Next Article in Journal
Mesenchymal Stromal Cells Perspective: New Potential Therapeutic for the Treatment of Neurological Diseases
Next Article in Special Issue
Ruthenium Complexes: An Alternative to Platinum Drugs in Colorectal Cancer Treatment
Previous Article in Journal
No Time Dependence of Ciprofloxacin Pharmacokinetics in Critically Ill Adults: Comparison of Individual and Population Analyses
Previous Article in Special Issue
DNA-Binding and Cytotoxicity of Copper(I) Complexes Containing Functionalized Dipyridylphenazine Ligands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Anticancer Diiron Vinyliminium Complexes: A Structure–Activity Relationship Study

by
Simona Braccini
1,†,
Giorgia Rizzi
1,†,‡,
Lorenzo Biancalana
1,
Alessandro Pratesi
1,
Stefano Zacchini
2,
Guido Pampaloni
1,
Federica Chiellini
1,*,§ and
Fabio Marchetti
1,*
1
Department of Chemistry and Industrial Chemistry, University of Pisa, Via G. Moruzzi 13, I-56124 Pisa, Italy
2
Department of Industrial Chemistry “Toso Montanari”, University of Bologna, Viale Risorgimento 4, I-40136 Bologna, Italy
*
Authors to whom correspondence should be addressed.
These authors contribute equally to this paper.
Present address: KU Leuven—KULAK, Department of Chemistry, Molecular Imaging and Photonics, Etienne Sabbelaan 53 Box 7657, 8500 Kortrijk, Belgium.
§
Deceased in March 2021.
Pharmaceutics 2021, 13(8), 1158; https://doi.org/10.3390/pharmaceutics13081158
Submission received: 27 June 2021 / Revised: 22 July 2021 / Accepted: 23 July 2021 / Published: 27 July 2021
(This article belongs to the Special Issue Beyond the Platinum in Metal-Based Cancer Therapy)

Abstract

:
A series of 16 novel diiron complexes of general formula [Fe2Cp2(CO)(μ-CO){μ-η13-C(R′)C(R″)CN(R)(Y)}]CF3SO3 (2–7), bearing different substituents on the bridging vinyliminium ligand, was synthesized in 69–95% yields from the reactions of diiron μ-aminocarbyne precursors with various alkynes. The products were characterized by elemental analysis, IR, 1H and 13C NMR spectroscopy; moreover the X-ray structures of 2c (R = Y = CH2Ph, R′ = R″ = Me) and 3a (R = CH2CH=CH2, Y = R′ = Me, R″ = H) were ascertained by single-crystal X-ray diffraction studies. NMR and UV–Vis methods were used to assess the D2O solubility, the stability in aqueous solution at 37 °C and the octanol–water partition coefficients of the complexes. A screening study evidenced a potent cytotoxicity of 2–7 against the A2780 cancer cell line, with a remarkable selectivity compared to the nontumoral Balb/3T3 cell line; complex 4c (R = Cy, Y = R′ = R″ = Me) revealed as the most performant of the series. The antiproliferative activity of a selection of complexes was also assessed on the cisplatin-resistant A2780cisR cancer cell line, and these complexes were capable of inducing a significant ROS production. Moreover, ESI-MS experiments indicated the absence of interaction of selected complexes with cytochrome c and the potentiality to inhibit the thioredoxin reductase enzyme (TrxR).

Graphical Abstract

1. Introduction

There is an urgent demand for the development of new, effective and targeted anticancer drugs, and in this regard transition metal complexes are at the forefront of research [1,2,3,4,5,6]. Indeed, this category of compounds offers peculiar properties associated with the presence of one or more transition elements, i.e., the availability of a variety of oxidation states, coordination environments and geometries, and the possibility of replacement and/or activation of ligands under suitable conditions [7,8,9]. Such arsenal of tools may provide an increased added pharmaceutical value with respect to common organic molecules, and in this light few platinum complexes have been successfully employed worldwide in clinical treatments against several types of tumors, in combination with other drugs [10,11]. On the other hand, the toxicity of the metal center may represent a serious issue, and this is certainly the case of platinum, a “heavy metal” belonging to the 5d series [12,13]. The severe side effects arising from this toxicity, along with a limited selectivity of action, and the tendency of the tumor cells to progressively acquire resistance to the treatment represent major inconveniences in the administration of platinum chemotherapics, besides their unquestionable efficacy [14,15,16]. Therefore, a huge effort has been devoted to evaluating potential drugs based on other transition metals [2]. Iron is an attractive element in this respect due to its bioavailability, which substantially limits the toxic effects of its compounds, and the feasible redox chemistry in physiological media, usually involving the +II and +III oxidation states [17,18]. A variety of monoiron complexes have been assessed for the anticancer potential both in vitro and in vivo [19,20,21,22]. Following the successful experience with ferroquine, a conjugate between ferrocene and the drug chloroquine, which entered phase II clinical trials as an antimalarial agent (Figure 1, structure I) [4,23,24], the anticancer properties of related ferrocene derivatives have been intensively investigated (an example in Figure 1, structure II) [25,26,27,28]. This family of compounds exhibits a substantial robustness supplied by the ferrocene skeleton and exerts a cytotoxic activity essentially by unbalancing cellular redox homeostasis via iron(II) to iron(III) oxidation [25,26,27,28,29]. Such a mode of action strikingly differs from that of platinum compounds, which instead induce cell death through DNA binding [10,30]. On the wave of our longtime experience with organo-iron synthetic chemistry [31,32,33,34,35], in the very last years we have contributed to unveil the anticancer potential of diiron complexes, which was almost unexplored [36,37,38,39,40]. This is surprising, in that diiron complexes hold a major advantage, compared to related monoiron species, due to the presence of adjacent metal atoms “working in concert”, thus allowing the easy construction and the stabilization of uncommon bridging ligands and the co-presence of other ligands [31,41,42,43]. In this context, commercial [Fe2Cp2(CO)4] (Cp = η5-C5H5) is a convenient starting material based on a [Fe+I-Fe+I] core, with the complementary electronic properties of the CO and Cp ligands providing robustness. Cationic complexes of type III (Figure 1), containing a tightly coordinated bridging vinyliminium ligand, originate from the stepwise coupling of one isocyanide with one alkyne [44]. Complexes III are rather stable in aqueous media and display a variable cytotoxicity, ranging from the micromolar range to inactivity; they appear to exert their action following the general behavior shown by cytotoxic iron compounds, i.e., through the interference with redox processes [36,37,38,39,40,45]. Notwithstanding, other ways that are not accessible to ferrocene derivatives might be viable, including binding to biological targets [36,37,38,39] and the auxiliary effect of slow carbon monoxide release [37,40,45]. The structural diversity offered by the choice of the isocyanide (R substituent) and alkyne (R′, R″) reagents enables to tune important physico-chemical properties of the complexes (e.g., water solubility, amphiphilicity), correlated to their activity. A wide range of alkynes has been explored in this regard, including the incorporation of specific bioactive fragments [46], whereas the variation of the R group has been quite restricted (compounds with R = Me or R = 2,6-C6H3Me2 = Xyl have been preferentially studied). This point is not trivial, on considering that related diiron complexes with a bridging aminocarbyne (iminium) ligand (structure IV, Figure 1) possess the ability to interact with bio-substrates, which seems strongly affected by both N-substituents, Y and R [37]. Herein, we report the synthesis and the full characterization of novel diiron vinyliminium complexes, containing unprecedented combinations of iminium substituents, and related studies aimed to assess the antiproliferative activity and to clarify the mechanism of action.

2. Experimental

2.1. Materials and Methods

The preparation and purification of complexes were carried out in air, and isolated products were stored in air. Solvents and organic reactants were purchased from Merck or TCI Europe. Compounds 1a–c,e [37] and 1d,f [47] were prepared according to the respective published procedures. Chromatography separations were carried out on columns of deactivated alumina (Merck, 4% w/w water). Infrared spectra of solutions were recorded on a Perkin Elmer Spectrum 100 FT-IR spectrometer with a CaF2 liquid transmission cell (2300–1500 cm−1 range). IR spectra were processed with Spectragryph software [48]. NMR spectra were recorded at 298 K on a Bruker Avance II DRX400 instrument equipped with a BBFO broadband probe. Chemical shifts (expressed in parts per million) are referenced to the residual solvent peaks (1H, 13C) [49] or to external standard (H3PO4, 31P). 1H and 13C NMR spectra were assigned with the assistance of 1H-13C (gs-HSQC and gs-HMBC) correlation experiments [50]. NMR signals due to a second isomeric form (where it has been possible to detect them) are italicized; integration values refer to the main isomer. Scheme 1, Scheme 2, Scheme 3, Scheme 4, Scheme 5, Scheme 6, Scheme 7, Scheme 8, Scheme 9, Scheme 10, Scheme 11, Scheme 12, Scheme 13, Scheme 14, Scheme 15 and Scheme 16 show the prevalent isomeric form detected by NMR in each case. Elemental analyses were performed on a Vario MICRO cube instrument.

