used in this study.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.3c05372.

  • Copies of 1H NMR and IR spectra from synthesized compounds; crystallographic data, HSs, and intermolecular interactions of compounds 3b, 3d, 3g, and 3h; molecular orbitals involved in the electronic transitions; potential energy curves of all Schiff bases; and cytotoxicity of all synthesized compounds on RAW264.7 cells (PDF)

Author Present Address

Facultad de Medicina, Universidad de las Américas, Quito, Ecuador

The authors declare no competing financial interest.

Supplementary Material

ao3c05372_si_001.pdf (4.5MB, pdf)

References

  1. Deshmukh P.; Soni P. K.; Kankoriya A.; Halve A. K.; Dixit R. 4-Aminoantipyrine: A Significant Tool for the Synthesis of Biologically Active Schiff Bases and Metal Complexes. Int. J. Pharm. Sci. Rev. Res. 2015, 34 (1), 162–170. [Google Scholar]
  2. Catalano A.; Sinicropi M. S.; Iacopetta D.; Ceramella J.; Mariconda A.; Rosano C.; Scali E.; Saturnino C.; Longo P. A Review on the Advancements in the Field of Metal Complexes with Schiff Bases as Antiproliferative Agents. Appl. Sci. 2021, 11 (13), 6027. 10.3390/app11136027. [DOI] [Google Scholar]
  3. Matela G. Schiff Bases and Complexes: A Review on Anti-Cancer Activity. Anticancer. Agents Med. Chem. 2020, 20 (16), 1908–1917. 10.2174/1871520620666200507091207. [DOI] [PubMed] [Google Scholar]
  4. Sakthivel A.; Jeyasubramanian K.; Thangagiri B.; Raja J. D. Recent Advances in Schiff Base Metal Complexes Derived from 4-Aminoantipyrine Derivatives and Their Potential Applications. J. Mol. Struct. 2020, 1222, 128885. 10.1016/j.molstruc.2020.128885. [DOI] [Google Scholar]
  5. Zhang J.; Xu L.; Wong W. Y. Energy Materials Based on Metal Schiff Base Complexes. Coord. Chem. Rev. 2018, 355, 180–198. 10.1016/j.ccr.2017.08.007. [DOI] [Google Scholar]
  6. Arumugam A.; Shanmugam R.; Munusamy S.; Muhammad S.; Algarni H.; Sekar M. Study of the Crystal Architecture, Optoelectronic Characteristics, and Nonlinear Optical Properties of 4-Amino Antipyrine Schiff Bases. ACS Omega 2023, 8, 15168–15180. 10.1021/acsomega.2c08305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Nibila T. A.; Ahamed T. S.; Soufeena P. P.; Muraleedharan K.; Periyat P.; Aravindakshan K. K. Synthesis, Structural Characterization, Hirshfeld Surface and DFT Based Reactivity, UV Filter and NLO Studies of Schiff Base Analogue of 4-Aminoantipyrine. Results Chem. 2020, 2, 100062. 10.1016/j.rechem.2020.100062. [DOI] [Google Scholar]
  8. Ebosie N. P.; Ogwuegbu M. O. C.; Onyedika G. O.; Onwumere F. C. Biological and Analytical Applications of Schiff Base Metal Complexes Derived from Salicylidene-4-Aminoantipyrine and Its Derivatives: A Review. J. Iran. Chem. Soc. 2021, 18 (12), 3145–3175. 10.1007/s13738-021-02265-1. [DOI] [Google Scholar]
  9. Kumari N.; Singh S.; Baral M.; Kanungo B. K. Schiff Bases: A Versatile Fluorescence Probe in Sensing Cations. J. Fluoresc. 2023, 33 (3), 859–893. 10.1007/s10895-022-03135-6. [DOI] [PubMed] [Google Scholar]
