Next Article in Journal
Green Analytical Method Using Single-Drop Microextraction Followed by Gas Chromatography for Nitro Compound Detection in Environmental Water and Forensic Rinse Water
Previous Article in Journal
Antioxidant and Anti-Inflammatory Activities of Astilboides tabularis (Hemsl.) Engl. Root Extract
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Cu(II), Cu(I) and Ag(I) Complexes of Phenoxy-Ketimine Schiff Base Ligands: Synthesis, Structures and Antibacterial Activity

by
Miriam Caviglia
1,
Zhenzhen Li
1,2,
Carlo Santini
1,
Jo’ Del Gobbo
1,
Cristina Cimarelli
1,
Miao Du
2,*,
Alessandro Dolmella
3,* and
Maura Pellei
1,*
1
School of Science and Technology, Chemistry Division, University of Camerino, Via Madonna Delle Carceri (ChIP), 62032 Camerino, Italy
2
College of Material and Chemical Engineering, Institute of New Energy Science and Technology, Zhengzhou University of Light Industry, Zhengzhou 450001, China
3
Department of Pharmaceutical and Pharmacological Sciences, University of Padova, Via Marzolo 5, 35131 Padova, Italy
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(9), 1893; https://doi.org/10.3390/molecules30091893
Submission received: 3 April 2025 / Revised: 17 April 2025 / Accepted: 22 April 2025 / Published: 24 April 2025

Abstract

:
Two phenoxy-ketimines ligands, 2-(1-(benzylimino)ethyl)phenol (HLBSMe) and 2-((benzylimino)(phenyl)methyl)phenol (HLBSPh), were synthesized and used as supporting ligands of new copper(II), copper(I), and silver(I) complexes. In order to confer different solubility properties to the metal complexes and to stabilize Cu and Ag in their +1 oxidation state, the lipophilic triphenylphosphine (PPh3) and the hydrophilic 1,3,5-triaza-7-phosphaadamantane (PTA) were selected as co-ligands in the syntheses of the Cu(I) and Ag(I) complexes. All compounds were characterized by CHN analysis, NMR, FT-IR spectroscopy, and electrospray ionization mass spectrometry (ESI-MS); the molecular structure of the copper(II) complex [Cu(LBSPh)2] was also determined by single-crystal X-ray diffraction. Finally, the antibacterial activity of the metal complexes, the Schiff base ligands and phosphane co-ligands, were assessed by determining the minimum inhibitory concentration (MIC) and minimum bactericidal concentration (MBC) against Gram-negative (Escherichia coli) and Gram-positive bacteria (Staphylococcus aureus).

1. Introduction

In recent years, the phenomenon of antibiotic resistance in hospitals, communities, and the environment has increasingly grown. Antibacterial resistance is a major global public health challenge, associated with an estimated 4.95 million deaths in 2019 [1,2]. Due to these developments, the World Health Organization (WHO) published the first list of drug-resistant bacterial priority pathogens classified as critical in 2017 [3]. Building on the 2017 edition, in 2024, the WHO published an updated list of drug-resistant bacteria posing the greatest threat to human health, aiming to guide the development of new treatments and strategies to prevent and mitigate the spread of antimicrobial resistance (AMR) [2]. Despite the current work, the global antibiotic pipeline is marked by limited innovation and limited global access to both new and existing treatments. This increase in antimicrobial resistance and the misuse of antimicrobials have led to the need to develop new antimicrobial compounds [4,5,6,7,8].
Medicinal inorganic chemistry offers possibilities for the design of therapeutic agents not readily available to organic compounds [7,9,10]. A successful strategy in antimicrobial chemotherapy has been the use of metallo-drugs, and this strategy has the potential to be used for treating multidrug-resistant infections more efficiently [4,11,12]. As a class of molecules, Schiff bases have been a topic of considerable interest, owing to their versatile metal chelating properties, and have been used to coordinate almost all metal ions [13,14,15,16,17,18,19,20,21,22,23,24]. Schiff bases and their metal complexes have been reported to exhibit a wide range of biological activities [15,25,26,27,28,29,30,31,32,33,34], including anticancer [35,36] and antimicrobial properties [37,38,39,40,41,42,43,44,45]. In most cases, the metallic derivatives of Schiff bases were found to exhibit higher antimicrobial activities than their parent ligands [27,46].
Numerous Schiff bases and their copper(II) derivatives have been reported to possess promising catalytic [47,48], anticancer [49,50,51,52], or biological properties [41,45,53,54,55,56,57,58,59,60,61] and have been successfully used as models in biological systems [62]. In particular, several Cu(II) complexes of N,O-donor Schiff base ligands were found to exhibit favorable antimicrobial activities [57,63,64,65,66,67,68,69,70,71].
Despite the enormous amount of work devoted to the synthesis and characterization of copper(II) Schiff bases complexes [65], there are relatively few reports devoted to the corresponding Cu(I) complexes, perhaps due to their tendency to undergo disproportionation to copper metal and copper(II) compounds in the absence of stabilizing co-ligands [72,73,74]. Schiff bases as chelating N-donors and phosphanes as P-donors were recently employed for the preparation of photoactive neutral [75,76] or cationic heteroleptic Cu(I) complexes [77,78,79]. On the other hand, in view of the wide spectra of antimicrobial activities against bacteria and fungi showed by silver(I) [80] and related to metal complexes [81,82,83], plenty of studies focused on the antimicrobial properties of silver(I) complexes with Schiff bases [84,85]. Recently, the Schiff bases Cu(I) and Ag(I) complexes incorporating phosphanes as co-ligands have awakened a new interest towards their biological properties [86].
Among the various types of Schiff bases [87,88,89,90,91,92,93,94,95,96], the phenoxy-imines and phenoxy-ketimines are of particular interest. These 2-imidoylphenols can exist in tautomeric forms due to intramolecular H-bonding between the hydroxyl H-atom and the N-atom (Scheme 1) [97,98,99].
Depending on the extent of the interaction, complete transfer of the proton from the hydroxyl group to the nitrogen atom may occur, giving conversion of the molecular form (benzenoid, enol-type) into the quinoid, keto-type tautomer [98,100]. Zwitterionic species characterized by an intramolecular charge-separation are possible intermediates in the tautomerization process [99]. The replacement of the hydrogen atom by the methyl, ethyl, or phenyl substituents in the C-C(H)=N moiety in phenoxy-imines Schiff bases shortens the intramolecular hydrogen bonds in related phenoxy-ketimines [101].
In recent years, metal complexes of N,O-bidentate phenoxy-ketimines Schiff base ligands have attracted much attention from researchers [47,48,102,103,104,105,106,107,108,109,110]. However, to our knowledge, copper(I)- and silver(I)-based complexes supported by N,O-bidentate phenoxy-ketimines, including phenoxy-imines Schiff base ligands, remain an unexplored research field.
Therefore, as part of our continuous investigation on the chemical and biological properties of copper- and silver-containing coordination compounds [111,112,113,114,115,116,117], we report here the syntheses, characterization, and biological evaluation of new Cu(II), Cu(I), and Ag(I) complexes containing phosphanes and the phenoxy-ketimines ligands, 2-(1-(benzylimino)ethyl)phenol (HLBSMe) (Oletimol), and 2-((benzylimino)(phenyl)methyl)phenol (HLBSPh) (Scheme 2 and Scheme 3). The ligands were selected to modify the steric and electronic properties of the resulting metal complexes using phenyl and methyl groups on the C-C(R)=N imino moiety, respectively. In designing the novel phenoxy-ketimines metal complexes, the lipophilic triphenylphosphine (PPh3) and the hydrophilic 1,3,5-triaza-7-phosphaadamantane (PTA) were selected as co-ligands, in order to stabilize copper and silver in their +1 oxidation state and to confer different solubility properties to the corresponding metal complexes.
The antimicrobial properties of the new copper and silver complexes as well as of the corresponding uncoordinated ligands and co-ligands were evaluated against Gram-negative (Escherichia coli) and Gram-positive (Staphylococcus aureus) bacteria. The antimicrobial data have been compared with the control antibiotic ciprofloxacin to assess their relative efficacy [66].

2. Results and Discussion

2.1. Synthesis and Characterization

The phenoxy-ketimines ligands HLBSMe and HLBSPh were prepared using a modified literature method [118,119,120], by the condensation reaction of the 2-acylphenols 2′−1-(2-hydroxyphenyl)ethan-1-one or (2-hydroxyphenyl)(phenyl)methanone, respectively, with stoichiometric amounts of benzylamine in methanol, and isolated as yellow solids in very good yield and high purity (Scheme 2).
The ligands were fully characterized by 1H-NMR (Table S1), 13C-NMR, FT-IR spectroscopy, and ESI-MS spectrometry. Both 1,3,5-triaza-phosphaadamantane copper(I) complexes [Cu(HLBSMe)(PTA)2]PF6 (3) and [Cu(HLBSPh)(PTA)2]PF6·2H2O (7) were synthesized in CH3CN using, as starting materials, the ligand HLBSMe or HLBSPh, the metal acceptor [Cu(CH3CN)4]PF6, and the PTA co-ligands in the stoichiometric ratio 1:1:2, respectively (Scheme 3). Analogously, triphenylphosphine copper(I) complex [Cu(HLBSPh)(PPh3)2]PF6·2CH3CN (6) was synthesized in CH3CN using, as starting material, the ligand HLBSPh, the metal acceptor [Cu(CH3CN)4]PF6, and the PPh3 co-ligands in the stoichiometric ratio 1:1:2 (Scheme 3).
Elemental analyses and spectroscopic studies such as FT-IR, 1H-, 31P-NMR, and ESI-MS confirm the stoichiometry of the synthesized complexes 3, 6, and 7. All the expected absorption bands were observed in the FT-IR spectra. The most characteristic feature of the complexation is the removal of the O⋯H⋯N hydrogen bond [101,121] and the presence of a very broad ν(OH) absorption band at about 3200 cm−1. The complexes exhibit weak bands in the range of 2878–3070 cm−1 typical of C-H stretching. The strong absorptions at 1571–1618 cm−1 are due to the stretching of the aromatic and imine bonds (C=C and C=N), while very intense absorptions in the range of 831–832 cm−1 are attributable to the stretching of the PF6 counterion. The 1H-NMR spectra, recorded in CD3CN for complexes 3, 6, and 7, confirm the stoichiometric ratio between the Schiff base ligand and phosphane co-ligands. They showed a single set of resonances for the skeletal ligands, indicating that the Schiff base ligand protons are equivalent. The signals of phenolic OH protons appeared as broad peaks in the range of 15.41–16.24 ppm, the signal of the CH2Ph protons is visible at δ 4.55–4.83 ppm, while the signal of the ketimine methyl CH3C=N proton for complex 3 appeared at δ 2.47 ppm, with a slight shift with respect to the signal of the free ligand due to the coordination to the copper acceptor. The aromatic hydrogens of ligand and triphenylphosphine co-ligands for [Cu(HLBSPh)(PPh3)2]PF6·2CH3CN (6) are detectable in the range of 6.71–7.69 ppm, while in the spectra of [Cu(HLBSMe)(PTA)2]PF6 (3) and [Cu(HLBSPh)(PTA)2]PF6·2H2O (7) recorded in CD3CN, the NCH2P protons of PTA co-ligands are visible as singlets at δ 4.06–4.09 ppm and the related NCH2N protons show characteristic AB quartets in the range of 4.52–4.60 ppm.
The 31P-NMR spectrum of [Cu(HLBSPh)(PPh3)2]PF6·2CH3CN (6) performed in CD3CN gives a broad singlet peak at −0.30 ppm, downfield shifted with respect to the value of the free triphenylphosphine in the same solvent (δ = −4.85 ppm). Likewise, a broad singlet is visible in the spectra of the PTA complexes 3 and 7, recorded in CD3CN, at −91.64 and −89.66 ppm, downfield shifted with respect to the value of the free PTA in CD3CN (δ −102.07 ppm). Moreover, the spectra of complexes 3, 6, and 7 show the distinctive septets due to the presence of the PF6 counterion at about −144.60 ppm with JF−P = 706 Hz. 1H-NMR and 31P{1H}-NMR data of complexes 3, 6, and 7 are summarized in Table S1. The ESI-MS study was performed by dissolving complexes 3, 6, and 7 in acetonitrile and recording the spectra in ion-positive and ion-negative modes. In the ESI-MS(+) spectrum of 3, the major peak at m/z 445 can be attributed to the species [Cu(HLBSMe)(PTA)]+. Similarly, the ESI-MS(+) spectrum of 7 showed at m/z 507 the peak due to the species [Cu(HLBSPh)(PTA)]+, confirming the formation and the stability of the complexes.
The copper(II) complexes [Cu(LBSMe)2] (5) and [Cu(LBSPh)2] (10) were synthesized using the copper(II) acetate salt Cu(CH3COO)2 and the ligands HLBSMe and HLBSPh, respectively, in the stoichiometric ratio 1:2 (Scheme 3). The reaction was performed in methanol at reflux for the synthesis of complex 5 and in acetonitrile at room temperature for the synthesis of 10. Isolated complexes are air-stable, soluble in CH3OH, CH3CN, CH2Cl2, CHCl3, and DMSO, and they have been characterized by means of FT-IR spectroscopy and ESI-MS spectrometry. Successful complexation was indicated by a slight shift of the imine bonds (ν(C=N)) in the IR spectra. In particular, the imine and aromatic bands in complexes 5 and 10 are observed at 1588–1597 cm−1 and 1569–1599 cm−1, respectively, whereas in the corresponding ligand precursors HLBSMe and HLBSPh, the bands appear at 1572–1610 and 1563–1604 cm−1, respectively. The ESI-MS spectra of complexes 5 and 10 were performed in CH3CN solution. In the ESI-MS(+) of complex 5, peaks at m/z 287 and 329 are due to the species [Cu(LBSMe)]+ and [Cu(LBSMe) + CH3CN]+, respectively; the major peaks at m/z 511 and 576, respectively, are due to the adducts [Cu(LBSMe)2 + H]+ and [Cu(LBSMe)2 + Na + CH3CN]+, respectively, confirming the presence of two ligands that coordinate copper. For complex 10, signals at m/z 349 and 390 are attributable to the fragments [Cu(LBSPh)]+ and [Cu(LBSPh) + CH3CN]+.
The silver complexes [Ag(HLBSMe)(PTA)]NO3 (4) and [Ag(HLBSPh)(PTA)]NO3 (9) were synthesized by a treatment of AgNO3 with one equivalent of 1,3,5-triaza-phosphadaamantane (PTA) and one equivalent of the corresponding HLBSMe or HLBSPh ligands in methanol and acetonitrile solution, respectively. Likewise, the triphenylphosphine silver(I) complex [Ag(HLBSPh)(PPh3)2]NO3 (8) was synthesized in CH3CN using as starting material the ligand HLBSPh, the metal acceptor silver nitrate, and the PPh3 co-ligands in the stoichiometric ratio 1:1:2. Isolated complexes are air-stable and have been characterized by means of 1H-, 31P-NMR and FT-IR spectroscopy, as well as ESI-MS spectrometry. In IR spectra, a slight shift of the imine bonds (ν(C=N)) was observed, indicating complexation. Complexes 4, 8, and 9 exhibit a series of bands in the 1479–1292 cm−1 region of the IR spectrum that can be attributed to the stretching modes of the NO3 group [122,123]. These bands are consistent with those previously reported for analogous silver(I) phosphane complexes [124,125,126]. However, the multiplicity of the observed bands prevents an unambiguous assignment of the nitrate group’s coordination mode—whether monodentate or bidentate—to the metal center.
The 1H-NMR spectra of 4, 8, and 9 in DMSO-d6 showed a set of resonances in agreement with the proposed structure. The slightly shifted resonance at δ 15.45–16.38 ppm corresponds to the OH phenolic proton. In the spectra 4 and 9, the NCH2P protons of PTA co-ligands are visible as singlets at δ 4.15–4.16 ppm and the related NCH2N protons show characteristic AB quartets in the range of 4.40–4.58 ppm. In the spectrum of complex 8, the aromatic hydrogens of the ligand and triphenylphosphine co-ligands are detectable in the range 6.71–7.63 ppm. In the 31P-NMR spectra of the PTA complexes 4 and 9 recorded in CDCl3, a singlet is visible at −93.60 and −94.43 ppm, downfield shifted with respect to the value of the free PTA (δ −102.50 ppm). Broad signals are also in the spectra recorded in CD3OD and CD3CN at 233K, indicating the impossibility to stop or slow the phosphane exchange process, to show doublets in which the coupling of 31P to the 107Ag and 109Ag are resolved. In the 31P-NMR spectrum of the triphenylphosphine complex 8 recorded in CDCl3 at room temperature, a very broad singlet is visible at 14.68 ppm, downfield shifted with respect to the value of the free PPh3 (δ −5.34 ppm). In the spectrum of 8 recorded in CD3CN at 233K, a doublet was observed. The related J(Ag-31P) coupling constant (478 Hz) is of the same order of magnitude of those reported for analogous silver(I) bis(triphenylphosphine) species [111,127,128,129,130]. 1H-NMR and 31P{1H}-NMR data of complexes 4, 8, and 9 are summarized in Table S1. The ESI-MS study was performed by dissolving complexes 4, 8, and 9 in methanol or acetonitrile and recording the spectra in ion-positive and ion-negative modes. In the ESI-MS(+) spectrum of 4 and 9, the peaks at m/z 491 and 553 can be attributed to the species [Ag(HLBSMe)(PTA)]+ and [Ag(HLBSPh)(PTA)]+, respectively, confirming the formation and the stability of the complexes.

