Next Article in Journal
Obtaining Granules from Waste Tannery Shavings and Mineral Additives by Wet Pulp Granulation
Next Article in Special Issue
Improvement in the Pharmacological Profile of Copper Biological Active Complexes by Their Incorporation into Organic or Inorganic Matrix
Previous Article in Journal
The Effect of Filtration on Physical and Chemical Properties of Osmo-Dehydrated Material
Previous Article in Special Issue
Mononuclear Perfluoroalkyl-Heterocyclic Complexes of Pd(II): Synthesis, Structural Characterization and Antimicrobial Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rare-Earth Metal Complexes of the Antibacterial Drug Oxolinic Acid: Synthesis, Characterization, DNA/Protein Binding and Cytotoxicity Studies

by
Ana-Madalina Maciuca
1,
Alexandra-Cristina Munteanu
1,*,
Mirela Mihaila
2,
Mihaela Badea
3,
Rodica Olar
3,
George Mihai Nitulescu
4,
Cristian V. A. Munteanu
5,
Marinela Bostan
2 and
Valentina Uivarosi
1,*
1
Department of General and Inorganic Chemistry, Faculty of Pharmacy, Carol Davila University of Medicine and Pharmacy, 6 Traian Vuia St, 020956 Bucharest, Romania
2
Center of Immunology, Stefan S. Nicolau Institute of Virology, 285 Mihai Bravu Ave, 030304 Bucharest, Romania
3
Department of Inorganic Chemistry, Faculty of Chemistry, University of Bucharest, 90-92 Panduri Str, 050663 Bucharest, Romania
4
Department of Pharmaceutical Chemistry, Faculty of Pharmacy, Carol Davila University of Medicine and Pharmacy, 6 Traian Vuia Str, 020956 Bucharest, Romania
5
Department of Bioinformatics and Structural Biochemistry, Institute of Biochemistry of the Romanian Academy (IBRA), 296 Spl. Independenţei, 060031 Bucharest, Romania
*
Authors to whom correspondence should be addressed.
Molecules 2020, 25(22), 5418; https://doi.org/10.3390/molecules25225418
Submission received: 22 October 2020 / Revised: 14 November 2020 / Accepted: 17 November 2020 / Published: 19 November 2020
(This article belongs to the Special Issue New Trends in Developing Complexes as Biological Active Species)

Abstract

:
“Drug repositioning” is a current trend which proved useful in the search for new applications for existing, failed, no longer in use or abandoned drugs, particularly when addressing issues such as bacterial or cancer cells resistance to current therapeutic approaches. In this context, six new complexes of the first-generation quinolone oxolinic acid with rare-earth metal cations (Y3+, La3+, Sm3+, Eu3+, Gd3+, Tb3+) have been synthesized and characterized. The experimental data suggest that the quinolone acts as a bidentate ligand, binding to the metal ion via the keto and carboxylate oxygen atoms; these findings are supported by DFT (density functional theory) calculations for the Sm3+ complex. The cytotoxic activity of the complexes, as well as the ligand, has been studied on MDA-MB 231 (human breast adenocarcinoma), LoVo (human colon adenocarcinoma) and HUVEC (normal human umbilical vein endothelial cells) cell lines. UV-Vis spectroscopy and competitive binding studies show that the complexes display binding affinities (Kb) towards double stranded DNA in the range of 9.33 × 104 − 10.72 × 105. Major and minor groove-binding most likely play a significant role in the interactions of the complexes with DNA. Moreover, the complexes bind human serum albumin more avidly than apo-transferrin.

Graphical Abstract

1. Introduction

In the context of the emergent resistance of bacteria strains and cancer cell lines to currently approved drugs, one strategy which has proved successful is that of “drug repositioning”. This refers to a process based on finding new applications for existing, failed, no longer in use or abandoned drugs [1]. The growing interest in synthesizing new metal complexes of different antibiotics is due to their new found pharmacological and toxicological properties [2,3].
Quinolones represent a group of synthetic antibacterial agents characterized by a broad spectrum of activity and good bioavailability [4,5] used in the treatment of infections with various localizations [6]. The mechanism is a complex one and implies binding to topoisomerase II (DNA-gyrase) [7,8] and topoisomerase IV [9,10]. This process induces bacteriostasis [11], double-strand breaks, chromosome fragmentation, formation of reactive oxygen species (ROS) [11,12] and eventually cell death. Mg2+ ions [13] play a key role in the formation of the quinolone-bacterial DNA-enzyme complex: they coordinate two oxygen atoms from the quinolone molecule and four water molecules in an octahedral manner [14], thus stabilizing the whole structure. Quinolones have a series of properties that make them suitable for repositioning as anti-cancer drugs, such as: immunomodulation [15], pro-apoptotic and antiproliferative properties [16,17,18,19,20,21] and anti-metastatic potential [22,23].
Oxolinic acid (5-ethyl-5,8-dihydro-8-oxo-l,3 dioxolo[4,5-g]quinoline-7-carboxylic acid, Figure 1A) is a first generation quinolone, developed by Warner-Lambert, approved by the FDA in the 1970s for the treatment of bacterial urinary infections not associated with bacteremia in adults [24]. The spectrum of activity is rather narrow and includes some Gram-negative bacteria [25,26,27]. Still, it has not proven itself to be more effective or better tolerated than other drugs such as nalidixic acid, ampicillin or trimethoprim-sulfamethoxazole [24,28,29]. Evidence shows that the efficacy of the treatment is not dose-related [24] and there have been some observations about the drug’s natural tendency to give rise to resistant mutants at a daily dose of less than 2 g/day [30]; these are aspects that have limited the use of oxolinic acid in humans. Nevertheless, it has been approved for use in the treatment of enteric infections in various animal species such as poultry, pigs, calves and fish [31] and for the management of pine wilt disease [32].
Several features in their structures (see Figure 1B) render quinolones good candidates for metal binding: the carboxyl moiety in position 3, the carbonyl moiety in position 4 and one or both nitrogen atoms from the piperazinyl ring in position 7, in the case of fluoroquinolones; the working conditions and the nature of the metal [33,34,35] determine which of the moieties will be involved in complex formation. So far, a considerable number of quinolone-metal complexes have been synthesized and characterized from the chemical and biological point of view; evidence that these complexes bind to calf thymus DNA [36,37,38,39] and that numerous complexes exhibit antibacterial [40,41,42,43], antitumoral [40,44,45], antifungal [40,45] and antiparasitic [46] properties have led to a growing interest in these complexes.
An intention of repurposing quinolones as anti-cancer agents has been observed and it stands on the evidence that some quinolone-metal complexes present antitumoral activity against leukemia [47] and breast cancer cells [44,48]; gold and ruthenium complexes have been found to be effective against metastatic cancer cells [49], without being toxic to normal cells [50], and present varied mechanisms of action [1].
Lanthanide (Ln3+) complexes with different drugs are appealing due to several properties [51]: the ability to bind to serum albumin [52], the development of stable MRI contrast agents [53,54,55] and the ability to replace cations such as Ca2+, Mg2+ or Zn2+ as co-factors in metaloenzymes with hydrolytic activity [56,57,58,59,60]. The ability to bind to DNA [61,62] and their luminescent properties [63,64,65] render Ln3+ and their complexes as potential candidates for nucleic acid imaging [66]. Lanthanides present a high affinity for ligands with donor atoms in the order O > N > S and coordination numbers that vary from 6 to 12 [67].
Up to date, several metal complexes of oxolinic acid have been reported [68,69,70,71,72,73,74,75,76,77,78,79,80]. Oxolinic acid can be considered a good candidate for the formation of complexes with Ln3+ metal ions, due to its adjacent oxygen-bearing moieties and the oxophilicity of the Ln3+ ion. We present herein the synthesis and physicochemical characterization of six novel complexes with some rare-earth metal cations and the results of the biological studies.

2. Results and Discussion

2.1. Syntheses of the Complexes

We report the synthesis of six new complexes of M3+ (M = Y3+, La3+, Sm3+, Eu3+, Gd3+, Tb3+) with oxolinic acid, in a molar ratio of 1:2. Briefly, the sodium salt of the ligand, together with the metal salt reacted at 150 °C in a microwave oven under continuous stirring (Scheme 1). Based on the experimental and some literature data [81,82], the following formulas were attributed to the complexes: [Y(oxo)(OH)(H2O)] (Y oxo), [La(oxo)(OH)(H2O)] (La oxo), [Sm(oxo)(OH)(H2O)] · H2O (Sm oxo), [Eu(oxo)(OH)(H2O)] (Eu oxo), [Gd(oxo)(OH)(H2O)]·0.5H2O (Gd oxo), [Tb(oxo)(OH)(H2O)] · 0.5 H2O (Tb oxo), where oxo is the deprotonated ligand.
The complexes were obtained as white powders and the solubility of the complexes was tested in various solvents of different polarities; the complexes are soluble in DMSO, slightly soluble in DMF and insoluble in methanol, ethanol, acetone, chloroform, diethyl ether, tetrahydrofuran and water.
Structural characterization included elemental analysis, spectroscopy studies (UV-VIS-NIR, FT-IR and high-resolution mass spectrometry) and thermal analysis.

2.2. Physicochemical Characterization of the Complexes

2.2.1. UV-Vis-NIR Spectra

UV-Vis-NIR spectra bring evidence of metal complex formation (Figure 2). The spectrum of the free ligand presents the characteristic absorption peaks in the 240–300 nm region, due to the absorption of the aromatic ring [83], and the 300–380 nm interval due to the n→π* (HOMO-LUMO) transitions [83]. Upon complexation, it can be observed that the peaks undergo a hypsochromic shift (~20 nm for the Y3+, La3+, Sm3+, Eu3+ complexes and 25 nm for the Gd3+ complex) due to metal quinolone interaction related to intra-ligand transitions [83]. Sm oxo presents specific absorption bands in the NIR region due to f-f transitions from the ground state 6H5/2 to 6F multiplet of Sm3+ ion (4f5) [84,85] (Figure 2). The wavelengths and corresponding absorbances are summarized in Table S1. The band at ~425 nm in the spectrum of Y oxo can be assigned to the ligand-metal charge transfer [86].

2.2.2. FT-IR Spectra

Coordination of the rare-earth metal ions by oxolinic acid was evaluated by means of Fourier transform infrared spectroscopy (Figures S1–S8) by comparing the spectra of the complexes with that of the free ligand; the obtained values are given in detail in Table 1. In order to determine the manner of coordination, specific peaks have been observed; changes in intensity and/or position are due to their implication in the chelation process [68,71,72,73]. The broad bands in the 3800–3300 cm−1 confirm the presence of water (lattice or coordinated) and a possible hydroxyl group coordinated to the metal. Two specific bands can be observed in the spectrum of the free ligand: at 1698 cm−1 due to the stretching vibration of the νC=O bond of the carboxylic acid moiety (νC=Oc) and at 1632 cm−1 due to the stretching vibration of the pyridone bond (νC=Op) [87]. The disappearance of the first band and the shift of the second one to lower values varying from 1594 cm−1 (Y oxo) to 1562 cm−1 (Tb oxo) are indicative of a bidentate manner of coordination of the ligand, through one oxygen atom from the carboxyl moiety and one from the carbonyl group. Due to the deprotonation process, two new bands appear in the spectra of the complexes, which can be attributed to the νO-C-O asymmetric (νO-C-Oas ~ 1630 cm−1) and νO-C-O symmetric (νO-C-Os-~1340 cm−1) stretching vibrations. In order to evaluate the binding mode of the carboxylate moiety to the metal, the Δ (νO-C-Oas - νO-C-Os) value was calculated and compared with the one corresponding to sodium oxalinate; these values vary from 292 (for Y oxo, La oxo, Gd oxo) to 295 (Tb oxo) (Table 1). The Δ values are higher than 200 and indicate that the carboxylate moiety coordinates the metal ion in a monodentate way. Overall, the oxolinic acid acts as a bidentate ligand through the carboxylic and the pyridonic oxygen atoms [88]. In the 500–400 cm−1 region of the spectra of the complexes (Table 1) new bands can be observed which can be attributed to the stretching vibration of the metal-oxygen bond [87], since they do not appear in the spectra of the free ligand or the sodium salt, confirming, once more, that the chelating process takes place. The complexes present similar wavenumbers for the studied vibrations, confirming that they have similar structures.

2.2.3. Mass Spectra

A typical isotopic distribution for ions of the type [M(oxo)2(DMSO)2]+ can be observed in the mass spectra (Figures S9–S13); due to the sample treatment (the solid powders were dissolved in DMSO and then diluted to a working concentration in methanol) used prior to the injection, the coordinated water molecule and hydroxyl group have been replaced by two DMSO molecules. Generally, we also observed peaks corresponding to a gradual loss of the DMSO molecules in the MS/MS spectra of some complexes such as Sm oxo, Eu oxo or Gd oxo (Figures S10–S12). Peak assignments are given in the Experimental section.

2.2.4. Thermal Behavior

The study of thermal behavior of complexes is useful for elucidating certain aspects about their composition, namely the number and nature of water molecules and the stoichiometry [89,90]. The analysis of the thermal decomposition curves brings additional confirmation concerning the structural composition of the complexes. The decomposition process is complex for all species and consists of water elimination as well as the oxidative degradation of the coordinated oxolinate anion (Table 2). The thermogravimetric (TG) curves confirm the number of water molecules from the structure of the metal complexes, both crystallization and coordination water molecules being eliminated together. In accordance with both derivative thermogravimetry (DTG) and differential thermal analysis (DTA) curves, the oxidative degradation of the organic part occurs in several overlapped exothermic processes, as can be seen in Figure 3 for Sm oxo. The overall mass loss is consistent with M2O3 stabilization as final residue, with the exception of Tb oxo, for which Tb4O7 is formed.