  10. Ismail A. H.; Al-Zaidi B. H.; Abd A. N.; Habubi N. F. Synthesis and Characterization of Novel Thin Films Derived from Pyrazole-3-One and Its Metal Complex with Bivalent Nickel Ion to Improve Solar Cell Efficiency. Chem. Pap. 2020, 74 (7), 2069–2078. 10.1007/s11696-019-01007-1. [DOI] [Google Scholar]
  11. Bhat A. P. I.; Inam F.; Bhat B. R. Nickel Catalyzed One Pot Synthesis of Biaryls under Air at Room Temperature. RSC Adv. 2013, 3 (44), 22191. 10.1039/c3ra43363c. [DOI] [Google Scholar]
  12. Okey N. C.; Obasi N. L.; Ejikeme P. M.; Ndinteh D. T.; Ramasami P.; Sherif E.-S. M.; Akpan E. D.; Ebenso E. E. Evaluation of Some Amino Benzoic Acid and 4-Aminoantipyrine Derived Schiff Bases as Corrosion Inhibitors for Mild Steel in Acidic Medium: Synthesis, Experimental and Computational Studies. J. Mol. Liq. 2020, 315, 113773. 10.1016/j.molliq.2020.113773. [DOI] [Google Scholar]
  13. Raczuk E.; Dmochowska B.; Samaszko-Fiertek J.; Madaj J. Different Schiff Bases—Structure, Importance and Classification. Molecules 2022, 27 (3), 787. 10.3390/molecules27030787. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Ceramella J.; Iacopetta D.; Catalano A.; Cirillo F.; Lappano R.; Sinicropi M. S. A Review on the Antimicrobial Activity of Schiff Bases: Data Collection and Recent Studies. Antibiotics 2022, 11 (2), 191. 10.3390/antibiotics11020191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Kasare M. S.; Dhavan P. P.; Shaikh A. H. I.; Jadhav B. L.; Pawar S. D. Novel Schiff Base Scaffolds Derived from 4-aminoantipyrine and 2-hydroxy-3-methoxy-5-(Phenyldiazenyl)Benzaldehyde: Synthesis, Antibacterial, Antioxidant and Anti-inflammatory. J. Mol. Recognit. 2022, 35 (9), e2976 10.1002/jmr.2976. [DOI] [PubMed] [Google Scholar]
  16. Teran R.; Guevara R.; Mora J.; Dobronski L.; Barreiro-Costa O.; Beske T.; Pérez-Barrera J.; Araya-Maturana R.; Rojas-Silva P.; Poveda A.; Heredia-Moya J. Characterization of Antimicrobial, Antioxidant, and Leishmanicidal Activities of Schiff Base Derivatives of 4-Aminoantipyrine. Molecules 2019, 24 (15), 2696. 10.3390/molecules24152696. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Çakmak R.; Başaran E.; Boğa M.; Erdoğan O. ¨.; Çınar E.; Çevik Ö. Schiff Base Derivatives of 4-Aminoantipyrine as Promising Molecules: Synthesis, Structural Characterization, and Biological Activities. Russ. J. Bioorganic Chem. 2022, 48 (2), 334–344. 10.1134/S1068162022020182. [DOI] [Google Scholar]
  18. Çakmak R.; Başaran E.; Şentürk M. Synthesis, Characterization, and Biological Evaluation of Some Novel Schiff Bases as Potential Metabolic Enzyme Inhibitors. Arch. Pharm. (Weinheim). 2022, 355 (4), 2100430. 10.1002/ardp.202100430. [DOI] [PubMed] [Google Scholar]
  19. Kobzar O. L.; Tatarchuk A. V.; Kachaeva M. V.; Pilyo S. G.; Sukhoveev O. V.; Sukhoveev V. V.; Brovarets V. S.; Vovk A. I. Azomethine Derivatives of P-Aminobenzoic Acid as Antioxidants and Xanthine Oxidase Inhibitors. Reports Natl. Acad. Sci. Ukr. 2020, (6), 74–82. 10.15407/dopovidi2020.06.074. [DOI] [Google Scholar]
  20. Mermer A.; Demirbas N.; Uslu H.; Demirbas A.; Ceylan S.; Sirin Y. Synthesis of Novel Schiff Bases Using Green Chemistry Techniques; Antimicrobial, Antioxidant, Antiurease Activity Screening and Molecular Docking Studies. J. Mol. Struct. 2019, 1181, 412–422. 10.1016/j.molstruc.2018.12.114. [DOI] [Google Scholar]