2.2. X-Ray Crystallography

Despite several efforts, we were able to get crystalline samples of only one of the aforementioned compounds. In particular, the structure of the [Cu(LBSPh)2] complex (10) was determined by single-crystal diffraction analysis on an item extracted from a batch of very good, dark green prismatic specimens, obtained by a slow recrystallization of the compound from a toluene solution. The investigation revealed that the neutral complex crystallizes in the monoclinic system; the solid-state structure was solved in the centrosymmetric I2/a space group (No. 15). The asymmetric unit (Figure 1) consists of one half of the molecule, that is, the copper ion, and a single molecule of the anionic 2-(benzylimino(phenyl)methyl)phenolate (LBSPh) ligand, which chelates the central Cu atom through the O, N atoms. The phenyl ring of the benzylimino residue of the ligand (C15/C20) is disordered over two possible arrangements; the alternate positions of disordered atoms have been named with a final ‘A’.
Upon chelation, the ligand and the metal form a six-membered metallacycle (Cu1, O1, C1, C6, C7, N1), showing an envelope conformation, with the Cu1 atom ‘at the flap’ and departing by 0.55 Å from the mean plane (p1) of the other five atoms, that are instead coplanar within 0.05 Å, given the sp2 nature of the C6-C7-N1 atoms of the ligand. In the chelating LBSPh molecule, the mean plane of the phenol ring (C1/C6) is almost coplanar with p1 (angle of 1.39°); instead, the mean planes passing through the phenyl rings of the benzyl moiety (C15/C20 or C15A/C20A, respectively, p2 and p2′) or emerging from the methine bridge (C8/C13, p3) are approximately orthogonal to the p1 plane, making with the latter angles of 79.1° (79.6° for p2′), and 75.2°, respectively. The p3 and p2 planes are also roughly orthogonal with respect to each other (angles of 69.9° and 72.2° for p2 and p2′, respectively).
The copper coordination environment is completed by a second LBSPh molecule (Figure 2), in which both the O1 and N1 atoms are trans-positioned to their symmetry-generated counterparts. The [Cu(LBSPh)2] complex (10) has a calix shape when observed through the O2N2 coordination plane and the four coordinating atoms around the central copper atom define a moderately distorted square planar environment (τ4 and τ4′ indexes of 0.21 and 0.18, respectively) [131,132], with O1-Cu1-O1* and N1-Cu1-N1* (* = 3/2 − x, y, 1 − z) angles of 160.77(9)° and 170.07(7)°, respectively—the first quite far from the ideal 180° value. Apparently, the departure is due to the engagement of the O1 atom in an intermolecular non-canonical hydrogen bond with the H18 atom of a nearby phenyl ring (see further below).
With respect to bond distances and angles, the 2-(benzylimino(phenyl)methyl)phenolate ligand does not show unusual metric values when compared to existing data. An investigation in the CCDC database [133] for molecules showing the benzyl(diphenylmethylene)amine moiety returned about 50 entries. A comparison of bond distances next to the C=N link with those found in this work showed that the object with the best overall similarity, that is, with the smaller sum of deviations for examined distances, is the (already reported) ligand itself [121], besides few other compounds [134,135]. Interestingly, molecules in which the C=N distance is close to the one reported here (1.299(2) Å) are those in which a phenolic moiety was also present [121,136,137,138]. The C=N bond participates in some degree of conjugation with the nearby phenyl rings, in known compounds as well as in the present complex. In fact, the C6-C7 and C7-C8 bonds are a little shorter than the ideal 1.54 Å (1.4654(19) and 1.5009(19) Å, compared with mean values of 1.494 and 1.492 Å for known molecules); the same is true for the N1-C14, C14-C15 distances in the benzyl residue (1.4787(17), 1.483(10) Å against corresponding mean values of 1.450 and 1.517 Å in reported compounds). These data suggest that bond conjugation in the ligand is little affected by coordination.
A second search in the CCDC repository was made to seek transition metal complexes showing a ligand similar to LBSPh. The search returned a very limited number of Mn, Co, Ni, and Cu compounds showing a tetradentate dimeric form of the LBSPh ligand, all prepared by the same research group [139,140,141,142,143]; to the best of our knowledge, then, 10 is the first Cu(II) complex of the LBSPh ligand reported to date. In known molecules, the O and N donor atoms are always trans to each other, opposite to what happens in the [Cu(LBSPh)2] complex. The Metal-O, Metal-N distances in these molecules fall in the ranges 1.82–1.91 Å, 1.85–2.03 Å, respectively, with differences basically due to the ionic radii of the elements. In fact, the mean values for the Cu-O, Cu-N distances in the Cu(II) known complexes [140] (1.879 and 1.949 Å, respectively) are in very good/good agreement with the 1.8839(11) and 1.9773(12) Å found here. A complete listing of bond lengths and angles for complex 10 is provided in the supporting information (Tables S2 and S3).
The examination of relevant nonbonding interactions (metric details listed in Table 1), mostly hydrophobic contacts, shows that the complex units are well packed, and that the unit cell does not show solvent accessible voids. Looking at the asymmetric unit (one half of the complex), just one intermolecular hydrogen bond connects H18 (or H18A) with the O1 atom of a nearby molecule at 3/2 − x, −1 + y, 1 − z (at 2.81/2.85 Å, respectively); this interaction also shows in the symmetry-generated half of the molecule at 1/2 − x, y, 1 − z (Figure 3; further packing diagrams highlighting the other contacts described here can be found in Figures S34–S36 of supporting information). Incidentally, the two symmetry-related halves of the complex have their C15/C20 (or C15A/C20A) rings in roughly eclipsed positions, so they are also coupled by somehow weak π····π interactions, with distances between planes encompassing the alternate arrangements of the rings of 4.20, 4.03 Å, respectively, (compared with 3.35 Å in graphite).
The H18····O1 contact creates a one-dimensional motif which propagates along the crystallographic b axis, and it is coupled to a second orthogonal chain due to the C-H····π contacts made by C1/C2 atoms and the H12 atom of another unit at 1 − x, 1 − y, 1 − z (Figure S36). The two chains cross-link each other and create a 2D grid in the ab plane (Figure S37). Finally, a third chain is formed by a second, T-shaped C-H····π contact involving the C15/C20 (or C15A/C20A) ring with the H11 atom of a nearby molecule at 1 − x, −1/2 + y, 1/2 − z (Figure S38; only C15····H11 approach shown). The H11 atom in fact points almost straight towards the C15/C20 ring centroid, with a contact length of 2.78 Å (also towards the C15A/C20A ring) and C11-H11-ring centroid angle of 168.8° and 165.0° for C15/C20, C15A/C20A rings, respectively. This chain propagates along the [011] plane and diagonally intersects the above described 2D motif, thus sealing a full 3D contact network.

2.3. In-Vitro Antibacterial Activity

First, the resazurin-based microtiter dilution assay (RMDA) method was used to evaluate the synergic antibacterial effect of Schiff bases (1 and 2), metal complexes (310), and phosphane co-ligands (11 and 12). Figure 4 and Figure 5 show the bacterial growth curves against Gram-negative (E. coli) and Gram-positive bacteria (S. aureus), respectively. The minimum inhibitory concentrations (MIC) were reported in Table 2 and half-maximal inhibitory concentrations (MIC50) values were calculated using computerized nonlinear regression analysis.
Generally, the growth of bacteria follows four phases: lag phase, logarithmic phase, stationary phase, and decline phase. When bacteria are exposed to external interferences or inhibitory factors, their growth may be significantly delayed or even completely suppressed, ultimately leading to growth cessation.
Based on the bactericidal growth curves, the free Schiff base ligands and phosphane co-ligands exhibited negligible toxicity against both Gram-negative and Gram-positive bacterial species, while all the metal complexes demonstrated different antibacterial activities. Especially, the silver complexes [Ag(HLBSMe)(PTA)]NO3 (4) and [Ag(HLBSPh)(PTA)]NO3 (9) showed a significant bactericidal effect. The MIC values of compound 4 against E. coli and S. aureus were 0.050 and 0.100 mg mL−1, respectively, while the MIC values of compound 9 were 0.025 and 0.050 mg mL−1, respectively. Interestingly, the MIC values of compound 4 were higher in comparison to the values of 9; however, the bacterial growth curves indicated that the growth rate of bacteria treated with compound 4 was slower when the concentrations were below 0.025 mg mL−1. Their different bacterial growth behaviors can lie in the structure of the Schiff base ligands with the presence of a methyl or a phenyl substituent in compounds 4 and 9, respectively. Previous studies have suggested that the bactericidal activity of silver compounds is primarily due to two distinct mechanisms, including the strong affinity to R-SH groups associated in amino acids or proteins and their ability to generate reactive oxygen species (ROS). Silver ions can readily bind to SH groups, such as those in cysteine residues, which can disrupt cellular metabolism and physiology by directly interfering with enzymatic functions or by destabilizing disulfide (S-S) bonds critical for maintaining protein structure. However, effective binding to SH groups requires silver ions to be sufficiently accessible and reactive [144,145].
The bactericidal activities of the compounds at different concentrations were further evaluated using the plate colony count method against E. coli (Figure 6) and S. aureus (Figure 7). The minimum bactericidal concentration (MBC) and half-maximal bactericidal concentration (MBC50) were calculated and included in Table 3.
The results demonstrate that the antimicrobial activities of the metal complexes 39 are significantly enhanced compared to the free Schiff base ligands, which can be explained by Tweedy’s chelation theory. Due to the electron donating properties of Schiff base ligands, a metal ion can share part of its charge through orbital overlap with Schiff base ligands, which reduces the polarity of the metal ion. Furthermore, this process increases the lipophilicity of the central metal atom, facilitating its penetration through the lipid layer of the microorganism, and blocking the interaction of enzymes on the cell membrane, thereby inhibiting the growth of the microorganism [146,147]. Subsequently, Cu(I) and Ag(I) coordination compounds were investigated to study the influence of the nature of a metal center in bioactivity. A comparison of the antimicrobial properties of compounds 3 and 4, 6 and 8, as well as 7 and 9 reveals that silver complexes exhibit superior antimicrobial effects against both E. coli and S. aureus strains compared to their copper-based counterparts. Among these metal complexes, compound 9 was found to be most active against the bacterial strains with MBC values of 0.100 and 0.200 mg mL−1 for E. coli and S. aureus, respectively. Additionally, compound 4 was prominently active with MBC values of 0.200 mg mL−1 against both tested bacterial strains. Compound 8 displayed moderate inhibitory effects with MBC50 values of 0.322 and 0.294 mg mL−1, however lower than the MBC50 values of 0.566 and 1.125 mg mL−1 for compound 6. As previously mentioned, Ag(I) ions have the ability to alter the bacteria cell wall structure and cell constituent by irreversibly binding to cysteine residues, thereby disrupting essential enzyme systems and ultimately inhibiting bacterial proliferation [144,148].
Furthermore, a possible correlation between the biological activity of the compounds and their lipophilicity and hydrophilicity was discussed. By comparing compounds 6 and 7, as well as 8 and 9, it was observed that the antibacterial activity of the complexes with PTA as a co-ligand was superior to those with PPh3 as a co-ligand. Notably, the lipophilic triphenylphosphine and the hydrophilic 1,3,5-triaza-7-phosphaadamantane have been chosen as co-ligands because they can stabilize copper and silver in +1 oxidation state and can give also a different hydrophilic–lipophilic balance to the corresponding complexes. The antibacterial data indicate that the PTA co-ligands confer the related silver complexes the most appropriate balance between hydrophilicity and lipophilicity [149,150,151,152], what is in fact a major challenge in the development of antimicrobial agents. The presence of lipophilic groups enhances penetration ability to the bacterial lipophilic outer membrane which contains murein, thereby facilitating interactions with the bacterial replicative system [144,145]. However, several studies have also highlighted the importance of water solubility as a key parameter in antibacterial activity.
The antibacterial performance of the synthesized compounds was systematically evaluated with respect to three critical determinants: (i) ligands chelation effects on bioavailability, (ii) properties of metal ions, and (iii) hydrophobic-lipophilic balance. Compared to the broad-spectrum antibiotic ciprofloxacin (Figure 8), although the synthesized metal complexes are less effective than the standard drug, compounds 4 and 9 have demonstrated significant potential as antibacterial agents. In addition, compared with other silver Schiff base metal complexes reported for antibacterial applications, the compounds 4 and 9 exhibits similar or lower MIC values [65,66,85,144,153,154,155,156].

3. Experimental Section

3.1. Materials and Instruments

All reagents were obtained from commercial suppliers and used as received. Melting Points (MP) were performed by an SMP3 Stuart Scientific Instrument (Bibby Sterilin Ltd., London, UK). Elemental analyses (C, H, N, S) (EA) were performed with a Fisons Instruments EA-1108 CHNS-O Elemental Analyzer (Thermo Fisher Scientific Inc., Waltham, MA, USA). Fourier-Transform InfraRed (FT-IR) spectra were recorded from 4000 to 700 cm−1 on a PerkinElmer Frontier Instrument (PerkinElmer Inc., Waltham, MA, USA), equipped with an Attenuated Total Reflection (ATR) unit using universal diamond top-plate as a sample holder. Abbreviations used in the analyses of the FT-IR spectra: br = broad, m = medium, mbr = medium broad, s = strong, sbr = strong broad, vs. = very strong, w = weak, wbr = weak broad. Nuclear Magnetic Resonance (NMR) spectra for the nuclei 1H, 13C and 31P were recorded with a Bruker 500 Ascend Spectrometer (Bruker BioSpin Corporation, Billerica, MA, USA; 500.13 MHz for 1H, 125.78 MHz for 13C, 202.46 MHz for 31P and 470.59 MHz for 19F). Tetramethylsilane (SiMe4) was used as external standard for the 1H- and 13C-NMR spectra, while 85% H3PO4 was used for the 31P-NMR spectra. The chemical shifts (δ) are reported in ppm, and coupling constants (J) are reported in hertz (Hz). Abbreviations used in the analyses of the NMR spectra: br = broad, d = doublet, dbr = broad doublet, m = multiplet, s = singlet, sbr = broad singlet, t = triplet. ElectroSpray Ionization Mass Spectra (ESI-MS) were recorded in positive- (ESI-MS(+)) or negative-ions (ESI-MS(−)) mode on a Waters Micromass ZQ Spectrometer equipped with a single quadrupole (Waters Corporation, Milford, MA, USA), using methanol or an acetonitrile mobile phase. The compounds were added to reagent grade methanol or acetonitrile to give approximately 0.1 mM solutions. These solutions were injected (1 µL) into the spectrometer fitted with an autosampler. The pump delivered the solutions to the mass spectrometer source at a flow rate of 200 μL/min and nitrogen was employed both as a drying and nebulizing gas. The capillary voltage was typically 2500 V. The temperature of the source was 100 °C, while the temperature of the desolvation was 400 °C. In the analyses of ESI-MS spectra, the confirmation of major peaks was supported by comparison of the observed and predicted isotope distribution patterns, the latter calculated using the IsoPro 3.1 computer software (T-Tech Inc., Norcross, GA, USA).