2.3. DFT Calculations

Since all attempts to isolate single crystals of the complexes have failed, the geometry optimization for Sm oxo was carried out via density functional theory (DFT) studies. Basis set 6-31G(d,p) was used for all H, C and O atoms and the small core potential ECP52MWB was used for Sm, as previously reported [91]. The fully optimized geometry of the complex is shown in Figure 4. The selected bond lengths, angles and charge densities are presented in Table 3. The coordination sphere around the metal center in the complexes involve the following atoms: O3, O25 (belonging to the deprotonated carboxylic group), O2, O24 (belonging to the carbonyl groups) and O46, O48 (belonging to a hydroxyl group and a water molecule, respectively). The M-O bond lengths are comparable with those previously reported for other complexes with quinolones [91,92,93,94,95]: the bond distances between Sm-O2 and Sm-O3, Sm-O24 and Sm-O25 are similar to values of 2.517 and 2.506 Å, 2.582 and 2.519 Å, respectively; still, it can be observed how the bond between Sm3+ and the oxygen atoms from the carboxyl moieties are shorter than the ones Sm3+ forms with the pyridonic oxygens. The bonds length for Sm-O46 is 2.557 Å, while the one for Sm-O48 is 2.664 Å. The O-Sm-O bond angles vary from 53.623° (O46-Sm-O25) to 134.706° (O3-Sm-O48) and indicate a distorted octahedral geometry. The dihedral angles present values different from 0° or 180°, varying from the only positive value, 87.515 (O46-Sm-O25-C28), to the largest negative one, −130.417 (O46-Sm-O2-C10), indicating that the Sm3+ cation and the donating atoms do not lie in the same plane; an exception is made by the O48-Sm-O3-C6 angle which presents a value of −176.277°, close to −180° meaning that the atoms involved are in the same plane. The charges accumulated on the oxygen atoms of de ligand structure vary from −0.530 (O25) to −0.594 (O24).
Furthermore, we have used the calculated vibrational frequencies to validate the optimized structures of the complexes and to aid the assignment of the experimental IR results. The calculated frequencies agree with the experimental data. The predicted IR spectra for the complex are shown in Figure S5 and the selected data, as well as the assignments, are given in Table 4. In the 3400–3200 cm−1 spectral region of the experimental IR spectrum of Sm oxo, the O-H stretches give weak bands, which can be found in the predicted spectrum at 3617 (weak) and 3309 (strong) cm−1. Moreover, strong bands at 1634 cm−1 (experimental spectrum) and 1758 and 1695 cm−1 (predicted spectrum), respectively, are due to the νO-C-Oas and νO-C-Os stretching vibrations and confirm the involvement of the carboxylic moiety in coordination. The strong bands at 1576 cm−1 (experimental) and 1608 cm−1 (predicted spectrum), respectively, also confirm the involvement of the pyridonic carbonyl in metal binding.

2.4. Biological Studies

2.4.1. Cytotoxicity Studies

For the evaluation of the cytotoxic effects of the complexes and ligand two human cancer cell lines were used, namely MDA-MB 231 (human breast adenocarcinoma) and LoVo (human colon adenocarcinoma); a normal cell line, human umbilical vein endothelial cells (HUVEC), was used as a reference. The effects were compared to the ones produced by adriamycin (ADR) and Cisplatin (Cis-Pt), respectively, which were used as positive controls. After 24 h and 48 h, respectively, cell viability was evaluated through MTS assay (data shown in Figures S16–S18) and the IC50 values were calculated (Table 5).
In terms of IC50 values, Sm oxo and Tb oxo displayed similar activity to the positive control, cisplatin, on LoVo cells. All complexes, however, displayed higher cytotoxic activity as compared to the free ligand. Interestingly, when tested on MDA-MB 231 cells, oxolinic acid showed higher activity in comparison with La oxo, Eu oxo, Gd oxo, Tb oxo. Although all compounds displayed lower activity than the positive control, ADR, the lowest IC50 values on MDA-MB 231 cells have been obtained for Y oxo (33.22 ± 14.92 µM) and Sm oxo (40.42 ± 6.27 µM). Altogether, the most promising results have been obtained for Sm oxo, which was moderately active on both tested cell lines. Regarding the effects shown on the HUVEC cell line, the IC50 values of the complexes are indicative of a lower toxicity compared to Cis-Pt.

2.4.2. Studies on DNA Binding

The stability of the complexes in solution has been investigated by means of UV-visible spectroscopy. The samples were prepared in Tris-HCl buffer as all biological interactions (with DNA and proteins) are tested in this medium. The recorded spectra are shown in Figures S14 and S15. No noticeable changes in absorbance or number and position of the peaks occur (data presented in Table S2). Since significant variation of the absorbances was not observed, the complexes can be considered stable in the given tested solution and time frame.

UV-Vis Spectra

Metal complexes interact with the DNA macromolecule through three major processes: (i) through electrostatic interactions with the negatively charged sugar-phosphate structure, (ii) via minor or major groove binding, through van der Waals forces, hydrogen bonds, electrostatic forces, hydrophobic interactions and (iii) by intercalation between the base-pairs, processes stabilized by the π-π* stacking interactions between the aromatic systems of the ligand and the DNA bases. Effective binding results in modifications of the UV-Vis spectra of the DNA sample, such as: hypochromic, hypsochromic (blue-shift) or bathochromic (red-shift) changes [84,96,97]. Other oxolinic acid metal complexes have been reported to possess DNA-binding properties [71,72,73], hence it has been hypothesized that DNA could act as a target for these new complexes as well.
Changes in the UV spectra of the seven studied compounds (Figure 5 and Figure S19) upon the addition of increasing concentrations of CT-DNA have been analyzed. The maximum shift observed for the band centered at ~330.5 nm is of 2 nm (Figure S19—Gd oxo). The values of the binding constant (Kb shown in Table 6 and Figure S20) are useful in evaluating the magnitude of the binding strength; a maximum value is observed for Eu oxo (10.72 (± 2.47) × 105 L∙mol−1) with a decrease in the Kb values in the following order: Eu oxo > Tb oxo > Gd oxo > Y oxo > La oxo (2.57 (± 0.17) × 105 L∙mol−1 ), dropping to Sm oxo which presents a value of 9.33 (± 1.46) × 104 L∙mol−1 (Table 6). The free ligand presents a Kb value in the range of 104, suggesting that complexation slightly improves the affinity of oxolinic acid for DNA. Similar changes are observed in the UV-vis spectra, suggesting similar binding modes for all complexes. The Kb values suggest a relatively moderate binding of the complexes to CT-DNA [71]. These results suggest that all complexes interact with the DNA macromolecule, although they do not give information about the type of binding.

Competitive Binding Assay with Ethidium Bromide (EB) through Fluorescence Spectroscopy

The manner of binding to the DNA macromolecule can be investigated through competitive binding studies with ethidium bromide (3,8-diamino-5-ethyl-6-phenylphenanthridinium bromide), a fluorescent dye; due to its aromatic ring, it can intercalate between two adjacent base pairs in the DNA structure and form a complex which emits fluorescent light; displacing EB results in a quenching effect [84,98]. The Stern-Volmer constant (KSV) and the dissociation constant (K50) have been calculated from the resulting data.
The emission spectra of the EB-DNA system recorded in the absence and presence of increasing concentrations of the studied compounds are presented in Figure S21. Quenching of fluorescence can be observed for all the compounds. While the spectra suggest that displacement of EB from EB-DNA-system takes place (Table 6) [72], the KSV values (Figure S22) suggest a moderate interaction, most likely through groove binding; still, an intercalative binding mode cannot be completely ruled out. La oxo presents the highest KSV value (1.27 (±0.03) × 104), affinities dropping in the order of Eu oxo > Sm oxo > Gd oxo = Tb oxo > oxolinic acid > Y oxo (8.38 (±0.29) × 103). Interestingly, K50 values (Figure S23) suggest that the strongest interaction (resulting in the most stable complex) occurs between Tb oxo and DNA, with a value of 20.28 ± 1.15 µM; values rise in the following order: Y oxo < La oxo < Eu oxo < Sm oxo < oxolinic acid < Gd oxo (47.75 ± 25.22 µM).

2.4.3. Interactions with Human Serum Albumin (HSA) and Apo-Transferrin (apo-Tf)

More evidence emerges indicating that the main transport proteins for metal complexes in the human body are albumin and transferrin. Human serum albumin (HSA) is the most abundant protein in the human circulatory system, being responsible for 80% of the colloid osmotic pressure [99]; amongst proteins, it possesses the highest affinity for drugs, influencing their pharmacokinetic and pharmacodynamic profiles. Apo-transferrin (apo-Tf) is mainly involved in the transport of iron (mainly Fe3+) and various xenobiotics [100,101]. It exhibits a significant tropism for tumor cells on account of the numerous specific receptors expressed on their surface [100,102,103] and because of this it has been investigated as a potential drug carrier for targeted delivery into tumor cells [104].
Insight into the nature of the interactions between proteins and complexes can be obtained by analyzing their fluorescence spectra; the tryptophane residue in position 214 (Trp-214) is highly sensitive to changes in the polarity of the medium [84,105]; due to the tryptophane residue, a strong peak is observed at ~340 nm when HSA samples are excited at 280 nm. Similarly, when excited at 295 nm, apo-Tf samples present a maximum emission peak at ~327 nm. These peaks can undergo changes upon interaction with different compounds.

Studies Regarding the Fluorescence Quenching Mechanism

The HSA peak recorded at 335 nm suffered slight hypsochromic shifts (Figure 6 and Figure S24), indicating a diminished exposure of the Trp-214 residue to the solvent. A moderate quenching effect can be observed varying from ~47% (Gd oxo) to ~58% (Eu oxo). The apo-Tf peak recorded at 327 nm, on the other hand, suffers no shift (Figure 6 and Figure S25) and the observed quenching effect is smaller than that in the case of HSA, varying from 35 % (Y oxo) to 50% (Eu oxo). Judging by this observation, Eu oxo seems to present the highest affinity for both proteins; it is followed, in both cases, by Tb oxo (52.63% and 47.62% for HAS and apo-Tf, respectively). Peak areas decrease with increasing concentrations of the tested compounds (Figures S26 and S27). The effects are more prevalent for the complexes than for the free ligand, as well as for HSA as compared to apo-Tf, indicating a stronger affinity of the compounds for the former protein.
More in-depth information was obtained by building and analyzing Stern–Volmer plots (F0/F vs. [Q]) and modified Stern-Volmer plots (F0/(F0-F) vs. 1/[Q]) for both HSA (Figure S28) and apo-Tf (Figure S29) interactions; the KSV (Stern–Volmer quenching constant), Kq (bimolecular quenching constant) and Ka (association binding constant) were calculated and interpreted.
A number of processes can explain the observed quenching effect, such as molecular collision, energy transfer, excited-state reaction or ground-state complex formation. However, three main processes are relevant for small molecules: static quenching (which involves the formation of a nonfluorescent complex between the quencher and the fluorophore in the ground state), dynamic quenching (the excited state of the fluorophore loses energy via a collisional process with the quencher) and combined static and dynamic quenching.
The calculated values for KSV and Kq (Table 7) give information about the static or dynamic nature of the quenching phenomenon. In the case of apo-Tf, the KSV constant reaches its maximum value for Tb oxo (19.45 (± 1.10) × 104 M−1) and decreases in the order Gd oxo > oxolinic acid > Eu oxo > La oxo > Sm oxo > Y oxo (6.52 (± 0.56) × 104 M−1), while for HSA the order in which the values decrease is Eu oxo (11.25 (± 0.92) × 104 M−1) > Y oxo > La oxo > Sm oxo > Gd oxo > oxolinic acid (1.95 (± 0.38) × 104 M−1). The Kq values follow the same pattern and are higher than diverse kinds of quenchers for biopolymers fluorescence (2.0 × 1010 M−1s−1), suggesting a static quenching mechanism [106].
The association binding constant, Ka (M−1), and the number of binding sites, n, were also calculated using Equation (7) (Table 7); Ka values indicate good HSA and apo-Tf binding, but a higher affinity towards HSA. Tb oxo (6.83(±1.33) × 105 M−1) and Y oxo (4.40(±1.09) × 104 M−1) present the highest affinities towards HSA and apo-Tf, respectively; in the case of both proteins, the lowest Ka values (Figures S30 and S31) correspond to the free ligand and are 10 times smaller than the ones for the complexes, indicating a rise in affinity that comes with coordination. The number of binding sites (n) has been found to be ~1 for both HSA and apo-Tf and all of the tested compounds.
Interestingly, plots of F0/F vs. [Q] (the Stern–Volmer plots) present a slightly downward curvature, concave towards the x-axis in the case of Sm oxo interacting with HSA (Figure S28) and oxolinic acid, Gd oxo and Tb oxo when interacting with apo-Tf (Figure S29). This could be explained by a selective quenching phenomenon of accessible versus inaccessible or partially accessible tryptophane residue [107].
Kd constants (Figures S30 and S31) have been calculated from Equation (8) and have been found to be less than 9 µM for both HSA and apo-Tf. Most of the calculated Hill coefficient (n) values are close to the value of 2, indicating positively cooperative binding. However, there are several exceptions (e.g., Eu oxo interacting with both proteins) where the coefficient is closer to the value of 1, indicating a negatively cooperative binding [108]. Overall, the values obtained for the binding affinities (in the micromolar range) are optimal and indicate that the complexes can be carried by the serum proteins throughout the human body. Moreover, the low micromolar values obtained for Kd constants are not high enough to cause significant drug retention and, therefore, to decrease the concentration and distribution of the free drug species to the cellular targets, and, hence, their efficacy.