  21. Erturk A. G. Synthesis, Structural Identifications of Bioactive Two Novel Schiff Bases. J. Mol. Struct. 2020, 1202, 127299. 10.1016/j.molstruc.2019.127299. [DOI] [Google Scholar]
  22. Baluja S.; Chanda S. Synthesis, Characterization and Antibacterial Screening of Some Schiff Bases Derived from Pyrazole and 4-Amino Antipyrine. Rev. Colomb. Ciencias Químico-Farmacéuticas 2016, 45 (2), 201. 10.15446/rcciquifa.v45n2.59936. [DOI] [Google Scholar]
  23. Anupama B.; Aruna A.; Manga V.; Sivan S.; Sagar M. V.; Chandrashekar R. Synthesis, Spectral Characterization, DNA/ Protein Binding, DNA Cleavage, Cytotoxicity, Antioxidative and Molecular Docking Studies of Cu(II)Complexes Containing Schiff Base-Bpy/Phen Ligands. J. Fluoresc. 2017, 27 (3), 953–965. 10.1007/s10895-017-2030-5. [DOI] [PubMed] [Google Scholar]
  24. Gowda J. I.; Nandibewoor S. T. Binding and Conformational Changes of Human Serum Albumin upon Interaction with 4-Aminoantipyrine Studied by Spectroscopic Methods and Cyclic Voltammetry. Spectrochim. Acta Mol. Biomol. Spectrosc. 2014, 124, 397–403. 10.1016/j.saa.2014.01.028. [DOI] [PubMed] [Google Scholar]
  25. Shanuja K.; Pandiyan R. P.; Arun T.; Raman N.; Shanuja S. K.; Gnanamani A. Preliminary Investigation of DNA Interaction and Antimicrobial Efficacy of 4-Amino-2,3-Dimethyl-1-Phenyl-3-Pyrazolin-5-One Incorporated Metal Complexes. Int. J. Inorg. Bioinorg. Chem. 2016, 6 (1), 23–34. [Google Scholar]
  26. Paulpandiyan R.; Raman N. DNA Binding Propensity and Nuclease Efficacy of Biosensitive Schiff Base Complexes Containing Pyrazolone Moiety: Synthesis and Characterization. J. Mol. Struct. 2016, 1125, 374–382. 10.1016/j.molstruc.2016.07.003. [DOI] [Google Scholar]
  27. Ani F. E.; Ibeji C. U.; Obasi N. L.; Kelani M. T.; Ukogu K.; Tolufashe G. F.; Ogundare S. A.; Oyeneyin O. E.; Maguire G. E. M.; Kruger H. G. Crystal, Spectroscopic and Quantum Mechanics Studies of Schiff Bases Derived from 4-Nitrocinnamaldehyde. Sci. Rep. 2021, 11 (1), 8151. 10.1038/s41598-021-87370-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Arroudj S.; Bouchouit M.; Bouchouit K.; Bouraiou A.; Messaadia L.; Kulyk B.; Figa V.; Bouacida S.; Sofiani Z.; Taboukhat S. Synthesis, Spectral, Optical Properties and Theoretical Calculations on Schiff Bases Ligands Containing o-Tolidine. Opt. Mater. (Amst). 2016, 56, 116–120. 10.1016/j.optmat.2015.12.046. [DOI] [Google Scholar]
  29. Aguilar-Llanos E.; Carrera-Pacheco S. E.; González-Pastor R.; Zúñiga-Miranda J.; Rodríguez-Pólit C.; Romero-Benavides J. C.; Heredia-Moya J. Synthesis and Evaluation of Biological Activities of Schiff Base Derivatives of 4-Aminoantipyrine and Cinnamaldehydes. Chem. Proc. 2022, 12 (1), 43. 10.3390/ecsoc-26-13684. [DOI] [Google Scholar]
  30. Li Y.; Liu Y.; Wang H.; Xiong X.; Wei P.; Li F. Synthesis, Crystal Structure, Vibration Spectral, and DFT Studies of 4-Aminoantipyrine and Its Derivatives. Molecules 2013, 18 (1), 877–893. 10.3390/molecules18010877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Farrugia L. J. WinGX and ORTEP for Windows : An Update. J. Appl. Crystallogr. 2012, 45 (4), 849–854. 10.1107/S0021889812029111. [DOI] [Google Scholar]