3.2. Synthesis

3.2.1. Synthesis of the Ligand HLBSMe (1)

The ligand (E)-2-(1-(benzylimino)ethyl)phenol, HLBSMe (1) was prepared by the reaction of 1-(2-hydroxyphenyl)ethan-1-one (1.100 mmol, 0.150 g) and benzylamine (1.000 mmol, 0.107 g) in methanol (30 mL). The mixture was stirred at room temperature for 4 h, monitored via TLC, until the complete consumption of acylphenol. The mixture was evaporated at reduced pressure and dissolved in CH2Cl2 (25 mL) and 1M solution of NaOH (4 × 8 mL). After extraction, the organic phase was anhydrified with Na2SO4, evaporated and dried at reduced pressure, giving the product HLBSMe in 80% yield. Solubility: CH3OH, CH3CH2OH, CH3CN, CHCl3, Acetone, EtOAc, DMSO. M.p.: 118–119 °C. FT-IR (cm−1): 3211wbr, 3108vw, 3061w, 3034w, 2918w, 2877w, 2814w (C-H); 1610s, 1572m (C=C and C=N); 1492m, 1451s, 1413m, 1376m, 1350m, 1330m, 1303s, 1263m, 1232m, 1181w, 1161m, 1129m, 1074w, 1055m, 1024m, 974m, 950m, 941m, 928m, 900m, 860m, 831m, 743vs, 733vs, 698vs, 639m, 584m, 559m, 525m, 501m. 1H-NMR (CDCl3, 293K): δ 2.51 (s, 3H, CH3), 4.87 (s, 2H, CH2), 6.83–7.61 (m, 9H, CHar), 15.78 (sbr, 1H, OH). 1H-NMR (CD3CN, 293K): δ 2.48 (s, 3H, CH3), 4.83 (s, 2H, CH2), 6.83–7.71 (m, 9H, CHar), 16.23 (sbr, 1H, OH). 1H-NMR (Acetone-d6, 293K): δ 2.52 (s, 3H, CH3), 4.88 (s, 2H, CH2), 6.80–7.72 (m, 9H, CHar), 15.87 (sbr, 1H, OH). 1H-NMR (DMSO-d6, 293K): δ 2.51 (s, 3H, CH3), 4.83 (s, 2H, CH2), 6.78–7.72 (m, 9H, CHar), 16.37 (sbr, 1H, OH). 13C{1H}-NMR (CDCl3, 293K): δ 14.7 (CH3), 53.4 (CH2), 117.2, 118.8, 119.5, 127.2, 127.5, 128.1, 128.8, 132.6, 138.5 (CHAr), 163.8 (COH), 172.2 (C=N). ESI-MS(+) (major positive ions, CH3CN), m/z (%): 226 (100) [LBSMe + H]+. Elemental analysis (%) calculated for C15H15NO: C 79.97, H 6.71, N 6.22; found: C 80.25, H 6.54, N 6.36.

3.2.2. Synthesis of the Ligand HLBSPh (2)

The ligand (E)-2-((benzylimino)(phenyl)methyl)phenol, HLBSPh (2) was prepared following the procedure described for ligand 1, using (2-hydroxyphenyl)(phenyl)methanone (1.100 mmol, 0.218 g) and was obtained in 72% yield. Solubility: CH3OH, CH3CH2OH, CH3CN, CHCl3, Acetone, EtOAc, DMSO. M.p.: 89–91 °C. FT-IR (cm−1): 30,653w, 3026w, 2988w, 2966w, 2922w, 2858w (C-H); 1604s, 1563m (C=C and C=N); 1498m, 1446m, 1378m, 1355w, 1350w, 1299m, 1266m, 1227m, 1203m, 1165m, 1153m, 1126m, 1094m, 1075m, 1049m, 1028m, 993m, 948m, 912m, 856w, 829m, 798m, 752vs, 711m, 697s, 676m, 627m, 588m. 1H-NMR (CDCl3, 293K): δ 4.59 (s, 2H, CH2), 6.68–7.56 (m, 14H, CHar), 15.64 (sbr, 1H, OH). 1H-NMR (CD3CN, 293K): δ 4.55 (s, 2H, CH2), 6.70–7.60 (m, 14H, CHar), 15.42 (sbr, 1H, OH). 13C{1H}-NMR (CD3CN, 293K): δ 55.3 (CH2), 117.5, 117.6, 120.0, 127.1, 127.4, 127.6, 128.6, 128.8, 129.2, 131.6, 132.5, 133.9, 139.4 (CHAr), 162.9 (COH), 174.8 (C=N). ESI-MS(+) (major positive ions, CH3CN), m/z (%): 288 (100) [LBSPh + H]+. Elemental analysis (%) calculated for C20H17NO: C 83.59, H 5.96, N 4.87; found: C 83.42, H 5.73, N 4.92.

3.2.3. Synthesis of [Cu(HLBSMe)(PTA)2]PF6 (3)

Tetrakis(acetonitrile)copper(I)hexafluorophosphate (0.500 mmol, 0.186 g) and 1,3,5-triaza-7-phosphadaamantane (1.000 mmol, 0.157 g) were dissolved in CH3CN (30 mL) and the reaction was stirred for 2 h at room temperature. Then, the ligand HLBSMe (0.500 mmol, 0.113 g) was added, and the reaction was stirred for 24 h at room temperature. The precipitate obtained was filtered and dried at reduced pressure giving the complex [Cu(HLBSMe)(PTA)2]PF6 in 66% yield. M.p.:191–193 °C. Solubility: CH3CN, DMSO. FT-IR (cm−1): 3200vbr (OH); 3062w, 3034w, 3268wbr, 2919wbr, 2879w (C-H); 1618m, 1574w (C=C and C=N); 1496w, 1449m, 1415w, 1376w, 1350w, 1294m, 1239m, 1161w, 1130w, 1107w, 1055w, 1040s, 969s, 949s, 895w; 831vs (PF6); 742vs, 733s, 698s, 639m, 583s, 556vs, 525m, 501m. 1H-NMR (CD3CN, 293K): δ 2.47 (s, 3H, CH3), 4.09 (s, 12H, NCH2P), 4.50–4.61 (AB q, 12H, NCH2N), 4.83 (s, 2H, CH2), 6.82–7.71 (m, 9H, CHar), 16.25 (sbr, 1H, OH). 31P{1H}-NMR (CD3CN, 293K): δ −144.62 (septet, JP−F = 706 Hz), −91.64 (sbr). ESI-MS(+) (major positive ions, CH3CN), m/z (%): 226 (100) [HLBSMe + H]+, 287 (20) [Cu(HLBSMe)]+, 377 (10) [Cu(PTA)2]+, 445 (100) [Cu(HLBSMe)(PTA)]+. ESI-MS(−) (major negative ions, CH3CN), m/z (%): 145 (100) [PF6]. Elemental analysis (%) calculated for C27H39CuF6N7OP3: C 43.35, H 5.25, N 13.11; found: C 44.14, H 5.39, N 13.05

3.2.4. Synthesis of [Ag(HLBSMe)(PTA)]NO3 (4)

Silver nitrate (0.500 mmol, 0.085 g) and 1,3,5-triaza-phosphadaamantane (0.500 mmol, 0.078 g) were dissolved in CH3OH (30 mL) and the reaction was stirred for 2 h at room temperature. Afterwards, the ligand HLBSMe (0.500 mmol, 0.113 g) was added, the reaction was stirred at room temperature for 24 h, then the solution was evaporated under reduced pressure. The product [Ag(HLBSMe)(PTA)]NO3 was obtained in 58% yield. M.p.: 206–208 °C. Solubility: CH3OH, CH2Cl2, CHCl3, CH3CN, DMSO. FT-IR (cm−1): 3210br (OH); 3060w, 3033w, 2917w, 2878w (C-H); 1616m, 1573w (C=C and C=N); 1494w; 1449m, 1425m, 1375m, 1334s, 1292s (NO3); 1240s, 1161m, 1129w, 1107m, 1074w, 1042w, 1013vs, 970vs, 947vs, 929m, 900m, 861w, 829w, 806s, 790sh, 743vs, 732vs, 698vs, 639m, 602s, 583s, 560s, 525m, 501m. 1H-NMR (DMSO-d6, 293K): δ 2.48 (s, 3H, CH3), 4.15 (s, 6H, NCH2P), 4.41–4.58 (AB q, 6H, NCH2N), 4.83 (s, 2H, CH2), 6.77–7.72 (m, 9H, CHar), 16.38 (s, 1H, OH). 31P{1H}-NMR (DMSO-d6, 293K): δ −86.03 (s). 31P{1H}-NMR (CDCl3, 293K): δ −93.60 (s). 31P{1H}-NMR (CD3OD, 233K): δ −82.00 (sbr). ESI-MS(+) (major positive ions, CH3OH), m/z (%): 226 (100) [HLBSMe + H]+, 248 (20) [HLBSMe + Na]+, 423 (50) [Ag(PTA)2]+, 491 (40) [Ag(HLBSMe)(PTA)]+. ESI-MS(−) (major negative ions, CH3OH), m/z (%): 145 (100) [AgNO3 + NO3]. Elemental Analysis (%) calculated for C21H27AgN5O4P: C 45.67, H 4.93, N 12.68; found: C 46.15, H 5.12, N 13.31.

3.2.5. Synthesis of [Cu(LBSMe)2] (5)

Copper(II) acetate (0.250 mmol, 0.045 g) and the ligand HLBSMe (0.500 mmol, 0.113 g) were solubilized in CH3OH (20 mL). The reaction was stirred at reflux for 4 h, giving a small amount of a dark green precipitate. After filtration, the mother liquors were dried at reduced pressure obtaining the green–blue solid [Cu(LBSMe)2] in 39% yield. M.p.:201–203 °C. Solubility: CH3OH, CH3CN, CH2Cl2, CHCl3, DMSO. FT-IR (cm−1): 3083w, 3058w, 3021w, 2974w, 2904w (C-H); 1597s, 1588s (C=C and C=N); 1536s; 1494m, 1469m, 1437s, 1366w, 1328s, 1262m, 1246m, 1215m, 1159m, 1132m, 1084w, 1055m, 1046m, 933m, 856m, 756s, 742s, 705s, 695s, 649m, 602m, 570m. ESI-MS(+) (major positive ions, CH3CN), m/z (%): 226 (100) [HLBSMe + H]+, 287 (40) [Cu(LBSMe)]+, 329 (30) [Cu(LBSMe) + CH3CN]+, 511 (90) [Cu(LBSMe)2 + H]+, 576 (80) [Cu(LBSMe)2 + Na + CH3CN]+. ESI-MS(−) (major negative ions, CH3CN), m/z (%): 240 (100) [LBSMe]. Elemental analysis (%) calculated for C30H28CuN2O2: C 70.36, H 5.51, N 5.47; found: C 69.79, H 5.41, N 5.25.

3.2.6. Synthesis of [Cu(HLBSPh)(PPh3)2]PF6·2CH3CN (6)

Tetrakis(acetonitrile)copper(I)hexafluorophosphate (0.500 mmol, 0.186 g) and triphenylphosphine (1.000 mmol, 0.262 g) were dissolved in CH3CN (30 mL) and the reaction was stirred for 3 h at room temperature. The ligand HLBSPh (0.500 mmol, 0.144 g) was added, and the reaction was stirred for 24 h at room temperature. The solvent was removed at reduced pressure and the solid obtained was washed with Et2O. A precipitate was recovered by filtration giving the complex [Cu(HLBSPh)2(PPh3)]PF6·2CH3CN in 83% yield. M.p.:161–164 °C Solubility: CH3OH, CH3CN, CHCl3, CH2Cl2, DMSO. FT-IR (cm−1): 3070w, 3050w, 3028w, 2938w, 2916w (C-H); 2301wbr, 2271wbr (CH3CN); 1605m, 1572w (C=C and C=N); 1497w, 1480m, 1450w, 1434s, 1369w, 1346w, 1328w, 1304m, 1258m, 1245w, 1181w, 1150w, 1112w, 1095m, 1072w, 1047w, 1035w, 999w, 972w, 923w, 879w, 859msh; 832vs (PF6); 775m, 766m, 744vs, 716m, 694vs, 646m, 618w, 557vs, 541m, 527m, 516s, 504vs. 1H-NMR (CD3CN 293K): δ 1.99 (s, 6H, CH3CN), 4.55 (s, 2H, CH2), 6.71–7.69 (m, 44H, CHar), 15.41 (s, 1H, OH). 1H-NMR (CDCl3 293K): δ 2.08 (s, 6H, CH3CN), 4.61 (s, 2H, CH2), 6.70–7.71 (m, 44H, CHar). 31P{1H}-NMR (CD3CN, 293K): δ −144.62 (septet, JP−F = 706 Hz), −0.30 (sbr). 31P{1H}-NMR (CDCl3, 293K): δ −144.25 (septet, JP−F = 713 Hz), −0.61 (s). ESI-MS(+) (major positive ions, CH3CN), m/z (%): 366 (100) [Cu(PPh3) + CH3CN]+, 589 (100) [Cu(PPh3)2]+. ESI-MS(−) (major negative ions, CH3CN), m/z (%): 145 (100) [PF6]. Elemental Analysis (%) calculated for C60H53CuF6N3OP3: C 65.36, H 4.85, N 3.81; found: C 65.25, H 4.88, N 3.66.

3.2.7. Synthesis of [Cu(HLBSPh)(PTA)2]PF6·2H2O (7)

Tetrakis(acetonitrile)copper(I)hexafluorophosphate (0.500 mmol, 0.186 g) and 1,3,5-triaza-7-phosphadaamantane (1.000 mmol, 0.157 g) were dissolved in CH3CN (30 mL) and the reaction was stirred for 2 h at room temperature. The ligand HLBSPh (0.500 mmol, 0.144 g) was added giving a yellow mixture. After 24 h at room temperature, the solvent was evaporated under reduced pressure obtaining the complex [Cu(HLBSPh)(PTA)2]PF6·2H2O in 72% yield. M.p.: 225–228 °C. Solubility: CH3OH, CH3CH2OH, CH2Cl2, CHCl3, CH3CN, DMSO. FT-IR (cm−1): 3345wbr, 3188wbr (OH); 3063wbr, 3028w, 2916w, 2878w (C-H); 1605s, 1571m (C=C and C=N); 1497m, 1441m, 1419m; C=N); 1346w, 1328w, 1303m, 1257m, 1242s, 1171w, 1150m, 1111m, 1081w, 1045w, 1015s, 970s, 949s, 923m, 876w; 832vs (PF6); 775s, 765s, 751vs, 715vs, 698vs, 646s, 611m, 582s, 555vs, 506s. 1H-NMR (CD3CN, 293K): δ 2.24 (s, 4H, H2O), 4.08 (s, 12H, NCH2N), 4.50–4.60 (AB q, 14H, NCH2P and CH2), 6.70–7.62 (m, 14H, CHar), 15.45 (s, 1H, OH). 31P{1H}-NMR (CD3CN, 293K): δ −144.58 (septet, JP−F = 706 Hz), −89.66 (br). ESI-MS(+) (major positive ions, CH3CN), m/z (%): 261 (90) [Cu(PTA) + CH3CN]+, 288 (20) [HLBSPh + H]+, 377 (10) [Cu(PTA)2]+, 507 (15) [Cu(HLBSPh)(PTA)]+. ESI-MS(−) (major negative ions, CH3CN), m/z (%): 145 (100) [PF6]. Elemental Analysis (%) calculated for C32H45CuF6N7O3P3: C 45.42, H 5.36, N 11.59; found: C 44.69, H 5.03, N 12.24.

3.2.8. Synthesis of [Ag(HLBSPh)(PPh3)2]NO3 (8)

Silver nitrate (0.500 mmol, 0.085 g) and triphenylphosphine (1.000 mmol, 0.262 g) were dissolved in CH3CN (30 mL) and the reaction was stirred for 2 h at room temperature. Afterwards, the ligand HLBSPh (0.500 mmol, 0.144 g) was added, and the reaction was stirred at room temperature for 24 h. The precipitate obtained was filtered and dried under reduced pressure giving the product [Ag(HLBSPh)(PPh3)2]NO3 in 40% yield. M.p.: 187–189 °C. Solubility: CH3OH, CH2Cl2, CHCl3, CH3CN. FT-IR (cm−1): 3059wbr, 3028w, 2918w (C-H); 1605m, 1571w (C=C and C=N); 1497w, 1489w; 1479m, 1436m, 1397s, 1346w, 1328w, 1296s (NO3); 1257m, 1179w, 1150m, 1112w, 1095m, 1072w, 1047m, 1025m, 997m, 971w, 918s, 864w, 849w, 833w, 825w, 818w, 775m, 766m, 751vs, 742vs, 715s, 693vs, 646m, 618w, 567w, 547w, 513ssh, 503vs. 1H-NMR (CDCl3, 293K): δ 4.58 (s, 2H, CH2), 6.68–7.70 (m, 44H, CHar), 15.67 (s, 1H, OH). 1H-NMR (DMSO-d6, 293K): δ 4.52 (s, 2H, CH2), 6.71–7.63 (m, 44H, CHar), 15.46 (s, 1H, OH). 31P{1H}-NMR (CDCl3, 293K): δ 14.68 (sbr). 31P{1H}-NMR (CD3CN, 233K): δ 8.69 (d, J(Ag-31P) = 478 Hz). ESI-MS(+) (major positive ions, CH3CN), m/z (%): 288 (10) [HLBSPh + H]+, 371 (30) [Ag(PPh3)]+, 410 (30) [Ag(PPh3) + CH3CN]+, 630 (100) [Ag(PPh3)2]+. ESI-MS(−) (major negative ions, CH3OH), m/z (%): 231 (100) [AgNO3 + NO3]. Elemental analysis (%) calculated for C56H47AgN2O4P2: C 68.51, H 4.83, N 2.85; found: C 69.43, H 4.87, N 2.90.

3.2.9. Synthesis of [Ag(HLBSPh)(PTA)]NO3 (9)

Silver nitrate (0.500 mmol, 0.085 g) and 1,3,5-triaza-7-phosphadaamantane (0.500 mmol, 0.078 g) were dissolved in CH3CN. After 2 h, the ligand HLBSPh (0.500 mmol, 0.145 g) was added and the reaction was stirred at room temperature for 22 h and refluxed for further 4 h. Successively, the solvent was removed under reduced pressure and the yellow solid product [Ag(HLBSPh)(PTA)]NO3·was obtained in 55% yield. M.p.: 232–234 °C. Solubility: CH3OH, CH3CN, CHCl3, CH2Cl2, DMSO. FT-IR (cm−1): 3200wbr (OH); 3064wbr, 3029w, 2916w (C-H); 1605s, 1571m (C=C and C=N); 1516w, 1497w, 1489w; 1450m, 1441m, 1425m, 1416m, 1374m, 1329s, 1294s (NO3); 1257s, 1241s, 1170w, 1150m, 1110m, 1081w, 1072w, 1045m, 1013s, 970vs, 947vs, 919s, 865w, 851m, 825m, 775s, 766s, 751vs, 715vs, 698vs, 646s, 605s, 582s, 564s, 547m, 506s. 1H-NMR (DMSO-d6, 293K): δ 4.16 (s, 6H, NCH2N), 4.42–4.58 (dd, 6H, NCH2P and CH2), 6.71–7.63 (m, 14H, CHar), 15.45 (sbr, 1H, OH). 31P{1H}-NMR (CDCl3, 293K): δ −94.43 (s). 31P{1H}-NMR (CD3CN, 233K): δ −86.09 (sbr). ESI-MS(+) (major positive ions, CH3CN), m/z (%): 158 (90) [PTA + H]+, 260 (80) [Ag(PTA)]+, 288 (100) [HLBSPh + H]+, 423 (50) [Ag(PTA)2]+, 553 (10) [Ag(HLBSPh)(PTA)]+. ESI-MS(−) (major negative ions, CH3CN), m/z (%): 231 (70) [AgNO3 + NO3]. Elemental analysis (%) calculated for C26H27AgN5O4P: C 50.83, H 4.76, N 11.50; found: C 51.27, H 5.12, N 11.91.

3.2.10. Synthesis of [Cu(LBSPh)2] (10)

Copper(II) acetate (1.000 mmol, 0.200 g) was dissolved in CH3OH obtaining an intense blue solution. At the same time, the ligand HLBSPh (2.000 mmol, 0.115 g) was solubilized in CH3CN. The two solutions were mixed, and it was stirred at room temperature for 1 h and then refluxed for 23 h. After 24 h, the dark green precipitate formed was filtered and dried under reduced pressure giving the complex [Cu(LBSPh)2] in 77% yield. A batch of good quality crystals of [Cu(LBSPh)2], suitable for X-ray analysis, was obtained by slow evaporation of a toluene solution of 10. M.p.: 228–230 °C. Solubility: CH3OH, CH3CH2OH, CH2Cl2, CH3Cl, EtOAc, CH3CN, DMSO, Acetone. FT-IR (cm−1): 3080w, 3044w, 3025w, 2940w (C-H); 1599s, 1568s (C=C and C=N); 1529vs; 1491m, 1456m, 1440vs, 1341vs, 1257s, 1241s, 1203w, 1178w, 1142vs, 1120m, 1075m, 1051m, 1039w, 1024m, 998w, 961m, 935s, 914m, 847s, 777m, 765m, 747vs, 720s, 702vs, 691vs, 641w, 629m, 618m. ESI-MS(+) (major positive ions, CH3CN), m/z (%): 288 (100) [HLBSPh + H]+, 349 (70) [Cu(LBSPh)]+, 390 (100) [Cu(LBSPh) + CH3CN]+, 636 (20) [2HLBSPh + H]+, 658 (20) [2HLBSPh + Na]+, 741 (25) [2HLBSPh + Na + 2CH3CN]+. Elemental analysis (%) calculated for C40H32CuN2O5: C 75.51, H 5.07, N 4.40; found: C 75.38, H 5.12, N 4.41.

3.3. Crystallographic Data Collection and Refinement

A batch of very good crystals were obtained by slow recrystallization of the [Cu(LBSPh)2] complex from a toluene solution. The specimens were of prismatic shape and intensely colored in green. The selected item was fixed to a glass capillary with Loctite glue and mounted on the head of a four-circle, kappa-geometry single-crystal OD Rigaku diffractometer. The data collection was performed under a graphite-monochromated Cu Kα radiation (λ = 1.54184 Å) at room temperature [295.1(3) K], and reflections were collected with the ω–scans technique. The preliminary screening of the sample allowed us to plan the data collection in the 1024 × 1024 pixel mode and 2 × 2 pixel binning. Intensity spots were recorded with an Atlas CCD area detector recently obtained by Prof. K Rissanen (University of Jyväskylä, Finland) in replacement of the former EOS camera. The crude diffraction data were collected working with the CrysAlisPro software, Version 1.171.43.141a [157], and treated for Lorentz and polarization effects. An empirical correction for absorption based on a multi-scan approach using equivalent reflections was applied by means of the SCALE3 ABSPACK algorithm, also accessible via the CrysAlisPro software. Two reference frames were collected every 50 frames to ensure crystal and equipment stability, showing no sign of sample deterioration or motion.
The complex crystallizes in the monoclinic system and the structure was solved in the I2/a space group; the unit cell parameters have been refined by working on the least-squares refinement of 18,133 strong reflections gathered during the whole experiment. The structure solution was performed by direct phasing and refined by full-matrix least-squares methods based on Fo2 through the OLEX2 program interface [158] with the SHELXT and SHELXL [159,160] programs. The asymmetric unit contains half of a molecule, with the copper ion lying on a proper two-fold axis, and the whole cell hosts four units of the complex, which are very efficiently packed (no solvent accessible voids). In the final stages of the refinement, non-H atoms were allowed to vibrate anisotropically, whereas the H atoms were placed in calculated positions and refined as a riding model, with their displacement parameters calculated as 1.2 times the Ueq of the ‘parent’ atom.
At this level, it turned out that the phenyl ring (C15/C20) belonging to the benzylimino moiety of the ligand can assume two possible arrangements (involved atoms labelled with a final ‘A’). The sum of the pertinent occupation factors for each disordered position were constrained to unity, with finally refined sofs of 0.57/0.43 for untagged and tagged atoms, respectively. A regular hexagon geometry with bond distances of 1.3900 Å was imposed on the disordered positions of the phenyl ring via the AFIX 6 constraint to ensure refinement convergence and to prevent some atoms from assuming non-positive definite displacement parameters. A summary of crystal and data collection parameters is given below in Table 4; full listings of bond lengths and angles are in Tables S2 and S3. Representations of the asymmetric unit of [Cu(LBSPh)2] and of the entire complex (Figure 1 and Figure 2) with and without the disordered rings, highlighting the selected numbering scheme, have been prepared with the Mercury 4.2.0 software [161]. Complete listings of atomic coordinates, bond lengths, bond angles, and anisotropic thermal parameters are available as supporting information, in the form of the .cif file, that has been deposited at the Cambridge Crystallographic Data Center (CCDC) with deposition number 2433689. These data can be obtained free of charge via www.ccdc.cam.ac.uk/structures (accessed on 21 April 2025).

3.4. Antibacterial Screening

The synthesized Schiff bases HLBSMe (1) and HLBSPh (2), the related metal complexes 310, and the free phosphane co-ligands triphenylphosphine (11) and, 3,5-triaza-7-phosphadaamantane (12), were evaluated for their in-vitro antibacterial activity against Gram-negative (Escherichia coli China-bio-00021) and Gram-positive bacteria (Staphylococcus aureus ATCC 25923), with ciprofloxacin serving as standard antibacterial agent. The antimicrobial properties of the compounds were evaluated using the standard broth microdilution method in 96-well cell culture plates following the reported method [162]. Firstly, the compounds were dissolved in PBS buffer containing 1% DMSO solution and stirred for 20 min until forming the homogeneous suspension (1.0 mg mL−1). Subsequently, a series of compounds solutions (50 µL) were prepared in a 96-well plate using the double dilution method, followed by the addition of bacterial suspensions. The final compound concentrations were 3.125, 6.25, 12.5, 25, 50, 100, and 200 μg mL−1, and the final bacterial concentration was 5 × 106 CFU mL−1. The group without an antimicrobial agent was set as the control. Then, the 96-well plate was incubated at 37 °C for 24 h and the optical density (OD600) values of each well was recorded at different times (0, 2, 4, 6, 8, 10, 12, 20, and 24 h) using a microplate reader (Thermo Scientific, Waltham, MA, USA). According to the principle that the turbidity of the culture increases with the proliferation of bacteria, the growth state of bacteria can be evaluated by detecting the absorbance value of the bacterial suspension at 600 nm. In addition, the synthesized compounds were evaluated for their in-vitro antibacterial activity against two bacterial strains using the plate count method. Typically, 100 µL of bacterial suspension (106 CFU mL−1) was mixed with 100 µL of compound solution at different concentrations (25, 50, 100, 200 and 400 μg mL−1). Subsequently, 100 µL of the mixed solution was spread on LB agar plates. After incubation at 37 °C for another 24 h, the surviving bacteria on each plate were counted. The antibacterial ratio was calculated using the following formula, where Cb and Ce are the numbers of colonies in the blank group and the experimental group, respectively, as follows:
Antibacterial   efficiency   =   C b C e C b × 100 %

Statistical Analysis

The collected data were analyzed using GraphPad Prism 10 or OriginPro 2021 (9.8) software. Statistical significance was defined as p < 0.05, and the results were presented to three decimal places.

4. Conclusions

In this study, we report the synthesis, characterization, and antibacterial evaluation of Cu(I) and Ag(I) complexes supported by two phenoxy-ketimines ligands. In particular, 2-(1-(benzylimino)ethyl)phenol (HLBSMe) and 2-((benzylimino)(phenyl)methyl)phenol (HLBSPh), characterized by the presence of methyl or phenyl substituents on the imine group, were employed for the preparation of the complexes 24 and 69, using the lipophilic PPh3 and the hydrophilic PTA as co-ligands to stabilize the metal in +1 oxidation state. The analogous copper(II) complexes 5 and 10 were synthesized using the copper(II) acetate salt Cu(CH3COO)2 and the ligands HLBSMe and HLBSPh, respectively. All species were fully characterized both in the solid state and in solution. The single-crystal XRD allowed us to describe the molecular structure of 10, the first reported complex of Cu(II) with the HLBSPh ligand, where the copper ion has a slightly distorted square planar environment, as well as to highlight the efficient crystal packing, even in the absence of strong canonical hydrogen bonds, largely driven by π····π and C-H····π nonbonding interactions. The antimicrobial properties of the new copper and silver complexes as well as of the corresponding uncoordinated ligands were evaluated against E. coli and S. aureus. The results revealed that all the complexes exhibited moderate to excellent antibacterial efficacy. Among them, the Ag(I) complexes displayed the highest antibacterial activity. Specifically, [Ag(HLBSPh)(PTA)]NO3 demonstrated the most promising antibacterial performance, with MIC values of 0.025 and 0.05 mg mL−1 against E. coli and S. aureus, respectively, while the corresponding MBC values were 0.100 and 0.200 mg mL−1.
This study provides new insights into the design of antimicrobial agents, highlighting the potential for rationally designing and synthesizing novel, highly efficient antibacterial compounds. To further enhance the antibacterial activity of the silver complexes of phenoxy-ketimine Schiff base ligands, future research should focus on optimizing the synthesis of the most promising candidates, including fine-tuning the electronic and steric effects of both primary and auxiliary ligands, as well as modifying the solubility properties of the related complexes. In addition, detailed mechanistic studies could reveal specific pathways and interactions, facilitating the design of more effective compounds. Future investigations should also prioritize an in-depth exploration of structure-activity relationships to elucidate the role of distinctive phosphanes and of different substituents on the phenoxy-ketimines ligands in biological activity.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30091893/s1, Figures S1–S35: FT-IR, 1H-, 13C-, and 31P-NMR spectra of compounds 110; Figures S36–S38: packing diagrams highlighting nonbonding interaction network for complex 10; Table S1: 1H- and 31P{1H}-NMR selected data; Tables S2 and S3: full listings of bond lengths and angles of the complex 10.

Author Contributions

Conceptualization, C.S. and M.P.; data curation, M.C., Z.L., J.D.G. and A.D.; formal analysis, M.C., Z.L., J.D.G. and A.D.; investigation, M.C., Z.L., C.C. and M.P.; methodology, M.C., Z.L., J.D.G., A.D. and M.P.; supervision, C.C., M.D. and M.P.; writing—original draft, Z.L., C.S., M.D., A.D. and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Unione Europea—NextGenerationEU (MUR-Fondo Promozione e Sviluppo—D.M. 737/2021, INVIRCuM, University of Camerino, FAR 2022 PNR, and NGEU PNRR, D.M. n. 351/2022 M4C1 I4.1) and by the University of Padova (PRID BIRD225980).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors. CCDC 2433689 contains the supplementary crystallographic data for this paper, available free of charge from the Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/structures (accessed on 21 April 2025).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hofer, U. The cost of antimicrobial resistance. Nat. Rev. Microbiol. 2019, 17, 3. [Google Scholar] [CrossRef] [PubMed]
  2. World Health Organization. WHO Bacterial Priority Pathogens List, 2024: Bacterial Pathogens of Public Health Importance to Guide Research, Development and Strategies to Prevent and Control Antimicrobial Resistance. Available online: https://www.who.int/publications/i/item/9789240093461 (accessed on 1 November 2024).
  3. World Health Organization. WHO Publishes List of Bacteria for Which New Antibiotics Are Urgently Needed. Available online: https://www.who.int/en/news-room/detail/27-02-2017-who-publishes-list-of-bacteria-for-which-new-antibiotics-are-urgently-needed (accessed on 1 November 2024).
  4. Frei, A.; Verderosa, A.D.; Elliott, A.G.; Zuegg, J.; Blaskovich, M.A.T. Metals to combat antimicrobial resistance. Nat. Rev. Chem. 2023, 7, 202–224. [Google Scholar] [CrossRef] [PubMed]
  5. Ruggieri, F.; Compagne, N.; Antraygues, K.; Eveque, M.; Flipo, M.; Willand, N. Antibiotics with novel mode of action as new weapons to fight antimicrobial resistance. Eur. J. Med. Chem. 2023, 256, 115413. [Google Scholar] [CrossRef] [PubMed]
  6. Miethke, M.; Pieroni, M.; Weber, T.; Brönstrup, M.; Hammann, P.; Halby, L.; Arimondo, P.B.; Glaser, P.; Aigle, B.; Bode, H.B.; et al. Towards the sustainable discovery and development of new antibiotics. Nat. Rev. Chem. 2021, 5, 726–749. [Google Scholar] [CrossRef]
  7. Lin, Y.; Betts, H.; Keller, S.; Cariou, K.; Gasser, G. Recent developments of metal-based compounds against fungal pathogens. Chem. Soc. Rev. 2021, 50, 10346–10402. [Google Scholar] [CrossRef]
  8. Collignon, P.; Beggs, J.J.; Walsh, T.R.; Gandra, S.; Laxminarayan, R. Anthropological and socioeconomic factors contributing to global antimicrobial resistance: A univariate and multivariable analysis. Lancet Planetary Health 2018, 2, E398–E405. [Google Scholar] [CrossRef]
  9. Barry, N.P.E.; Sadler, P.J. Exploration of the medical periodic table: Towards new targets. Chem. Commun. 2013, 49, 5106–5131. [Google Scholar] [CrossRef]
  10. Mjos, K.D.; Orvig, C. Metallodrugs in Medicinal Inorganic Chemistry. Chem. Rev. 2014, 114, 4540–4563. [Google Scholar] [CrossRef]
  11. Cvijan, B.B.; Jacic, J.K.; Bajcetic, M. The Impact of Copper Ions on the Activity of Antibiotic Drugs. Molecules 2023, 28, 5133. [Google Scholar] [CrossRef]
  12. Frei, A.; Zuegg, J.; Elliott, A.G.; Baker, M.; Braese, S.; Brown, C.; Chen, F.; Dowson, C.G.; Dujardin, G.; Jung, N.; et al. Metal complexes as a promising source for new antibiotics. Chem. Sci. 2020, 11, 2627–2639. [Google Scholar] [CrossRef]
  13. Ghanghas, P.; Choudhary, A.; Kumar, D.; Poonia, K. Coordination metal complexes with Schiff bases: Useful pharmacophores with comprehensive biological applications. Inorg. Chem. Commun. 2021, 130, 108710. [Google Scholar] [CrossRef]
  14. Alkis, M.E.; Kelestemür, Ü.; Alan, Y.; Turan, N.; Buldurun, K. Cobalt and ruthenium complexes with pyrimidine based Schiff base: Synthesis, characterization, anticancer activities and electrochemotherapy efficiency. J. Mol. Struct. 2021, 1226, 129402. [Google Scholar] [CrossRef]
  15. More, M.S.; Joshi, P.G.; Mishra, Y.K.; Khanna, P.K. Metal complexes driven from Schiff bases and semicarbazones for biomedical and allied applications: A review. Mater. Today Chem. 2019, 14, 100195. [Google Scholar] [CrossRef] [PubMed]
  16. Magyari, J.; Holló, B.B.; Vojinovic-Jesic, L.S.; Radanovic, M.M.; Armakovic, S.; Armakovic, S.J.; Molnár, J.; Kincses, A.; Gajdács, M.; Spengler, G.; et al. Interactions of Schiff base compounds and their coordination complexes with the drug cisplatin. New J. Chem. 2018, 42, 5834–5843. [Google Scholar] [CrossRef]
  17. Al Zoubi, W.; Al-Hamdani, A.A.S.; Kaseem, M. Synthesis and antioxidant activities of Schiff bases and their complexes: A review. Appl. Organomet. Chem. 2016, 30, 810–817. [Google Scholar] [CrossRef]
  18. Al Zoubi, W.; Ko, Y.G. Organometallic complexes of Schiff bases: Recent progress in oxidation catalysis. J. Organomet. Chem. 2016, 822, 173–188. [Google Scholar] [CrossRef]
  19. Jia, Y.; Li, J.B. Molecular Assembly of Schiff Base Interactions: Construction and Application. Chem. Rev. 2015, 115, 1597–1621. [Google Scholar] [CrossRef]
  20. Qin, W.L.; Long, S.; Panunzio, M.; Biondi, S. Schiff Bases: A Short Survey on an Evergreen Chemistry Tool. Molecules 2013, 18, 12264–12289. [Google Scholar] [CrossRef]
  21. Gupta, K.C.; Sutar, A.K. Catalytic activities of Schiff base transition metal complexes. Coord. Chem. Rev. 2008, 252, 1420–1450. [Google Scholar] [CrossRef]
  22. Nakayama, Y.; Saito, J.; Bando, H.; Fujita, T. Propylene Polymerization Behavior of Fluorinated Bis(phenoxy-imine) Ti Complexes with an MgCl2-Based Compound (MgCl2-Supported Ti-Based Catalysts). Macromol. Chem. Phys. 2005, 206, 1847–1852. [Google Scholar] [CrossRef]
  23. Vigato, P.A.; Tamburini, S. The challenge of cyclic and acyclic Schiff bases and related derivatives. Coord. Chem. Rev. 2004, 248, 1717–2128. [Google Scholar] [CrossRef]
  24. Cozzi, P.G. Metal-Salen Schiff base complexes in catalysis: Practical aspects. Chem. Soc. Rev. 2004, 33, 410–421. [Google Scholar] [CrossRef] [PubMed]
  25. Akitsu, T. Novelties in Schiff Bases; IntechOpen: London, UK, 2024; p. 160. [Google Scholar]
  26. Younus, H.A.; Saleem, F.; Hameed, A.; Al-Rashida, M.; Al-Qawasmeh, R.A.; El-Naggar, M.; Rana, S.; Saeed, M.; Khan, K.M. Part-II: An update of Schiff bases synthesis and applications in medicinal chemistry-a patent review (2016–2023). Expert Opin. Ther. Patents 2023, 33, 841–864. [Google Scholar] [CrossRef] [PubMed]
  27. Nath, B.D.; Islam, M.M.; Karim, M.R.; Rahman, S.; Shaikh, M.A.A.; Georghiou, P.E.; Menelaou, M. Recent Progress in Metal-Incorporated Acyclic Schiff-Base Derivatives: Biological Aspects. ChemistrySelect 2022, 7, e202104290. [Google Scholar] [CrossRef]
  28. Soroceanu, A.; Bargan, A. Advanced and Biomedical Applications of Schiff-Base Ligands and Their Metal Complexes: A Review. Crystals 2022, 12, 1436. [Google Scholar] [CrossRef]
  29. Omidi, S.; Kakanejadifard, A. A review on biological activities of Schiff base, hydrazone, and oxime derivatives of curcumin. RSC Adv. 2020, 10, 30186–30202. [Google Scholar] [CrossRef]
  30. Parveen, S. Recent advances in anticancer ruthenium Schiff base complexes. Appl. Organomet. Chem. 2020, 34, e5687. [Google Scholar] [CrossRef]
  31. Khan, A.M.; Abid, O.U.R.; Mir, S. Assessment of biological activities of chitosan Schiff base tagged with medicinal plants. Biopolymers 2020, 111, e23338. [Google Scholar] [CrossRef]
  32. Hameed, A.; al-Rashida, M.; Uroos, M.; Ali, S.A.; Khan, K.M. Schiff bases in medicinal chemistry: A patent review (2010–2015). Expert Opin. Ther. Patents 2017, 27, 63–79. [Google Scholar] [CrossRef]
  33. Nayab, P.S.; Akrema; Ansari, I.A.; Shahid, M.; Rahisuddin. New phthalimide-appended Schiff bases: Studies of DNA binding, molecular docking and antioxidant activities. Luminescence 2017, 32, 829–838. [Google Scholar] [CrossRef]
  34. Przybylski, P.; Huczynski, A.; Pyta, K.; Brzezinski, B.; Bartl, F. Biological Properties of Schiff Bases and Azo Derivatives of Phenols. Curr. Org. Chem. 2009, 13, 124–148. [Google Scholar] [CrossRef]
  35. Lv, L.; Zheng, T.P.; Tang, L.; Wang, Z.R.; Liu, W.K. Recent advances of Schiff base metal complexes as potential anticancer agents. Coord. Chem. Rev. 2025, 525, 216327. [Google Scholar] [CrossRef]
  36. Kadhum, A.M.; Mallah, S.H.; Waheeb, A.S.; Salman, A.W.; Zafar, A.; Ahmad, N.S.; Siraj, S.; Iqbal, M.A. Advancement in Schiff base complexes for treatment of colon cancer. Rev. Inorg. Chem. 2024, 1–21. [Google Scholar] [CrossRef]
  37. Thakur, S.; Jaryal, A.; Bhalla, A. Recent advances in biological and medicinal profile of Schiff bases and their metal complexes: An updated version (2018–2023). Results Chem. 2024, 7, 101350. [Google Scholar] [CrossRef]
  38. Jorge, J.; Santos, K.F.D.; Timoteo, F.; Vasconcelos, R.R.P.; Caceres, O.I.A.; Granja, I.J.A.; De Souza, D.M., Jr.; Frizon, T.E.A.; Botteselle, G.D.; Braga, A.L.; et al. Recent Advances on the Antimicrobial Activities of Schiff Bases and their Metal Complexes: An Updated Overview. Curr. Med. Chem. 2024, 31, 2330–2344. [Google Scholar] [CrossRef] [PubMed]
  39. 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, 191. [Google Scholar] [CrossRef]
  40. Sovari, S.N.; Zobi, F. Recent Studies on the Antimicrobial Activity of Transition Metal Complexes of Groups 6–12. Chemistry 2020, 2, 418–452. [Google Scholar] [CrossRef]
  41. Malik, M.A.; Dar, O.A.; Gull, P.; Wani, M.Y.; Hashmi, A.A. Heterocyclic Schiff base transition metal complexes in antimicrobial and anticancer chemotherapy. MedChemComm 2018, 9, 409–436. [Google Scholar] [CrossRef]
  42. Hassan, M.A.; Omer, A.M.; Abbas, E.; Baset, W.M.A.; Tamer, T.M. Preparation, physicochemical characterization and antimicrobial activities of novel two phenolic chitosan Schiff base derivatives. Sci. Rep. 2018, 8, 11416. [Google Scholar] [CrossRef]
  43. Low, M.L.; Maigre, L.; Dorlet, P.; Guillot, R.; Pagès, J.M.; Crouse, K.A.; Policar, C.; Delsuc, N. Conjugation of a New Series of Dithiocarbazate Schiff Base Copper(II) Complexes with Vectors Selected to Enhance Antibacterial Activity. Bioconjug. Chem. 2014, 25, 2269–2284. [Google Scholar] [CrossRef]
  44. da Silva, C.M.; da Silva, D.L.; Modolo, L.V.; Alves, R.B.; de Resende, M.A.; Martins, C.V.B.; de Fátima, Â. Schiff bases: A short review of their antimicrobial activities. J. Adv. Res. 2011, 2, 1–8. [Google Scholar] [CrossRef]
  45. Bagihalli, G.B.; Avaji, P.G.; Patil, S.A.; Badami, P.S. Synthesis, spectral characterization, in vitro antibacterial, antifungal and cytotoxic activities of Co(II), Ni(II) and Cu(II) complexes with 1,2,4-triazole Schiff bases. Eur. J. Med. Chem. 2008, 43, 2639–2649. [Google Scholar] [CrossRef] [PubMed]
  46. Ashraf, T.; Ali, B.; Qayyum, H.; Haroone, M.S.; Shabbir, G. Pharmacological aspects of Schiff base metal complexes: A critical review. Inorg. Chem. Commun. 2023, 150, 110449. [Google Scholar] [CrossRef]
  47. Routaray, A.; Nath, N.; Maharana, T.; Sahoo, P.K.; Das, J.P.; Sutar, A.K. Salicylaldimine Copper(II) complex catalyst: Pioneer for ring opening Polymerization of Lactide. J. Chem. Sci. 2016, 128, 883–891. [Google Scholar] [CrossRef]
  48. John, A.; Katiyar, V.; Pang, K.; Shaikh, M.M.; Nanavati, H.; Ghosh, P. Ni(II) and Cu(II) complexes of phenoxy-ketimine ligands: Synthesis, structures and their utility in bulk ring-opening polymerization (ROP) of L-lactide. Polyhedron 2007, 26, 4033–4044. [Google Scholar] [CrossRef]
  49. Felemban, M.F.; Tayeb, F.J.; Alqarni, A.; Ashour, A.A.; Shafie, A. Recent advances in Schiff base coinage metal complexes as anticancer agents: A comprehensive review (2021–2025). Dyes Pigment. 2025, 237, 112710. [Google Scholar] [CrossRef]
  50. Babic, S.; Marjanovic, J.S.; Divac, V.M.; Klisuric, O.R.; Milivojevic, D.; Bogojeski, J.V.; Rakovic, I.; Zaric, M.; Jovanovic, M.; Zaric, R.Z.; et al. Molecular docking study and in vitro evaluation of apoptotic effect of biogenic-amine-based N, O-Cu(II) complexes as potent antitumor agents. J. Coord. Chem. 2025, 78, 1007–1026. [Google Scholar] [CrossRef]
  51. Richa; Kumar, V.; Kataria, R. Phenanthroline and Schiff Base associated Cu(II)-coordinated compounds containing N, O as donor atoms for potent anticancer activity. J. Inorg. Biochem. 2024, 251, 112440. [Google Scholar] [CrossRef] [PubMed]
  52. Santini, C.; Pellei, M.; Gandin, V.; Porchia, M.; Tisato, F.; Marzano, C. Advances in Copper Complexes as Anticancer Agents. Chem. Rev. 2014, 114, 815–862. [Google Scholar] [CrossRef]
  53. Panova, E.V.; Voronina, J.K.; Safin, D.A. Copper(II) Chelates of Schiff Bases Enriched with Aliphatic Fragments: Synthesis, Crystal Structure, In Silico Studies of ADMET Properties and a Potency against a Series of SARS-CoV-2 Proteins. Pharmaceuticals 2023, 16, 286. [Google Scholar] [CrossRef]
  54. Mohan, N.; Vidhya, C.V.; Suni, V.; Ameer, J.M.; Kasoju, N.; Mohanan, P.V.; Sreejith, S.S.; Kurup, M.R.P. Copper(II) salen-based complexes as potential anticancer agents. New J. Chem. 2022, 46, 12540–12550. [Google Scholar] [CrossRef]
  55. Yusuf, T.L.; Oladipo, S.D.; Zamisa, S.; Kumalo, H.M.; Lawal, I.A.; Lawal, M.M.; Mabuba, N. Design of New Schiff-Base Copper(II) Complexes: Synthesis, Crystal Structures, DFT Study, and Binding Potency toward Cytochrome P450 3A4. ACS Omega 2021, 6, 13704–13718. [Google Scholar] [CrossRef] [PubMed]
  56. Kargar, H.; Behjatmanesh-Ardakani, R.; Torabi, V.; Sarvian, A.; Kazemi, Z.; Chavoshpour-Natanzi, Z.; Mirkhani, V.; Sahraei, A.; Tahir, M.N.; Ashfaq, M. Novel copper(II) and zinc(II) complexes of halogenated bidentate N,O-donor Schiff base ligands: Synthesis, characterization, crystal structures, DNA binding, molecular docking, DFT and TD-DFT computational studies. Inorg. Chim. Acta 2021, 514, 120004. [Google Scholar] [CrossRef]
  57. Guo, Y.N.; Hu, X.B.; Zhang, X.L.; Pu, X.H.; Wang, Y. The synthesis of a Cu(II) Schiff base complex using a bidentate N2O2 donor ligand: Crystal structure, photophysical properties, and antibacterial activities. RSC Adv. 2019, 9, 41737–41744. [Google Scholar] [CrossRef]
  58. Lian, W.J.; Wang, X.T.; Xie, C.Z.; Tian, H.; Song, X.Q.; Pan, H.T.; Qiao, X.; Xu, J.Y. Mixed-ligand copper(II) Schiff base complexes: The role of the co-ligand in DNA binding, DNA cleavage, protein binding and cytotoxicity. Dalton Trans. 2016, 45, 9073–9087. [Google Scholar] [CrossRef] [PubMed]
  59. Ei-Sherif, A.A.; Eldebss, T.M.A. Synthesis, spectral characterization, solution equilibria, in vitro antibacterial and cytotoxic activities of Cu(II), Ni(II), Mn(II), Co(II) and Zn(II) complexes with Schiff base derived from 5-bromosalicylaldehyde and 2-aminomethylthiophene. Spectrochim. Acta Part A 2011, 79, 1803–1814. [Google Scholar] [CrossRef]
  60. Li, Y.P.; Wu, Y.B.; Zhao, J.; Yang, P. DNA-binding and cleavage studies of novel binuclear copper(II) complex with 1,1′-dimethyl-2,2′-biimidazole ligand. J. Inorg. Biochem. 2007, 101, 283–290. [Google Scholar] [CrossRef] [PubMed]
  61. Lv, J.; Liu, T.T.; Cai, S.L.; Wang, X.; Liu, L.; Wang, Y.M. Synthesis, structure and biological activity of cobalt(II) and copper(II) complexes of valine-derived Schiff bases. J. Inorg. Biochem. 2006, 100, 1888–1896. [Google Scholar] [CrossRef]
  62. Costamagna, J.; Ferraudi, G.; Matsuhiro, B.; Campos-Vallette, M.; Canales, J.; Villagrán, M.; Vargas, J.; Aguirre, M.J. Complexes of macrocycles with pendant arms as models for biological molecules. Coord. Chem. Rev. 2000, 196, 125–164. [Google Scholar] [CrossRef]
  63. Lei, Y.; Yang, Q.W.; Hu, Z.H.; Wang, S. Synthesis, structural studies and antimicrobial activity of copper(II) complexes derived from 2,4-difluoro-6-(((2-(pyrrolidin-1-yl)ethyl)imino) methyl)phenol. Polyhedron 2024, 259, 117071. [Google Scholar] [CrossRef]
  64. Sonawane, H.R.; Vibhute, B.T.; Aghav, B.D.; Deore, J.V.; Patil, S.K. Versatile applications of transition metal incorporating quinoline Schiff base metal complexes: An overview. Eur. J. Med. Chem. 2023, 258, 115549. [Google Scholar] [CrossRef] [PubMed]
  65. Kumar, R.; Singh, A.A.; Kumar, U.; Jain, P.; Sharma, A.K.; Kant, C.; Faizi, M.S.H. Recent advances in synthesis of heterocyclic Schiff base transition metal complexes and their antimicrobial activities especially antibacterial and antifungal. J. Mol. Struct. 2023, 1294, 136346. [Google Scholar] [CrossRef]
  66. Devi, J.; Sharma, S.; Kumar, S.; Kumar, B.; Kumar, D.; Jindal, D.K.; Das, S. Synthesis, characterization, in vitro antimicrobial and cytotoxic studies of Co(II), Ni(II), Cu(II), and Zn(II) complexes obtained from Schiff base ligands of 1, 2, 3, 4-tetrahydro-naphthalen-1-ylamine. Appl. Organomet. Chem. 2022, 36, e6760. [Google Scholar] [CrossRef]
  67. Peewasan, K.; Merkel, M.P.; Zarschler, K.; Stephan, H.; Anson, C.E.; Powell, A.K. Tetranuclear Cu(II)-chiral complexes: Synthesis, characterization and biological activity. RSC Adv. 2019, 9, 24087–24091. [Google Scholar] [CrossRef] [PubMed]
  68. Dar, O.A.; Lone, S.A.; Malik, M.A.; Wani, M.Y.; Ahmad, A.; Hashmi, A.A. New transition metal complexes with a pendent indole ring: Insights into the antifungal activity and mode of action. RSC Adv. 2019, 9, 15151–15157. [Google Scholar] [CrossRef]
  69. Kargar, H.; Aghaei-Meybodi, F.; Behjatmanesh-Ardakani, R.; Elahifard, M.R.; Torabi, V.; Fallah-Mehrjardi, M.; Tahir, M.N.; Ashfaq, M.; Munawar, K.S. Synthesis, crystal structure, theoretical calculation, spectroscopic and antibacterial activity studies of copper(II) complexes bearing bidentate Schiff base ligands derived from 4-aminoantipyrine: Influence of substitutions on antibacterial activity. J. Mol. Struct. 2021, 1230, 129908. [Google Scholar] [CrossRef]
  70. Kargar, H.; Aghaei-Meybodi, F.; Elahifard, M.R.; Tahir, M.N.; Ashfaq, M.; Munawar, K.S. Some new Cu(II) complexes containing O,N-donor Schiff base ligands derived from 4-aminoantipyrine: Synthesis, characterization, crystal structure and substitution effect on antimicrobial activity. J. Coord. Chem. 2021, 74, 1534–1549. [Google Scholar] [CrossRef]
  71. Alturiqi, A.S.; Alaghaz, A.; Zayed, M.E.; Ammar, R.A. Synthesis, characterization, biological activity, and corrosion inhibition in acid medium of unsymmetrical tetradentate N2O2 Schiff base complexes. J. Chin. Chem. Soc. 2018, 65, 1060–1074. [Google Scholar] [CrossRef]
  72. Regojevic, M.S.; Zoric, M.Z.; Radnovic, N.D.; Bogdanovic, M.G.; Holló, B.B.; Rodic, M.V.; Raievic, V.; Borisev, I.D.; Vojinovic-Jesic, L.S.; Hozjan, M.; et al. Anion-directed synthesis of copper(I/II) complexes with a Schiff base derived from 2-(diphenylphosphino)benzaldehyde and aminoguanidine. J. Mol. Struct. 2025, 1336, 14. [Google Scholar] [CrossRef]
  73. Khalaji, A.D.; Weil, M.; Hadadzadeh, H.; Daryanavard, M. Two different 1D-chains in the structures of the copper(I) coordination polymers based on bidentate Schiff-base building units and thiocyanate anions as bridging ligands. Inorg. Chim. Acta 2009, 362, 4837–4842. [Google Scholar] [CrossRef]
  74. Morshedi, M.; Amirnasr, M.; Triki, S.; Khalaji, A.D. New (NS)2 Schiff base with a flexible spacer: Synthesis and structural characterization of its first coordination polymer [Cu2(l-I)2(l-(thio)2dapte)]n (1). Inorg. Chim. Acta 2009, 362, 1637–1640. [Google Scholar] [CrossRef]
  75. Ferraro, V.; Fuhr, O.; Bizzarri, C.; Braese, S. Substituted Pyrrole-based Schiff Bases: Effect on the Luminescence of Neutral Heteroleptic Cu(I) Complexes. Eur. J. Inorg. Chem. 2024, 27, e202400080. [Google Scholar] [CrossRef]
  76. Crestani, M.G.; Manbeck, G.F.; Brennessel, W.W.; McCormick, T.M.; Eisenberg, R. Synthesis and Characterization of Neutral Luminescent Diphosphine Pyrrole- and Indole-Aldimine Copper(I) Complexes. Inorg. Chem. 2011, 50, 7172–7188. [Google Scholar] [CrossRef]
  77. Lv, J.; Wu, X.Y.; Wang, R.; Wu, Y.Q.; Xu, S.X.; Zhao, F.; Wang, Y.B. Schiff base-type Cu(I) complexes containing naphthylpyridyl-methanimine ligands featuring higher light-absorption capability: Synthesis, structures, and photophysical properties. Polyhedron 2022, 224, 116002. [Google Scholar] [CrossRef]
  78. Lv, J.; Lu, Y.F.; Wang, J.L.; Zhao, F.; Wang, Y.B.; He, H.F.; Wu, Y.Q. Schiff base-type copper(I) complexes exhibiting high molar extinction coefficients: Synthesis, characterization and DFT studies. J. Mol. Struct. 2022, 1249, 131638. [Google Scholar] [CrossRef]
  79. Lv, J.; Li, Q.Q.; Wang, J.L.; Xu, S.X.; Zhao, F.; He, H.F.; Wang, Y.B. Orange-red emissive Cu(I) complexes bearing Schiff base ligands: Synthesis, structures, and photophysical properties. J. Mol. Struct. 2022, 1252, 132180. [Google Scholar] [CrossRef]
  80. Klasen, H.J. A historical review of the use of silver in the treatment of burns. II. Renewed interest for silver. Burns 2000, 26, 131–138. [Google Scholar] [CrossRef] [PubMed]
  81. Cai, L.; Huang, Y.Q.; Duan, Y.Y.; Liu, Q.; Xu, Q.L.; Jia, J.; Wang, J.M.; Tong, Q.; Luo, P.P.; Wen, Y.J.; et al. Schiff-base silver nanocomplexes formation on natural biopolymer coated mesoporous silica contributed to the improved curative effect on infectious microbes. Nano Res. 2021, 14, 2735–2748. [Google Scholar] [CrossRef]
  82. Azócar, M.I.; Gómez, G.; Levín, P.; Paez, M.; Muñoz, H.; Dinamarca, N. Review: Antibacterial behavior of carboxylate silver(I) complexes. J. Coord. Chem. 2014, 67, 3840–3853. [Google Scholar] [CrossRef]
  83. Nomiya, K.; Tsuda, K.; Sudoh, T.; Oda, M. Ag(I)-N bond-containing compound showing wide spectra in effective antimicrobial activities: Polymeric silver(I) imidazolate. J. Inorg. Biochem. 1997, 68, 39–44. [Google Scholar] [CrossRef]
  84. Khan, E.; Hanif, M.; Akhtar, M.S. Schiff bases and their metal complexes with biologically compatible metal ions; biological importance, recent trends and future hopes. Rev. Inorg. Chem. 2022, 42, 307–325. [Google Scholar] [CrossRef]
  85. Njogu, E.M.; Omondi, B.; Nyamori, V.O. Silver(I)-pyridinyl Schiff base complexes: Synthesis, characterisation and antimicrobial studies. J. Mol. Struct. 2017, 1135, 118–128. [Google Scholar] [CrossRef]
  86. Adeleke, A.A.; Oladipo, S.D.; Zamisa, S.J.; Sanusi, I.A.; Omondi, B. DNA/BSA binding studies and in vitro anticancer and antibacterial studies of isoelectronic Cu(I)- and Ag(I)-pyridinyl Schiff base complexes incorporating triphenylphosphine as co-ligands. Inorg. Chim. Acta 2023, 558, 121760. [Google Scholar] [CrossRef]
  87. Raczuk, E.; Dmochowska, B.; Samaszko-Fiertek, J.; Madaj, J. Different Schiff Bases-Structure, Importance and Classification. Molecules 2022, 27, 787. [Google Scholar] [CrossRef] [PubMed]
  88. Ren, M.; Xu, Z.L.; Bao, S.S.; Wang, T.T.; Zheng, Z.H.; Ferreira, R.A.S.; Zheng, L.M.; Carlos, L.D. Lanthanide salen-type complexes exhibiting single ion magnet and photoluminescent properties. Dalton Trans. 2016, 45, 2974–2982. [Google Scholar] [CrossRef]
  89. Kitamura, F.; Sawaguchi, K.; Mori, A.; Takagi, S.; Suzuki, T.; Kobayashi, A.; Kato, M.; Nakajima, K. Coordination Structure Conversion of Hydrazone-Palladium(II) Complexes in the Solid State and in Solution. Inorg. Chem. 2015, 54, 8436–8448. [Google Scholar] [CrossRef] [PubMed]
  90. Chew, S.T.; Lo, K.M.; Sinniah, S.K.; Sim, K.S.; Tan, K.W. Synthesis, characterization and biological evaluation of cationic hydrazone copper complexes with diverse diimine co-ligands. RSC Adv. 2014, 4, 61232–61247. [Google Scholar] [CrossRef]
  91. Su, W.; Qian, Q.Q.; Li, P.Y.; Lei, X.L.; Xiao, Q.; Huang, S.; Huang, C.S.; Cui, J.G. Synthesis, Characterization, and Anticancer Activity of a Series of Ketone-N4-Substituted Thiosemicarbazones and Their Ruthenium(II) Arene Complexes. Inorg. Chem. 2013, 52, 12440–12449. [Google Scholar] [CrossRef]
  92. Whiteoak, C.J.; Salassa, G.; Kleij, A.W. Recent advances with π-conjugated salen systems. Chem. Soc. Rev. 2012, 41, 622–631. [Google Scholar] [CrossRef]
  93. Krishnamoorthy, P.; Sathyadevi, P.; Butorac, R.R.; Cowley, A.H.; Bhuvanesh, N.S.P.; Dharmaraj, N. Copper(I) and nickel(II) complexes with 1:1 vs. 1:2 coordination of ferrocenyl hydrazone ligands: Do the geometry and composition of complexes affect DNA binding/cleavage, protein binding, antioxidant and cytotoxic activities? Dalton Trans. 2012, 41, 4423–4436. [Google Scholar] [CrossRef]
  94. Bagherzadeh, M.; Zare, M. Synthesis and characterization of NaY zeolite-encapsulated Mn-hydrazone Schiff base: An efficient and reusable catalyst for oxidation of olefins. J. Coord. Chem. 2012, 65, 4054–4066. [Google Scholar] [CrossRef]
  95. Chellan, P.; Land, K.M.; Shokar, A.; Au, A.; An, S.H.; Clavel, C.M.; Dyson, P.J.; de Kock, C.; Smith, P.J.; Chibale, K.; et al. Exploring the Versatility of Cycloplatinated Thiosemicarbazones as Antitumor and Antiparasitic Agents. Organometallics 2012, 31, 5791–5799. [Google Scholar] [CrossRef]
  96. Prabhakaran, R.; Kalaivani, P.; Poornima, P.; Dallemer, F.; Paramaguru, G.; Padma, V.V.; Renganathan, R.; Huang, R.; Natarajan, K. One pot synthesis of structurally different mono and dimeric Ni(II) thiosemicarbazone complexes and N-arylation on a coordinated ligand: A comparative biological study. Dalton Trans. 2012, 41, 9323–9336. [Google Scholar] [CrossRef]
  97. Minkin, V.I.; Tsukanov, A.V.; Dubonosov, A.D.; Bren, V.A. Tautomeric Schiff bases: Iono-, solvato-, thermo- and photochromism. J. Mol. Struct. 2011, 998, 179–191. [Google Scholar] [CrossRef]
  98. Hansen, P.E.; Filarowski, A. Characterisation of the PT-form of o-hydroxy acylarornatic Schiff bases by NMR spectroscopy and DFT calculations. J. Mol. Struct. 2004, 707, 69–75. [Google Scholar] [CrossRef]
  99. Dominiak, P.M.; Grech, E.; Barr, G.; Teat, S.; Mallinson, P.; Wozniak, K. Neutral and ionic hydrogen bonding in Schiff bases. Chem. Eur. J. 2003, 9, 963–970. [Google Scholar] [CrossRef] [PubMed]
  100. Król-Starzomska, I.; Filarowski, A.; Rospenk, M.; Koll, A.; Melikova, S. Proton transfer equilibria in Schiff bases with steric repulsion. J. Phys. Chem. A 2004, 108, 2131–2138. [Google Scholar] [CrossRef]
  101. Filarowski, A.; Glowiaka, T.; Koll, A. Strengthening of the intramolecular O⋯H⋯N hydrogen bonds in Schiff bases as a result of steric repulsion. J. Mol. Struct. 1999, 484, 75–89. [Google Scholar] [CrossRef]
  102. Jinno, S.; Senoo, T.; Mori, K. Access to ortho-Hydroxyphenyl Ketimines via Imine Anion-Mediated Smiles Rearrangement. Org. Lett. 2022, 24, 4140–4144. [Google Scholar] [CrossRef]
  103. Guerrero-Corella, A.; Esteban, F.; Iniesta, M.; Martín-Somer, A.; Parra, M.; Díaz-Tendero, S.; Fraile, A.; Alemán, J. 2-Hydroxybenzophenone as a Chemical Auxiliary for the Activation of Ketiminoesters for Highly Enantioselective Addition to Nitroalkenes under Bifunctional Catalysis. Angew. Chem. Int. Ed. 2018, 57, 5350–5354. [Google Scholar] [CrossRef]
  104. Al-Qaisi, F.; Genjang, N.; Nieger, M.; Repo, T. Synthesis, structure and catalytic activity of bis(phenoxyiminato)iron(III) complexes in coupling reaction of CO2 and epoxides. Inorg. Chim. Acta 2016, 442, 81–85. [Google Scholar] [CrossRef]
  105. Sibaouih, A.; Ryan, P.; Axenov, K.V.; Sundberg, M.R.; Leskelä, M.; Repo, T. Efficient coupling of CO2 and epoxides with bis(phenoxyiminato) cobalt(III)/Lewis base catalysts. J. Mol. Catal. A Chem. 2009, 312, 87–91. [Google Scholar] [CrossRef]
  106. Marcazzan, P.; Patrick, B.O.; James, B.R. Catalyst poisoning in catalyzed imine hydrogenation: A novel zwitterionic Rh(I)/o-hydroxy-substituted imine complex. J. Mol. Catal. A Chem. 2006, 257, 26–30. [Google Scholar] [CrossRef]
  107. Cimarelli, C.; Palmieri, G.; Volpini, E. Synthesis of enantiopure 2-aminoalkylphenols by stereoselective addition of Grignard reagents to chiral 2-imidoylphenols. Tetrahedron Asymmetry 2002, 13, 2011–2018. [Google Scholar] [CrossRef]
  108. Mondal, B.; Chakraborty, S.; Munshi, P.; Walawalkar, M.G.; Lahiri, G.K. Ruthenium-(II)/-(III) terpyridine complexes incorporating imine functionalities. Synthesis, structure, spectroscopic and electrochemical properties. J. Chem. Soc. Dalton Trans. 2000, 2327–2335. [Google Scholar] [CrossRef]
  109. Keerthi, K.D.; Santra, B.K.; Lahiri, G.K. Ruthenium(II) bipyridine complexes with modified phenolic Schiff base ligands. Synthesis, spectroscopic characterization and redox properties. Polyhedron 1998, 17, 1387–1396. [Google Scholar] [CrossRef]
  110. Choudhary, N.; Hughes, D.L.; Kleinkes, U.; Larkworthy, L.F.; Leigh, G.J.; Maiwald, M.; Marmion, C.J.; Sanders, J.R.; Smith, G.W.; Sudbrake, C. New tetradentate Schiff bases, their oxovanadium(IV) complexes, and some complexes of bidentate Schiff bases with vanadium(III). Polyhedron 1997, 16, 1517–1528. [Google Scholar] [CrossRef]
  111. Pellei, M.; Santini, C.; Bagnarelli, L.; Caviglia, M.; Sgarbossa, P.; De Franco, M.; Zancato, M.; Marzano, C.; Gandin, V. Novel Silver Complexes Based on Phosphanes and Ester Derivatives of Bis(pyrazol-1-yl)acetate Ligands Targeting TrxR: New Promising Chemotherapeutic Tools Relevant to SCLC Management. Int. J. Mol. Sci. 2023, 24, 4091. [Google Scholar] [CrossRef]
  112. Del Bello, F.; Pellei, M.; Bagnarelli, L.; Santini, C.; Giorgioni, G.; Piergentili, A.; Quaglia, W.; Battocchio, C.; Iucci, G.; Schiesaro, I.; et al. Cu(I) and Cu(II) Complexes Based on Lonidamine-Conjugated Ligands Designed to Promote Synergistic Antitumor Effects. Inorg. Chem. 2022, 61, 4919–4937. [Google Scholar] [CrossRef]
  113. Pellei, M.; Bagnarelli, L.; Luciani, L.; Del Bello, F.; Giorgioni, G.; Piergentili, A.; Quaglia, W.; De Franco, M.; Gandin, V.; Marzano, C.; et al. Synthesis and Cytotoxic Activity Evaluation of New Cu(I) Complexes of Bis(pyrazol-1-yl) Acetate Ligands Functionalized with an NMDA Receptor Antagonist. Int. J. Mol. Sci. 2020, 21, 2616. [Google Scholar] [CrossRef]
  114. Morelli, M.B.; Amantini, C.; Santoni, G.; Pellei, M.; Santini, C.; Cimarelli, C.; Marcantoni, E.; Petrini, M.; Del Bello, F.; Giorgioni, G.; et al. Novel antitumor copper(II) complexes designed to act through synergistic mechanisms of action, due to the presence of an NMDA receptor ligand and copper in the same chemical entity. New J. Chem. 2018, 42, 11878–11887. [Google Scholar] [CrossRef]
  115. Gandin, V.; Ceresa, C.; Esposito, G.; Indraccolo, S.; Porchia, M.; Tisato, F.; Santini, C.; Pellei, M.; Marzano, C. Therapeutic potential of the phosphino Cu(I) complex (HydroCuP) in the treatment of solid tumors. Sci. Rep. 2017, 7, 13936. [Google Scholar] [CrossRef]
  116. Tisato, F.; Marzano, C.; Peruzzo, V.; Tegoni, M.; Giorgetti, M.; Damjanovic, M.; Trapananti, A.; Bagno, A.; Santini, C.; Pellei, M.; et al. Insights into the cytotoxic activity of the phosphane copper(I) complex Cu(thp)(4) PF6. J. Inorg. Biochem. 2016, 165, 80–91. [Google Scholar] [CrossRef] [PubMed]
  117. Papini, G.; Bandoli, G.; Dolmella, A.; Gioia Lobbia, G.; Pellei, M.; Santini, C. New homoleptic carbene transfer ligands and related coinage metal complexes. Inorg. Chem. Commun. 2008, 11, 1103–1106. [Google Scholar] [CrossRef]
  118. Räisänen, M.T.; Elo, P.; Kettunen, M.; Klinga, M.; Leskelä, M.; Repo, T. Practical method for 2-hydroxyphenylketimine synthesis. Synth. Commun. 2007, 37, 1765–1777. [Google Scholar] [CrossRef]
  119. Cimarelli, C.; Palmieri, G.; Volpini, E. An improved solvent-free preparation of 2-imidoylphenols. Org. Prep. Proced. Int. 2001, 33, 369–371. [Google Scholar] [CrossRef]
  120. Cimarelli, C.; Palmieri, G. Alkylation of dianions derived from 2-(1-iminoalkyl) phenols: Synthesis of functionalized 2-acyl phenols. Tetrahedron 1998, 54, 15711–15720. [Google Scholar] [CrossRef]
  121. Filarowski, A.; Koll, A.; Głowiak, T. Steric Modification of the Intramolecular Hydrogen Bond in 2-(Methylimino-phenyl-methyl)-phenols. Monatshefte Chem. 1999, 130, 1097–1108. [Google Scholar] [CrossRef]
  122. Mihaylov, M.Y.; Zdravkova, V.R.; Ivanova, E.Z.; Aleksandrov, H.A.; St Petkov, P.; Vayssilov, G.N.; Hadjiivanov, K.I. Infrared spectra of surface nitrates: Revision of the current opinions based on the case study of ceria. J. Catal. 2021, 394, 245–258. [Google Scholar] [CrossRef]
  123. Nakamoto, K. Applications in Coordination Chemistry. In Infrared and Raman Spectra of Inorganic and Coordination Compounds: Part B: Applications in Coordination, Organometallic, and Bioinorganic Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 2008; pp. 1–273. [Google Scholar]
  124. Effendy; Healy, P.C.; Marchetti, F.; Pettinari, C.; Pettinari, R.; Tombesi, A.; Skelton, B.W.; White, A.H. Synthesis and structural characterization of some 1:1 and 1:2 adducts of silver(I) salts with hindered Pmes3, PPhmes2 and PPh2mes bases (Ph = phenyl, mes=2,4,6-trimethylpheny1)). Inorg. Chim. Acta 2022, 535, 120857. [Google Scholar] [CrossRef]
  125. Meijboom, R.; Bowen, R.J.; Berners-Price, S.J. Coordination complexes of silver(I) with tertiary phosphine and related ligands. Coord. Chem. Rev. 2009, 253, 325–342. [Google Scholar] [CrossRef]
  126. Barron, P.F.; Dyason, J.C.; Healy, P.C.; Engelhardt, L.M.; Skelton, B.W.; White, A.H. Lewis Base Adducts of Group 11 Metal Compounds. Part 24. Co-ordination of Triphenylphosphine with Silver Nitrate. A Solid-state Cross-polarization Magic Angle Spinning 31P Nuclear Magnetic Resonance, Crystal Structure, and Infrared Spectroscopic Study of Ag(PPh3)nNO3 (n = 1–4). J. Chem. Soc.-Dalton Trans. 1986, 9, 1965–1970. [Google Scholar]
  127. Pellei, M.; Del Gobbo, J.; Caviglia, M.; Karade, D.V.; Gandin, V.; Marzano, C.; Poyil, A.N.; Dias, H.V.R.; Santini, C. Synthesis and cytotoxicity studies of Cu(I) and Ag(I) complexes based on sterically hindered β-diketonates with different degrees of fluorination. Dalton Trans. 2023, 52, 12098–12111. [Google Scholar] [CrossRef] [PubMed]
  128. Dias, H.V.R.; Flores, J.A.; Pellei, M.; Morresi, B.; Gioia Lobbia, G.; Singh, S.; Kobayashi, Y.; Yousufuddin, M.; Santini, C. Silver(I) and copper(I) complexes supported by fully fluorinated 1,3,5-triazapentadienyl ligands. Dalton Trans. 2011, 40, 8569–8580. [Google Scholar] [CrossRef]
  129. Pellei, M.; Alidori, S.; Papini, G.; Gioia Lobbia, G.; Gorden, J.D.; Dias, H.V.R.; Santini, C. Silver(I)-organophosphane complexes of electron withdrawing CF3- or NO2-substituted scorpionate ligands. Dalton Trans. 2007, 4845–4853. [Google Scholar] [CrossRef]
  130. Dias, H.V.R.; Alidori, S.; Gioia Lobbia, G.; Papini, G.; Pellei, M.; Santini, C. Small Scorpionate Ligands:  Silver(I)-Organophosphane Complexes of 5-CF3-Substituted Scorpionate Ligand Combining a B−H···Ag Coordination Motif. Inorg. Chem. 2007, 46, 9708–9714. [Google Scholar] [CrossRef]