Studies on Conformational Changes of HSA and Apo-Tf Due to the Interaction with the Tested Compounds

Structural changes due to interaction with drug molecules can be studied by measuring the synchronous spectra; using the established wavelength intervals Δλ = 15 nm and Δλ = 60 nm, where Δλ = λem − λex, the tyrosine or tryptophane residues, respectively, can be highlighted. The tryptophane residue is sensitive to any change in hydrophobicity in its vicinity; any shift of the peak is due to a change in polarity of the medium surrounding the aminoacid: a blue shift is due to a reduction in exposure to the solvent, implying a more hydrophobic environment, whilst a red shift is determined by an increased exposure to the solvent and a more polar environment [105,109].
The synchronous spectra at Δλ = 15 nm and Δλ = 60 nm for HSA and apo-Tf, respectively, are shown in Figures S32–S35, respectively. There are no changes observed in the synchronous spectra recorded at Δλ = 15 nm for either of the two proteins. On the other hand, bathochromic shifts can be observed in the synchronous Δλ = 60 nm for both proteins; these changes imply that the proteins adopt a conformation which increases the exposure of the tryptophane residue to the solvent, creating a more polar environment in its vicinity.

3. Materials and Methods

3.1. Materials

All reagents and solvents were of analytical reagent grade and were used without further purification. Oxolinic acid, YCl3·6H2O, LaCl3, EuCl3·6H2O, GdCl3·6H2O, SmCl3·6H2O, TbCl3·6H2O, human transferrin, human serum albumin and double-stranded calf-thymus DNA were purchased from Sigma Aldrich Chemical Co. (Schnelldorf, Germany). The following abbreviations were used to designate signal intensity: b = broad, w = weak, m = medium, s = sharp.

3.2. Synthesis

The synthesis method requires two steps: (1) synthesis of the sodium salt of the oxolinic acid and (2) synthesis of the lanthanide complexes. For (1), oxolinic acid (0.9 mmol; 0.2351 g) were dissolved in NaOH solution (900 µL, 1 M), under continuous magnetic stirring and heating (50 °C); the solvent was removed, yielding a residual white solid of sodium oxolinate. Step (2) involved a microwave-assisted method: in a microwave vial, the sodium oxalinate, 0.3 mmols of hexahydrate (for the Y3+, 0.0910 g, Eu3+, 0.1099 g, Gd3+, 0.1115 g, Sm3+, 0.1094 g, and Tb3+, 0.1120 g, complexes) or anhydrous (for the La3+ complex, 0.0735 g) metal chloride and distilled water (2 mL) were added. The vile was kept in the microwave apparatus for 30 min, under stirring, at 165 °C. The yellow precipitates formed were filtered, washed with water (3 × 10 mL) and then dried in the dessicator.
  • Y oxo—Y(C26H20N2O10)(OH)(H2O), MW = 644.37 g/mol; Elemental analysis found (calculated): %C 48.97 (48.46), %H 3.53 (3.60), N% 3.89 (4.35); Molar conductance (DMSO, Ω−1∙cm2∙mol−1): 2.1; MS (ESI+): m/z: 761.04 ([Y(oxo)2(DMSO)2]+); UV-vis (nm): 1945, 345, 265; FT-IR (cm−1): 3724 (w, ν(O-H) coordinated water molecule), 3055 (w, (ν(O-H) lattice water), 2973 (w, ν(C-H)), 2924 (ν(C-H)), 1630 (s, ν(O-C-O)as), 1594 (s, ν(C=O)pyridone), 1338 (s, ν(O-C-O)s).
  • La oxo—La(C26H20N2O10)(OH)(H2O), MW = 694.37 g/mol; Elemental analysis found (calculated): %C 44.95 (44.97), %H 3.08 (3.34), N% 4.16 (4.03); Molar conductance (DMSO, Ω−1∙cm2∙mol−1): 6.6; MS (ESI+): m/z: 815.05 ([La(oxo)2(DMSO)2]+); UV-vis (nm): 1910, 345, 266; FT-IR (cm−1): 3392 (w, ν(O-H) coordinated water molecule), 3060 (w, ν(C-H)), 2982 (w, ν(C-H)), 1632 (s, ν(O-C-O)as), 1577 (s, ν(C=O)pyridone), 1340 (s, ν(O-C-O)s).
  • Sm oxo—[Sm(C26H20N2O10)(OH)(H2O) · H2O, MW = 723.84 g/mol; Elemental analysis found (calculated): %C 41.13 (43.14), %H 3.19 (3.48), N% 4.47 (3.87); Molar conductance (DMSO, Ω−1∙cm2∙mol−1): 2.7; MS (ESI+): m/z: 828.05 ([Sm(oxo)2(DMSO)2]+), 750.04 ([Sm(oxo)2(DMSO)]+), 672.02 ([Sm(oxo)2]+); UV-vis (nm): 1940, 1930, 1555, 1500, 1390, 1245, 1090, 345, 255; FT-IR (cm−1): 3345 (w, ν(O-H) coordinated water molecule), 2917 (w, ν(C-H)), 1634 (s, ν(O-C-O)as), 1576 (s, ν(C=O)pyridone), 1340 (s, ν(O-C-O)s).
  • Eu oxo—Eu(C26H20N2O10)(OH)(H2O), MW = 707.43 g/mol; Elemental analysis found (calculated): %C 43.91 (44.14), %H 2.92 (3.28), N% 4.24 (3.96); Molar conductance (DMSO, Ω−1∙cm2∙mol−1): 3.5; MS (ESI+): m/z: 829.05 ([Eu(oxo)2(DMSO)2]+), 751.04 ([Eu(oxo)2(DMSO)]+), 671.02 ([Eu(oxo)2]+); UV-vis (nm): 1945, 345, 270; FT-IR (cm−1): 3389 (w, ν(O-H) coordinated water molecule), 2984 (w, ν(C-H)), 1633 (s, ν(O-C-O)as), 1579 (s, ν(C=O)pyridone), 1340 (s, ν(O-C-O)s).
  • Gd oxo—[Gd(C26H20N2O10)(OH)(H2O)] · 0.5H2O, MW = 721.72 g/mol; Elemental analysis found (calculated): %C 43.52 (43.26), %H 3.17 (3.35), N% 4.17 (3.88); Molar conductance (DMSO, Ω−1∙cm2∙mol−1): 2.1; MS (ESI+): m/z: 834.05 ([Gd(oxo)2(DMSO)2]+), 756.04 ([Gd(oxo)2(DMSO)]+), 676.03 ([Gd(oxo)2]+); UV-vis (nm): 1950, 1940, 340, 260; FT-IR (cm−1): 3378 (w, ν(O-H) coordinated water molecule), 3059 (w, (ν(O-H) lattice water), 1633 (s, ν(O-C-O)as), 1579 (s, ν(C=O)pyridone), 1341 (s, ν(O-C-O)s).
  • Tb oxo—[Tb(C26H20N2O10)(OH)(H2O)] ·0.5H2O, MW = 723.40 g/mol; Elemental analysis found (calculated): %C 43.31 (43.16), %H 3.39 (3.35), N% 4.18 (3.87); Molar conductance (DMSO, Ω−1∙cm2∙mol−1): 3.2; MS (ESI+): m/z: 835.07 ([Tb(oxo)2(DMSO)2]+); UV-vis: 1945, 380; FT-IR (cm−1): 3391 (w, ν(O-H) coordinated water molecule), 3057 (w, (ν(O-H) lattice water), 2976 (w, ν(C-H)), 2921 (w, ν(C-H)), 1631 (s, ν(O-C-O)as), 1562 (s, ν(C=O)pyridone), 1336 (s, ν(O-C-O)s).

3.3. Physico-Chemical Characterization

Elemental analysis of C, H and N was performed on a PE 2400 analyzer (Perkin Elmer) (Billerica, MA, USA). Infrared spectra were recorded on a FT-IR VERTEX 70 spectrometer (Bruker: Billerica, MA, USA), using KBr pellets. UV-visible-NIR spectra were recorded on solid probes, by diffuse reflectance method, using spectralon as a reference sample and without dilution, in the range of 200–2000 nm, on a V-670 spectrophotometer (Jasco: Tokyo, Japan). The conductivity was measured with a Consort C830 (Turnhout, Belgium) conductometer with an SK10T platinum electrode embedded in glass (cell constant 1.0 cm−1) for 10−3 M solutions in DMSO. Fluorescence spectra were recorded using a Jasco FP 6500 spectrofluorometer. Thermal behavior (TG and DTA) was analyzed using a Labsys 1200 Setaram instrument, with a sample average mass of 12 mg, over the temperature range of 20–1000 °C, with a heating rate of 10 °C/min; measurements were carried out in a synthetic air atmosphere with a flow rate of 16.66 cm3/min by using alumina crucible. For high-resolution mass spectrometry (HRMS) the complexes were dissolved in DMSO and 1:10 (v:v) dilutions in MeOH were injected and analyzed on LTQ-Orbitrap Velos Pro with a nanoESI interface. The instrument was controlled using LTQ Tune v2.7 with manual data acquisition.

3.4. Geometry Optimization Study

With the aid of the Gaussian 09 software (version A.02) [110], the optimized structure of Sm oxo were obtained using the B3LYP DFT functional; the small core potential ECP52MWB was used for Sm and 6-31G(d,p) basis set was used for all H, C, O atoms. Doublet spin multiplicity was used, and the self-consistent field (SCF) method was employed in order to obtain the optimized geometries corresponding to the minimum on the potential energy surface. The optimized geometries were computed under no symmetry restrictions and the absence of imaginary frequencies in the vibrational analysis was verified in order to confirm that each structure corresponds to a true minimum. Gauss View 6.0.16 [111] has been used for data visualization and Mercury 4.0 [112] software for image capture.3.4. Biological studies.

3.4.1. Cytotoxicity

The cytotoxicity assay is based on the ability of metabolically active cells to reduce MTS, a yellow tetrazolium salt, to the colored formazan, soluble in the culture medium; the concentration of formazan can be spectrophotometrically determined. A total of 15 × 103 cells/well were cultured in 100 µL culture medium; culture supernatants were discarded and then cells were treated for 24 h and 48 h, respectively, with increasing concentrations of drugs. At the end of the incubation time, 20 µL reagent were added to each well, reagent containing MTS, [3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxy-phenyl)-2-(4-sulfophenyl)-2H-tetrazolium, inner salt], and phenazine ethosulfate (PES); PES is a cationic dye with high chemical stability that binds to MTS and forms a stable solution. The plates were incubated with the coloring solution for 4 h at 37 °C, while being mildly agitated every 15 min. The reduced MTS to formazan was spectrophotometrically measured at λ = 492 nm. The test was run on two cancer cell lines, LoVo and MDA-MB 231, and on a normal cell line, HUVEC. Data were expressed as cell viability in comparison with untreated cells considered 100% viable, using the following formula:
Cell   viability   ( % ) = absorbance   of   treated   cells     absorbance   of   culture   medium absorbance   of   untreated   cells   absorbance   of   culture   medium   ×   100
Cell viability data were obtained in triplicate (n = 3) and expressed as mean values ± standard deviations (SD). Statistical analysis was carried out using one-way analysis of variance (ANOVA) and Tukey’s test; a p value ≤ 0.05 was considered statistically significant. A parallel assay for the evaluation of DMSO cytotoxicity was conducted using the same experimental conditions; no cell cytotoxicity was observed for concentrations lower than 1%.
All assays were performed in 96-well microtiter plates with flat bottom (Falcon), using CellTiter 96 Aqueous One Solution Cell Proliferation Assay (Promega), an MTS-based colorimetric assay. The formazan was spectrophotometrically measured using the Tecan GENios Spectrophotometer.

3.4.2. Studies on DNA Binding

The stability of the new complexes was tested before the study of interactions with DNA and proteins. An amount of 10−3 M solutions of the tested compounds were prepared in DMSO; further dilution of the samples was carried out in Tris-HCl buffer in order to obtain solutions of 40 μM. The UV-visible spectra were recorded at different time intervals (t = 0, 10, 20, 30, 60, 120 min.), in order to observe the possible variation of the absorbances; a control sample containing only DMSO and buffer was used.