2.2. Synthesis and Characterization of Diiron Complexes

General procedure. A solution of 1a–f (ca. 0.5 mmol) in acetonitrile (ca. 10 mL) was treated with Me3NO (1.2 eq.). The resulting mixture was stirred for 1 h, during which time progressive color darkening was noticed. The complete conversion of the starting material into the corresponding CO/NCMe substitution product was clearly checked by IR spectroscopy [51,52,53]. The volatiles were removed under vacuum to afford a dark-brown residue, which was dissolved into dichloromethane (ca. 20 mL) and treated with the appropriate alkyne (1.5–2.0 eq.). The mixture was stirred at room temperature for 48 h, then it was charged on an alumina column. Elution with CH2Cl2 and then CH2Cl2/THF (1:1 v/v) allowed to remove the excess of alkyne and impurities. The fraction corresponding to the product was separated using the appropriate eluent (vide infra), then the solvent was evaporated under reduced pressure. The residue was dissolved in the minimum volume of CH2Cl2, and subsequent addition of hexane (20–30 mL) gave a powder which was dried under vacuum.
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2HC1N(CH2Ph)2}]CF3SO3, 2a (Scheme 1)
Scheme 1. Structure of the cation of 2a.
Scheme 1. Structure of the cation of 2a.
Pharmaceutics 13 01158 sch001
From 1a and propyne (THF solution, ca. 1 mol/L). Brownish-red solid, yield 81%. Eluent for chromatography: MeCN. Anal. calcd. for C31H28F3Fe2NO5S: C, 53.55; H, 4.06; N, 2.01; S, 4.61. Found: C, 53.41; H, 4.12; N, 1.97; S, 4.55. IR (CH2Cl2): ῦ/cm−1 = 1991vs (CO), 1808s (μ-CO), 1644m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 7.44, 7.21, 7.10 (m, 10 H, Ph); 5.86, 5.22 (d, 2 H, 2JHH = 14.7 Hz, CH2); 5.66, 5.32 (s, 10 H, Cp); 4.90 (s, 1 H, C2H); 4.65, 4.33 (d, 2 H, 2JHH = 14.7 Hz, CH2); 4.04 (s, 3 H, C3Me). 13C{1H} NMR (acetone-d6): δ/ppm = 257.1 (μ-CO); 229.4 (C1); 210.7 (CO); 210.2 (C3); 132.7, 132.3, 129.1, 129.0, 128.9, 128.8 (Ph); 91.2, 88.0 (Cp); 64.2, 59.4 (CH2); 52.9 (C2), 41.4 (C3Me).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Ph)C2HC1N(CH2Ph)2}]CF3SO3, 2b (Scheme 2)
Scheme 2. Structure of the cation of 2b.
Scheme 2. Structure of the cation of 2b.
Pharmaceutics 13 01158 sch002
From 1a and phenylacetylene. Brown solid, yield 81%. Eluent for chromatography: MeCN. Anal. calcd. for C36H30F3Fe2NO5S: C, 57.09; H, 3.99; N, 1.85; S, 4.23. Found: C, 56.88; H, 4.03; N, 1.88; S, 4.30. IR (CH2Cl2): ῦ/cm−1 = 1992vs (CO), 1810s (μ-CO), 1642m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 7.79, 7.57, 7.46, 7.40, 7.31, 7.10 (m, 15 H, Ph); 5.94, 4.48 (d, 2 H, 2JHH = 14.67 Hz, CH2); 5.50, 5.34 (s, 10 H, Cp); 5.43, 4.72 (d, 2 H, 2JHH = 13.69 Hz, CH2); 4.75 (s, 1 H, C2H). 13C{1H} NMR (acetone-d6): δ/ppm = 256.6 (μ-CO); 227.8 (C1); 210.0 (CO); 156.4 (ipso-Ph); 132.8, 132.2, 129.2, 129.1, 128.4, 127.3, 127.0 (Ph); 92.2, 88.2 (Cp); 63.9, 60.0 (CH2); 53.5 (C2). C3 signal overlapped with solvent.
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2(Me)C1N(CH2Ph)2}]CF3SO3, 2c (Scheme 3)
Scheme 3. Structure of the cation of 2c.
Scheme 3. Structure of the cation of 2c.
Pharmaceutics 13 01158 sch003
From 1a and 2-butyne. Brown solid, yield 84%. Eluent for chromatography: MeCN. Anal. calcd. for C32H30F3Fe2NO5S: C, 54.18; H, 4.26; N, 1.97; S, 4.52. Found: C, 54.08; H, 4.32; N, 1.90; S, 4.57. IR (CH2Cl2): ῦ/cm−1 = 1987vs (CO), 1805s (μ-CO), 1628m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 7.38, 7.05 (m, 10 H, Ph); 5.70, 5.45 (d, 2 H, 2JHH = 14.7 Hz, CH2); 4.78, 4.27 (d, 2 H, 2JHH = 15.3 Hz, CH2); 5.64, 5.29 (s, 10 H, Cp); 4.00 (s, 3 H, C3Me); 2.19 (s, 3 H, C2Me). 13C{1H} NMR (acetone-d6): δ/ppm = 258.2 (μ-CO); 230.4 (C1); 210.9 (CO); 203.7 (C3); 132.6, 131.6, 128.9, 128.3 (Ph); 91.6, 88.6 (Cp); 65.7 (C2); 63.1, 58.4 (CH2); 36.8 (C3Me); 15.9 (C2Me). Crystals suitable for X-ray analysis were obtained by slow evaporation of the solvent from an acetone solution of 2c.
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2HC1NMe(CH2CH = CH2)}]CF3SO3, 3a (Scheme 4)
Scheme 4. Structure of the cation of 3a.
Scheme 4. Structure of the cation of 3a.
Pharmaceutics 13 01158 sch004
From 1b and propyne (THF solution, ca. 1 mol/L). Brown solid, yield 76%. Eluent for chromatography: THF/MeCN 1:1. Anal. calcd. for C21H22F3Fe2NO5S: C, 44.32; H, 3.90; N, 2.46; S, 5.63. Found: C, 44.21; H, 3.96; N, 2.37; S, 5.70. IR (CH2Cl2): ῦ/cm−1 = 1990vs (CO), 1807s (μ-CO), 1669m (C1 = N), 1643w (C = C). 1H NMR (acetone-d6): δ/ppm = 6.04, 5.75 (m, 1 H, CH = CH2); 5.61, 5.51, 5.49, 5.39 (m, 2 H, CH = CH2); 5.55, 5.22, 5.21 (s, 10 H, Cp); 4.98, 4.90, 4.17 (m, 3JHH = 6.4 Hz, NCH2); 4.70, 4.67 (s, 1 H, C2H); 3.99 (s, 3 H, C3Me); 3.92, 3.27 (s, 3 H, NMe). Isomer ratio (E/Z) = 1.2. 13C{1H} NMR (acetone-d6): δ/ppm = 257.3, 256.6 (μ-CO); 227.3, 226.9 (C1); 210.7, 208.4 (CO); 205.5 (C3); 130.7, 129.7 (CH = CH2); 121.5, 121.3 (CH = CH2); 90.8, 87.8, 87.7 (Cp); 66.7, 60.4 (NCH2); 52.4, 52.1 (C2); 47.3, 41.8 (NMe); 41.2 (C3Me). Crystals suitable for X-ray analysis were obtained by slow diffusion of diethyl ether into a solution of 3a in CH2Cl2, at −30 °C.
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Ph)C2HC1NMe(CH2CH = CH2)}]CF3SO3, 3b (Scheme 5)
Scheme 5. Structure of the cation of 3b.
Scheme 5. Structure of the cation of 3b.
Pharmaceutics 13 01158 sch005
From 1b and phenylacetylene. Brown solid, yield 79%. Eluent for chromatography: MeCN. Anal. calcd. for C26H24F3Fe2NO5S: C, 49.47; H, 3.83; N, 2.22; S, 5.08. Found: C, 49.55; H, 3.81; N, 2.29; S, 5.13. IR (CH2Cl2): ῦ/cm−1 = 1993vs (CO), 1809s (μ-CO), 1666m (C1 = N), 1643w (C = C). 1H NMR (CDCl3): δ/ppm = 7.56, 7.41 (m, 5 H, Ph); 5.85 (br, 1 H, CH = CH2); 5.64, 5.54, 5.35, 5.31 (m, 2 H, CH = CH2); 5.23, 5.21, 4.97 (s, 10 H, Cp); 4.34, 4.17 (m, 2 H, NCH2); 4.76, 4.69 (s, 1 H, C2H); 3.89, 3.30 (s, 3 H, NMe). Isomer ratio (Z/E) = 1.4. 13C{1H} NMR (acetone-d6): δ/ppm = 256.3, 255.6 (μ-CO); 226.1, 225.1 (C1); 210.1, 210.0 (CO); 204.4, 204.3 (C3); 156.4, 156.3 (ipso-Ph); 130.7, 129.8 (CH = CH2); 128.4, 128.3, 127.4, 127.3, 127.0, 126.9 (Ph); 121.6, 121.4 (CH = CH2); 91.7, 87.9, 87.8 (Cp); 67.0, 60.3 (NCH2); 53.0, 52.8 (C2); 47.7, 41.8 (NMe).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2(Me)C1NMe(CH2CH = CH2)}]CF3SO3, 3c (Scheme 6)
Scheme 6. Structure of the cation of 3c.
Scheme 6. Structure of the cation of 3c.
Pharmaceutics 13 01158 sch006
From 1b and 2-butyne. Brown solid, yield 94%. Eluent for chromatography: MeCN. Anal. calcd. for C22H24F3Fe2NO5S: C, 45.31; H, 4.15; N, 2.40; S, 5.50. Found: C, 45.17; H, 4.21; N, 2.31; S, 5.55. IR (CH2Cl2): ῦ/cm−1 = 1986vs (CO), 1805s (μ-CO), 1656m (C1 = N), 1639w (C = C). 1H NMR (acetone-d6): δ/ppm = 6.05, 5.80 (m, 1 H, CH = CH2); 5.62, 5.55, 5.41, 5.38 (m, 2 H, CH = CH2); 5.52, 5.51, 5.20, 5.19 (s, 10 H, Cp); 5.05–4.87, 4.20, 4.03 (m, 2 H, NCH2); 3.91 (s, 3 H, C3Me); 3.89, 3.21 (s, 3 H, NMe); 2.04 (s, 3 H, C2Me). Isomer ratio (Z/E) = 1.2. 13C{1H} NMR (acetone-d6): δ/ppm = 258.1, 257.4 (μ-CO); 227.3, 226.5 (C1); 210.8, 210.6 (CO); 202.1, 201.9 (C3); 130.8, 129.8 (CH = CH2); 121.7, 121.5 (CH = CH2); 91.0, 88.2, 88.1 (Cp); 65.5 (C2); 65.4, 60.5 (NCH2); 45.3, 41.3 (NMe); 36.2 (C3Me); 15.4, 14.8 (C2Me).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2HC1NMe(Cy)}]CF3SO3, 4a (Scheme 7)
Scheme 7. Structure of the cation of 4a.
Scheme 7. Structure of the cation of 4a.
Pharmaceutics 13 01158 sch007
From 1c and propyne (THF solution, ca. 1 mol/L). Brown solid, yield 90%. Eluent for chromatography: MeCN. Anal. calcd. for C24H28F3Fe2NO5S: C, 47.16; H, 4.62; N, 2.29; S, 5.25. Found: C, 47.02; H, 4.70; N, 2.23; S, 5.31. IR (CH2Cl2): ῦ/cm−1 = 1989vs (CO), 1805s (μ-CO), 1661m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 5.53, 5.52, 5.20, 5.17 (s, 10 H, Cp); 4.81, 3.60 (m, 1 H, CHCy); 4.74, 4.69 (s, 1 H, C2H); 3.98, 3.96 (s, 3 H, C3Me); 3.90, 3.20 (s, 3 H, NMe); 2.26, 2.03–1.51, 1.43–1.06 (m, 10 H, CH2Cy). Isomer ratio (E/Z) = 1.2. 13C{1H} NMR (acetone-d6): δ/ppm = 257.3, 256.4 (μ-CO); 225.5, 224.6 (C1); 210.9, 210.7 (CO); 208.2, 207.9 (CO); 204.9 (C3); 90.7, 87.7, 87.5 (Cp); 75.2, 68.2 (CHCy); 52.0, 51.3 (C2); 42.9, 38.2 (NMe); 41.2, 41.1 (C3Me); 29.9, 29.8, 29.7, 25.1, 24.7, 24.6, 24.5 (CH2Cy).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Ph)C2HC1NMe(Cy)}]CF3SO3, 4b (Scheme 8)
Scheme 8. Structure of the cation of 4b.
Scheme 8. Structure of the cation of 4b.
Pharmaceutics 13 01158 sch008