  131. Okuniewski, A.; Rosiak, D.; Chojnacki, J.; Becker, B. Coordination polymers and molecular structures among complexes of mercury(II) halides with selected 1-benzoylthioureas. Polyhedron 2015, 90, 47–57. [Google Scholar] [CrossRef]
  132. Yang, L.; Powell, D.R.; Houser, R.P. Structural variation in copper(I) complexes with pyridylmethylamide ligands:: Structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 9, 955–964. [Google Scholar] [CrossRef]
  133. Allen, F.H. The Cambridge Structural Database: A quarter of a million crystal structures and rising. Acta Cryst. 2002, B58, 380–388. [Google Scholar] [CrossRef]
  134. Wang, D.J.; Fan, L.; Wang, G.H. Crystal structure of N-(diphenylmethylene)diphenylmethanamine, C26H21N. Z. Krist.-New Cryst. Struct. 2009, 224, 186–188. [Google Scholar]
  135. Liu, X.L.; Gao, A.; Ding, L.; Xu, J.; Zhao, B.G. Aminative Umpolung Synthesis of Aryl Vicinal Diamines from Aromatic Aldehydes. Org. Lett. 2014, 16, 2118–2121. [Google Scholar] [CrossRef] [PubMed]
  136. Vorontsova, N.V.; Bystrova, G.S.; Antonov, D.Y.; Vologzhanina, A.V.; Godovikov, I.A.; Il’in, M.M. Novel ligands based on bromosubstituted hydroxycarbonyl 2.2 paracyclophane derivatives: Synthesis and application in asymmetric catalysis. Tetrahedron Asymmetry 2010, 21, 731–738. [Google Scholar] [CrossRef]
  137. Blackwell, J.M.; Piers, W.E.; Parvez, M.; McDonald, R. Solution and solid-state characteristics of imine adducts with tris(pentafluorophenyl)borane. Organometallics 2002, 21, 1400–1407. [Google Scholar] [CrossRef]
  138. Rozenberg, V.; Danilova, T.Y.; Sergeeva, E.; Vorontsov, E.; Starikova, Z.; Lysenko, K.; Belokon’, Y. Regioselective Fries Rearrangement and Friedel−Crafts Acylation as Efficient Routes to Novel Enantiomerically Enriched ortho-Acylhydroxy [2.2]paracyclophanes. Eur. J. Org. Chem. 2000, 2000, 3295–3303. [Google Scholar] [CrossRef]
  139. Ito, M.; Kasuga, N.C.; Matsuse, R.; Hirotsu, M. Crystal structures and circular dichroism of {2,20-[(1S,2S)-1,2-diphenylethane-1,2-diylbis(nitrilo-phenylmethanylylidene)]diphenolato} nickel(II) and its ethanol solvate. Acta Cryst. 2024, E80, 1259–1265. [Google Scholar]
  140. Hirotsu, M.; Kuwamura, N.; Kinoshita, I.; Kojima, M.; Yoshikawa, Y.; Ueno, K. Steric, geometrical and solvent effects on redox potentials in salen-type copper(II) complexes. Dalton Trans. 2009, 7678–7683. [Google Scholar] [CrossRef]
  141. Hirotsu, M.; Kojima, M.; Nakajima, K.; Kashino, S.; Yoshikawa, Y. Stereochemistry and electrochemistry of cobalt(II) and cobalt(III) complexes containing optically active tetradentate Schiff base ligands. Bull. Chem. Soc. Jpn. 1996, 69, 2549–2557. [Google Scholar] [CrossRef]
  142. Hirotsu, M.; Nakajima, K.; Kojima, M.; Yoshikawa, Y. Manganese(III) Complexes Containing Optically Active Tetradentate Schiff Base Ligands. Effect of Phenyl Substituents. Inorg. Chem. 1995, 34, 6173–6178. [Google Scholar] [CrossRef]
  143. Hirotsu, M.; Kojima, M.; Nakajima, K.; Kashino, S.; Yoshikawa, Y. Steric Control of Redox Potentials of Cobalt(II) Schiff Base Complexes with Phenyl Substituents. Chem. Lett. 1994, 23, 2183–2186. [Google Scholar] [CrossRef]
  144. Cifuentes-Vaca, O.L.; Andrades-Lagos, J.; Campanini-Salinas, J.; Laguna, A.; Vásquez-Velásquez, D.; Gimeno, M.C. Silver(I) and copper(I) complexes with a Schiff base derived from 2-aminofluorene with promising antibacterial activity. Inorg. Chim. Acta 2019, 489, 275–279. [Google Scholar] [CrossRef]
  145. Berthon, G. Critical evaluation of the stability constants of metal complexes of amino acids* with polar side chains (TechnicaI report). Pure Appl. Chem. 1995, 67, 1117–1240. [Google Scholar] [CrossRef]
  146. Kumar, M.; Singh, A.K.; Singh, S.; Singh, A.K.; Rao, P.K.; Yadav, R.K.; Singh, A.P.; Tripathi, U.N. Exploration of iron(III) complexes with bidentate N, O-donor Schiff base ligands through synthesis, characterization, DFT, and antibacterial studies. J. Mol. Struct. 2025, 1319, 139496. [Google Scholar] [CrossRef]
  147. Salah, N.; Adly, O.M.I.; Ibrahim, M.A.; Abdelaziz, M.; Abdelrhman, E.M. New Metal Complexes Incorporating Schiff Base Ligand Based on Pyridine Moiety: Synthesis, Spectral Characterization, DFT, Biological Evaluation, and Molecular Docking. Appl. Organomet. Chem. 2025, 39, e7751. [Google Scholar] [CrossRef]
  148. Oladipo, S.D.; Mocktar, C.; Omondi, B. In vitro biological studies of heteroleptic Ag(I) and Cu(I) unsymmetrical N,N’-diarylformamidine dithiocarbamate phosphine complexes; the effect of the metal center. Arab. J. Chem. 2020, 13, 6379–6394. [Google Scholar] [CrossRef]
  149. Reddy, T.S.; Privér, S.H.; Ojha, R.; Mirzadeh, N.; Velma, G.R.; Jakku, R.; Hosseinnejad, T.; Luwor, R.; Ramakrishna, S.; Wlodkowic, D.; et al. Gold(I) complexes of the type [AuL{κC-2-C6H4P(S)Ph2}] [L = PTA, PPh3, PPh2(C6H4-3-SO3Na) and PPh2(2-py)]: Synthesis, characterisation, crystal structures, and In Vitro and In Vivo anticancer properties. Eur. J. Med. Chem. 2025, 281, 117007. [Google Scholar] [CrossRef] [PubMed]
  150. Singh, K.; Singh, V.K.; Mishra, R.; Sharma, A.; Pandey, A.; Srivastava, S.K.; Chaurasia, H. Design, Synthesis, DFT, docking Studies, and antimicrobial evaluation of novel benzimidazole containing sulphonamide derivatives. Bioorg. Chem. 2024, 149, 107473. [Google Scholar] [CrossRef]
  151. Yan, Y.; Xia, X.X.; Fatima, A.; Zhang, L.; Yuan, G.J.; Lian, F.X.; Wang, Y. Antibacterial Activity and Mechanisms of Plant Flavonoids against Gram-Negative Bacteria Based on the Antibacterial Statistical Model. Pharmaceuticals 2024, 17, 292. [Google Scholar] [CrossRef]
  152. Hasan, A.; Varna, D.; Chakraborty, I.; Angaridis, P.A.; Raptis, R.G. Synthesis, structure and antibacterial properties of a mononuclear Ag(I) complex, [Ag(OBz)(PTA)2] (OBz =benzoate, PTA =1,3,5-triaza-7-phospadamantane). Results Chem. 2022, 4, 100580. [Google Scholar] [CrossRef]
  153. Pervaiz, M.; Sadiq, A.; Sadiq, S.; Saeed, Z.; Imran, M.; Younas, U.; Bukhari, S.M.; Khan, R.R.M.; Rashid, A.; Adnan, A. Design and synthesis of Schiff base Homobimetallic-Complexes as promising antimicrobial agents. Inorg. Chem. Commun. 2022, 137, 109206. [Google Scholar] [CrossRef]
  154. Ejidike, I.P. Cu(II) Complexes of 4-[(1E)-N-{2-[(Z)-Benzylidene-amino]ethyl}ethanimidoyl]benzene-1,3-diol Schiff Base: Synthesis, Spectroscopic, In-Vitro Antioxidant, Antifungal and Antibacterial Studies. Molecules 2018, 23, 1581. [Google Scholar] [CrossRef]
  155. Calu, L.; Badea, M.; Korosin, N.C.; Chifiriuc, M.C.; Bleotu, C.; Stanica, N.; Silvestro, L.; Maurer, M.; Olar, R. Spectral, thermal and biological characterization of complexes with a Schiff base bearing triazole moiety as potential antimicrobial species. J. Therm. Anal. Calorim. 2018, 134, 1839–1850. [Google Scholar] [CrossRef]
  156. Zhang, Y.B.; Liu, Q.; Jing, H.R.; Cai, Y.J.; Wang, Q.; Li, Y.G. Synthesis, characterization, and antimicrobial activity of two Schiff base silver(I) complexes derived from 4-carboxybenzaldehyde. J. Coord. Chem. 2017, 70, 1066–1076. [Google Scholar] [CrossRef]
  157. CrysAlisPro Versions 1.171.43.141a; Rigaku Oxford Diffraction: Oxford, UK, 2024.
  158. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  159. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  160. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
  161. Macrae, C.F.; Bruno, I.J.; Chisholm, J.A.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P.A. Mercury CSD 2.0—New features for the visualization and investigation of crystal structures. J. Appl. Crystallogr. 2008, 41, 466–470. [Google Scholar] [CrossRef]
  162. Liu, J.M.; Guo, C.P.; Liu, Z.Z.; Cheng, F.; Zhang, S.; Zhang, Z.H. Simultaneous sterilization and biosensing of pathogenic bacteria via copper phthalocyanine-based COF embedded with Cu-N4 single atomic sites and silver nanoparticles. Chem. Eng. J. 2024, 494, 153139. [Google Scholar] [CrossRef]
Scheme 1. Tautomerism of 2-imidoylphenols.
Scheme 1. Tautomerism of 2-imidoylphenols.
Molecules 30 01893 sch001
Scheme 2. Synthesis of HLBSMe (1) and HLBSPh (2) ligands.
Scheme 2. Synthesis of HLBSMe (1) and HLBSPh (2) ligands.
Molecules 30 01893 sch002
Scheme 3. Synthesis of compounds 310.
Scheme 3. Synthesis of compounds 310.
Molecules 30 01893 sch003
Figure 1. (Left): a drawing of the asymmetric unit of [Cu(LBSPh)2] (10), showing the numbering scheme; disorder not shown. (Right): same view highlighting the disorder of the phenyl ring C15/C20. Hydrogen atoms omitted; thermal ellipsoids at 50% probability.
Figure 1. (Left): a drawing of the asymmetric unit of [Cu(LBSPh)2] (10), showing the numbering scheme; disorder not shown. (Right): same view highlighting the disorder of the phenyl ring C15/C20. Hydrogen atoms omitted; thermal ellipsoids at 50% probability.
Molecules 30 01893 g001
Figure 2. Two orthogonal views of the [Cu(LBSPh)2] complex. (Left): overall view of the complex looking through the coordination plane, highlighting its calix shape; (right): view of the Cu(II) coordination plane. Hydrogen atoms omitted; thermal ellipsoids at 50% probability.
Figure 2. Two orthogonal views of the [Cu(LBSPh)2] complex. (Left): overall view of the complex looking through the coordination plane, highlighting its calix shape; (right): view of the Cu(II) coordination plane. Hydrogen atoms omitted; thermal ellipsoids at 50% probability.
Molecules 30 01893 g002
Figure 3. Nonbonding contacts for [Cu(LBSPh)2]; projection down the crystallographic c axis, highlighting the one-dimensional motif generated by the O1····H18, O1····H18A hydrogen bonds (in cyan) and propagating along the crystallographic b axis. The π····π pairing of the C15/C20, C15A/C20A rings with the corresponding symmetry-generated counterparts is also visible. Other contacts omitted for clarity.
Figure 3. Nonbonding contacts for [Cu(LBSPh)2]; projection down the crystallographic c axis, highlighting the one-dimensional motif generated by the O1····H18, O1····H18A hydrogen bonds (in cyan) and propagating along the crystallographic b axis. The π····π pairing of the C15/C20, C15A/C20A rings with the corresponding symmetry-generated counterparts is also visible. Other contacts omitted for clarity.
Molecules 30 01893 g003
Figure 4. Bacterial growth curves of E. coli after treatment with Schiff base ligands (1 and 2), metal complexes (310), and phosphane co-ligands (11 and 12).
Figure 4. Bacterial growth curves of E. coli after treatment with Schiff base ligands (1 and 2), metal complexes (310), and phosphane co-ligands (11 and 12).
Molecules 30 01893 g004
Figure 5. Bacterial growth curves of S. aureus after treatment with Schiff base ligands (1 and 2), metal complexes (310), and phosphane co-ligands (11 and 12).
Figure 5. Bacterial growth curves of S. aureus after treatment with Schiff base ligands (1 and 2), metal complexes (310), and phosphane co-ligands (11 and 12).
Molecules 30 01893 g005
Figure 6. (a) The bacterial viability of compounds 112 (from left to right) with different concentration and (b) corresponding bacterial colony images of E. coli. Colonies were quantified with ImageJ software (1.50d) and presented as mean values from duplicate independent experiments. Error bars denote standard deviation. The significance level between the control and treated samples was established at p < 0.05.
Figure 6. (a) The bacterial viability of compounds 112 (from left to right) with different concentration and (b) corresponding bacterial colony images of E. coli. Colonies were quantified with ImageJ software (1.50d) and presented as mean values from duplicate independent experiments. Error bars denote standard deviation. The significance level between the control and treated samples was established at p < 0.05.
Molecules 30 01893 g006
Figure 7. (a) The bacterial viability of compounds 112 (from left to right) with different concentrations and (b) corresponding bacterial colony images of S. aureus. Colonies were quantified with ImageJ software (1.50d) and presented as mean values from duplicate independent experiments. Error bars denote standard deviation. The significance level between the control and treated samples was established at p < 0.05.
Figure 7. (a) The bacterial viability of compounds 112 (from left to right) with different concentrations and (b) corresponding bacterial colony images of S. aureus. Colonies were quantified with ImageJ software (1.50d) and presented as mean values from duplicate independent experiments. Error bars denote standard deviation. The significance level between the control and treated samples was established at p < 0.05.
Molecules 30 01893 g007
Figure 8. Bacterial growth curves of (a) E. coli and (b) S. aureus after treatment with ciprofloxacin. (c) Corresponding bacterial viability of E. coli (green) and S. aureus (yellow) after treatment with ciprofloxacin. (d) The summary of antimicrobial activity of ciprofloxacin.
Figure 8. Bacterial growth curves of (a) E. coli and (b) S. aureus after treatment with ciprofloxacin. (c) Corresponding bacterial viability of E. coli (green) and S. aureus (yellow) after treatment with ciprofloxacin. (d) The summary of antimicrobial activity of ciprofloxacin.
Molecules 30 01893 g008
Table 1. Relevant nonbonding interactions for [Cu(LBSPh)2] (10).
Table 1. Relevant nonbonding interactions for [Cu(LBSPh)2] (10).
A AtomD AtomP Atom aA····D (Å)A····D–P (°)Symmetry Op. b
O1H18C182.81171.83/2 − x, y, 1 − z
O1H18AC182.85159.93/2 − x, y, 1 − z
C15/C20H11C11 c2.78 d168.8 d1 − x, −1/2 + y, 1/2 − z
C15A/C20A H11C11 c2.78 d165.0 d1 − x, −1/2 + y, 1/2 − z
C1H12C12 c2.91147.71 − x, 1 − y, 1 − z
C2H12C12 c2.91145.71 − x, 1 − y, 1 − z
C15/C20C15/C20 e 4.20 f26.6 g1/2 − x, y, 1 − z
C15A/C20AC15A/C20A e 4.03 f23.4 g1/2 − x, y, 1 − z
a Atom to which D atom is bound; b symmetry operations related to asymmetric unit = x, y, z; c C-H····π interactions; d distances and angles referring to selected ring centroids; e π····π interactions; f distance between ring centroids; g angle between mean planes encompassing phenyl rings.
Table 2. MIC and MIC50 values of Schiff base ligands (1 and 2), metal complexes (310), and co-ligands (11 and 12) in-vitro antimicrobial screening.
Table 2. MIC and MIC50 values of Schiff base ligands (1 and 2), metal complexes (310), and co-ligands (11 and 12) in-vitro antimicrobial screening.
E. coliS. aureus
N.Tested CompoundsMICMIC50MICMIC50
1HLBSMe////
2HLBSPh////
3[Cu(HLBSMe)(PTA)2]PF6/0.057 (0.076)/0.116 (0.149)
4[Ag(HLBSMe)(PTA)]NO30.050 (0.090)0.012 (0.021)0.100 (0.181)0.061 (0.110)
5[Cu(LBSMe)2]/0.110 (0.210)/0.381 (0.738)
6[Cu(HLBSPh)(PPh3)2]PF6·2CH3CN////
7[Cu(HLBSPh)(PTA)2]PF6·2H2O/0.114 (0.130)/0.084 (0.100)
8[Ag(HLBSPh)(PPh3)2]NO3/0.098 (0.100)/0.080 (0.079)
9[Ag(HLBSPh)(PTA)]NO30.025 (0.040)0.010 (0.016)0.050 (0.081)0.036 (0.058)
10[Cu(LBSPh)2]////
11PPh3////
12PTA////
MIC and MIC50 calculated in mg mL−1 (mM in parentheses); “/”: more than 0.381 mg mL−1 and 0.738 mM.
Table 3. MBC and MBC50 values of Schiff base ligands (1 and 2), metal complexes (310), and co-ligands (11 and 12) in-vitro antimicrobial screening.
Table 3. MBC and MBC50 values of Schiff base ligands (1 and 2), metal complexes (310), and co-ligands (11 and 12) in-vitro antimicrobial screening.
E. coliS. aureus
N.Tested CompoundsMBCMBC50MBCMBC50
1HLBSMe////
2HLBSPh////
3[Cu(HLBSMe)(PTA)2]PF6/0.262 (0.350)/0.642 (0.858)
4[Ag(HLBSMe)(PTA)]NO30.200 (0.362)0.048 (0.086)0.200 (0.362)0.062 (0.112)
5[Cu(LBSMe)2]/0.243 (0.474)/0.578 (1.128)
6[Cu(HLBSPh)(PPh3)2]PF6·2CH3CN/0.566 (0.513)/1.125 (1.020)
7[Cu(HLBSPh)(PTA)2]PF6·2H2O/0.319 (0.377)/0.551 (0.651)
8[Ag(HLBSPh)(PPh3)2]NO3/0.322 (0.327)/0.294 (0.299)
9[Ag(HLBSPh)(PTA)]NO30.100 (0.163)0.046 (0.075)0.200 (0.326)0.068 (0.110)
10[Cu(LBSPh)2]/0.282 (0.443)/0.875 (1.375)
11PPh3////
12PTA////
MBC and MBC50 calculated in mg mL−1 (mM in parentheses); “/”: more than 1.125 mg mL−1 and 1.375 mM.
Table 4. Summary of crystal and data collection parameters for [Cu(LBSPh)2] (10).
Table 4. Summary of crystal and data collection parameters for [Cu(LBSPh)2] (10).
Empirical FormulaC40H32CuN2O2
Formula weight636.21
Temperature/K295.1(3)
RadiationCu Kα (λ = 1.54184)
Crystal systemmonoclinic
Space groupI2/a
a20.4267(4)
b8.7751(2)
c18.3240(3)
α90
β101.135(2)
γ90
Volume/Å33222.68(11)
Z4
ρcalc g/cm31.311
μ/mm−11.246
F (000)1324.0
Crystal size/mm30.42 × 0.4 × 0.2
2θ range for data collection/°8.824 to 155.422
Index ranges−20 ≤ h ≤ 25, −11 ≤ k ≤ 11, −23 ≤ l ≤ 23
Reflections collected36,327
Independent reflections3416 [Rint = 0.0377]
Data/restraints/parameters3416/0/247
Goodness-of-fit on F21.034
Final R indexes [I ≥ 2σ (I)]R1 = 0.0331, wR2 = 0.0927
Largest diff. peak/hole/e Å−30.17/−0.31
Goodness–of–fit = [Σ (w (Fo2Fc2)2]/(NobsvnsNparams)]1/2, based on all data; R1 = Σ ||Fo| − |Fc||/Σ |Fo|; wR2 = [Σ w (Fo2Fc2)2w (Fo2)2]1/2.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Caviglia, M.; Li, Z.; Santini, C.; Del Gobbo, J.; Cimarelli, C.; Du, M.; Dolmella, A.; Pellei, M. New Cu(II), Cu(I) and Ag(I) Complexes of Phenoxy-Ketimine Schiff Base Ligands: Synthesis, Structures and Antibacterial Activity. Molecules 2025, 30, 1893. https://doi.org/10.3390/molecules30091893

AMA Style

Caviglia M, Li Z, Santini C, Del Gobbo J, Cimarelli C, Du M, Dolmella A, Pellei M. New Cu(II), Cu(I) and Ag(I) Complexes of Phenoxy-Ketimine Schiff Base Ligands: Synthesis, Structures and Antibacterial Activity. Molecules. 2025; 30(9):1893. https://doi.org/10.3390/molecules30091893

Chicago/Turabian Style

Caviglia, Miriam, Zhenzhen Li, Carlo Santini, Jo’ Del Gobbo, Cristina Cimarelli, Miao Du, Alessandro Dolmella, and Maura Pellei. 2025. "New Cu(II), Cu(I) and Ag(I) Complexes of Phenoxy-Ketimine Schiff Base Ligands: Synthesis, Structures and Antibacterial Activity" Molecules 30, no. 9: 1893. https://doi.org/10.3390/molecules30091893

APA Style

Caviglia, M., Li, Z., Santini, C., Del Gobbo, J., Cimarelli, C., Du, M., Dolmella, A., & Pellei, M. (2025). New Cu(II), Cu(I) and Ag(I) Complexes of Phenoxy-Ketimine Schiff Base Ligands: Synthesis, Structures and Antibacterial Activity. Molecules, 30(9), 1893. https://doi.org/10.3390/molecules30091893

Article Metrics

Back to TopTop