UV-Vis Spectra

DNA binding studies were carried out by recording the UV-Vis spectra of the 1.8 mL samples containing 20 µM of the studied complex or free ligand and increasing amounts (0–40 µM with an increment of 5 µM) of DNA stock solution. The samples were prepared in Tris-HCl buffers (5 mM Tris-HCl/50 mM NaCl with a pH adjusted to 7.4 using a 0.1 M HCl solution). The concentration of the DNA stock solution was determined by measuring the absorption intensity at 260 nm and using a value of 6600 M−1 cm−1 for the molar extinction coefficient. Another necessary step was confirming that the DNA stock solution was sufficiently free of protein by calculating the ratio between the absorption at 260nm and 280nm, respectively, a ratio which had to be between 1.8 and 2 [98]; the value obtained was 1.8. These studies were carried out using a Jasco 650 spectrophotemeter, in 1 cm quartz cells, at room temperature. Three control samples were used, one containing DNA and DMSO, one with the studied complex and the last one with DMSO.
Equation (2) was used to calculate the binding constant (Kb).
The equation used to evaluate the data is:
[ D N A ] ε a ε f = [ D N A ] ε b ε f + 1 K b ( ε b ε f )
where [DNA] is the concentration of DNA in the sample, expressed in nucleobases, εa, εb, εf, are the apparent absorption coefficient, the extinction coefficient for the complex in totally bound form and the extinction coefficient for the free complex, respectively, and Kb is the intrinsic binding constant. Kb was determined from the ratio of the slope (1/(εbεf)) to the intercept (1/Kb(εbεf)) [113].

Competitive Binding Assay with Ethidium Bromide (EB) through Fluorescence Spectroscopy

Competitive binding studies were carried out on a Jasco FP 6500 fluorometer at room temperature, using quartz cells of 1 cm path-length. The fluorescence spectra were recorded on 1.6 mL samples containing 10 µM CT-DNA and 2 µM EB in the absence and presence of increasing amounts (0–40 µM with an increment of 5 µM) of complexes and free ligands, respectively, after being kept in contact for 10 min at room temperature, under continuous stirring and in the dark. The excitation wavelength used was 500 nm and the spectra was recorded in the 520–800 nm range. Two control samples were used, the first one containing DMSO, DNA and EB and the second one just the complex.
Data from the assay can be interpreted using the classical Stern-Volmer equation (Equation (3)) [114]:
F 0 F = 1 +   K S V · [ Q ]
where F0 and F are the fluorescence intensities of the EB-DNA samples in the absence and presence of the tested compounds and [Q] is the concentration of the compound; KSV is the Stern-Volmer constant and can be calculated from the plot of F0/F vs. [Q] as the slope.
The ratio of binding relative to the control sample exposed only to DMSO was plotted against the logarithm of concentration. The data points were fitted using the modified Hill function with offset (Origin®) (Equation (4)):
y = A 1 + ( A 2 A 1 ) × x n K 50 n + x n
where y is the ratio of binding (F/F0) and x is the concentration, A1 is the minimum of y values, A2 is the maximum of y values, K50 is the concentration corresponding to 50% binding and n is the Hill coefficient [115].

3.4.3. Interactions with Human Serum Albumin (HSA) and Apo-Transferrin (apo-Tf)

Studies on Fluorescence Quenching Mechanism

Fluorescence spectra of HSA and apo-Tf in the presence and absence of the studied compounds were recorded using a Jasco FP 6500 spectrofluorometer, in quartz cells of 1 cm path-length. Some 2.5 mL samples containing 2.5 µM HSA or 1 µM apo-Tf, respectively, were titrated by successively adding 5 µL of the tested compound, corresponding to an increment of 1 µM, ranging from 0–8 µM. The samples were shaken for 3 min at room temperature and the corresponding spectra were recorded in the emission wavelength ranges 300–500 nm (HSA) and 305–500 nm (apo-Tf), using the excitation wavelengths of 280 nm (HSA) and 295 nm (apo-Tf). Experiments were carried out in Tris-HCl buffer at pH = 7.4 (5 mM Tris-HCl/50 mM NaCl for HSA and 5 mM Tris-HCl/50 mM NaCl/25 mM NaHCO3 for apo-Tf, respectively). Solutions containing the protein and DMSO and, respectively, the tested compounds alone, at the concentrations used in this study, were used as controls. The modified Stern-Volmer equation has been used to calculate the Stern-Volmer constant (Equation (5)):
F 0 F 0 F = 1 f a   K S V [ Q ] +   1 f a
where F0 and F are the fluorescence intensities in the absence and presence, respectively, of the tested compounds, fa is the fraction of fluorophore accessible to the quencher (the tested compounds), [Q] is the concentration of the quencher and KSV (M−1) is the Stern-Volmer constant; from the plot of F0/(F0-F) against 1/[Q], KSV and fa can be determined as the slope and the y intercept, respectively (Equation (6)):
K S V =   K q τ 0
where Kq (M−1 s−1) is the bimolecular quenching rate constant and τ0 is the lifetime of the fluorophore in the absence of the studied compounds [109,116]. The value of Kq is useful when trying to determine whether the type of mechanism underlying the quenching phenomenon is static or dynamic [98].
The affinity constant (Ka) and number of binding sites (n) have been calculated using the following equation (Equation (7)):
lg F 0 F F =   lg K a + n lg Q
In order to calculate the binding affinity, Kd, the following quadratic function can be used (Equation (8)):
F 0 F F 0 F b =   [ P ] t + [ Q ] + K d ( [ P ] t + [ Q ] + K d ) 2 4 [ P ] t [ Q ] 2 [ P ] t
where F and F0 are the fluorescence intensities of the protein in the presence and absence of the quencher, respectively, Fb is the fluorescence of a fully bound protein–quencher, [P]t is the total concentration of the protein, and [Q] is the concentration of added quencher. To quantify the binding affinities, each binding isotherm was plotted as a function of bound to free protein (F/F0) versus the total concentration of the quencher in solution. Then, the isotherms were fitted to the following nonlinear regression function (Equation (9)):
F F 0 =   F 1 + ( F 2 F 1 ) K d n K d n + [ Q ] n
where n denotes the Hill coefficient, and F2 and F1 are the vertical and horizontal asymptotes, respectively. The Kd in Equation (9) is derived from the quadratic binding function (Equation (8)) [115]. Data were obtained in triplicate, averaged and expressed as mean ± standard deviation (SD). All graphs fitting models were defined and developed applying OriginPro 2018.

Studies on Conformational Changes of HSA and Apo-Tf Conformation Due to the Interaction with the Tested Compounds

These studies were carried out by measuring the synchronous fluorescence spectra of the two proteins in the presence and absence of the studied complexes and free ligands; the spectra were recorded on the same samples as for the binding studies, but using Δλ = 60 nm and Δλ = 15 nm, respectively, in the 250–400 nm interval. Two control samples were used, one containing the studied protein and DMSO and the other containing the complex.

4. Conclusions

Six new neutral complexes of oxolinic acid with rare-earth metal ions have been synthesized and further characterized. Structural characterization has comprised data obtained by means of elemental analysis, spectroscopy studies (UV-VIS-NIR, FT-IR, high-resolution mass spectrometry) and thermal analysis. The cytotoxic effects of the complexes, as well as the free ligand, were tested on two cancer cell lines, MDA-MB 231 (human breast adenocarcinoma), and LoVo (human colon adenocarcinoma) and the normal human umbilical vein endothelial cells (HUVEC) cell line; the complexes display very low cytotoxic effects towards the HUVEC cell line and satisfactory anticancer activities, some of the complexes presenting IC50 values being lower than the positive controls.
Regarding the DNA interaction potential, Kb values are consistent with moderate interactions; competitive binding assay also suggests a moderate capacity for displacing EB from EB-DNA complex. HSA and apo Tf binding studies reveal good binding affinities for both proteins, with relatively high binding constants. The complexes displayed an overall higher affinity towards HSA, the interaction occurring via a static quenching mechanism.
The potential use as anticancer agents of the rare-earth metal complexes could be further explored in future studies by obtaining ternary complexes with oxolinic acid and various ligands suitable for DNA intercalation. This would be an opportunity to explore ways of altering the DNA-interaction mechanism, with the aim of improving their antitumoral properties. Another direction of expanding this study is to investigate the magnetic properties of the lanthanide complexes and the possibility to be used as MRI contrast agents for theranostic applications.

Supplementary Materials

The following are available online, Table S1: Wavelengths and absorbance (A) values observed in the UV-Vis-NIR spectra of oxolinic acid and complexes, Table S2: Absorbances of tested compounds in UV-vis, in DMSO-TrisHCl buffer, [compound]= 40μM, Figure S1: FT-IR spectrum of oxolinic acid, Figure S2: FT-IR spectrum of sodium oxolinate, Figure S3: FT-IR spectrum of Y oxo, Figure S4: FT-IR spectrum of La oxo, Figure S5: Experimental vs. predicted IR spectrum of Sm oxo, Figure S6: FT-IR spectrum of Eu oxo, Figure S7: FT-IR spectrum of Gd oxo, Figure S8: FT-IR spectrum of Tb oxo, Figure S9: MS spectrum for Y oxo (left) and La oxo (right), Figure S10: MS, MS/MS and MS/MS/MS spectra for Sm oxo, Figure S11: MS, MS/MS and MS/MS/MS spectra for Eu oxo, Figure S12: MS, MS/MS and MS/MS/MS spectra for Gd oxo, Figure S13: MS spectrum for Tb oxo, Figure S14: Stability assay-UV-vis spectra of the complexes in DMSO-Tris Cl buffer mixture, Figure S15: The variation of the absorbances of the studied complexes during the stability assay, in DMSO-Tris-HCl buffer mixture, Figure S16: Cell viability (%) after 24h of incubation with tested compounds: A. on MDA-MB 231 cell line, B. LoVo cell line, Figure S17: Cell viability (%) after 48h of incubation with tested compounds: A. on MDA-MB 231 cell line, B. LoVo cell line, Figure S18: Normal cell viability (%) after A. 24h and B. 48h of incubation with tested compounds, Figure S19: Absorption spectra of the tested compounds in the absence and presence of increasing amounts of DNA. [compound] = 20 μM; [DNA] = 0, 5, 10, 15, 20, 25, 30, 35, 40 μM. The arrows show the absorption changes upon increasing the DNA concentration, Figure S20: Representation of the DNA-binding constant (Kb) of the studied compounds, Figure S21: Fluorescence spectra of the EB - DNA system in the absence and presence of increasing amounts of the tested compounds. λex = 500 nm, [EB] = 2 μM, [DNA] = 10 μM, [compound] = 0, 5, 10, 15, 20, 25, 30, 35, 40 μM. Arrows indicate the changes in fluorescence intensities upon increasing the concentrations of the tested compounds, Figure S22: Graphical representation of the KSV constants of the studied compounds, Figure S23: Representation of the K50 constant of the studied compounds, Figure S24: Changes in fluorescence intensity of free HSA vs. compound-HSA systems; the black arrows indicate a decrease of the intensity upon the addition of increasing amounts of compound; [HSA] = 2.5 μM, [compound] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM, Figure S25: Changes in fluorescence intensity of free apo Tf vs. compound-apo Tf systems; the black arrows indicate a decrease of the intensity upon the addition of increasing amounts of compound. [apo-Tf] = 1 μM, [compound] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM, Figure S26: Changes in the peak areas for the compound-HSA systems. [HSA] = 2.5 μM, [compounds] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM, Figure S27: Variation of the apo-Tf fluorescence peak area upon adding increasing amounts of studied compounds. [apo-Tf] = 1 μM, [compound] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM, Figure S28: Classical (left) and modified (right) Stern-Volmer plots for each of the HSA interaction studies, Figure S29: Classical (left) and modified (right) Stern-Volmer plots for the apo-Tf interaction assay, Figure S30: Ka and Kd for each of the studied compound-HSA systems, Figure S31: Representation of Ka and Kd constants for each of the apo-Tf interaction systems, Figure S32: Synchronous spectra for the HSA interaction systems recorded at Δλ= 15 nm. [HSA] = 2.5 μM, [Sm oxo] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM, Figure S33: Synchronous spectra for the HSA interaction systems recorded at Δλ = 60 nm. [HSA] = 2.5 μM, [Sm oxo] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM, Figure S34: Synchronous spectra of the tested compounds - apo-Tf systems recorded at Δλ = 15 nm. [apo-Tf] = 1 μM, [compound] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM, Figure S35: Synchronous spectra of the tested compounds-apo-Tf interaction systems recorded at Δλ = 60 nm. [apo-Tf] = 1 μM, [compound] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM.

Author Contributions

Conceptualization, A.-M.M. and V.U.; Data curation, A.-M.M., A.-C.M., M.M., M.B. (Mihaela Badea), R.O., G.M.N. and C.M.; Formal analysis, A.-M.M., A.-C.M., M.M., R.O., G.M.N., C.M. and M.B. (Marinela Bostan); Funding acquisition, A.-C.M., M.B. (Mihaela Badea), C.M. and V.U.; Investigation, A.-M.M. and G.M.N.; Methodology, A.-M.M., M.M., C.V.A.M., M.B. (Marinela Bostan) and V.U.; Project administration, A.-C.M., C.V.A.M. and V.U.; Resources, M.B. (Mihaela Badea), R.O. and V.U.; Software, A.-C.M.; Supervision, A.-C.M. and V.U. All authors have read and agreed to the published version of the manuscript.

Funding

This paper received financial support from the National Council for Higher Education, Ministry of National Education (Romania) through grants number PN-III-P1-1.1-PD-2019-1242 and PN-III-P1-1.1-PD-2019-1312, contract no. PD 219/2020.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yadav, V.; Talwar, P. Repositioning of fluoroquinolones from antibiotic to anti-cancer agents: An underestimated truth. Biomed. Pharmacother. 2019, 111, 934–946. [Google Scholar] [CrossRef]
  2. Guerra, W.; Silva-Caldeira, P.P.; Terenzi, H.; Pereira-Maia, E.C. Impact of metal coordination on the antibiotic and non-antibiotic activities of tetracycline-based drugs. Coord. Chem. Rev. 2016, 327, 188–199. [Google Scholar] [CrossRef]
  3. Möhler, J.S.; Kolmar, T.; Synnatschke, K.; Hergert, M.; Wilson, L.A.; Ramu, S.; Elliott, A.G.; Blaskovich, M.A.T.; Sidjabat, H.E.; Paterson, D.L. Enhancement of antibiotic-activity through complexation with metal ions-Combined ITC, NMR, enzymatic and biological studies. J. Inorg. Biochem. 2017, 167, 134–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Zhang, G.-F.; Zhang, S.; Pan, B.; Liu, X.; Feng, L.-S. 4-Quinolone derivatives and their activities against Gram positive pathogens. Eur. J. Med. Chem. 2018, 143, 710–723. [Google Scholar] [CrossRef] [PubMed]
  5. Kocsis, B.; Domokos, J.; Szabo, D. Chemical structure and pharmacokinetics of novel quinolone agents represented by avarofloxacin, delafloxacin, finafloxacin, zabofloxacin and nemonoxacin. Ann. Clin. Microbiol. Antimicrob. 2016, 15, 34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Dhiman, P.; Arora, N.; Thanikachalam, P.V.; Monga, V. Recent advances in the synthetic and medicinal perspective of quinolones: A review. Bioorg. Chem. 2019, 92, 103291. [Google Scholar] [CrossRef] [PubMed]
  7. Nagaraja, V.; Godbole, A.A.; Henderson, S.R.; Maxwell, A. DNA topoisomerase I and DNA gyrase as targets for TB therapy. Drug Discov. Today 2017, 22, 510–518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Charrier, C.; Salisbury, A.-M.; Savage, V.J.; Duffy, T.; Moyo, E.; Chaffer-Malam, N.; Ooi, N.; Newman, R.; Cheung, J.; Metzger, R.; et al. Novel bacterial topoisomerase inhibitors with potent broad-spectrum activity against drug-resistant bacteria. Antimicrob. Agents Chemother. 2017, 61, e02100-16. [Google Scholar] [CrossRef] [Green Version]
  9. Kato, J.; Nishimura, Y.; Imamura, R.; Niki, H.; Hiraga, S.; Suzuki, H. New topoisomerase essential for chromosome segregation in E. coli. Cell 1990, 63, 393–404. [Google Scholar] [CrossRef]
  10. Idowu, T.; Schweizer, F. Ubiquitous nature of fluoroquinolones: The oscillation between antibacterial and anticancer activities. Antibiotics 2017, 6, 26. [Google Scholar] [CrossRef] [Green Version]
  11. Drlica, K.; Malik, M.; Kerns, R.J.; Zhao, X. Quinolone-mediated bacterial death. Antimicrob. Agents Chemother. 2008, 52, 385–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Dwyer, D.J.; Kohanski, M.A.; Hayete, B.; Collins, J.J. Gyrase inhibitors induce an oxidative damage cellular death pathway in Escherichia coli. Mol. Syst. Biol. 2007, 3, 91. [Google Scholar] [CrossRef] [PubMed]
  13. Blower, T.R.; Williamson, B.H.; Kerns, R.J.; Berger, J.M. Crystal structure and stability of gyrase-fluoroquinolone cleaved complexes from Mycobacterium tuberculosis. Proc. Natl. Acad. Sci. USA. 2016, 113, 1706–1713. [Google Scholar] [CrossRef] [Green Version]
  14. Wohlkonig, A.; Chan, P.F.; Fosberry, A.P.; Homes, P.; Huang, J.; Kranz, M.; Leydon, V.R.; Miles, T.J.; Pearson, N.D.; Perera, R.L.; et al. Structural basis of quinolone inhibition of type IIA topoisomerases and target-mediated resistance. Nat. Struct. Mol. Biol. 2010, 17, 1152–1153. [Google Scholar] [CrossRef] [PubMed]
  15. Dalhoff, A.; Shalit, I. Immunomodulatory effects of quinolones. Lancet. Infect. Dis. 2003, 3, 359–371. [Google Scholar] [CrossRef]
  16. Mukherjee, P.; Mandal, E.R.; Das, S.K. Evaluation of antiproliferative activity of enoxacin on a human breast cancer cell line. Int. J. Hum. Genet. 2005, 5, 57–63. [Google Scholar] [CrossRef]
  17. Yu, M.; Li, R.; Zhang, J. Repositioning of antibiotic levofloxacin as a mitochondrial biogenesis inhibitor to target breast cancer. Biochem. Biophys. Res. Commun. 2016, 471, 639–645. [Google Scholar] [CrossRef]
  18. Perucca, P.; Savio, M.; Cazzalini, O.; Mocchi, R.; Maccario, C.; Sommatis, S.; Ferraro, D.; Pizzala, R.; Pretali, L.; Fasani, E. Structure–activity relationship and role of oxygen in the potential antitumour activity of fluoroquinolones in human epithelial cancer cells. J. Photochem. Photobiol. B Biol. 2014, 140, 57–68. [Google Scholar] [CrossRef]
  19. Sousa, E.J.; Graca, I.; Baptista, T.; Vieira, F.Q.; Palmeira, C.; Henrique, R.; Jeronimo, C. Enoxacin inhibits growth of prostate cancer cells and effectively restores microRNA processing. Epigenetics 2013, 8, 548–558. [Google Scholar] [CrossRef] [Green Version]
  20. Aranha, O.; Grignon, R.; Fernandes, N.; McDonnell, T.J.; Wood, D.P.; Sarkar, F.H. Suppression of human prostate cancer cell growth by ciprofloxacin is associated with cell cycle arrest and apoptosis. Int. J. Oncol. 2003, 22, 787–794. [Google Scholar] [CrossRef]
  21. Aranha, O.; Wood, D.P.; Sarkar, F.H. Ciprofloxacin mediated cell growth inhibition, S/G2-M cell cycle arrest, and apoptosis in a human transitional cell carcinoma of the bladder cell line. Clin. Cancer Res. 2000, 6, 891–900. [Google Scholar] [PubMed]
  22. Chen, T.-C.; Hsu, Y.-L.; Tsai, Y.-C.; Chang, Y.-W.; Kuo, P.-L.; Chen, Y.-H. Gemifloxacin inhibits migration and invasion and induces mesenchymal–epithelial transition in human breast adenocarcinoma cells. J. Mol. Med. 2014, 92, 53–64. [Google Scholar] [CrossRef] [PubMed]
  23. Kan, J.-Y.; Hsu, Y.-L.; Chen, Y.-H.; Chen, T.-C.; Wang, J.-Y.; Kuo, P.-L. Gemifloxacin, a fluoroquinolone antimicrobial drug, inhibits migration and invasion of human colon cancer cells. Biomed Res. Int. 2013, 2013. [Google Scholar] [CrossRef] [PubMed]
  24. Gleckman, R.; Alvarez, S.; Joubert, D.W.; Matthews, S.J. Drug therapy reviews: Oxolinic acid. Am. J. Hosp. Pharm. 1979, 36, 1077–1079. [Google Scholar] [CrossRef]
  25. Pianotti, R.S.; Mohan, R.R.; Schwartz, B.S. Biochemical effects of oxolinic acid on Proteus vulgaris. J. Bacteriol. 1968, 95, 1622–1626. [Google Scholar] [CrossRef] [Green Version]
  26. Klein, D.; Matsen, J.M. In vitro susceptibility comparisons and recommendations for oxolinic acid. Antimicrob. Agents Chemother. 1976, 9, 649–654. [Google Scholar] [CrossRef] [Green Version]
  27. Neussel, H.; Linzenmeier, G. In vitro investigations with oxolinic acid, a new chemotherapeutic agent. Chemotherapy 1973, 18, 253–261. [Google Scholar] [CrossRef]
  28. Mannisto, P.T.; PT, M. Comparison of oxolinic acid, trimethoprim, and trimethoprim-sulfamethoxazole in the treatment and long-term control of urinary tract infection. Curr. Ther. Res. Clin. Exp. 1976, 20, 645–654. [Google Scholar]
  29. Ghatikar, K.N. A multicentric trial of a new synthetic antibacterial in urinary infections. Curr. Ther. Res. Clin. Exp. 1974, 16, 130. [Google Scholar]
  30. Clark, H.; Brown, N.K.; Wallace, J.F.; Turck, M. Emergence of resistant organisms as a function of dose in oxolinic acid therapy. Am. J. Med. Sci. 1971, 261, 145–148. [Google Scholar] [CrossRef]
  31. Barrón, D.; Jiménez-Lozano, E.; Bailac, S.; Barbosa, J. Simultaneous determination of flumequine and oxolinic acid in chicken tissues by solid phase extraction and capillary electrophoresis. Anal. Chim. Acta 2003, 477, 21–27. [Google Scholar] [CrossRef]
  32. Kwon, H.R.; Choi, G.J.; Choi, Y.H.; Jang, K.S.; Sung, N.; Kang, M.S.; Moon, Y.; Lee, S.K.; Kim, J. Suppression of pine wilt disease by an antibacterial agent, oxolinic acid. Pest Manag. Sci. Former. Pestic. Sci. 2010, 66, 634–639. [Google Scholar] [CrossRef] [PubMed]
  33. Sadeek, S.; El-Shwiniy, W.; El-Attar, M. Synthesis, characterization and antimicrobial investigation of some moxifloxacin metal complexes. Spectrochim. Acta. A Mol. Biomol. Spectrosc. 2011, 84, 99–110. [Google Scholar] [CrossRef] [PubMed]
  34. Li, Y.-X.; Chen, Z.-F.; Xiong, R.-G.; Xue, Z.; Ju, H.-X.; You, X.-Z. A mononuclear complex of norfloxacin with silver (I) and its properties. Inorg. Chem. Commun. 2003, 6, 819–822. [Google Scholar] [CrossRef]
  35. Gao, F.; Yang, P.; Xie, J.; Wang, H. Synthesis, characterization and antibacterial activity of novel Fe (III), Co (II), and Zn (II) complexes with norfloxacin. J. Inorg. Biochem. 1995, 60, 61–67. [Google Scholar]
  36. Song, G.; Yan, Q.; He, Y. Studies on interaction of norfloxacin, Cu2+, and DNA by spectral methods. J. Fluoresc. 2005, 15, 673. [Google Scholar] [CrossRef]
  37. Drevensek, P.; Turel, I.; Poklar Ulrih, N. Influence of copper(II) and magnesium(II) ions on the ciprofloxacin binding to DNA. J. Inorg. Biochem. 2003, 96, 407–415. [Google Scholar] [CrossRef]
  38. Guo, D.-S.; Jing, B.-Y.; Yuan, X.-Y. Influence of Mg+2 and Cu+2 on the interaction between quinolone and calf thymus DNA. J. Fluoresc. 2011, 21, 113–118. [Google Scholar] [CrossRef]
  39. Robles, J.; Martín-Polo, J.; Alvarez-Valtierra, L.; Hinojosa, L.; Mendoza-Díaz, G. A Theoretical-experimental study on the structure and activity of certain quinolones and the interaction of their Cu(II)-complexes on a DNA model. Met. Based. Drugs 2000, 7, 301–311. [Google Scholar] [CrossRef] [Green Version]
  40. Uivarosi, V. Metal complexes of quinolone antibiotics and their applications: An update. Molecules 2013, 18, 11153–11197. [Google Scholar] [CrossRef]
  41. Rusu, A.; Hancu, G.; Cristina Munteanu, A.; Uivarosi, V. Development perspectives of silver complexes with antibacterial quinolones: Successful or not? J. Organomet. Chem. 2017, 839, 19–30. [Google Scholar] [CrossRef]
  42. Saha, D.K.; Padhye, S.; Anson, C.E.; Powell, A.K. Hydrothermal synthesis, crystal structure, spectroscopy, electrochemistry and antimycobacterial evaluation of the copper (II) ciprofloxacin complex: [Cu(cf)2(BF4)2]·6H2O. Inorg. Chem. Commun. 2002, 5, 1022–1027. [Google Scholar] [CrossRef]
  43. El-Halim, H.F.A.; Mohamed, G.G.; El-Dessouky, M.M.I.; Mahmoud, W.H. Ligational behaviour of lomefloxacin drug towards Cr(III), Mn(II), Fe(III), Co(II), Ni(II), Cu(II), Zn(II), Th(IV) and UO2(VI) ions: Synthesis, structural characterization and biological activity studies. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2011, 82, 8–19. [Google Scholar] [CrossRef] [PubMed]
  44. Patitungkho, S.; Adsule, S.; Dandawate, P.; Padhye, S.; Ahmad, A.; Sarkar, F.H. Synthesis, characterization and anti-tumor activity of moxifloxacin-copper complexes against breast cancer cell lines. Bioorg. Med. Chem. Lett. 2011, 21, 1802–1806. [Google Scholar] [CrossRef] [PubMed]
  45. Batista, D.d.G.J.; da Silva, P.B.; Stivanin, L.; Lachter, D.R.; Silva, R.S.; Felcman, J.; Louro, S.R.W.; Teixeira, L.R.; Soeiro, M.d.N.C. Co(II), Mn(II) and Cu(II) complexes of fluoroquinolones: Synthesis, spectroscopical studies and biological evaluation against Trypanosoma cruzi. Polyhedron 2011, 30, 1718–1725. [Google Scholar] [CrossRef]
  46. Regiel-Futyra, A.; Dąbrowski, J.M.; Mazuryk, O.; Śpiewak, K.; Kyzioł, A.; Pucelik, B.; Brindell, M.; Stochel, G. Bioinorganic antimicrobial strategies in the resistance era. Coord. Chem. Rev. 2017, 351, 76–117. [Google Scholar] [CrossRef]
  47. Efthimiadou, E.K.; Thomadaki, H.; Sanakis, Y.; Raptopoulou, C.P.; Katsaros, N.; Scorilas, A.; Karaliota, A.; Psomas, G. Structure and biological properties of the copper (II) complex with the quinolone antibacterial drug N-propyl-norfloxacin and 2, 2′-bipyridine. J. Inorg. Biochem. 2007, 101, 64–73. [Google Scholar] [CrossRef]
  48. Shingnapurkar, D.; Butcher, R.; Afrasiabi, Z.; Sinn, E.; Ahmed, F.; Sarkar, F.; Padhye, S. Neutral dimeric copper–sparfloxacin conjugate having butterfly motif with antiproliferative effects against hormone independent BT20 breast cancer cell line. Inorg. Chem. Commun. 2007, 10, 459–462. [Google Scholar] [CrossRef]
  49. Lazarević, T.; Rilak, A.; Bugarčić, Ž.D. Platinum, palladium, gold and ruthenium complexes as anticancer agents: Current clinical uses, cytotoxicity studies and future perspectives. Eur. J. Med. Chem. 2017, 142, 8–31. [Google Scholar] [CrossRef]
  50. Gouvea, L.R.; Garcia, L.S.; Lachter, D.R.; Nunes, P.R.; de Castro Pereira, F.; Silveira-Lacerda, E.P.; Louro, S.R.W.; Barbeira, P.J.S.; Teixeira, L.R. Atypical fluoroquinolone gold (III) chelates as potential anticancer agents: Relevance of DNA and protein interactions for their mechanism of action. Eur. J. Med. Chem. 2012, 55, 67–73. [Google Scholar] [CrossRef]
  51. Măciucă, A.-M.; Munteanu, A.-C.; Uivarosi, V. Quinolone complexes with lanthanide ions: An insight into their analytical applications and biological activity. Molecules 2020, 25, 1347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Montgomery, C.P.; New, E.J.; Parker, D.; Peacock, R.D. Enantioselective regulation of a metal complex in reversible binding to serum albumin: Dynamic helicity inversion signalled by circularly polarised luminescence. Chem. Commun. (Camb.) 2008, 4261–4263. [Google Scholar] [CrossRef] [PubMed]
  53. Hermann, P.; Kotek, J.; Kubíček, V.; Lukeš, I. Gadolinium(III) complexes as MRI contrast agents: Ligand design and properties of the complexes. Dalt. Trans. 2008, 3027–3047. [Google Scholar] [CrossRef] [PubMed]
  54. Xiao, Y.-D.; Paudel, R.; Liu, J.; Cong, M.; Zhang, Z.-S.; Zhou, S.-K. MRI contrast agents: Classification and application (Review). Int. J. Mol. Med. 2016, 38, 1319–1326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Kim, H.-K.; Lee, G.H.; Chang, Y. Gadolinium as an MRI contrast agent. Future Med. Chem. 2018, 10, 639–661. [Google Scholar] [CrossRef] [PubMed]
  56. Franklin, S.J. Lanthanide-mediated DNA hydrolysis. Curr. Opin. Chem. Biol. 2001, 5, 201–208. [Google Scholar] [CrossRef]
  57. Komiyama, M.; Takeda, N.; Shigekawa, H. Hydrolysis of DNA and RNA by lanthanide ions: Mechanistic studies leading to new applications. Chem. Commun. 1999, 1443–1451. [Google Scholar] [CrossRef]
  58. Hussain, A.; Chakravarty, A.R. Photocytotoxic lanthanide complexes. J. Chem. Sci. 2012, 124, 1327–1342. [Google Scholar] [CrossRef] [Green Version]
  59. Liu, C.; Wang, M.; Zhang, T.; Sun, H. DNA hydrolysis promoted by di-and multi-nuclear metal complexes. Coord. Chem. Rev. 2004, 248, 147–168. [Google Scholar] [CrossRef]
  60. Sreedhara, A.; Cowan, J.A. Catalytic hydrolysis of DNA by metal ions and complexes. JBIC J. Biol. Inorg. Chem. 2001, 6, 337–347. [Google Scholar] [CrossRef]
  61. Xu, M.; Chen, F.-J.; Huang, L.; Xi, P.; Zeng, Z. Binding of rare earth metal complexes with an ofloxacin derivative to bovine serum albumin and its effect on the conformation of protein. J. Lumin. 2011, 131, 1557–1565. [Google Scholar] [CrossRef]
  62. Xu, M.; Zhang, Y.; Xu, Z.; Zeng, Z. Crystal structure, biological studies of water-soluble rare earth metal complexes with an ofloxacin derivative. Inorganica Chim. Acta 2012, 384, 324–332. [Google Scholar] [CrossRef]
  63. Fu, P.K.-L.; Turro, C. Energy transfer from nucleic acids to Tb (III): Selective emission enhancement by single DNA mismatches. J. Am. Chem. Soc. 1999, 121, 1–7. [Google Scholar] [CrossRef]
  64. Xu, L.; Zhou, L.; Chen, X.; Shen, X.; Wang, J.; Zhang, J.; Pei, R. Luminescence sensitization of Tb3+-DNA complexes by Ag+. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2017, 180, 85–90. [Google Scholar] [CrossRef]
  65. White, G.F.; Litvinenko, K.L.; Meech, S.R.; Andrews, D.L.; Thomson, A.J. Multiphoton-excited luminescence of a lanthanide ion in a protein complex: Tb3+ bound to transferrin. Photochem. Photobiol. Sci. 2004, 3, 47–55. [Google Scholar] [CrossRef]
  66. Lim, S.; Franklin, S.J. Lanthanide-binding peptides and the enzymes that might have been. Cell. Mol. Life Sci. C. 2004, 61, 2184–2188. [Google Scholar] [CrossRef]
  67. Cotton, S. Coordination chemistry of the lanthanides. In Lanthanide and Actinide Chemistry; Wiley Online Books; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2006; pp. 35–60. ISBN 9780470010082. [Google Scholar]
  68. Tarushi, A.; Lafazanis, K.; Kljun, J.; Turel, I.; Pantazaki, A.A.; Psomas, G.; Kessissoglou, D.P. First- and second-generation quinolone antibacterial drugs interacting with zinc(II): Structure and biological perspectives. J. Inorg. Biochem. 2013, 121, 53–65. [Google Scholar] [CrossRef]
  69. Skyrianou, K.C.; Perdih, F.; Papadopoulos, A.N.; Turel, I.; Kessissoglou, D.P.; Psomas, G. Nickel–quinolones interaction: Part 5—Biological evaluation of nickel (II) complexes with first-, second-and third-generation quinolones. J. Inorg. Biochem. 2011, 105, 1273–1285. [Google Scholar] [CrossRef]
  70. Zampakou, M.; Akrivou, M.; Andreadou, E.G.; Raptopoulou, C.P.; Psycharis, V.; Pantazaki, A.A.; Psomas, G. Structure, antimicrobial activity, DNA-and albumin-binding of manganese (II) complexes with the quinolone antimicrobial agents oxolinic acid and enrofloxacin. J. Inorg. Biochem. 2013, 121, 88–99. [Google Scholar] [CrossRef]
  71. Skyrianou, K.C.; Perdih, F.; Turel, I.; Kessissoglou, D.P.; Psomas, G. Nickel–quinolones interaction. Part 2—Interaction of nickel (II) with the antibacterial drug oxolinic acid. J. Inorg. Biochem. 2010, 104, 161–170. [Google Scholar] [CrossRef]
  72. Tarushi, A.; Psomas, G.; Raptopoulou, C.P.; Kessissoglou, D.P. Zinc complexes of the antibacterial drug oxolinic acid: Structure and DNA-binding properties. J. Inorg. Biochem. 2009, 103, 898–905. [Google Scholar] [CrossRef] [PubMed]
  73. Psomas, G.; Tarushi, A.; Efthimiadou, E.K.; Sanakis, Y.; Raptopoulou, C.P.; Katsaros, N. Synthesis, structure and biological activity of copper (II) complexes with oxolinic acid. J. Inorg. Biochem. 2006, 100, 1764–1773. [Google Scholar] [CrossRef] [PubMed]
  74. Mendoza-Díaz, G.; Martínez-Aguilera, L.M.R.; Moreno-Esparza, R.; Pannell, K.H.; Cervantes-Lee, F. Some mixed-ligand complexes of copper (II) with drugs of the quinolone family and (N-N) donors. Crystal structure of [Cu (phen)(Cnx)(H2O)] NO3· H2O. J. Inorg. Biochem. 1993, 50, 65–78. [Google Scholar] [CrossRef]
  75. Tarushi, A.; Efthimiadou, E.K.; Christofis, P.; Psomas, G. Neutral mononuclear dioxomolybdenum (VI) and dioxouranium (VI) complexes of oxolinic acid: Characterization and biological evaluation. Inorganica Chim. Acta 2007, 360, 3978–3986. [Google Scholar] [CrossRef]
  76. Tarushi, A.; Psomas, G.; Raptopoulou, C.P.; Psycharis, V.; Kessissoglou, D.P. Structure and DNA-binding properties of bis (quinolonato) bis (pyridine) zinc (II) complexes. Polyhedron 2009, 28, 3272–3278. [Google Scholar] [CrossRef]
  77. Irgi, E.P.; Geromichalos, G.D.; Balala, S.; Kljun, J.; Kalogiannis, S.; Papadopoulos, A.; Turel, I.; Psomas, G. Cobalt (II) complexes with the quinolone antimicrobial drug oxolinic acid: Structure and biological perspectives. RSC Adv. 2015, 5, 36353–36367. [Google Scholar] [CrossRef]
  78. Al-Saif, F.A.; Alibrahim, K.A.; Alfurhood, J.A.; Refat, M.S. Synthesis, spectroscopic, thermal, biological, morphological and molecular docking studies of the different quinolone drugs and their cobalt (II) complexes. J. Mol. Liq. 2018, 249, 438–453. [Google Scholar] [CrossRef]
  79. Tarushi, A.; Christofis, P.; Psomas, G. Synthesis, characterization and interaction with DNA of mononuclear metal complexes with oxolinic acid. Polyhedron 2007, 26, 3963–3972. [Google Scholar] [CrossRef]
  80. Kim, Y.-S.; Kim, K.M.; Song, R.; Jun, M.J.; Sohn, Y.S. Synthesis, characterization and antitumor activity of quinolone–platinum (II) conjugates. J. Inorg. Biochem. 2001, 87, 157–163. [Google Scholar] [CrossRef]
  81. Ahmed, I.; Atta, A.H.; Refat, M.S. Complexation of gadolinium (III) and terbium (III) with nalidixicacid (NDX): Molar conductivity, thermal and spectral investigation. Int. J. Electrochem. Sci. 2014, 9, 5187–5203. [Google Scholar]
  82. Na, B.; Zhang, X.; Shi, W.; Zhang, Y.; Wang, B.; Gao, C.; Gao, S.; Cheng, P. Six-coordinate lanthanide complexes: Slow relaxation of magnetization in the dysprosium (III) complex. Chem. Eur. J. 2014, 20, 15975–15980. [Google Scholar] [CrossRef] [PubMed]
  83. Neugebauer, U.; Szeghalmi, A.; Schmitt, M.; Kiefer, W.; Popp, J.; Holzgrabe, U. Vibrational spectroscopic characterization of fluoroquinolones. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2005, 61, 1505–1517. [Google Scholar] [CrossRef] [PubMed]
  84. Munteanu, A.-C.; Badea, M.; Olar, R.; Silvestro, L.; Dulea, C.; Negut, C.-D.; Uivarosi, V. Synthesis and structural investigation of new bio-relevant complexes of lanthanides with 5-hydroxyflavone: DNA binding and protein interaction studies. Molecules 2016, 21, 1737. [Google Scholar] [CrossRef] [Green Version]
  85. Sastri, V.R.; Perumareddi, J.R.; Rao, V.R.; Rayudu, G.V.S.; Bünzli, J.-C. Modern Aspects of Rare Earths and Their Complexes; Elsevier: Amsterdam, The Netherlands, 2003; ISBN 0080536689. [Google Scholar]
  86. Sadeek, S.A.; El-Shwiniy, W.H. Metal complexes of the fourth generation quinolone antimicrobial drug gatifloxacin: Synthesis, structure and biological evaluation. J. Mol. Struct. 2010, 977, 243–253. [Google Scholar] [CrossRef]
  87. Sadeek, S.A.; El-Shwiniy, W.H.; Zordok, W.A.; El-Didamony, A.M. Spectroscopic, structure and antimicrobial activity of new Y(III) and Zr(IV) ciprofloxacin. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2011, 78, 854–867. [Google Scholar] [CrossRef] [PubMed]
  88. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds, Part B: Applications in Coordination, Organometallic, and Bioinorganic Chemistry; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2009; ISBN 047174493X. [Google Scholar]
  89. Uivarosi, V.; Barbuceanu, S.F.; Aldea, V.; Arama, C.-C.; Badea, M.; Olar, R.; Marinescu, D. Synthesis, spectral and thermal studies of new rutin vanadyl complexes. Molecules 2010, 15, 1578–1589. [Google Scholar] [CrossRef] [PubMed]
  90. Badea, M.; Olar, R.; Marinescu, D.; Uivarosi, V.; Aldea, V.; Nicolescu, T. Thermal stability of new vanadyl complexes with flavonoid derivatives as potential insulin-mimetic agents. J. Therm. Anal. Calorim. 2010, 99, 823–827. [Google Scholar] [CrossRef]
  91. Emelina, T.B.; Freidzon, A.Y.; Bagaturyants, A.A.; Karasev, V.E. Electronic structure and energy transfer in europium(III)–ciprofloxacin complexes: A theoretical study. J. Phys. Chem. A 2016, 120, 7529–7537. [Google Scholar] [CrossRef]
  92. Joshi, R.; Pandey, N.; Tilak, R.; Yadav, S.K.; Mishra, H.; Pokharia, S. New triorganotin(IV) complexes of quinolone antibacterial drug sparfloxacin: Synthesis, structural characterization, DFT studies and biological activity. Appl. Organomet. Chem. 2018, 32, e4324. [Google Scholar] [CrossRef]
  93. Kumar, M.; Kumar, G.; Mogha, N.K.; Jain, R.; Hussain, F.; Masram, D.T. Structure, DNA/proteins binding, docking and cytotoxicity studies of copper(II) complexes with the first quinolone drug nalidixic acid and 2,2′-dipyridylamine. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 212, 94–104. [Google Scholar] [CrossRef]
  94. Andrews, P.C.; Beck, T.; Fraser, B.H.; Junk, P.C.; Massi, M. Synthesis and structural characterisation of cationic, neutral and hydroxo-bridged lanthanoid (La, Gd, Ho, Yb, Y) bis 5-(2-pyridyl)tetrazolate complexes. Polyhedron 2007, 26, 5406–5413. [Google Scholar] [CrossRef]
  95. Clarkson, I.; Dickins, R.; de Sousa, A. Non-radiative deactivation of the excited states of europium, terbium and ytterbium complexes by proximate energy-matched OH, NH and CH oscillators: An improved luminescence method for establishing solution hydration states. J. Chem. Soc. Perkin Trans. 1999, 2, 493–504. [Google Scholar] [CrossRef]
  96. Tomczyk, M.D.; Walczak, K.Z. l, 8-Naphthalimide based DNA intercalators and anticancer agents. A systematic review from 2007 to 2017. Eur. J. Med. Chem. 2018, 159, 393–422. [Google Scholar] [CrossRef] [PubMed]
  97. Buchtik, R.; Trávníček, Z.; Vančo, J.; Herchel, R.; Dvořák, Z. Synthesis, characterization, DNA interaction and cleavage, and in vitro cytotoxicity of copper (II) mixed-ligand complexes with 2-phenyl-3-hydroxy-4 (1H)-quinolinone. Dalt. Trans. 2011, 40, 9404–9412. [Google Scholar] [CrossRef] [Green Version]
  98. Wu, S.-S.; Yuan, W.-B.; Wang, H.-Y.; Zhang, Q.; Liu, M.; Yu, K.-B. Synthesis, crystal structure and interaction with DNA and HSA of (N, N′-dibenzylethane-1, 2-diamine) transition metal complexes. J. Inorg. Biochem. 2008, 102, 2026–2034. [Google Scholar] [CrossRef]
  99. Yin, B.-T.; Yan, C.-Y.; Peng, X.-M.; Zhang, S.-L.; Rasheed, S.; Geng, R.-X.; Zhou, C.-H. Synthesis and biological evaluation of α-triazolyl chalcones as a new type of potential antimicrobial agents and their interaction with calf thymus DNA and human serum albumin. Eur. J. Med. Chem. 2014, 71, 148–159. [Google Scholar] [CrossRef]
  100. Du, H.; Xiang, J.; Zhang, Y.; Tang, Y. A spectroscopic and molecular modeling study of sinomenine binding to transferrin. Bioorg. Med. Chem. Lett. 2007, 17, 1701–1704. [Google Scholar] [CrossRef]
  101. Kratz, F.; Beyer, U.; Roth, T.; Tarasova, N.; Collery, P.; Lechenault, F.; Cazabat, A.; Schumacher, P.; Unger, C.; Falken, U. Transferrin conjugates of doxorubicin: Synthesis, characterization, cellular uptake, and in vitro efficacy. J. Pharm. Sci. 1998, 87, 338–346. [Google Scholar] [CrossRef]
  102. Li, H.; Qian, Z.M. Transferrin/transferrin receptor-mediated drug delivery. Med. Res. Rev. 2002, 22, 225–250. [Google Scholar] [CrossRef]
  103. Wagner, E.; Curiel, D.; Cotten, M. Delivery of drugs, proteins and genes into cells using transferrin as a ligand for receptor-mediated endocytosis. Adv. Drug Deliv. Rev. 1994, 14, 113–135. [Google Scholar] [CrossRef] [Green Version]
  104. Kawabata, H. Transferrin and transferrin receptors update. Free Radic. Biol. Med. 2019, 133, 46–54. [Google Scholar] [CrossRef] [PubMed]
  105. Hu, Y.-J.; Liu, Y.; Wang, J.-B.; Xiao, X.-H.; Qu, S.-S. Study of the interaction between monoammonium glycyrrhizinate and bovine serum albumin. J. Pharm. Biomed. Anal. 2004, 36, 915–919. [Google Scholar] [CrossRef] [PubMed]
  106. Deepa, S.; Mishra, A.K. Fluorescence spectroscopic study of serum albumin–bromadiolone interaction: Fluorimetric determination of bromadiolone. J. Pharm. Biomed. Anal. 2005, 38, 556–563. [Google Scholar] [CrossRef]
  107. Lakowicz, J.R. Quenching of fluorescence. In Principles of Fluorescence Spectroscopy; Lakowicz, J.R., Ed.; Springer: Boston, MA, USA, 2006; pp. 277–330. ISBN 978-0-387-31278-1. [Google Scholar]
  108. Weiss, J.N. The Hill equation revisited: Uses and misuses. FASEB J. 1997, 11, 835–841. [Google Scholar] [CrossRef] [PubMed]
  109. Zhang, G.; Wang, Y.; Zhang, H.; Tang, S.; Tao, W. Human serum albumin interaction with paraquat studied using spectroscopic methods. Pestic. Biochem. Physiol. 2007, 87, 23–29. [Google Scholar] [CrossRef]
  110. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. (Eds.) Gaussian 09, Revision E.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  111. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. (Eds.) Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  112. Macrae, C.F.; Sovago, I.; Cottrell, S.J.; Galek, P.T.A.; McCabe, P.; Pidcock, E.; Platings, M.; Shields, G.P.; Stevens, J.S.; Towler, M.; et al. Mercury 4.0: From visualization to analysis, design and prediction. J. Appl. Crystallogr. 2020, 53, 226–235. [Google Scholar] [CrossRef] [Green Version]
  113. Navarro, M.; Hernández, C.; Colmenares, I.; Hernández, P.; Fernández, M.; Sierraalta, A.; Marchán, E. Synthesis and characterization of [Au(dppz)2]Cl3. DNA interaction studies and biological activity against Leishmania (L) mexicana. J. Inorg. Biochem. 2007, 101, 111–116. [Google Scholar] [CrossRef]
  114. Lakowicz, J.R.; Weber, G. Quenching of fluorescence by oxygen. Probe for structural fluctuations in macromolecules. Biochemistry 1973, 12, 4161–4170. [Google Scholar] [CrossRef]
  115. Munteanu, A.-C.; Badea, M.; Olar, R.; Silvestro, L.; Mihaila, M.; Brasoveanu, L.I.; Musat, M.G.; Andries, A.; Uivarosi, V. Cytotoxicity studies, DNA interaction and protein binding of new Al (III), Ga (III) and In (III) complexes with 5-hydroxyflavone. Appl. Organomet. Chem. 2018, 32, e4579. [Google Scholar] [CrossRef] [Green Version]
  116. Fu, X.-B.; Lin, Z.-H.; Liu, H.-F.; Le, X.-Y. A new ternary copper (II) complex derived from 2-(2′-pyridyl) benzimidazole and glycylglycine: Synthesis, characterization, DNA binding and cleavage, antioxidation and HSA interaction. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 122, 22–33. [Google Scholar] [CrossRef]
Figure 1. (A) Oxolinic acid; (B) Metal coordination sites in the general structure of quinolones.
Figure 1. (A) Oxolinic acid; (B) Metal coordination sites in the general structure of quinolones.
Molecules 25 05418 g001
Scheme 1. Synthesis scheme for complexes of oxolinic acid; M = Y3+, La3+, Sm3+, Eu3+, Gd3+, Tb3+.
Scheme 1. Synthesis scheme for complexes of oxolinic acid; M = Y3+, La3+, Sm3+, Eu3+, Gd3+, Tb3+.
Molecules 25 05418 sch001
Figure 2. UV-Visible spectra of oxolinic acid, sodium oxolinate and M oxo (M = Y3+, La3+, Sm3+, Eu3+, Gd3+, Tb3+).
Figure 2. UV-Visible spectra of oxolinic acid, sodium oxolinate and M oxo (M = Y3+, La3+, Sm3+, Eu3+, Gd3+, Tb3+).
Molecules 25 05418 g002
Figure 3. TG, DTG and DTA curves for Sm oxo.
Figure 3. TG, DTG and DTA curves for Sm oxo.
Molecules 25 05418 g003
Figure 4. Optimized molecular geometry of Sm oxo at the B3LYP/ ECP52MWB (Sm)/ 6-31G(d,p) (H, C, O atoms) level in Gaussian 09 software package; the Sm atom isshown in gree, the O atoms in red, the N atoms in blue, C atoms in black and H atoms in white.
Figure 4. Optimized molecular geometry of Sm oxo at the B3LYP/ ECP52MWB (Sm)/ 6-31G(d,p) (H, C, O atoms) level in Gaussian 09 software package; the Sm atom isshown in gree, the O atoms in red, the N atoms in blue, C atoms in black and H atoms in white.
Molecules 25 05418 g004
Figure 5. Left: Absorption spectra of oxo and Sm oxo in the absence and presence of increasing amounts of DNA. [compound] = 20 μM; [DNA] = 0, 5, 10, 15, 20, 25, 30, 35, 40 μM. The arrow shows the absorption changes upon the increase of the DNA concentration. Right: Fluorescence spectra of the EB-DNA system in the absence and presence of increasing amounts of the tested compounds. λex = 500 nm, [EB] = 2 μM, [DNA] = 10 μM, [compound] = 0, 5, 10, 15, 20, 25, 30, 35, 40 μM. Arrows indicate the changes in fluorescence intensities upon the increase of the concentrations of the tested compounds.
Figure 5. Left: Absorption spectra of oxo and Sm oxo in the absence and presence of increasing amounts of DNA. [compound] = 20 μM; [DNA] = 0, 5, 10, 15, 20, 25, 30, 35, 40 μM. The arrow shows the absorption changes upon the increase of the DNA concentration. Right: Fluorescence spectra of the EB-DNA system in the absence and presence of increasing amounts of the tested compounds. λex = 500 nm, [EB] = 2 μM, [DNA] = 10 μM, [compound] = 0, 5, 10, 15, 20, 25, 30, 35, 40 μM. Arrows indicate the changes in fluorescence intensities upon the increase of the concentrations of the tested compounds.
Molecules 25 05418 g005
Figure 6. The fluorescence intensities of the free proteins (apo-Tf and HSA) decrease upon the addition of increasing amounts of Sm oxo. [apo-Tf] = 1 μM, [HSA] = 2.5 μM, [Sm oxo] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM. Black arrows indicate that the peak areas decrease with increasing concentrations of the complex. The effects are more prevalent for Sm oxo than for the free ligand, as well as for HSA as compared to apo-Tf, indicating a stronger affinity of the compounds for the former protein.
Figure 6. The fluorescence intensities of the free proteins (apo-Tf and HSA) decrease upon the addition of increasing amounts of Sm oxo. [apo-Tf] = 1 μM, [HSA] = 2.5 μM, [Sm oxo] = 0, 1, 2, 3, 4, 5, 6, 7, 8 μM. Black arrows indicate that the peak areas decrease with increasing concentrations of the complex. The effects are more prevalent for Sm oxo than for the free ligand, as well as for HSA as compared to apo-Tf, indicating a stronger affinity of the compounds for the former protein.
Molecules 25 05418 g006
Table 1. Assignments of the recorded bands in the IR spectra of the studied compounds.
Table 1. Assignments of the recorded bands in the IR spectra of the studied compounds.
OxoOxoNa+Y oxoLa oxoSm oxoEu oxoGd oxoTb oxoAssignments
--3725 vw
3399 w
3622 w
3359 w
3345 wb3389 wb3378 wb3391 wbν(O-H);
COOH, H2O
3061 w3280 w,b3056 w3057 w3059 w3059 w3060 w3057 wν(C-H) aromatic
2983 w
2930 w
2975 w
2910 w
2974 w
2925 w
2909 w2916 w2984 w-2976 w
2921 w
ν (C-H) aliphatic
1698 ms------ ν(C=O) COOH
-1634 vs1631 s16321634163316331632 sν as (COO−)
1632 ms1592 s1594 s1577 s1576 s1578 s1579 s1582 sν(C=O) pyridone
1438 vs1498 ms
1473 s
1497 s
1469 s
1496 w
1469 vs
1495 m
1465 s
1493 s
1479 s
1465 vs
1496 s
1466 vs
1496 s
1468 vs
ν (-CH);
δ (-CH2)
1384 m1391 s1387 m1413 m1412 m1416 m1416 m 1387 mδ (C-H) methyl group
1338 vs1339 s1341 s1340 s13401341 m1336 sν s(COO-)
-297292292294293292295Δ = ν as (COO-)- ν s(COO-)
1147 m1188 ms1192 m11931191 m1194 m1194 m1193 mν (C-N)
1075 m
1093 m
1091 mw1088 w1089 w 1084 w1092 w1092 w1088 wν (O-CH2-O)
772 m
754 m
768 m
742
770 s
744
771
744
772 s
745 m
773 ms
744 m
773 ms
744 m
770 ms
743 m
δ (COO-)
645 s639 m649 m640 m646 m643 m644 m647 mring deformation
606 m607 m611 m608 m610 m609 m609 m610 m
556 m547 m559 m553 m557 m556 m556 m558 m
-513 m------
499 m499 m500 m497 s500 m---
448 w459 w---453 m445 m459 mw
431 w-428 m431 w----
418 m416 m416 m---413 ms412 m
---408 s407 s410 s--
--401 m400 s----
--512 s-513 s512 s511 s511 sν (M-O)
---485 s484 m483 s483 s499 s
--468 w456 m456 m---
-----421 s423 m422 m
Table 2. Thermal decomposition data (in air flow) for the complexes.
Table 2. Thermal decomposition data (in air flow) for the complexes.
ComplexStepThermal EffectTemperature Range/°CΔmexp/%Δmcalc/%
Y oxo1Endothermic125–2802.402.80
2Exothermic280–90079.9079.70
Residue (1/2 Y2O3)17.7017.50
La oxo1Endothermic75–1502.702.60
2Exothermic150–90073.5073.90
Residue (1/2 La2O3)23.8023.50
Sm oxo1Endothermic75–2204.605.00
2Exothermic220–95071.4070.90
Residue (1/2 Sm2O3)24.0024.10
Eu oxo1Endothermic125–2402.602.50
2Exothermic240–87572.4072.60
Residue (1/2 Eu2O3)25.0024.90
Gd oxo1Endothermic110–2503.803.70
2Exothermic250–88070.8071.20
Residue (1/2 Gd2O3)25.4025.10
Tb oxo1Endothermic60–1603.603.70
2Exothermic280–98070.5070.50
Residue (1/4 Tb4O7)25.9025.80
Table 3. Geometric parameters: bond lengths, bond angles, dihedral angles, charge density, total energy of Sm oxo as resulted from DFT calculations at the B3LYP/ ECP52MWB (Sm)/ 6-31G(d,p) (H, C, O atoms) level in Gaussian 09 software package.
Table 3. Geometric parameters: bond lengths, bond angles, dihedral angles, charge density, total energy of Sm oxo as resulted from DFT calculations at the B3LYP/ ECP52MWB (Sm)/ 6-31G(d,p) (H, C, O atoms) level in Gaussian 09 software package.
Bond Length (Å)Sm oxoChargeSm oxo
Sm-O22.517Sm1.092
Sm-O32.506O2−0.586
Sm-O242.582O3−0.536
Sm-O252.519O4−0.504
Sm-O462.557O24−0.594
Sm-O482.664O25−0.530
Bond angle (°)Sm oxoO26−0.498
O2-Sm-O369.233O46−0.480
O24-Sm-O2568.477O48−0.593
O46-Sm-O48102.286N5−0.435
O46-Sm-O2553.623N7−0.437
O3-Sm-O48134.706C50.116
O2-Sm-O24110.618C90.116
Sm-O25-C28119.094C12−0.023
Sm-O3-C6144.027Total energy (a.u.)−2045.743
Dihedral angleSm oxo
O46-Sm-O25-C2887.515
O48-Sm-O3-C6−176.277
O48-Sm-O24-C32−116.492
O46-Sm-O2-C10−130.417
Sm-O25-C28-O26−101.279
Sm-O3-C6-O4169.368
Table 4. IR selected data for the Sm oxo complex (experimental vs. calculated for the optimized structure.
Table 4. IR selected data for the Sm oxo complex (experimental vs. calculated for the optimized structure.
AssignmentsExperimental DataB3LYP/6-31G(d,p)/
ECP52MWB
ν(O-H);
COOH, H2O
3345 w, b3617 w
3309 s
ν(C-H) aromatic 3059 w3134 w
3097 w
ν(C-H) aliphatic2916 w3060 w
3059 w
3043 w
3034 w
ν as(COO-)16341758 s
1695 s
ν(C=O) pyridone1576 s1608 s
ν(-CH);
δ(-CH2)
1495 m
1465 s
1527 s
1510 m
δ(C-H) methyl group1412 m1433 w
1428 w
ν s(COO-)1340 s1360 s
ν(C-N)1191 m1224 w
ν (O-CH2-O)1084 w1073 w
δ(COO-)772 m
745 m
784 w
757 w
ring deformation646 m637 m
610 m614 w
557 m517 w
500 m503 w
407 s-
ν (M-O)513 s513 w
484 m503 w
456 m425 w
513 s
484 m
456 m
513 w
503 w
425 w
Table 5. IC50 values (μM) for oxo, M oxo and cisplatin/adryamicin.
Table 5. IC50 values (μM) for oxo, M oxo and cisplatin/adryamicin.
CompoundIC50 (μM)
LoVo
(Human Colon Adenocarcinoma)
MDA-MB 231
(Human Breast Adenocarcinoma)
HUVEC
(Normal, Human Umbilical Vein Endothelial Cells)
oxo>200 μM43.49 ± 4.94>200 μM
Y oxo90.41 ± 9.3033.22 ± 14.92>200 μM
La oxo52.10 ± 5.54>100 μM>200 μM
Sm oxo41.51 ± 15.2840.42 ± 6.27>200 μM
Eu oxo106.60 ± 16.0473.65 ± 19.80>200 μM
Gd oxo56.49 ± 3.8373.79 ± 18.50>200 μM
Tb oxo40.59 ± 7.43>200 μM>200 μM
Cis-Pt40.15 ± 13.94-28.46 ± 6.28
ADR-7.85 ± 0.70-
IC50 value represents the concentration of the tested compound required to inhibit 50% of cell growth with respect to control sample, in the absence of the compound. Cytotoxicity was assessed by MTS assay after 48 h of incubation with each compound. Data obtained are based on the average of three independent experiments and the reported errors are the corresponding standard deviations.
Table 6. Binding constants (Kb), which characterize the interactions of the tested compounds with DNA, and quenching constants (KSV) for the EB-DNA-tested compounds systems; binding constants characterizing the DNA-tested compounds interactions calculated from the EB displacement assays.
Table 6. Binding constants (Kb), which characterize the interactions of the tested compounds with DNA, and quenching constants (KSV) for the EB-DNA-tested compounds systems; binding constants characterizing the DNA-tested compounds interactions calculated from the EB displacement assays.
Kb (L∙mol−1)Ksv (×104)EB Displacement Assay
Compound K50 (µM)n
oxo(0.46 ± 0.14) × 104(1.05 ± 0.02) × 10442.07 ± 18.482.98 ± 1.14
Y-oxo(4.02 ± 1.41) × 105(8.38 ± 0.29) × 10328.07 ± 4.441.92 ± 0.45
La-oxo(2.57 ± 0.17) × 105(1.27 ± 0.03) × 10437.05 ± 8.863.44 ± 0.94
Sm-oxo(9.33 ± 1.46) × 104(1.21 ± 0.03) × 10441.55 ± 19.322.44 ± 0.82
Eu-oxo(10.72 ± 2.47) × 105(1.26 ± 0.02) × 10440.15 ± 11.822.65 ± 0.73
Gd-oxo(5.77 ± 1.07) × 105(1.18 ± 0.01) × 10447.75 ± 25.221.66 ± 0.47
Tb-oxo(9.84 ± 3.83) × 105(1.18 ± 0.00) × 10420.28 ± 1.152.68 ± 0.19
Table 7. Binding constants and thermodynamic parameters calculated for the interaction of the complexes with apo-Tf and HSA, respectively.
Table 7. Binding constants and thermodynamic parameters calculated for the interaction of the complexes with apo-Tf and HSA, respectively.
ComplexT
(K)
Ksv
(M−1)
Kq
(M−1∙s−1)
Ka
(M−1)
n
(Number of Binding Sites)
Kd
(μM)
n
(Hill Coefficient)
Apo-Tf
oxo298(17.45 ± 0.95) × 104
(fa = 0.36±0.01)
17.45 × 1012(0.73 ± 0.07) × 1030.67 ± 0.008.11 ± 7.341.89 ± 1.35
Y-oxo298(6.52 ± 0.56) × 104
(fa = 1.15 ± 0.07)
6.52 × 1012(4.40 ± 1.09) × 1040.94 ± 0.018.34 ± 1.711.58 ± 0.23
La-oxo298(14.13 ± 1.46) × 104
(fa = 0.68 ± 0.03)
14.13 × 1012(0.53 ± 0.23) × 1040.77 ± 0.038.89 ± 2.111.71 ± 0.21
Sm-oxo298(9.34 ± 0.66) × 104
(fa = 0.85 ± 0.05)
9.34 × 1012(4.16 ± 1.09) × 1040.95 ± 0.025.73 ± 0.552.03 ± 0.32
Eu-oxo298(16.55 ± 0.94) × 104
(fa = 0.81 ± 0.02)
16.55 × 1012(2.74 ± 0.41) × 1040.88 ± 0.018.27 ± 4.170.95 ± 0.21
Gd-oxo298(17.51 ± 1.06) × 104
(fa = 0.73 ± 0.02)
17.51 × 1012(0.58 ± 0.25) × 1040.76 ± 0.038.80 ± 6.571.24 ± 0.45
Tb-oxo298(19.45 ± 1.10) × 104
(fa = 0.77 ± 0.01)
19.45 × 1012(1.28 ± 0.17) × 1040.81 ± 0.015.57 ± 1.221.57 ± 0.33
HSA
oxo298(1.95 ± 0.38) × 104
(fa = 1.92 ± 0.35)
1.95 × 1012(1.08 ± 0.22) × 1040.90 ± 0.017.26 ± 2.661.62 ± 0.39
Y-oxo298(10.53 ± 0.95) × 104
(fa = 1.07 ± 0.05)
10.53 × 1012(2.04 ± 0.74) × 1051.44 ± 0.036.30 ± 2.471.52 ± 0.47
La-oxo298(6.82 ± 0.74) × 104
(fa = 1.26 ± 0.11)
6.82 × 1012(5.80 ± 2.01) × 1051.14 ± 0.026.22 ± 1.051.74 ± 0.21
Sm-oxo298(6.56 ± 0.90) × 104
(fa = 1.44 ± 0.15)
6.56 × 1012(4.28 ± 2.12) × 1051.11 ± 0.044.69 ± 0.871.55 ± 0.37
Eu-oxo298(11.25 ± 0.92) × 104
(fa = 0.94 ± 0.04)
11.25 × 1012(4.75 ± 0.80) × 1051.11 ± 0.015.56 ± 0.861.33 ± 0.18
Gd-oxo298(5.77 ± 0.84) × 104
(fa = 1.39 ± 0.15)
5.77 × 1012(1.98 ± 0.82) × 1051.06 ± 0.035.32 ± 0.691.81 ± 0.26
Tb-oxo298(2.84 ± 0.70) × 104
(fa = 2.13 ± 0.45)
2.84 × 1012(6.83 ± 1.33) × 1051.18 ± 0.015.97 ± 0.891.58 ± 0.20
Sample Availability: Samples of all compounds synthesized in this study are available from the authors.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Maciuca, A.-M.; Munteanu, A.-C.; Mihaila, M.; Badea, M.; Olar, R.; Nitulescu, G.M.; Munteanu, C.V.A.; Bostan, M.; Uivarosi, V. Rare-Earth Metal Complexes of the Antibacterial Drug Oxolinic Acid: Synthesis, Characterization, DNA/Protein Binding and Cytotoxicity Studies. Molecules 2020, 25, 5418. https://doi.org/10.3390/molecules25225418

AMA Style

Maciuca A-M, Munteanu A-C, Mihaila M, Badea M, Olar R, Nitulescu GM, Munteanu CVA, Bostan M, Uivarosi V. Rare-Earth Metal Complexes of the Antibacterial Drug Oxolinic Acid: Synthesis, Characterization, DNA/Protein Binding and Cytotoxicity Studies. Molecules. 2020; 25(22):5418. https://doi.org/10.3390/molecules25225418

Chicago/Turabian Style

Maciuca, Ana-Madalina, Alexandra-Cristina Munteanu, Mirela Mihaila, Mihaela Badea, Rodica Olar, George Mihai Nitulescu, Cristian V. A. Munteanu, Marinela Bostan, and Valentina Uivarosi. 2020. "Rare-Earth Metal Complexes of the Antibacterial Drug Oxolinic Acid: Synthesis, Characterization, DNA/Protein Binding and Cytotoxicity Studies" Molecules 25, no. 22: 5418. https://doi.org/10.3390/molecules25225418

Article Metrics

Back to TopTop