From 1c and phenylacetylene. Brownish-green solid, yield 93%. Eluent for chromatography: MeCN. Anal. calcd. for C29H30F3Fe2NO5S: C, 51.73; H, 4.49; N, 2.08; S, 4.76. Found: C, 51.68; H, 4.53; N, 2.12; S, 4.69. IR (CH2Cl2): ῦ/cm−1 = 1992vs (CO), 1809s (μ-CO), 1658m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 7.83, 7.57, 7.44 (m, 5 H, Ph); 5.44, 5.40, 5.28, 5.27 (s, 10 H, Cp); 4.95, 3.83 (m, 1 H, CHCy); 4.77, 4.73 (s, 1 H, C2H); 4.02, 3.35 (s, 3 H, NMe); 2.3–1.1 (m, 10 H, CH2Cy). Isomer ratio (E/Z) = 1.2. 13C{1H} NMR (acetone-d6): δ/ppm = 256.7, 255.7 (μ-CO); 224.3, 223.6 (C1); 210.3, 210.1 (CO); 204.5, 204.1 (C3); 156.4 (ipso-Ph); 128.3, 127.4, 127.0 (Ph); 91.7, 87.9, 87.7 (Cp); 75.5, 68.0 (CHCy); 52.4, 51.8 (C2); 43.2, 38.1 (NMe); 29.9, 29.8, 25.2, 24.7–24.5 (CH2Cy).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2(Me)C1NMe(Cy)}]CF3SO3, 4c (Scheme 9)
Scheme 9. Structure of the cation of 4c.
Scheme 9. Structure of the cation of 4c.
Pharmaceutics 13 01158 sch009
From 1c and 2-butyne. Brownish-green solid, yield 95%. Eluent for chromatography: MeCN. Anal. calcd. for C25H30F3Fe2NO5S: C, 48.02; H, 4.84; N, 2.24; S, 5.13. Found: C, 47.89; H, 4.90; N, 2.18; S, 5.19. IR (CH2Cl2): ῦ/cm−1 = 1986vs (CO), 1805s (μ-CO), 1646m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 5.51, 5.20, 5.15 (s, 10 H, Cp); 4.86, 3.51 (br, 1 H, CHCy); 3.93, 3.16 (s, 3 H, NMe); 3.91 (s, 3 H, C3Me); 2.2, 1.9–1.1 (m, 10 H, CH2Cy); 2.04, 2.02 (s, 3 H, C2Me). Isomer ratio (Z/E) = 2. 13C{1H} NMR (acetone-d6): δ/ppm = 258.3, 257.5 (μ-CO); 226.3, 224.4 (C1); 211.3, 210.7 (CO); 202.2, 201.9 (C3); 91.1, 88.2, 88.0 (Cp); 73.9, 68.0 (CHCy); 64.8, 64.2 (C2); 40.9, 36.4 (NMe); 37.3, 36.2 (C3Me); 30.4, 30.0, 25.3, 24.7, 24.6, 24.5 (CH2Cy); 15.7, 15.0 (C2Me).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2HC1NMe(CH2Ph)}]CF3SO3, 5a (Scheme 10)
Scheme 10. Structure of the cation of 5a.
Scheme 10. Structure of the cation of 5a.
Pharmaceutics 13 01158 sch010
From 1d and propyne (THF solution, ca. 1 mol/L). Brown solid, yield 88%. Eluent for chromatography: MeCN. Anal. calcd. for C25H24F3Fe2NO5S: C, 48.49; H, 3.91; N, 2.26; S, 5.18. Found: C, 48.58; H, 3.85; N, 2.21; S, 5.11. IR (CH2Cl2): ῦ/cm−1 = 1991vs (CO), 1808s (μ-CO), 1664m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 7.39–7.27, 7.07 (m, 5 H, Ph); 5.49, 5.31, 4.80, 4.42 (d, 2 H, 2JHH = 14.2 Hz, CH2); 5.25, 5.01 (s, 10 H, Cp); 4.74, 4.65 (s, 1 H, C2H); 3.88 (s, 3 H, C3Me); 3.62, 3.00 (s, 3 H, NMe). Isomer ratio (E/Z) = 1.2. 13C{1H} NMR (acetone-d6): δ/ppm = 256.8 (μ-CO); 227.5, 226.7 (C1); 210.2, 210.0 (CO); 209.1, 208.8 (C3); 132.4, 131.7, 129.4, 129.3, 129.2, 129.1 (Ph); 90.5, 90.4, 87.9, 87.8 (Cp); 68.4, 61.9 (CH2); 52.1, 52.6 (C2); 47.9, 42.1 (NMe); 42.4, 42.3 (C3Me).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2(Me)C1NMe(CH2Ph)}]CF3SO3, 5b (Scheme 11)
Scheme 11. Structure of the cation of 5b.
Scheme 11. Structure of the cation of 5b.
Pharmaceutics 13 01158 sch011
From 1d and 2-butyne. Brown solid, yield 70%. Eluent for chromatography: MeCN. Anal. calcd. for C26H26F3Fe2NO5S: C, 49.31; H, 4.14; N, 2.21; S, 5.06. Found: C, 49.20; H, 4.19; N, 2.16; S, 5.09. IR (CH2Cl2): ῦ/cm−1 = 1986vs (CO), 1805s (μ-CO), 1650m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 7.49–7.44, 7.32 (m, 5 H, Ph); 5.57, 5.55, 5.25, 5.23 (s, 10 H, Cp); 4.87, 4.57 (d, 2JHH = 14.7 Hz, 2 H, CH2); 3.95 (s, 3 H, C3Me); 3.81, 3.06 (s, 3 H, NMe); 2.14 (s, 3 H, C2Me). Isomer ratio (Z/E) = 1.5. 13C{1H} NMR (acetone-d6): δ/ppm = 258.3 (μ-CO); 228.5, 226.9 (C1); 211.0, 210.7 (CO); 202.6, 202.2 (C3); 133.4, 132.4 (ipso-Ph); 129.2, 129.1, 129.0, 128.9, 128.8, 128.7 (Ph); 91.2, 88.3, 88.2 (Cp); 65.9, 61.3 (C2); 65.6, 65.5 (CH2); 45.4, 41.6 (NMe); 36.4, 36.3 (C3Me); 15.5, 15.0 (C2Me).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Et)C2(Et)C1NMe(CH2Ph)}]CF3SO3, 5c (Scheme 12)
Scheme 12. Structure of the cation of 5c.
Scheme 12. Structure of the cation of 5c.
Pharmaceutics 13 01158 sch012
From 1d and 3-hexyne. Red-brown solid, yield 69%. Eluent for chromatography: MeCN. Anal. calcd. for C28H30F3Fe2NO5S: C, 50.85; H, 4.57; N, 2.12; S, 4.85. Found: C, 50.78; H, 4.64; N, 2.09; S, 4.94. IR (CH2Cl2): ῦ/cm−1 = 1986vs (CO), 1804s (μ-CO), 1649m (C1 = N). 1H NMR (CDCl3): δ/ppm = 7.41–7.29, 7.08 (m, 5 H, Ph); 5.59, 5.40, 4.85, 4.35 (d, 2JHH ≈ 14 Hz, 2 H, NCH2); 5.28, 5.02 (s, 10 H, Cp); 4.19, 4.09, 2.53, 2.28, 2.07 (br, 4 H, CH2CH3); 3.69, 2.90 (s, 3 H, NMe); 1.95, 1.72, 1.44, 1.28, 1.14 (br, 6 H, CH2CH3). Isomer ratio (trans-Z/cis-E) = 1.8. 13C{1H} NMR (CDCl3): δ/ppm = 258.9, 257.8 (μ-CO); 228.3, 226.7 (C1); 211.9, 211.4 (CO); 210.9, 210.3 (C3); 132.4, 131.3, 129.4, 129.3, 129.2, 129.1, 128.9 (Ph); 90.7, 90.5, 88.0, 87.8 (Cp); 68.5, 61.8 (NCH2); 66.9 (C2); 46.4, 42.1 (NMe); 43.2, 42.9 (C3-CH2); 24.1, 23.0 (C2-CH2); 20.2, 19.3, 14.9, 13.4 (CH2CH3).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2HC1NMe(4-C6H4OMe)}]CF3SO3, 6a (Scheme 13) [52]
Scheme 13. Structure of the cation of 6a.
Scheme 13. Structure of the cation of 6a.
Pharmaceutics 13 01158 sch013
From 1e and propyne (THF solution, ca. 1 mol/L). Brown solid, yield 73%. Eluent for chromatography: MeCN. Anal. calcd. for C25H24F3Fe2NO6S: C, 47.27; H, 3.81; N, 2.21; S, 5.05. Found: C, 47.13; H, 3.91; N, 2.13; S, 5.09. IR (CH2Cl2): ῦ/cm−1 = 1991vs (CO), 1811s (μ-CO), 1642m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 7.64, 7.24, 7.20, 6.90 (m, 4 H, C6H4); 5.60, 5.55, 5.32, 4.99 (s, 10 H, Cp); 4.85, 4.72 (s, 1 H, C2H); 4.41, 3.69 (s, 3 H, NMe); 4.02, 3.95 (s, 3 H, C3Me); 3.81 (s, 3 H, OMe). Isomer ratio (E/Z) = 2. 13C{1H} NMR (acetone-d6): δ/ppm = 255.6, 254.8 (μ-CO); 230.4, 229.0 (C1); 210.7, 210.4 (CO); 209.5, 208.8 (C3); 160.3, 159.8 (ipso-C6H4); 139.4, 136.5, 126.1, 122.8, 114.6, 114.3 (C6H4); 91.0, 88.1, 87.8 (Cp); 55.2, 55.1 (OMe); 53.7, 53.4 (C2); 53.6, 46.2 (NMe); 41.4, 41.3 (C3Me).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Ph)C2HC1NMe(4-C6H4OMe)}]CF3SO3, 6b (Scheme 14)
Scheme 14. Structure of the cation of 6b.
Scheme 14. Structure of the cation of 6b.
Pharmaceutics 13 01158 sch014
From 1e and phenylacetylene. Brown solid, yield 71%. Eluent for chromatography: MeCN. Anal. calcd. for C30H26F3Fe2NO6S: C, 51.68; H, 3.76; N, 2.01; S, 4.60. Found: C, 51.55; H, 3.88; N, 2.06; S, 4.71. IR (CH2Cl2): ῦ/cm−1 = 1994vs (CO), 1814s (μ-CO), 1639m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 7.84, 7.68, 7.56, 7.43, 7.28, 6.98 (m, 9 H, Ph + C6H4); 5.57, 5.33, 5.27, 5.18 (s, 10 H, Cp); 4.93, 4.72 (s, 1 H, C2H); 4.51, 3.80 (s, 3 H, NMe); 3.99, 3.83 (s, 3 H, OMe). Isomer ratio (E/Z) = 4. 13C{1H} NMR (acetone-d6): δ/ppm = 254.7 (μ-CO); 229.2 (C1); 210.0 (CO); 204.9 (C3); 159.9 (ipso-C6H4); 156.4 (ipso-Ph); 128.8, 128.4, 127.2, 127.0, 126.9, 126.5, 122.7, 114.7, 114.4 (Ph + C6H4); 92.0, 91.9, 88.3, 87.9 (Cp); 55.3, 55.2 (OMe); 54.2, 54.0 (C2); 54.3, 46.0 (NMe).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2(Me)C1NMe(4-C6H4OMe)}]CF3SO3, 6c (Scheme 15)
Scheme 15. Structure of the cation of 6c.
Scheme 15. Structure of the cation of 6c.
Pharmaceutics 13 01158 sch015
From 1e and 2-butyne. Brown solid, yield 83%. Eluent for chromatography: MeCN. Anal. calcd. for C30H26F3Fe2NO6S: C, 51.68; H, 3.76; N, 2.01; S, 4.60. Found: C, 51.55; H, 3.88; N, 2.06; S, 4.71. IR (CH2Cl2): ῦ/cm−1 = 1986vs (CO), 1810s (μ-CO), 1626m (C1 = N). 1H NMR (acetone-d6): δ/ppm = 7.59, 7.20, 7.10, 6.92 (m, 4 H, C6H4); 5.57, 5.45, 5.25, 4.96 (s, 10 H, Cp); 4.44, 3.56 (s, 3 H, NMe); 3.93, 3.78 (s, 3 H, OMe); 3.89, 3.82 (s, 3 H, C3Me); 2.26, 1.50 (s, 3 H, C2Me). Isomer ratio (Z/E) = 2.2. 13C{1H} NMR (acetone-d6): δ/ppm = 257.3, 255.4 (μ-CO); 229.9, 229.1 (C1); 212.3, 211.8 (CO); 204.2, 203.6 (C3); 161.3, 160.7 (ipso-C6H4); 138.4, 138.2; 127.7, 123.0, 115.4, 115.1 (C6H4); 92.5, 92.2, 89.6, 89.2 (Cp); 67.8, 67.2 (C2); 56.2, 56.1 (OMe); 52.7, 46.5 (NMe); 37.9, 37.2 (C3Me); 15.8, 15.3 (C2Me).
[Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2(Me)C1NMe(2-naphthyl)}]CF3SO3, 7 (Scheme 16)
Scheme 16. Structure of the cation of 7.
Scheme 16. Structure of the cation of 7.
Pharmaceutics 13 01158 sch016
From 1f and 2-butyne. Brown-red solid, yield 90%. Eluent for chromatography: MeCN. Anal. calcd. for C29H26F3Fe2NO5S: C, 52.04; H, 3.92; N, 2.09; S, 4.79. Found: C, 51.89; H, 3.99; N, 2.01; S, 4.94. IR (CH2Cl2): ῦ/cm−1 = 1987vs (CO), 1811s (μ-CO), 1637m (C1 = N), 1616w-m (arom C = C). 1H NMR (acetone-d6): δ/ppm = 8.32–7.61, 7.28 (m, 7 H, arom CH); 5.67, 5.51, 5.36, 5.01 (s, 10 H, Cp); 4.67, 3.75 (s, 3 H, NMe); 3.97, 3.93 (s, 3 H, C3Me); 2.26, 1.55 (s, 3 H, C2Me). Isomer ratio (Z/E) = 1.5. 13C{1H} NMR (acetone-d6): δ/ppm = 256.0, 254.5 (μ-CO); 230.9 (C1); 211.4 (CO); 203.9 (C3); 141.0 (ipso-Naph); 133.4, 132.7, 132.6, 129.7, 129.4, 128.5, 128.4, 128.1, 127.7, 127.5, 124.6, 122.8, 118.9 (arom CH); 91.7, 91.3, 88.9, 88.3 (Cp); 67.1 (C2); 51.4, 45.6 (NMe); 37.1, 36.4 (C3Me); 15.0, 14.5 (C2Me).

2.3. X-ray Crystallography

Crystal data and collection details for 2c and 3a are reported in Table 1. The data were recorded on a Bruker APEX II diffractometer equipped with a PHOTON100 detector using Mo–Kα radiation. The data were corrected for Lorentz polarization and absorption effects (empirical absorption correction SADABS) [54]. The structures were solved by direct methods and refined by full-matrix least-squares based on all data using F2 [55]. Hydrogen atoms were fixed at calculated positions and refined by a riding model. All non-hydrogen atoms were refined with anisotropic displacement parameters. The crystals of 2c appeared to be non-merohedrally twinned. The TwinRotMat routine of PLATON [56] was used to determine the twinning matrix and to write the reflection data file (.hkl) containing the twin components. Refinement was performed using the instruction HKLF 5 in SHELXL and one BASF parameter, which refined as 0.213 (5). Moreover, the whole molecule 2c is disordered and, therefore, it has been split into two positions and refined anisotropically employing one occupancy factor per disordered group. Because of this twinning and high disorder, several restraints were used during the refinement. The details are included in the CIF file. Therefore, even if the connectivity and overall geometry of 2c are certain, bonding distances and angles must be taken with care.

2.4. Solubility and Stability in Water and Determination of Octanol–Water Partition Coefficients

(a)
Solubility in D2O. A suspension of the selected diiron compound (3–5 mg) in a D2O solution (ca. 1 mL) containing Me2SO2 (3.36 × 10−3 M) as internal standard [57] was vigorously stirred at 21 °C for 1.5 h. The resulting saturated solution was filtered over celite, transferred into an NMR tube and then analyzed by 1H NMR spectroscopy (delay time = 3 s; number of scans = 20). The concentration (solubility) was calculated by the relative integral (related to Cp and/or NMe signals) with respect to Me2SO2 (δ = 3.14 ppm).
(b)
Stability in DMSO-d6/D2O solution. The selected diiron compound (ca. 4 mg) was added of 1 mL of D2O (3a,c) or DMSO-d6/D2O (2:1 v/v) mixture (2a,c, 3b, 4a–c, 5a–c, 6a–c, 7), containing Me2SO2 (3.36 × 10−3 M). The resulting mixture was stirred for 30 min and then filtered over celite; the filtrated solution was transferred into an NMR tube. The sample was analyzed by 1H NMR (time = 0) and then heated at 37 °C for 72 h. After cooling to room temperature, the solution was separated from a small amount of solid by filtration over celite, and the new 1H NMR spectrum was recorded (delay time = 3 s; number of scans = 20). In every case, no newly formed organometallic species was detected. The amount of starting material in solution (% with respect to the initial spectrum) was calculated by the relative integral (vs. Cp and/or NMe signals) with respect to Me2SO2 as internal standard. Data related to the spectra recorded at time = 0 are reported in the following.
2a.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 7.43, 7.15, 6.99 (m, 10 H, Ph); 5.83, 5.14 (d, 2 H, CH2); 5.48, 5.08 (s, 10 H, Cp); 4.54, 4.27 (d, 2 H, CH2); 3.87 (s, 3 H, C3Me).
2c.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 7.4–6.9 (m, 10 H, Ph); 4.73 (d, 1 H, CH2); 5.42, 5.01 (s, 10 H, Cp); 3.82 (s, 3 H, C3Me); 2.17 (s, 3 H, C2Me).
3a.1H NMR (D2O): δ/ppm = 5.92–5.82, 5.62–5.54, 5.38–5.33 (m, CH = CH2); 5.29, 5.29, 4.92, 4.90 (s, Cp); 3.68, 3.67 (s, C3Me); 3.65, 3.01 (NMe). Isomer ratio = 1.1.
3b.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 7.73, 7.59, 7.47 (m, 5 H, Ph); 5.97 (br, CH = CH2); 5.65, 5.54, 5.35, 5.31 (m, CH = CH2); 5.27, 5.25, 5.15 (s, 10 H, Cp); 4.12 (m, NCH2); 3.85, 3.23 (s, 3 H, NMe). Isomer ratio = 1.5.
3c.1H NMR (D2O): δ/ppm = 5.99, 5.62, 5.52, 5.36 (m, CH = CH2); 5.26, 5.24, 4.92, 4.90 (s, Cp); 3.68, 3.65, 3.05 (s, C3Me + NMe); 2.04 (s, C2Me).
4a.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 5.35, 5.34, 5.17, 4.97 (s, 10 H, Cp); 3.77, 3.75 (s, 3 H, C3Me); 3.73, 3.00 (s, 3 H, NMe); 3.43 (m, 1 H, CHCy); 2.08–0.85 (m, 10 H, CH2Cy). Isomer ratio = 1.9.
4c.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 5.35, 5.34, 5.01, 4.97 (s, 10 H, Cp); 3.77, 3.75, 3.73 (s, NMe + C3Me); 2.1, 1.8–1.1 (m, 10 H, CH2Cy); 1.87, 1.85 (s, 3 H, C2Me). Isomer ratio = 1.8.
5a.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 7.48, 7.36, 7.21 (m, 5 H, Ph); 5.45, 5.43, 5.07 (s, 10 H, Cp); 3.88 (s, 3 H, C3Me); 3.61, 3.03 (s, 3 H, NMe). Isomer ratio = 1.2.
5b.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 7.49–7.36, 7.19 (m, 5 H, Ph); 5.40, 5.39, 5.06, 5.01 (s, 10 H, Cp); 4.50 (d, 1 H, CH2); 3.79 (s, 3 H, C3Me); 3.65, 3.06 (s, 3 H, NMe); 1.94, 1.91 (C2Me). Isomer ratio = 1.7.
5c.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 7.51, 7.43, 7.14 (m, 5 H, Ph); 5.53, 5.30, 4.51 (d, 2 H, NCH2); 5.42, 5.40, 5.12, 5.09 (s, 10 H, Cp); 4.18, 4.13, 2.46, 2.13 (br, 4 H, CH2CH3); 3.72, 2.94 (s, NMe); 2.11, 1.69, 1.13, 0.98 (br, CH2CH3). Isomer ratio = 1.7.
6a.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 7.52, 7.21, 7.10, 6.93 (m, 4 H, C6H4); 5.43, 5.37, 5.14, 4.85 (s, 10 H, Cp); 4.63 (s, 1 H, C2H); 4.41, 3.69 (s, 3 H, NMe); 4.23, 3.93 (s, 3 H, C3Me); 3.87, 3.79 (s, 3 H, OMe). Isomer ratio = 2.
6b.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 7.73, 7.58, 7.43, 7.23, 7.16, 6.98 (m, 9 H, Ph + C6H4); 5.37, 5.19, 5.12, 5.03 (s, 10 H, Cp); 4.32, 3.96 (s, 3 H, NMe); 3.79, 3.64 (s, 3 H, OMe). Isomer ratio = 3.
6c.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 7.94, 7.21, 7.04, 6.94 (m, 4 H, C6H4); 5.45, 5.32, 5.10, 4.84 (s, 10 H, Cp); 4.30, 3.46 (s, 3 H, NMe); 3.93, 3.78 (s, 3 H, OMe); 3.74 (s, 3 H, C3Me); 2.00, 1.44 (s, 3 H, C2Me). Isomer ratio = 2.3.
7.1H NMR (D2O/DMSO-d6 2:1): δ/ppm = 8.23–7.21 (m, 7 H, arom CH); 5.51, 5.34, 5.16, 4.86 (s, 10 H, Cp); 4.47, 3.61 (s, 3 H, NMe); 3.82, 3.78 (s, 3 H, C3Me); 2.09, 1.44 (s, 3 H, C2Me). Isomer ratio = 2.
(c)
Octanol–water partition coefficients (Log Pow). Partition coefficients (Pow; IUPAC: KD partition constant [58]), defined as Pow = corg/caq, where corg and caq are molar concentrations of the selected compound in the organic and aqueous phase, respectively, were determined by the shake-flask method and UV–Vis measurements [36,59,60]. Deionized water and 1-octanol were vigorously stirred for 24 h to enable saturation of both phases, then separated by centrifugation. A stock solution of the selected diiron compound (ca. 2 mg; 2a–c, 3b, 4a–c, 5a, 5c, 6b,c, 7) was prepared by first adding acetone (50 μL, to help solubilization), followed by water-saturated octanol (2.5 mL). The solution was diluted with water-saturated octanol (ca. 1:3 v/v ratio, cFe2 ≈ 10−4 M, so that 1.5 ≤ A ≤ 2.0 at λmax) and the UV–Vis spectrum was recorded (A0org). An aliquot of the solution (Vorg = 1.2 mL) was transferred into a test tube and octanol-saturated water (Vorg = Vaq = 1.2 mL) was added. The mixture was vigorously stirred for 15 min at 21 °C and then centrifuged (5000 rpm, 10 min). The UV–Vis spectrum of the organic phase was recorded (Aforg) and the partition coefficient was calculated as Pow = Aforg/(A0org − Aforg), where A0org and Aforg are the absorbance in the organic phase before and after partition with the aqueous phase, respectively [59]. An inverse procedure was followed for 3a, 3c, 5b and 6a, starting from a solution of the compound in octanol-saturated water. The partition coefficient was calculated as Pow = (A0aq − Afaq)/Afaq where A0aq and Afaq are the absorbance in the aqueous phase before and after partition with the organic phase, respectively. The wavelength of the maximum absorption of each compound (ca. 300 nm) was used for UV–Vis quantitation. The procedure was repeated three times for each sample (from the same stock solution); results are given as mean ± standard deviation. Naphthoquinone was used as a reference compound (Log P = 1.8 ± 0.2; literature [61]: 1.71).

2.5. Cell Culture and Cytotoxicity Studies

In vitro cytotoxicity investigations were carried out by using human ovarian carcinoma cisplatin-sensitive A2780 (ECACC93112519), human ovarian carcinoma cisplatin-resistant A2780cisR (ECACC 93112517) and mouse embryo fibroblasts Balb/3T3 clone A31 (ATCC CCL-163) cell lines. A2780 and A2780cisR were purchased from the European Collection of Authenticated Cell Cultures (ECACC), and Balb/3T3 clone A31 cell line from the American Type Culture Collection (ATCC).
A2780 and A2780cisR cells were routinely cultured in RPMI 1640 (Merck) containing 2 mM of L-glutamine (Merck), 1% of penicillin/streptomycin solution (Merck—10 000 U mL−1: 10 mg mL−1), 10% of fetal bovine serum (Merck—FBS) and antimycotic (InvivoGen, USA), and Balb/3T3 clone A31 in Dulbecco’s modified Eagle medium (Merck—DMEM) supplemented with 4 mM of L-glutamine, 1% of penicillin/streptomycin solution, 10% of calf serum (Merck) and antimycotic. The acquired resistance of A2780cisR cells was maintained by routine supplementation of media with 1 μM of cisplatin (Merck). The cultures were maintained at 37 °C and in a 5% CO2-enriched atmosphere.
A2780, A2780cisR and Balb/3T3 clone A31 cells were seeded in 96-well tissue culture polystyrene plates at a concentration of 3 × 103, 6 × 103 and 1 × 103 cells per well, respectively. After overnight incubation, the cells were treated with different concentrations (0–100 µM) of the selected compounds for 72 h at 37 °C in a 5% atmosphere of CO2. Stock solutions of compounds were prepared in DMSO and sequentially diluted in medium (final DMSO concentration of 0.5%). Cells incubated with cisplatin (0−100 μM) were used as positive control. At the end of the incubation time, cell viability was assessed by means of WST-1 tetrazolium salt reagent (Roche). Briefly, cells were incubated for 4 h with the tetrazolium salt reagent diluted 1:10, at 37 °C and 5% CO2. Measurements of formazan dye absorbance, which directly correlates with the number of viable cells, were carried out with a micro-plate reader (Biorad) at 450 nm, using 655 nm as reference wavelength. The 50% inhibitory concentration (IC50) refers to compound concentration at which 50% of cell death is observed with respect to the control. For each tested compound, assay was performed on triplicate. The concentration effect curves were generated by nonlinear regression curves (GraphPad Prism Software, 2021) and the data reported as mean ± standard deviation.

2.6. ROS Determination

The intracellular production of reactive oxygen species (ROS) upon treatment of the compounds 3b, 4a, 4c, 5b and 6c was measured by using the DCFH-DA (2′,7′-dichlorodihydrofluorescein diacetate, Merck) assay, based on cellular uptake of the non-fluorescent diacetate following deacetylation by esterases (2′,7′-dichlorodihydrofluorescein, DCFH) and oxidation to the fluorescent dichlorofluorescein (2′,7′-dichlorofluorescein, DCF) [62]. A2780 cells were seeded at concentration of 4 × 104 cells/well/90 µL of complete growth medium into 96-well plates. After overnight incubation, the cells were treated following manufacturer protocol. 100 µL of a solution containing the fluorogenic probe were added to the culture medium and, after 1 h of incubation with 5% CO2 at 37 °C, the cells were exposed to a final concentration of 20 µM of the tested complexes; H2O2 100 µM was used as a positive control. Stock solutions of compounds were prepared as described above; cells incubated with equal amounts of DMSO in supplemented RPMI were used as control. The fluorescence was measured up to 24 h with an excitation wavelength of 485 nm and with a 535 nm emission filter by Multilabel Counter (PerkinElmer). For each tested compound, assay was performed on triplicate. The data were reported as mean ± standard deviation, statistical differences were analyzed using one-way analysis of variance (ANOVA) and a Tukey test was used for post hoc analysis. A p-value < 0.05 is considered statistically significant.

2.7. Biomolecules Binding Studies

(a)
Sample preparation. Cytochrome c (Cyt c) was commercially available and used as received; the TrxR dodecapeptide (TrxR-pept) was synthesized as reported in the literature [63,64]. The stock solutions of the selected iron-based complexes were prepared in DMSO up to a final concentration of 10−2 M. The stock solution of TrxR-pept was prepared in LC-MS grade water by dissolving the required amount of lyophilized peptide to reach a final concentration of 10−3 M. The stock solution of Cyt c 10−3 M was prepared by dissolving the required amount of protein in 2 mM ammonium acetate solution at pH 6.8. In the interaction tests with TrxR-pept, opportune aliquots of each complex and TrxR-pept stock solutions were mixed and diluted with LC-MS grade water to 10−4 M final concentration and a TrxR-pept/complex ratio of 1:1. For each iron-based complex/Cyt c pair, appropriate aliquots of the respective stock solutions were mixed and subsequently diluted with 2 mM ammonium acetate solution (pH 6.8) to a final protein concentration of 10−4 M and a protein-to-metal molar ratio of 1:2. All the sample mixtures were incubated for 24 h at 37 °C. Subsequently, opportune dilutions were performed as detailed:
the TrxR-pept containing solutions were further diluted with LC-MS grade water to a final TrxR-pept concentration of 10−5 M and added with 0.1% v/v of formic acid just before infusion in the mass spectrometer;
the protein-containing solutions were diluted with 2 mM ammonium acetate solution (pH 6.8) to a final protein concentration of 10−6 M and added with 0.1% v/v of formic acid just before infusion.
(b)
ESI-MS instrumental parameters. The ESI mass spectra were acquired using a TripleTOF® 5600+ high-resolution mass spectrometer (Sciex, Framingham, MA, USA), with a DuoSpray® interface operating with an ESI probe. Respective ESI mass spectra were acquired through direct infusion at 7 μL min−1 flow rate. The ESI source parameters were optimized for each biomolecule and were as follows: for TrxR dodecapeptide positive polarity, ionspray voltage floating 5500 V, temperature 0, ion source gas 1 (GS1) 35 L min−1; ion source gas 2 (GS2) 0; curtain gas (CUR) 20 L min−1, declustering potential (DP) 300 V, collision energy (CE) 10 V, range 1070–1600 m/z; for Cyt c positive polarity, ionspray voltage floating 5500 V, temperature 0, ion source gas 1 (GS1) 35 L min−1; ion source gas 2 (GS2) 0; curtain gas (CUR) 20 L min−1, declustering potential (DP) 180 V, collision energy (CE) 10 V, range 500–1800 m/z. For acquisition, Analyst TF software 1.7.1 (Sciex, Framingham, MA, USA) was used and deconvoluted spectra were obtained by using the Bio Tool Kit micro-application v.2.2 embedded in PeakViewTM software v.2.2 (Sciex, Framingham, MA, USA).

3. Results and Discussion

3.1. Synthesis and Structural Characterization of Diiron Complexes

A series of diiron complexes with novel vinyliminium ligands, 2–7, was synthesized from the respective aminocarbyne precursors, 1a–f, by means of a two-step procedure (see Scheme 17 and Table 2). This consists of the preliminary CO-NCMe substitution promoted by trimethylamine-N-oxide, followed by replacement of the labile acetonitrile ligand with the appropriate alkyne in dichloromethane solution at room temperature. Three main alkynes were selected, i.e., propyne, phenylacetylene and 2-butyne, which previously demonstrated to provide enhanced cytotoxicity to other vinyliminium complexes (Figure 1) [39]. In addition, 3-hexyne was employed to build 5c, in order to outline the possible effect of the length of alkyl chains (upon comparison with 5b) [39]. Note that 1a–f are available from [Fe2Cp2(CO)4] in high yields via multigram scale synthesis [36,47].
The products 2–7 were purified by alumina chromatography and finally isolated as air-stable triflate salts in 69–95% yields. They were fully characterized by elemental analysis, IR and NMR spectroscopy (Figures S1–S32). The IR spectra (in CH2Cl2 solution) share a common pattern with three main bands; two of them are ascribable to the terminal and bridging carbonyl stretching vibrations and fall in the ranges 1986–1992 cm−1 and 1804–1814 cm−1, respectively. The third band is related to the iminium (C1 = N) function and reveals a double bond character; however, its wavenumber is rather sensitive to the electronic properties of all vinyliminium substituents (i.e., R, Y, R′ and R″). For instance, it occurs at 1644 and 1664 cm−1 in 2a and 5a, respectively, the latter complex differing from the former for a CH3 group instead of CH2Ph on nitrogen. The substitution of the C2 carbon determines a significant decrease of the C1 = N wavenumber, as well as a minor decrease for the CO signal [e.g., for 4a: ῦ/cm−1 = 1989 (CO), 1805 (μ-CO), 1661 (C1 = N); for 4c: ῦ/cm−1 = 1986 (CO), 1805 (μ-CO), 1646 (C1 = N)]. The allyl moiety in 3a–c manifests itself with a weak infrared absorption at ca. 1640 cm−1 (double carbon–carbon bond). The NMR spectra of 2a–c (in acetone-d6) show a single set of resonances, and in particular the singlets due to the Cp rings, in the 1H spectra, have been detected at ca. 5.6 and 5.3 ppm, pointing out their mutual cis arrangement. In fact, cis/trans isomerism was previously recognized in some analogous complexes, the trans form being associated to a diagnostic upfield shift of one Cp resonance (δ ≤ 4.5 ppm) [65]. Consistently, the Cp signals of the remaining complexes 3–7 (apart from 5c, vide infra) fall within the range 4.97–5.67 ppm, thus indicating cis configuration (acetone-d6 or CDCl3 solutions). However, two sets of signals are generally observed in the spectra of 3–7 with a ratio variable between 1.2 and 2, related to E-Z isomerism generated by the two possible orientations of the iminium substituents (R ≠ Y). The E isomer is prevalent in 3a, 4b, 6a and 6b, exhibiting high-field signals for the N-methyl [e.g., at 3.27 (1H) and 41.8 (13C) ppm in the case of 3a]; the opposite Z isomer prevails in 3b, 3c, 4c, 5b, 5c, 6c and 7 [e.g., for 3b: δ(NMe) = 3.89 (1H) and 47.7 (13C) ppm]. Analogous features were generally recognized on related vinyliminium complexes with R = CH2Ph or Xyl and Y = Me (see Introduction and Figure 1) [39,66]. In the case of 5c, a comparison of the NMR data with the library of data available in the literature points out the existence in CDCl3 solution of trans-Z and cis-E isomers, in ca. 2:1 ratio; in detail, the Cp and NMe groups resonate at 4.85, 4.35 and 3.69 ppm in the former isomer, and at 5.59, 5.40 and 2.90 in the latter (1H spectra).
Salient 13C NMR features of 2–7 concern the carbon nuclei constituting the C3 bridging chain, which have been recognized at 223.6–230.4 ppm (C1), 52.0–67.1 ppm (C2) and 201.9–210.9 ppm (C3). The typical low-field values for C1 and C3 reflect the (amino)alkylidene and bridging alkylidene nature of these centers, respectively [31]; on the other hand, the interval found for C2 indicates a partial alkenic nature, and the presence of an alkyl substituent causes a downfield shift by 10–15 ppm (e.g., δ = 52.9 ppm for 2a and 65.7 ppm for 2c).
The structures of 2c and 3a were confirmed by single-crystal X-ray diffraction studies: a view of the cation of 3a is shown in Figure 2, with relevant bonding parameters in the caption, while a view of the cation of 2c is supplied as Supplementary Materials (Figure S33). In both structures, the Cp ligands are cis oriented, while the N-substituents in 3a are arranged according to the E configuration. In 3a, the bridging C3 carbon is slightly asymmetric between the two irons [Fe(2)-C(3) 2.036(4) and Fe(1)-C(3) 1.940(4) Å], while the bridging carbonyl exhibits a lower asymmetry [Fe(2)-C(12) 1.934(4) and Fe(1)-C(12) 1.911(4) Å]. The N(1)-C(1) [1.285(4) Å], C(1)-C(2) [1.419(5) Å] and C(2)-C(3) [1.409(5) Å] distances evidence an extensive charge delocalization, as already found in similar complexes [39,44,66], suggesting that the representation of the bridging ligand as vinyliminium is appropriate but not univocal. In particular, a comparison between the C(2)-C(3) distance and C(6)-C(7) [1.487(5) Å] and C(7)-C(8) [1.319(5) Å] outlines that the former is intermediate between a single and a double Csp2-Csp2 bond.

3.2. Solubility and Stability in Water and Determination of Octanol–Water Partition Coefficients

The water solubility of complexes was measured by means of 1H NMR spectroscopy (Table 3). In D2O, most of the complexes showed an appreciable solubility, falling in the millimolar range, and an isomer ratio slightly different with respect to what is found in acetone-d6 or CDCl3. The solubilities of 3a–c, 4a–c and 6a (ranging from 1.02 to 4.02 g∙L−1) are comparable to that of cisplatin (ca. 3 g∙L−1) and other common platinum drugs [67,68,69,70]. On the other hand, 2b,c, containing more than one phenyl group, and 7, containing a naphthyl, are almost insoluble in D2O.
Next, the stability of the complexes was assessed by 1H NMR spectroscopy in DMSO-d6/D2O solutions, stored at 37 °C for 72 h (Table 3). In general, the complexes exhibited a considerable stability, with an average of 83% of starting material detected at the end of the experiment. Especially those complexes with R′ = Me and R″ = H appear robust (mean percentage = 90%). The slow decomposition of the complexes is featured by the formation of a minor amount of precipitate suggesting extensive rupture of the diiron scaffold [71], while new organometallic species were not detected in solution. The octanol–water partition coefficients were measured by means of a UV–Vis technique (Table 3, see Experimental for details). Most complexes display a substantial amphiphilic character, with Log Pow values ranging from −0.77 to 1.30. In general, the presence of the phenyl group on the C3 carbon leads to a significant increase in lipophilicity, while the allyl group on the iminium moiety determines the opposite effect. Complex 2b, containing three phenyl units, displays the highest Log Pow value (1.4) and was excluded from the biological tests due to insufficient solubility.

3.3. Cytotoxicity

Firstly, the antiproliferative activity of the diiron complexes was assessed on human ovarian carcinoma A2780 cancer cells and, to estimate the selectivity, also on nontumoral Balb/3T3 clone A31 cells (Table 4 and Figure S34). Cisplatin was employed as a reference. In general, diiron complexes exhibit strong cytotoxicity against the considered cancer cell line, with IC50 values falling in the low micromolar range. Furthermore, an impressive selectivity was recognized with respect to the Balb/3T3 cell line: the selectivity index (S.I.) ranges from 7 to 52, while the value for cisplatin is only 2 (Table 4).
The cytotoxicity of the complexes against the A2780 cancer cell line approximately correlates with the lipophilicity, and actually the most hydrophilic compounds 3a and 3c (Log Pow < −0.7), bearing R = CH2CH = CH2, display a relatively lower potency (IC50 ≈ 17 μM). On the other hand, 4a, 4c, 5a, 5b, 6a and 6c, bearing R = cyclohexyl, benzyl or 4-methoxyphenyl, show a potent activity despite the negative Log Pow values.
The analogous complexes 5b and 5c, differing to each other for the alkyl chain on C2 and C3, display comparable activity. The activity of 3a (IC50 = 17.7 ± 0.8 μM, Log Pow = −0.75) is enhanced compared to that previously determined on the same cell line for the analogous complex [Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2HC1NMe2}]CF3SO3 (IC50 = 35 ± 3 μM, Log Pow = −0.3) [39], indicating that the replacement of one methyl N-substituent with the allyl may improve the anticancer activity. Conversely, IC50 values of 0.50 ± 0.06 μM and 0.90 ± 0.06 μM were previously found for, respectively, [Fe2Cp2(CO)(μ-CO){μ-η13-C3(Me)C2(Me)C1NMe(Xyl)}]CF3SO3 (Log Pow = 0.0) and [Fe2Cp2(CO)(μ-CO){μ-η13-C3(Ph)C2HC1NMe(Xyl)}]CF3SO3 (Log Pow = 0.4) [39], thus outlining the beneficial effect provided by a xylyl ring among possible aryl groups (see data related to 7 and 6b).
Note that here R = cyclohexyl, benzyl and 4-methoxyphenyl supply the highest selectivity indexes, and especially the combination with R′ = R″ = Me or Et is convenient (average S.I. ≥ 30). It is worth mentioning that R = cyclohexyl and R = 4-methoxyphenyl contribute to the best cytotoxicity profiles also concerning the series of aminocarbyne complexes 1a–f [37]. Complex 4c, derived from the assembly of cyclohexylisocyanide and 2-butyne (Table 2), appears as the most performant within the series of vinyliminium complexes 2–7. The favorable effect on the anticancer activity provided by the decoration of metal complexes with the cyclohexyl (Cy) moiety has been documented in the literature, attributed to the compact and hydrophobic structure of Cy facilitating the entrance of the drug into the tumor cell [72,73,74].
A selection of the most promising compounds was further investigated. First, the cytotoxicity of 3b, 4a, 4c, 5b and 6c was measured on the cisplatin-resistant A2780cisR cell line. Overall, a loss of activity was observed with respect to the cisplatin-sensitive cell line (A2780), nevertheless 4a, 4c and 5b maintain a higher level of activity than that of cisplatin under the same conditions.

3.4. ROS Production

Previous findings pointed out that diiron vinyliminium complexes exert their anticancer activity mainly by unbalancing cell redox homeostasis, and this phenomenon may be ascribable to several routes possibly taking place inside the cells: (1) complex reduction by neutralization of the net cationic charge; (2) fragmentation into monoiron derivatives; and (3) disassembly of the diiron core releasing the two Fe+I centers [36,39,40,45]. According to this premise, we considered of interest to evaluate the ability of selected complexes, 3b, 4a, 4c, 5b and 6c, to induce intracellular ROS production. Thus, fluorescence measurements were carried out with the DCFH-DA assay, by exposing A2780 cells to, respectively, the diiron complexes, the reference drug cisplatin and H2O2 (positive control). All the diiron complexes elicited an appreciable ROS production, significantly higher than that of cisplatin, especially after ca. 20 h and progressively increasing up to 24 h (Figure 3). The benzyl complex 5b revealed the most effective of the series in inducing the ROS production, and this outcome may be associated with the high cytotoxicity exhibited by this complex against the cancer cells, despite its relatively low hydrophobicity.

3.5. Protein Binding Studies

Recent results on the anticancer activity of the aminocarbyne complex 1e highlighted a possible mechanism involving the inhibition of the selenoenzyme thioredoxin reductase (TrxR) [36]; more specifically, substantial inhibition of this enzyme was detected in human pancreatic PSN-1 cancer cells. Former studies by other authors demonstrated a similar mode of action by ferrocenyl drug candidates [75]. Therefore, we became interested in elucidating the potentiality of the new vinyliminium complexes to inhibit TrxR.
During the past years, high-resolution electrospray ionization mass spectrometry (HR-ESI-MS) has emerged as a valuable and powerful tool to study the formation of adducts derived from the interaction of metallodrugs with biomolecules [76,77,78]. Moreover, this technique allows the identification of the precise nature of the fragments attached to protein side chains and, thus, to infer some relevant mechanistic information [79,80,81].
Since the amount of native TrxR required for the ESI-MS experiment, albeit low, can hardly be found commercially available at a reasonable cost, for the present study we opted for a model system that has been already successfully employed [82]. The model is a synthetic dodecapeptide (TrxR-pept) mimicking the C-terminal tryptic portion of the TrxR enzyme, consisting in the amino acidic sequence {Ac-SGGDILQSG[CU]G-NH2} and featured by the presence of the peculiar {-Cys-Sec-} motif (Cys = cysteine; Sec = selenocysteine).
First, we verified the suitability of the model system by analyzing its interaction with complex 1e, for which TrxR inhibition was previously recognized by enzymatic assay (vide infra). The resulting ESI mass spectrum is shown in Figure 4. It contains three main peaks at m/z 1183.3859, due to unreacted TrxR-pept, and at m/z 1205.3771 and 1221.3413, corresponding to the adducts of TrxR-pept with sodium and potassium, respectively. In addition, three more peaks of lower intensity are ascribable to the metalated dodecapeptide. Notably, the signal at m/z 1237.3218 is in good accordance with the presence of a [TrxR-pept + Fe − H]+ species. Moreover, the signals at m/z 1253.3236 and 1259.2970 were assigned to the iron-bound TrxR-pept with the additional presence of one oxygen atom and one sodium ion, respectively. Intriguingly, the degradation of diiron bis-cyclopentadienyl carbonyl complexes inside the tumor cell, releasing the iron atoms, has been hypothesized to represent a major mechanism for the antiproliferative activity [36,37,39,45].
At this stage of the study, we were not able to ascertain the oxidation state of the iron center nor its coordinative surrounding; however, it is reasonable to assume that the binding of iron is made possible by the coordination of sulfur and selenium atoms belonging to the {-Cys-Sec-} unit. This hypothesis is in alignment with the significant TrxR inhibitory activity manifested by 1e [37], which suggests a direct involvement of the TrxR catalytic site.
Following the coherent results obtained for 1e, we extended the same method to test the TrxR inhibition ability of selected vinyliminium complexes. Thus, 2c, 3b, 4a, 4c, 5b, 6b, 6c and 7 were incubated at 37 °C for 24 h with TrxR-pept in 1:1 complex-to-peptide ratio.
The recorded ESI mass spectra share a common pattern resembling that obtained for 1e (see Figure S35 for a representative spectrum). More in detail, in addition to the signals of unreacted TrxR-pept and its adducts with sodium and potassium, one relevant signal at m/z 1237.3172 was recognized, attributed to [TrxR-pept + Fe − H]+. This outcome suggests that diiron vinyliminium complexes may actually act as inhibitors of the thioredoxin reductase enzyme, analogously to diiron aminocarbyne complexes such as 1e.
Then, we analyzed the interaction of the vinyliminium complexes 2c, 3b, 4a, 4c, 5b, 6b, 6c and 7 with a small model protein, i.e., cytochrome c (Cyt c). The resulting ESI mass spectra indicated the substantial chemical inertness of the cationic part of the complexes towards Cyt c. In fact, the spectra only show the signal due to the unreacted protein (12358.451 Da) and two other peaks at 12508.414 and 12659.366 Da, corresponding to the adducts of Cyt c with one and two CF3SO3 anions, respectively (Figure S36). This absence of reactivity confirms the strong robustness of 2–7, and represents a partial indication that the anticancer activity of these complexes is selectively directed to the TrxR inhibition, the possible side-reactions with other cellular or plasma proteins being unlikely. Accordingly, previous fluorescence experiments on the interaction of analogous diiron complexes with bovine serum albumin (BSA) pointed out the occurrence of a reversible binding event [39,46], potentially adequate to BSA-mediated transport and diffusion of the drug [83,84]. Note that the establishment of unselective, strong covalent interactions between a metal drug and proteins is, in principle, one of the routes for off-target reactions causing undesired side effects.

4. Conclusions

Diiron complexes based on the {Fe2Cp2(CO)2} scaffold and containing a bridging vinyliminium ligand display some notable properties for a potential drug, i.e., the presence of a biocompatible metal element, straightforward synthesis from inexpensive precursors, appreciable water solubility and/or amphiphilicity, and remarkable stability in aqueous media. The wide structural variability is guaranteed by the cooperativity provided by the bimetallic system. Following preliminary studies on other similar compounds, here we have synthesized a series of new diiron vinyliminium complexes, which have been investigated for their in vitro anticancer activity. In particular, a variety of alkyl and aryl groups (R, Y) on the iminium function has been screened for the first time. Overall, the results point out a significant and fine influence of all vinyliminium substituents (R, Y, R′ and R″) on the physico-chemical properties of the complexes and their cytotoxicity, and especially the choice of R and Y is crucial to provide an optimal profile, with both electronic and steric factors presumably playing a role. Complex 4c, bearing a cyclohexyl moiety on the nitrogen, emerges for its high antiproliferative activity (associated with a slightly negative Log Pow value) towards A2780 and A2780cisR cancer cells and impressive selectivity compared to Balb/3T3 healthy cells (selectivity indexes are 52 and 8, respectively). In accordance with previous findings, targeted studies reveal the probable interference of the complexes with the cellular redox balance, triggering the production of reactive oxygen species. Moreover, mass spectrometry experiments suggest that, although the complexes seem sufficiently robust to resist degradation by common proteins, the peculiar structure of the catalytic site of TrxR might accelerate the intracellular disassembly of the diiron scaffold, resulting in the enzyme inhibition via incorporation of iron. Advanced studies are in course to develop diiron vinyliminium complexes as a suitable class of organometallic anticancer candidates.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/pharmaceutics13081158/s1: Figures S1–S32: NMR spectra of complexes; Figure S33: X-ray structure of 2c; Figure S34: Dose–response cell viability curves; Figures S35–S36: Representative high-resolution ESI mass spectra for the interaction of 3b with the TrxR dodecapeptide (TrxR-pept) and Cyt c. CCDC reference numbers 2087828 (2c) and 2087829 (3a) contain the supplementary crystallographic data for the X-ray studies reported in this paper. These data can be obtained free of charge at www.ccdc.cam.ac.uk/conts/retrieving.html (or from the Cambridge Crystallographic Data Centre, 12, Union Road, Cambridge CB2 1EZ, UK; e-mail: [email protected]).

Author Contributions

Conceptualization, F.M. and F.C.; methodology, S.B., L.B., G.R., A.P. and S.Z.; investigation, S.B., L.B., G.R., A.P. and S.Z.; data curation, S.B., L.B., G.R., A.P., S.Z., F.C. and F.M.; writing—original draft preparation, F.M.; writing—review and editing, S.B., L.B., A.P., S.Z., G.P. and F.M.; supervision, F.C. and F.M.; funding acquisition, L.B., A.P., G.P., F.C. and F.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Pisa (PRA_2020_39).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article or Supplementary Materials.

Acknowledgments

A.P. thanks Luigi Messori (Dept. of Chemistry “U. Schiff,” University of Florence) for making available the TripleTOF® 5600+ mass spectrometer.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. Gibson, D. Platinum(IV) anticancer agents; are we en route to the holy grail or to a dead end? J. Inorg. Biochem. 2021, 217, 111353. [Google Scholar] [CrossRef]
  2. Johnstone, T.C.; Suntharalingam, K.; Lippard, S.J. The Next Generation of Platinum Drugs: Targeted Pt(II) Agents, Nanoparticle Delivery, and Pt(IV) Prodrugs. Chem. Rev. 2016, 116, 3436–3486. [Google Scholar] [CrossRef] [Green Version]
  3. Murray, B.S.; Dyson, P.J. Recent progress in the development of organometallics for the treatment of cancer. Curr. Opin. Chem. Biol. 2020, 56, 28–34. [Google Scholar] [CrossRef] [PubMed]
  4. Chellan, P.; Sadler, P.J. Enhancing the Activity of Drugs by Conjugation to Organometallic Fragments. Chem. Eur. J. 2020, 26, 8676–8688. [Google Scholar] [CrossRef] [PubMed]
  5. Štarha, P.; Trávníček, Z. Non-platinum complexes containing releasable biologically active ligands. Coord. Chem. Rev. 2019, 395, 130–145. [Google Scholar] [CrossRef]
  6. Bratsos, I.; Gianferrara, T.; Alessio, E.; Hartinger, C.G.; Jakupec, M.A.; Keppler, B.K. Ruthenium and Other Non-Platinum Anticancer Compounds. In Bioinorganic Medicinal Chemistry; Alessio, E., Ed.; Wiley-VCH: Weinheim, Germany, 2011; pp. 151–174. [Google Scholar]
  7. Boros, E.; Dyson, P.J.; Gasser, G. Classification of Metal-Based Drugs according to Their Mechanisms of Action. Chem 2020, 6, 41–60. [Google Scholar] [CrossRef]
  8. Anthony, E.J.; Bolitho, E.M.; Bridgewater, H.E.; Carter, O.W.L.; Donnelly, J.M.; Imberti, C.; Lant, E.C.; Lermyte, F.; Needham, R.J.; Palau, M.; et al. Metallodrugs are unique: Opportunities and challenges of discovery and development. Chem. Sci. 2020, 11, 12888–12917. [Google Scholar] [CrossRef]
  9. Haas, K.L.; Franz, K.J. Application of Metal Coordination Chemistry To Explore and Manipulate Cell Biology. Chem. Rev. 2009, 109, 4921–4960. [Google Scholar] [CrossRef] [Green Version]
  10. Riddell, I.A.; Lippard, S.J. Cisplatin and Oxaliplatin: Our Current Understanding of Their Actions. Met. Ions Life Sci. 2018, 18, 1–42. [Google Scholar]
  11. Yu, C.; Wang, Z.; Sun, Z.; Zhang, L.; Zhang, W.; Xu, Y.; Zhang, J.-J. Platinum-Based Combination Therapy: Molecular Rationale, Current Clinical Uses, and Future Perspectives. J. Med. Chem. 2020, 63, 13397–13412. [Google Scholar] [CrossRef] [PubMed]
  12. Green, A.J.; Planchart, A. The neurological toxicity of heavy metals: A fish perspective. CBPC 2018, 208, 12–19. [Google Scholar] [CrossRef] [PubMed]
  13. Campbell, J.M.; Bateman, E.; Peters, M.D.J.; Bowen, J.M.; Keefe, D.M.; Stephenson, M.D. Fluoropyrimidine and platinum toxicity pharmacogenetics: An umbrella review of systematic reviews and meta-analyses. Pharmacogenomics 2016, 17, 435–451. [Google Scholar] [CrossRef] [PubMed]
  14. Oun, R.; Moussa, Y.E.; Wheate, N.J. The side effects of platinum-based chemotherapy drugs: A review for chemists. Dalton Trans. 2018, 47, 6645–6653. [Google Scholar] [CrossRef] [PubMed]
  15. Siddik, Z.H. Cisplatin: Mode of cytotoxic action and molecular basis of resistance. Oncogene 2003, 22, 7265–7279. [Google Scholar] [CrossRef] [Green Version]
  16. Goldstein, R.S.; Mayor, G.H. Minireview. The nephrotoxicity of cisplatin. Life Sci. 1983, 32, 685–690. [Google Scholar] [CrossRef]
  17. Sanchez, M.; Sabio, L.; Galvez, N.; Capdevila, M.; Dominguez-Vera, J.M. Iron chemistry at the service of life. IUBMB Life 2017, 69, 382–388. [Google Scholar] [CrossRef] [Green Version]
  18. Crichton, R. Iron Metabolism-From Molecular Mechanisms to Clinical Consequences, 4th ed.Wiley: Hoboken, NJ, USA, 2016. [Google Scholar]
  19. Wani, W.A.; Baig, U.; Shreaz, S.; Shiekh, R.A.; Iqbal, P.F.; Jameel, E.; Ahmad, A.; Mohd-Setapar, S.H.; Mushtaque, M.; Hun, L.T. Recent advances in iron complexes as potential anticancer agents. New J. Chem. 2016, 40, 1063–1090. [Google Scholar] [CrossRef]
  20. Basu, U.; Roy, M.; Chakravarty, A.R. Recent advances in the chemistry of iron-based chemotherapeutic agents. Coord. Chem. Rev. 2020, 417, 213339. [Google Scholar] [CrossRef]
  21. Simenel, A.A.; Morozova, E.A.; Snegur, L.V.; Zykova, S.I.; Kachala, V.V.; Ostrovskaya, L.A.; Bluchterova, N.V.; Fomina, M.M. Simple route to ferrocenylalkyl nucleobases. Antitumor activity in vivo. Appl. Organomet. Chem. 2009, 23, 219–224. [Google Scholar] [CrossRef]
  22. Simenel, A.A.; Samarina, S.V.; Snegur, L.V.; Starikova, Z.A.; Ostrovskaya, L.A.; Bluchterova, N.V.; Fomina, M.M. o-Carboxybenzoylferrocene. Bioactivity and chemical Modifications. Appl. Organomet. Chem. 2008, 22, 276–280. [Google Scholar] [CrossRef]
  23. Wani, W.A.; Jameel, E.; Baig, U.; Mumtazuddin, S.; Hun, L.T. Ferroquine and its derivatives: New generation of antimalarial agents. Eur. J. Med. Chem. 2015, 101, 534–551. [Google Scholar] [CrossRef] [PubMed]
  24. Peter, S.; Aderibigbe, B.A. Ferrocene-based compounds with antimalaria/anticancer activity. Molecules 2019, 24, 3604. [Google Scholar] [CrossRef] [Green Version]
  25. Patra, M.; Gasser, G. The medicinal chemistry of ferrocene and its derivatives. Nat. Chem. Rev. 2017, 1, 0066. [Google Scholar] [CrossRef]
  26. Sansook, S.; Hassell-Hart, S.; Ocasio, C.; Spencer, J. Ferrocenes in medicinal chemistry; a personal perspective. J. Organomet. Chem. 2020, 905, 121017. [Google Scholar] [CrossRef]
  27. Braga, S.S.; Silva, A.M.S. A New Age for Iron: Antitumoral Ferrocenes. Organometallics 2013, 32, 5626–5639. [Google Scholar] [CrossRef]
  28. Jaouen, G.; Vessieres, A.; Top, S. Ferrocifen type anti cancer drugs. Chem. Soc. Rev. 2015, 44, 8802–8817. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, Y.; Dansette, P.M.; Pigeon, P.; Top, S.; McGlinchey, M.J.; Mansuy, D.; Jaouen, G. A new generation of ferrociphenols leads to a great diversity of reactive metabolites, and exhibits remarkable antiproliferative properties. Chem. Sci. 2018, 9, 70–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Gibson, D. The mechanism of action of platinum anticancer agents—What do we really know about it? Dalton Trans. 2009, 10681–10689. [Google Scholar] [CrossRef]
  31. Marchetti, F. Constructing Organometallic Architectures from Aminoalkylidyne Diiron Complexes. Eur. J. Inorg. Chem. 2018, 3987–4003. [Google Scholar] [CrossRef] [Green Version]
  32. Busetto, L.; Marchetti, F.; Mazzoni, R.; Salmi, M.; Zacchini, S.; Zanotti, V. [3+2+1] cycloaddition involving alkynes, CO and bridging vinyliminium ligands in diiron complexes: A dinuclear version of the Dötz reaction? Chem. Commun. 2010, 46, 3327–3329. [Google Scholar] [CrossRef]
  33. Marchetti, F.; Zacchini, S.; Zanotti, V. Photochemical Alkyne Insertions into the Iron − Thiocarbonyl Bond of [Fe2(CS)(CO)3(Cp)2]. Organometallics 2016, 35, 2630–2637. [Google Scholar] [CrossRef]
  34. Marchetti, F.; Zacchini, S.; Zanotti, V. Carbon monoxide–isocyanide coupling promoted by acetylide addition to a diiron complex. Chem. Commun. 2015, 51, 8101–8104. [Google Scholar] [CrossRef] [PubMed]
  35. Boni, A.; Funaioli, T.; Marchetti, F.; Pampaloni, G.; Pinzino, C.; Zacchini, S. Reversible Reductive Dimerization of Diiron μ-Vinyl Complex via C-C Coupling: Characterization and Reactivity of the Intermediate Radical Species. Organometallics 2011, 30, 4115–4122. [Google Scholar] [CrossRef]
  36. Agonigi, G.; Biancalana, L.; Lupo, M.G.; Montopoli, M.; Ferri, N.; Zacchini, S.; Binacchi, F.; Biver, T.; Campanella, B.; Pampaloni, G.; et al. Exploring the Anticancer Potential of Diiron Bis-cyclopentadienyl Complexes with Bridging Hydrocarbyl Ligands: Behavior in Aqueous Media and In Vitro Cytotoxicity. Organometallics 2020, 39, 645–657. [Google Scholar] [CrossRef]
  37. Biancalana, L.; De Franco, M.; Ciancaleoni, G.; Zacchini, S.; Pampaloni, G.; Gandin, V.; Marchetti, F. Easily Available and Amphiphilic Diiron Cyclopentadienyl Complexes Exhibit In Vitro Anticancer Activity in 2D and 3D Human Cancer Cells via Redox Modulation Triggered by CO Release. Chem. Eur. J. 2021, 27, 10169–10185. [Google Scholar] [CrossRef]
  38. Schoch, S.; Batchelor, L.K.; Funaioli, T.; Ciancaleoni, G.; Zacchini, S.; Braccini, S.; Chiellini, F.; Biver, T.; Pampaloni, G.; Dyson, P.J.; et al. Diiron Complexes with a Bridging Functionalized Allylidene Ligand: Synthesis, Structural Aspects, and Cytotoxicity. Organometallics 2020, 39, 361–373. [Google Scholar] [CrossRef]
  39. Rocco, D.; Batchelor, L.K.; Agonigi, G.; Braccini, S.; Chiellini, F.; Schoch, S.; Biver, T.; Funaioli, T.; Zacchini, S.; Biancalana, L.; et al. Anticancer Potential of Diiron Vinyliminium Complexes. Chem. Eur. J. 2019, 25, 14801–14816. [Google Scholar] [CrossRef] [PubMed]
  40. Agonigi, G.; Batchelor, L.K.; Ferretti, E.; Schoch, S.; Bortoluzzi, M.; Braccini, S.; Chiellini, F.; Biancalana, L.; Zacchini, S.; Pampaloni, G.; et al. Mono-, Di- and Tetra-iron Complexes with Selenium or Sulphur Functionalized Vinyliminium Ligands: Synthesis, Structural Characterization and Antiproliferative Activity. Molecules 2020, 25, 1656. [Google Scholar] [CrossRef] [Green Version]
  41. Mazzoni, R.; Salmi, M.; Zanotti, V. C-C Bond Formation in Diiron Complexes. Chem. Eur. J. 2012, 18, 10174–10194. [Google Scholar] [CrossRef] [PubMed]
  42. Ritleng, V.; Chetcuti, M.J. Hydrocarbyl Ligand Transformations on Heterobimetallic Complexes. Chem. Rev. 2007, 107, 797–858. [Google Scholar] [CrossRef]
  43. García, M.E.; García-Vivó, D.; Ramos, A.; Ruiz, M.A. Phosphinidene-bridged binuclear complexes. Coord. Chem. Rev. 2017, 330, 1–36. [Google Scholar] [CrossRef] [Green Version]
  44. Ciancaleoni, G.; Zacchini, S.; Zanotti, V.; Marchetti, F. DFT Mechanistic Insights into the Alkyne Insertion Reaction Affording Diiron μ-Vinyliminium Complexes and New Functionalization Pathways. Organometallics 2018, 37, 3718–3731. [Google Scholar] [CrossRef]
  45. Rocco, D.; Busto, N.; Pérez-Arnaiz, C.; Biancalana, L.; Zacchini, S.; Pampaloni, G.; Garcia, B.; Marchetti, F. Antiproliferative and bactericidal activity of diiron and monoiron cyclopentadienyl carbonyl complexes comprising a vinyl-aminoalkylidene unit. Appl. Organomet. Chem. 2020, 34, e5923. [Google Scholar] [CrossRef]
  46. Schoch, S.; Mouna, H.; Pereira, S.A.P.; Saraiva, M.L.M.F.S.; Braccini, S.; Chiellini, F.; Biver, T.; Zacchini, S.; Pampaloni, G.; Dyson, P.J.; et al. A Simple Strategy to Incorporate Bioactive Fragments in Anticancer Diiron Bis-Cyclopentadienyl Complexes. Organometallics 2021. [Google Scholar] [CrossRef]
  47. Agonigi, G.; Bortoluzzi, M.; Marchetti, F.; Pampaloni, G.; Zacchini, S.; Zanotti, V. Regioselective Nucleophilic Additions to Diiron Carbonyl Complexes Containing a Bridging Aminocarbyne Ligand: A Synthetic, Crystallographic and DFT Study. Eur. J. Inorg. Chem. 2018, 960–971. [Google Scholar] [CrossRef]
  48. Menges, F. “Spectragryph—Optical Spectroscopy Software”, Version 1.2.5, @2016–2017. Available online: http://www.effemm2.de/spectragryph.
  49. Fulmer, G.R.; Miller, A.J.M.; Sherden, N.H.; Gottlieb, H.E.; Nudelman, A.; Stoltz, B.M.; Bercaw, J.E.; Goldberg, K.I. NMR Chemical Shifts of Trace Impurities: Common Laboratory Solvents, Organics, and Gases in Deuterated Solvents Relevant to the Organometallic Chemist. Organometallics 2010, 29, 2176–2179. [Google Scholar] [CrossRef] [Green Version]
  50. Willker, W.; Leibfritz, D.; Kerssebaum, R.; Bermel, W. Gradient selection in inverse heteronuclear correlation spectroscopy. Magn. Reson. Chem. 1993, 31, 287–292. [Google Scholar] [CrossRef]
  51. Albano, V.G.; Busetto, L.; Monari, M.; Zanotti, V. Reactions of acetonitrile di-iron μ-aminocarbyne complexes; synthesis and structure of [Fe2(μ-CNMe2)(μ-H)(CO)2(Cp)2]. J. Organomet. Chem. 2000, 606, 163–168. [Google Scholar] [CrossRef]
  52. Busetto, L.; Marchetti, F.; Zacchini, S.; Zanotti, V. Unprecedented Zwitterionic Iminium-Chalcogenide Bridging Ligands in Diiron Complexes. Organometallics 2006, 25, 4808–4816. [Google Scholar] [CrossRef]
  53. IR Data of Acetonitrile Adducts (ῦ/cm−1, Acetonitrile Solution): 1984vs (CO), 1817s (μ-CO), 1531w-m (μ-CN) (from 1a); 1984vs (CO), 1813s (μ-CO), 1571m (μ-CN) (from 1b); 1981vs (CO), 1812s (μ-CO), 1551w (μ-CN) (from 1c); 1984vs (CO), 1814s (μ-CO), 1563w-m (μ-CN) (from 1d) [51]; 1984vs (CO), 1814s (μ-CO) (from 1e) [52]; 1986vs (CO), 1815s (μ-CO), 1527w (μ-CN) (from 1f)
  54. Sheldrick, G.M. SADABS-2008/1—Bruker AXS Area Detector Scaling and Absorption Correction; Bruker AXS: Madison, WI, USA, 2008. [Google Scholar]
  55. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  56. Spek, A.L. PLATON, A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2005. [Google Scholar]
  57. Rundlöf, T.; Mathiasson, M.; Bekiroglu, S.; Hakkarainen, B.; Bowden, T.; Arvidsson, T. Survey and qualification of internal standards for quantification by 1H NMR spectroscopy. J. Pharm. Biomed. Anal. 2010, 52, 645–651. [Google Scholar] [CrossRef]
  58. Rice, N.M.; Irving, H.M.N.H.; Leonard, M.A. Nomenclature for liquid-liquid distribution (solvent extraction). Pure Appl. Chem. 1993, 65, 2373–2396. [Google Scholar] [CrossRef]
  59. OECD Guidelines for Testing of Chemicals; OECD: Paris, France, 1995; Volume 107.
  60. Dearden, J.C.; Bresnen, G.M. The Measurement of Partition Coefficients. Quant. Struct. Act. Relat. 1988, 7, 133–144. [Google Scholar] [CrossRef]
  61. Currie, D.J.; Lough, C.E.; Silver, R.F.; Holmes, H.L. Partition Coefficients Of Some Conjugated Heteroenoid Compounds and 1,4-Naphthoquinones. Can. J. Chem. 1966, 44, 1035–1043. [Google Scholar] [CrossRef]
  62. Rosenkranz, A.R.; Schmaldienst, S.; Stuhlmeier, K.M.; Chen, W.; Knapp, W.; Zlabinger, G.J. A microplate assay for the detection of oxidative products using 2′,7′-dichlorofluorescin-diacetate. J. Immunol. Methods 1992, 156, 39–45. [Google Scholar] [CrossRef]
  63. Fabbrini, M.G.; Cirri, D.; Pratesi, A.; Ciofi, L.; Marzo, T.; Guerri, A.; Nistri, S.; Dell’Accio, A.; Gamberi, T.; Severi, M.; et al. A Fluorescent Silver(I) Carbene Complex with Anticancer Properties: Synthesis, Characterization, and Biological Studies. ChemMedChem 2019, 14, 182–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zoppi, C.; Massai, L.; Cirri, D.; Gabbiani, C.; Pratesi, A.; Messori, L. Protein metalation by two structurally related gold(I) carbene complexes: An ESI MS study. Inorg. Chim. Acta 2021, 520, 120297. [Google Scholar] [CrossRef]
  65. Albano, V.G.; Busetto, L.; Marchetti, F.; Monari, M.; Zacchini, S.; Zanotti, V. Stereochemistry of the insertion of disubstituted alkynes into the metal aminocarbyne bond in diiron complexes. J. Organomet. Chem. 2004, 689, 528–538. [Google Scholar] [CrossRef]
  66. Albano, V.G.; Busetto, L.; Marchetti, F.; Monari, M.; Zacchini, S.; Zanotti, V. Diiron μ-Vinyliminium Complexes from Acetylene Insertion into a Metal—Aminocarbyne Bond. Organometallics 2003, 22, 1326–1331. [Google Scholar] [CrossRef]
  67. Zhao, J.; Gou, S.; Liu, F. Potent Anticancer Activity and Possible Low Toxicity of Platinum(II) Complexes with Functionalized 1,1-Cyclobutanedicarboxylate as a Leaving Ligand. Chem. Eur. J. 2014, 20, 15216–15225. [Google Scholar] [CrossRef]
  68. Gust, R.; Krauser, R.; Schmid, B.; Schönenberger, H. Breast cancer inhibiting diastereomeric diacetato[1,2-bis(4-fluorophenyl)ethylenediamine]platinum(II) derivatives: Synthesis and studies on the relationship between reactivity and antitumor activity. Inorg. Chim. Acta 1996, 250, 203–218. [Google Scholar] [CrossRef]
  69. Gust, R.; Krauser, R.; Schmid, B.; Schönenberger, H. Synthesis and Antitumor Activity of [1,2-Bis(4-fluorophenyl)ethylenediamine][dicarboxylato]platinum(II) Complexes. Arch. Pharm. Pharm. Med. Chem. 1998, 331, 27–35. [Google Scholar] [CrossRef]
  70. Marloye, M.; Berger, G.; Gelbcke, M.; Dufrasne, F. A survey of the mechanisms of action of anticancer transition metal complexes. Future Med. Chem. 2016, 8, 2263–2286. [Google Scholar] [CrossRef] [PubMed]
  71. We Previously Found that the Precipitate Formed upon Maintaining Various Complexes Based on the {Fe2Cp2(CO)x} Core (x = 2 or 3) in Aqueous Solution Always Corresponds to Some Form of iron(III) oxide [37, 39]
  72. Marchetti, F.; Di Nicola, C.; Pettinari, R.; Pettinari, C.; Aiello, I.; La Deda, M.; Candreva, A.; Morelli, S.; De Bartolo, L.; Crispini, A. Zinc(II) Complexes of Acylpyrazolones Decorated with a Cyclohexyl Group Display Antiproliferative Activity Against Human Breast Cancer Cells. Eur. J. Inorg. Chem. 2020, 1027–1039. [Google Scholar] [CrossRef]
  73. Akhmetova, V.R.; Akhmadiev, N.S.; Abdullin, M.F.; Dzhemileva, L.U.; Dyakonov, V.A. Synthesis of new N,N-Pd(Pt) complexes based on sulfanyl pyrazoles, and investigation of their in vitro anticancer activity. RSC Adv. 2020, 10, 15116–15123. [Google Scholar] [CrossRef] [Green Version]
  74. Zanda, E.; Busto, N.; Biancalana, L.; Zacchini, S.; Biver, T.; Garcia, B.; Marchetti, F. Anticancer and antibacterial potential of robust Ruthenium(II) arene complexes regulated by choice of α-diimine and halide ligands. Chem. Biol. Interact. 2021, 344, 109522. [Google Scholar] [CrossRef] [PubMed]
  75. Scalcon, V.; Salmain, M.; Folda, A.; Top, S.; Pigeon, P.; Shirley, L.H.Z.; Jaouen, G.; Bindoli, A.; Vessières, A.; Rigobello, M.P. Tamoxifen-like metallocifens target the thioredoxin system determining mitochondrial impairment leading to apoptosis in Jurkat cells. Metallomics 2017, 9, 949–959. [Google Scholar] [CrossRef] [Green Version]
  76. Casini, A.; Mastrobuoni, G.; Ang, W.H.; Gabbiani, C.; Pieraccini, G.; Moneti, G.; Dyson, P.J.; Messori, L. ESI–MS Characterisation of Protein Adducts of Anticancer Ruthenium(II)-Arene PTA (RAPTA) Complexes. ChemMedChem 2007, 2, 631–635. [Google Scholar] [CrossRef]
  77. Pratesi, A.; Cirri, D.; Ciofi, L.; Messori, L. Reactions of Auranofin and Its Pseudohalide Derivatives with Serum Albumin Investigated through ESI-Q-TOF MS. Inorg. Chem. 2018, 57, 10507–10510. [Google Scholar] [CrossRef]
  78. Pelosi, C.; Saitta, F.; Zerino, C.; Canil, G.; Biver, T.; Pratesi, A.; Duce, C.; Fessas, D.; Gabbiani, C.; Tiné, M.R. Thermodynamic Evaluation of the Interactions between Anticancer Pt(II) Complexes and Model Proteins. Molecules 2021, 26, 2376. [Google Scholar] [CrossRef]
  79. Zoppi, C.; Messori, L.; Pratesi, A. ESI MS studies highlight the selective interaction of Auranofin with protein free thiols. Dalton Trans. 2020, 49, 5906–5913. [Google Scholar] [CrossRef]
  80. Massai, L.; Zoppi, C.; Cirri, D.; Pratesi, A.; Messori, L. Reactions of Medicinal Gold(III) Compounds With Proteins and Peptides Explored by Electrospray Ionization Mass Spectrometry and Complementary Biophysical Methods. Front. Chem. 2020, 8, 581648. [Google Scholar] [CrossRef] [PubMed]
  81. Massai, L.; Pratesi, A.; Bogojeski, J.; Banchini, M.; Pillozzi, S.; Messori, L.; Bugarčić, Ž.D. Antiproliferative properties and biomolecular interactions of three Pd(II) and Pt(II) complexes. J. Inorg. Biochem. 2016, 165, 1–6. [Google Scholar] [CrossRef] [PubMed]
  82. Pratesi, A.; Gabbiani, C.; Michelucci, E.; Ginanneschi, M.; Papini, A.M.; Rubbiani, R.; Ott, I.; Messori, L. Insights on the mechanism of thioredoxin reductase inhibition by gold N-heterocyclic carbene compounds using the synthetic linear selenocysteine containing C-terminal peptide hTrxR(488-499): An ESI-MS investigation. J. Inorg. Biochem. 2014, 136, 161–169. [Google Scholar] [CrossRef]
  83. Tŏpala, T.; Pascual-Álvarez, A.; Moldes-Tolosa, M.Á.; Bodoki, A.; Castiñei ras, A.; Torres, J.; del Pozo, C.; Borrás, J.; Alzuet-Piña, G. New sulfonamide complexes with essential metal ions [Cu(II), Co(II), Ni(II) and Zn(II)]. Effect of the geometry and the metal ion on DNA binding and nuclease activity. BSA protein interaction. J. Inorg. Biochem. 2020, 202, 110823. [Google Scholar] [CrossRef] [PubMed]
  84. Elsadek, B.; Kratz, F. Impact of albumin on drug delivery-New applications on the horizon. J. Control. Release 2012, 157, 4–28. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of selected iron complexes with reported biological (anticancer) activity. a) Ferrocene derivatives: (I) ferroquine, antimalarial drug investigated in clinical trials; (II) ferrocifens (R=H, OH), derived from conjugation of ferrocene with the anticancer drug tamoxifen; (III) Diiron μ-vinyliminium complexes with structural variability from combination of isocyanide (CNR, red) and alkyne (R′C≡CR″, blue) reagents (R=Me, 2,6-C6H3Me2); (IV) Diiron μ-aminocarbyne complexes able to transfer Y+ (green) to bio-substrates.
Figure 1. Structures of selected iron complexes with reported biological (anticancer) activity. a) Ferrocene derivatives: (I) ferroquine, antimalarial drug investigated in clinical trials; (II) ferrocifens (R=H, OH), derived from conjugation of ferrocene with the anticancer drug tamoxifen; (III) Diiron μ-vinyliminium complexes with structural variability from combination of isocyanide (CNR, red) and alkyne (R′C≡CR″, blue) reagents (R=Me, 2,6-C6H3Me2); (IV) Diiron μ-aminocarbyne complexes able to transfer Y+ (green) to bio-substrates.
Pharmaceutics 13 01158 g001
Scheme 17. General synthetic pathway to diiron vinyliminium complexes (see Table 2 for details).
Scheme 17. General synthetic pathway to diiron vinyliminium complexes (see Table 2 for details).
Pharmaceutics 13 01158 sch017
Figure 2. View of the cation of 3a. H atoms have been omitted for clarity. Displacement ellipsoids are at the 50% probability level. Selected bond lengths (Å) and angles (°): Fe(1)-Fe(2) 2.5426(7), Fe(1)-C(11) 1.763(3), Fe(1)-C(12) 1.911(4), Fe(2)-C(12) 1.934(4), Fe(2)-C(1) 1.851(3), Fe(2)-C(2) 2.055(3), Fe(2)-C(3) 2.036(4), Fe(1)-C(3) 1.940(4), C(1)-N(1) 1.285(4), C(1)-C(2) 1.419(5), C(2)-C(3) 1.409(5), N(1)-C(5) 1.466(5), N(1)-C(6) 1.481(4), C(6)-C(7) 1.487(5), C(7)-C(8) 1.319(5), C(11)-O(11) 1.144(4), C(12)-O(12) 1.172(4), Fe(1)-C(11)-O(11) 178.2(4), Fe(1)-C(12)-Fe(2) 82.79(15), Fe(1)-C(3)-Fe(2) 79.47(14), Fe(1)-C(3)-C(2) 120.7(3), C(3)-C(2)-C(1) 116.2(3), C(2)-C(1)-Fe(2) 76.6(2), C(1)-N(1)-C(5) 121.7(3), C(1)-N(1)-C(6) 121.3(3), C(5)-N(1)-C(6) 117.0(3), N(1)-C(6)-C(7) 112.7(3), C(6)-C(7)-C(8) 122.5(4).
Figure 2. View of the cation of 3a. H atoms have been omitted for clarity. Displacement ellipsoids are at the 50% probability level. Selected bond lengths (Å) and angles (°): Fe(1)-Fe(2) 2.5426(7), Fe(1)-C(11) 1.763(3), Fe(1)-C(12) 1.911(4), Fe(2)-C(12) 1.934(4), Fe(2)-C(1) 1.851(3), Fe(2)-C(2) 2.055(3), Fe(2)-C(3) 2.036(4), Fe(1)-C(3) 1.940(4), C(1)-N(1) 1.285(4), C(1)-C(2) 1.419(5), C(2)-C(3) 1.409(5), N(1)-C(5) 1.466(5), N(1)-C(6) 1.481(4), C(6)-C(7) 1.487(5), C(7)-C(8) 1.319(5), C(11)-O(11) 1.144(4), C(12)-O(12) 1.172(4), Fe(1)-C(11)-O(11) 178.2(4), Fe(1)-C(12)-Fe(2) 82.79(15), Fe(1)-C(3)-Fe(2) 79.47(14), Fe(1)-C(3)-C(2) 120.7(3), C(3)-C(2)-C(1) 116.2(3), C(2)-C(1)-Fe(2) 76.6(2), C(1)-N(1)-C(5) 121.7(3), C(1)-N(1)-C(6) 121.3(3), C(5)-N(1)-C(6) 117.0(3), N(1)-C(6)-C(7) 112.7(3), C(6)-C(7)-C(8) 122.5(4).
Pharmaceutics 13 01158 g002
Figure 3. Fluorescence kinetic measurements of intracellular reactive oxygen species (ROS). A2780 cells incubated (A) up to 24 h and (B) for 24 h with 20 μM of diiron complexes at 37 °C and 5% CO2. Results are expressed as the mean of three analyses with standard deviation. Values marked with * are significantly different from the control (* p < 0.05).
Figure 3. Fluorescence kinetic measurements of intracellular reactive oxygen species (ROS). A2780 cells incubated (A) up to 24 h and (B) for 24 h with 20 μM of diiron complexes at 37 °C and 5% CO2. Results are expressed as the mean of three analyses with standard deviation. Values marked with * are significantly different from the control (* p < 0.05).
Pharmaceutics 13 01158 g003
Figure 4. High-resolution ESI mass spectrum of 1e incubated for 24h at 37 °C with 10−5 M TrxR-pept solution in water; 1:1 peptide to complex ratio. 0.1% v/v of formic acid was added just before infusion. Unlabeled peaks: a, m/z 1205.3771; b, m/z 1237.3218; c, m/z 1253.3236.
Figure 4. High-resolution ESI mass spectrum of 1e incubated for 24h at 37 °C with 10−5 M TrxR-pept solution in water; 1:1 peptide to complex ratio. 0.1% v/v of formic acid was added just before infusion. Unlabeled peaks: a, m/z 1205.3771; b, m/z 1237.3218; c, m/z 1253.3236.
Pharmaceutics 13 01158 g004
Table 1. Crystal data and measurement details for 2c and 3a.
Table 1. Crystal data and measurement details for 2c and 3a.
2c3a
FormulaC32H30F3Fe2NO5SC21H22F3Fe2NO5S
FW709.33569.15
T, K100(2)100(2)
λ, Å0.710730.71073
Crystal systemMonoclinicMonoclinic
Space groupP21/cP21/c
a, Å17.2348(14)9.9306(6)
b, Å8.9679(7)12.5791(7)
c, Å19.8255(14)17.5345(10)
β104.380(3)90.604(2)
Cell volume, Å32968.2(4)2190.3(2)
Z44
Dc, g∙cm−31.5871.726
μ, mm−11.1101.480
F(000)14561160
Crystal size, mm0.18 × 0.15 × 0.110.21 × 0.18 × 0.13
θ limits,°2.121–25.0811.993–25.994
Reflections collected28,21739,294
Independent reflections5184 [Rint = 0.0817]4267 [Rint = 0.1414]
Data/restraints/parameters5184/1041/5134267/0/300
Goodness on fit on F21.2831.142
R1 (I > 2σ(I))0.12070.0480
wR2 (all data)0.25120.0964
Largest diff. peak and hole, e Å−30.679/−1.1560.607/−0.504
Table 2. Diiron vinyliminium complexes (2–7) obtained by aminocarbyne precursors (1a–f) according to Scheme 17.
Table 2. Diiron vinyliminium complexes (2–7) obtained by aminocarbyne precursors (1a–f) according to Scheme 17.
PrecursorRYR′R″Product
1aCH2PhCH2PhMeH2a
1aCH2PhCH2PhPhH2b
1aCH2PhCH2PhMeMe2c
1bCH2CH = CH2MeMeH3a
1bCH2CH = CH2MePhH3b
1bCH2CH = CH2MeMeMe3c
1cC6H11 = CyMeMeH4a
1cCyMePhH4b
1cCyMeMeMe4c
1dCH2PhMeMeH5a
1dCH2PhMeMeMe5b
1dCH2PhMeEtEt5c
1e4-C6H4OMeMeMeH6a
1e4-C6H4OMeMePhH6b
1e4-C6H4OMeMeMeMe6c
1f2-naphthylMeMeMe7
Table 3. Solubility of diiron complexes in D2O at 21 °C (1H NMR, Me2SO2 internal standard), residual % in DMSO-d6/D2O (2:1 v/v) after 72 h at 37 °C (1H NMR, Me2SO2 internal standard) and octanol–water partition coefficients at 21 °C (Log Pow, UV–Vis). a Stability evaluated in neat D2O.
Table 3. Solubility of diiron complexes in D2O at 21 °C (1H NMR, Me2SO2 internal standard), residual % in DMSO-d6/D2O (2:1 v/v) after 72 h at 37 °C (1H NMR, Me2SO2 internal standard) and octanol–water partition coefficients at 21 °C (Log Pow, UV–Vis). a Stability evaluated in neat D2O.
CompoundSolubility/mol·L−1Solubility/g·L−1Stability %Log Pow
2a4.33 × 10−40.30810.45 ± 0.04
2b<1 × 10−4<0.08=1.4 ± 0.2
2c<1 × 10−4<0.07870.56 ± 0.08
3a6.43 × 10−33.6693 a−0.75 ± 0.01
3b2.34 × 10−31.4868−0.09 ± 0.01
3c6.89 × 10−34.0285 a−0.77 ± 0.01
4a3.49 × 10−32.1395−0.29 ± 0.03
4b1.52 × 10−31.02650.41 ± 0.02
4c2.63 × 10−31.6495−0.17 ± 0.01
5a5.41 × 10−40.3485−0.26 ± 0.01
5b1.52 × 10−30.9690−0.43 ± 0.01
5c1.10 × 10−30.73840.13 ± 0.01
6a2.22 × 10−31.4194−0.46 ± 0.01
6b4.98 × 10−40.35760.37 ± 0.04
6c1.06 × 10−30.6986−0.21 ± 0.01
7<1 × 10−4<0.07660.38 ± 0.02
Table 4. IC50 values (μM) determined for diiron complexes and cisplatin on human ovarian carcinoma (A2780), human ovarian carcinoma cisplatin-resistant (A2780cisR) and murine embryonic fibroblast (Balb/3T3 clone A31) cell lines after 72 h exposure. Values are given as mean ± standard deviation. Selectivity index (S.I.) calculated as the ratio between IC50 for Balb/3T3 and IC50 for A2780.
Table 4. IC50 values (μM) determined for diiron complexes and cisplatin on human ovarian carcinoma (A2780), human ovarian carcinoma cisplatin-resistant (A2780cisR) and murine embryonic fibroblast (Balb/3T3 clone A31) cell lines after 72 h exposure. Values are given as mean ± standard deviation. Selectivity index (S.I.) calculated as the ratio between IC50 for Balb/3T3 and IC50 for A2780.
CompoundA2780A2780cisRBalb/3T3S.I.
2a2.0 ± 0.7=14 ± 47
2c1.91 ± 0.15=18 ± 39
3a17.7 ± 0.8=>100>5.6
3b6.0 ± 0.829 ± 3>100>16
3c17 ± 3=>100>5.9
4a3.3 ± 0.411.6 ± 1.3>100>30
4b2.6 ± 0.3=35 ± 313
4c1.8 ± 0.211.4 ± 0.994 ± 1652
5a7.1 ± 0.4=>100>14
5b5.1 ± 0.312 ± 4>100>19
5c3.75 ± 0.14=90 ± 824
6a3.5 ± 0.5=94 ± 427
6b3.8 ± 0.4=39 ± 310
6c3.8 ± 0.417 ± 3>100>26
72.2 ± 0.2=48 ± 322
cisplatin0.40 ± 0.0726 ± 30.8 ± 0.22
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Braccini, S.; Rizzi, G.; Biancalana, L.; Pratesi, A.; Zacchini, S.; Pampaloni, G.; Chiellini, F.; Marchetti, F. Anticancer Diiron Vinyliminium Complexes: A Structure–Activity Relationship Study. Pharmaceutics 2021, 13, 1158. https://doi.org/10.3390/pharmaceutics13081158

AMA Style

Braccini S, Rizzi G, Biancalana L, Pratesi A, Zacchini S, Pampaloni G, Chiellini F, Marchetti F. Anticancer Diiron Vinyliminium Complexes: A Structure–Activity Relationship Study. Pharmaceutics. 2021; 13(8):1158. https://doi.org/10.3390/pharmaceutics13081158

Chicago/Turabian Style

Braccini, Simona, Giorgia Rizzi, Lorenzo Biancalana, Alessandro Pratesi, Stefano Zacchini, Guido Pampaloni, Federica Chiellini, and Fabio Marchetti. 2021. "Anticancer Diiron Vinyliminium Complexes: A Structure–Activity Relationship Study" Pharmaceutics 13, no. 8: 1158. https://doi.org/10.3390/pharmaceutics13081158

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

Article Metrics

Back to TopTop