  32. Elgrishi N.; Rountree K. J.; McCarthy B. D.; Rountree E. S.; Eisenhart T. T.; Dempsey J. L. A Practical Beginner’s Guide to Cyclic Voltammetry. J. Chem. Educ. 2018, 95 (2), 197–206. 10.1021/acs.jchemed.7b00361. [DOI] [Google Scholar]
  33. Sandford C.; Edwards M. A.; Klunder K. J.; Hickey D. P.; Li M.; Barman K.; Sigman M. S.; White H. S.; Minteer S. D. A Synthetic Chemist’s Guide to Electroanalytical Tools for Studying Reaction Mechanisms. Chem. Sci. 2019, 10 (26), 6404–6422. 10.1039/C9SC01545K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Gowda J. I.; Nandibewoor S. T. Electrochemical Behavior of 4-Aminophenazone Drug at a Graphite Pencil Electrode and Its Application in Real Samples. Ind. Eng. Chem. Res. 2012, 51 (49), 15936–15941. 10.1021/ie302501f. [DOI] [Google Scholar]
  35. Hamurcu F.; Mamaş S.; Ozdemir U. O.; Gündüzalp A. B.; Senturk O. S. New Series of Aromatic/ Five-Membered Heteroaromatic Butanesulfonyl Hydrazones as Potent Biological Agents: Synthesis, Physicochemical and Electronic Properties. J. Mol. Struct. 2016, 1118, 18–27. 10.1016/j.molstruc.2016.03.092. [DOI] [Google Scholar]
  36. Andres T.; Eckmann L.; Smith D. K. Voltammetry of Nitrobenzene with Cysteine and Other Acids in DMSO. Implications for the Biological Reactivity of Reduced Nitroaromatics with Thiols. Electrochim. Acta 2013, 92, 257–268. 10.1016/j.electacta.2013.01.047. [DOI] [Google Scholar]
  37. Obasi L. N.; Kaior G. U.; Rhyman L.; Alswaidan I. A.; Fun H.-K.; Ramasami P. Synthesis, Characterization, Antimicrobial Screening and Computational Studies of 4-[3-(4-Methoxy-Phenyl)-Allylideneamino]-1,5-Dimethyl-2-Phenyl-1,2-Dihydro-Pyrazol-3-One. J. Mol. Struct. 2016, 1120, 180–186. 10.1016/j.molstruc.2016.05.037. [DOI] [Google Scholar]
  38. Guimarães C.; Bento R. R. F.; Faria J. L. B.; Ramos R. J.; Silva L. E.; Freire P. T. C.; Gusmão G.; Teixeira A. M. R. Vibrational Spectra of (4E)-4-((E)-3-Phenyl-Allylideneamino)-1,2-Dihydro-2,3-Dimethyl-1-Phenylpirazol-5-One. J. Mol. Struct. 2011, 1006 (1–3), 589–595. 10.1016/j.molstruc.2011.10.009. [DOI] [Google Scholar]
  39. Eltayeb N. E.; Şen F.; Lasri J.; Hussien M. A.; Elsilk S. E.; Babgi B. A.; Gökce H.; Sert Y. Hirshfeld Surface Analysis, Spectroscopic, Biological Studies and Molecular Docking of (4E)-4-((Naphthalen-2-Yl)Methyleneamino)-1,2-Dihydro-2,3-Dimethyl-1-Phenylpyrazol-5-One. J. Mol. Struct. 2020, 1202 (xxxx), 127315. 10.1016/j.molstruc.2019.127315. [DOI] [Google Scholar]
  40. El-Nahass M. M.; Kamel M. A.; El-Menyawy E. M. Spectroscopic Analysis of 2-(2,3-Dihydro-1,5-Dimethyl-3-Oxo-2-Phenyl-1H-Pyrazol-4-Ylimino)-2-(4-Nitro-Phenyl) Acetonitrile. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2011, 79 (3), 618–624. 10.1016/j.saa.2011.03.044. [DOI] [PubMed] [Google Scholar]
  41. Roriz B. C.; Buccini D. F.; Dos Santos B. F.; Silva S. R. d. S.; Domingues N. L. d. C.; Moreno S. E. Synthesis and Biological Activities of a Nitro-Shiff Base Compound as a Potential Anti-Inflammatory Agent. Eur. J. Pharm. Sci. 2020, 148, 105300. 10.1016/j.ejps.2020.105300. [DOI] [PubMed] [Google Scholar]
  42. Aguzue O. C.; Adedayo A.; Phillip O. A. Mechanochemical Synthesis and Potentiation of the Antimicrobial Activity of 4-[3-(4-Methoxyphenyl)-Allylideneamino]-1,5-Dimethyl-2-Phenylpyrazol-3-One by Metal Chelation. J. Sci. Math. Lett. 2020, 8 (2), 15–21. 10.37134/jsml.vol8.2.3.2020. [DOI] [Google Scholar]
  43. Römling U.; Balsalobre C. Biofilm Infections, Their Resilience to Therapy and Innovative Treatment Strategies. J. Int. Med. 2012, 272 (6), 541–561. 10.1111/joim.12004. [DOI] [PubMed] [Google Scholar]
  44. Olsen I. Biofilm-Specific Antibiotic Tolerance and Resistance. Eur. J. Clin. Microbiol. Infect. Dis. 2015, 34 (5), 877–886. 10.1007/s10096-015-2323-z. [DOI] [PubMed] [Google Scholar]
  45. de Lima Pimenta A.; Chiaradia-Delatorre L. D.; Mascarello A.; de Oliveira K. A.; Leal P. C.; Yunes R. A.; de Aguiar C. B. N. M.; Tasca C. I.; Nunes R. J.; Smânia A. Synthetic Organic Compounds with Potential for Bacterial Biofilm Inhibition, a Path for the Identification of Compounds Interfering with Quorum Sensing. Int. J. Antimicrob. Agents 2013, 42 (6), 519–523. 10.1016/j.ijantimicag.2013.07.006. [DOI] [PubMed] [Google Scholar]
  46. Pandolfi F.; D’Acierno F.; Bortolami M.; De Vita D.; Gallo F.; De Meo A.; Di Santo R.; Costi R.; Simonetti G.; Scipione L. Searching for New Agents Active against Candida Albicans Biofilm: A Series of Indole Derivatives, Design, Synthesis and Biological Evaluation. Eur. J. Med. Chem. 2019, 165, 93–106. 10.1016/j.ejmech.2019.01.012. [DOI] [PubMed] [Google Scholar]
  47. Park J. S.; Ryu E.-J.; Li L.; Choi B.-K.; Kim B. M. New Bicyclic Brominated Furanones as Potent Autoinducer-2 Quorum-Sensing Inhibitors against Bacterial Biofilm Formation. Eur. J. Med. Chem. 2017, 137, 76–87. 10.1016/j.ejmech.2017.05.037. [DOI] [PubMed] [Google Scholar]
  48. Vashistha A.; Sharma N.; Nanaji Y.; Kumar D.; Singh G.; Barnwal R. P.; Yadav A. K. Quorum Sensing Inhibitors as Therapeutics: Bacterial Biofilm Inhibition. Bioorg. Chem. 2023, 136, 106551. 10.1016/j.bioorg.2023.106551. [DOI] [PubMed] [Google Scholar]
  49. CrysAlisPro Software System; Rigaku Corporation: Oxford, UK, 2015.
  50. Sheldrick G. M. SHELXT - Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71 (1), 3–8. 10.1107/S2053273314026370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Sheldrick G. M. A Short History of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2008, 64 (1), 112–122. 10.1107/S0108767307043930. [DOI] [PubMed] [Google Scholar]
  52. Spackman P. R.; Turner M. J.; Mckinnon J. J.; Wolff S. K.; Grimwood D. J.; Jayatilaka D.; Spackman M. A. CrystalExplorer: a program for Hirshfeld surface analysis, visualization and quantitative analysis of molecular crystals. J. Appl. Crystallogr. 2021, 54, 1006–1011. 10.1107/S1600576721002910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Balouiri M.; Sadiki M.; Ibnsouda S. K. Methods for in Vitro Evaluating Antimicrobial Activity: A Review. J. Pharm. Anal. 2016, 6 (2), 71–79. 10.1016/j.jpha.2015.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao3c05372_si_001.pdf (4.5MB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES