Acessibilidade / Reportar erro

Reduction Potential of RuIII-Based Complexes with Potential Antitumor Activity and Thermodynamics of their Hydrolysis Reactions and Interactions with Possible Biological Targets: a Theoretical Investigation

Abstract

In this article density functional theory (DFT)-based calculations were employed to investigate the electrochemistry of the antitumor ruthenium complexes trans-tetrachloro(dimethylsulfoxide)imidazole ruthenate(III) (NAMI-A) and trans-[tetrachlorobis(1.-indazole)ruthenate(III)] (KP1019), their hydrolysis products as well as their interactions with biological S-donors and N-donors targets as cysteine, glutathione and guanine nucleobase. The compounds exhibit different electrochemical behavior upon hydrolysis. While the reduction potential of NAMI-A increases up to 0.8 V upon hydrolysis, the reduction potential of KP1019 remains almost constant after the first hydrolysis. NAMI-A and KP1019 complexes have thermodynamic preference to be reduced prior to undergoing hydrolysis and, strong preference to undergo successive hydrolysis instead of interacting with the S-donor and N-donor ligands. Interaction with S-donor ligands in the unprotonated form is highly unfavorable, with the free energy in solution (ΔGsol) ≥ 18 kcal mol-1. For both complexes, the interaction with the guanine and glutathione are of the same magnitude (ΔGsol ca. –0.6 kcal mol-1) meaning that these ligands can compete for binding to the metallodrug.

Keywords:
density functional theory calculations; ruthenium-based metallodrugs; reduction potential; competitive biological reactions; solvent effects; electronic structure calculations


Introduction

Since the discovery of cisplatin and its use in the clinic therapy as one of the most effective anticancer drugs,11 Rosenberg, B.; VanCamp, L.; Trosko, J. E.; Mansour, H. V.; Nature 1969, 222, 385.,22 Rosenberg, B.; Van Camp, L.; Krigas, T.; Nature 1965, 205, 698. the research and development of metal-based drugs have received great attention from the scientific community.33 Clarke, M. J.; Coord. Chem. Rev. 2003, 236, 209.

4 Mjos, K. D.; Orvig, C.; Chem. Rev. 2014, 114, 4540.
-55 Murray, B. S.; Babak, M. V.; Hartinger, C. G.; Dyson, P. J.; Coord. Chem. Rev. 2016, 306, 86. Cancer is certainly among the leading causes of death in the world and, therefore, it has been the main target for the development of new drugs, especially those with antimetastatic activity. In spite of the great efficacy of platinum-based drugs against ovarian, bladder and testicular cancers, they display limited activity against some of the most common tumors, such as colon and breast cancers.66 Johnstone, T. C.; Suntharalingam, K.; Lippard, S. J.; Chem. Rev. 2016, 116, 3436.

7 Pasini, A.; Zunino, F.; Angew. Chem., Int. Ed. Engl. 1987, 26, 615.
-88 Boulikas, T.; Vougiouka, M.; Oncol. Rep. 2003, 10, 1663. In addition, the high toxicity and incidence of drug resistance (acquired or intrinsic) remain the main challenges in their clinical application.66 Johnstone, T. C.; Suntharalingam, K.; Lippard, S. J.; Chem. Rev. 2016, 116, 3436. These facts stimulated the research of new metal-based drugs against cancer, aiming at improving clinical efficacy, reducing overall toxicity, and broadening the spectrum of activity.44 Mjos, K. D.; Orvig, C.; Chem. Rev. 2014, 114, 4540.,55 Murray, B. S.; Babak, M. V.; Hartinger, C. G.; Dyson, P. J.; Coord. Chem. Rev. 2016, 306, 86.,99 Smith, N. A.; Sadler, P. J.; Philos. Trans. R. Soc., A 2013, 371, 20120519. In this regard, ruthenium complexes have been shown to be a very attractive platform, thanks to their low toxicity and low drug resistance. In addition, ruthenium has a range of oxidation states (RuII, RuIII and RuIV) accessible under physiological conditions, which is unique among the platinum-group metals.33 Clarke, M. J.; Coord. Chem. Rev. 2003, 236, 209.,44 Mjos, K. D.; Orvig, C.; Chem. Rev. 2014, 114, 4540. This is a very significant feature since the activities of most metal-based anticancer drugs are dependent on their oxidation states.

There are currently two antitumor compounds of ruthenium(III) (Figure 1) in the clinical phase, thanks to their broad antineoplastic activity: trans-[tetrachloro(dimethylsulfoxide)imidazole ruthenate(III)] ([Im]trans-[RuCl4(Im)(DMSO)]), also known as NAMI-A,1010 Sava, G.; Pacor, S.; Mestroni, G.; Alessio, E.; Clin. Exp. Metastasis 1992, 10, 273. and trans-[tetrachlorobis(1H-indazole)ruthenate(III)] ([IndH]trans-[RuCl4(Ind)2]), known as KP1019.1111 Hartinger, C. G.; Jakupec, M. A.; Zorbas-Seifried, S.; Groessl, M.; Egger, A.; Berger, W.; Zorbas, H.; Dyson, P. J.; Keppler, B. K.; Chem. Biodiversity 2008, 5, 2140. NAMI-A is able to effectively inhibit the development and growth of lung metastasis in experimental models of solid tumors in vivo,1212 Sava, G.; Bergamo, A.; Zorzet, S.; Gava, B.; Casarsa, C.; Cocchietto, M.; Furlani, A.; Scarcia, V.; Serli, B.; Iengo, E.; Alessio, E.; Mestroni, G.; Eur. J. Cancer 2002, 38, 427.,1313 Leijen, S.; Burgers, S. A.; Baas, P.; Pluim, D.; Tibben, M.; van Werkhoven, E.; Alessio, E.; Sava, G.; Beijnen, J. H.; Schellens, J. H.; Invest. New Drugs 2015, 33, 201. whereas KP1019 shows cytotoxic activity by inducing apoptosis in a large line of cancer cells, especially colorectal tumors.1111 Hartinger, C. G.; Jakupec, M. A.; Zorbas-Seifried, S.; Groessl, M.; Egger, A.; Berger, W.; Zorbas, H.; Dyson, P. J.; Keppler, B. K.; Chem. Biodiversity 2008, 5, 2140.,1414 Sava, G.; Zorzet, S.; Turrin, C.; Francesca, V.; Soranzo, M. R.; Zabucchi, G.; Cocchietto, M.; Bergamo, A.; DiGiovine, S.; Pezzoni, G.; Sartor, L.; Garbisa, S.; Clin. Cancer Res. 2003, 9, 1898.,1515 Gava, B.; Zorzet, S.; Spessotto, P.; Cocchietto, M.; Sava, G.; J. Pharmacol. Exp. Ther. 2006, 317, 284.

Figure 1
Structures of (a) NAMI-A and (b) KP1019.

Both complexes were considered as promising antitumor agents. However, NAMI-A is no longer in clinical trials due to side effects and antitumor activity less than expected,1616 Alessio, E.; Eur. J. Inorg. Chem. 2017, 2017, 1549. whereas the indazolium counter ion in KP1019 has been replaced by the sodium (KP-1339/IT139) ion in order to increase the solubility and was selected for further development.1717 Trondl, R.; Heffeter, P.; Kowol, C. R.; Jakupec, M. A.; Berger, W.; Kepler, B. K.; Chem. Sci. 2014, 5, 2925. These results evidenced the need to know the individual chemical characteristics of the different ruthenium-based compounds for therapeutic purposes. Although both complexes underwent clinical trials phase, their mechanism of action is not yet fully understood. It is well known that RuIII compounds are more inert toward ligand-exchange reactions than RuII.33 Clarke, M. J.; Coord. Chem. Rev. 2003, 236, 209. Therefore, metallodrugs based on RuIII need to be activated by reduction, which will accelerate the hydrolysis reactions and interactions with biological targets as was first proposed by Clarke.33 Clarke, M. J.; Coord. Chem. Rev. 2003, 236, 209. Consequently, RuIII-based drugs must have accessible reduction potentials in the biological medium. Electrochemical and chemical reduction studies of RuIII metallodrugs are, therefore, of great importance in order to understand their biological activities and thus have been the subject of several studies.1818 Costa, G.; Balducci, G.; Alessio, E.; Tavagnacco, C.; Mestroni, G.; J. Electroanal. Chem. 1990, 296, 57.

19 Alessio, E.; Balducci, G.; Lutman, A.; Mestroni, G.; Calligaris, M.; Attia, W. M.; Inorg. Chim. Acta 1993, 203, 205.

20 Hartmann, M.; Lipponer, K.-G.; Kepler, B. K.; Inorg. Chim. Acta 1998, 267, 137.

21 Frasca, D. R.; Clarke, M. J.; J. Am. Chem. Soc. 1999, 121, 8523.

22 Ravera, M.; Baracco, S.; Cassino, C.; Zanello, P.; Osella, D.; Dalton Trans. 2004, 2347.

23 Reisner, E.; Arion, V. B.; da Silva, M. F. C. G.; Lichtenecker, R.; Eichinger, A.; Kepler, B. K.; Kukushkin, V. Y.; Pombeiro, A. J. L.; Inorg. Chem. 2004, 43, 7083.

24 Schluga, P.; Hartinger, C. G.; Egger, A.; Reisner, E.; Galanski, M.; Jakupec, M. A.; Kepler, B. K.; Dalton Trans. 2006, 1796.

25 Ravera, M.; Cassino, C.; Baracco, S.; Osella, D.; Eur. J. Inorg. Chem. 2006, 740.

26 Cebrián-Losantos, B.; Reisner, E.; Kowol, C. R.; Roller, A.; Shova, S.; Arion, V. B.; Kepler, B. K.; Inorg. Chem. 2008, 47, 6513.
-2727 Reisner, E.; Arion, V. B.; Kepler, B. K.; Pombeiro, A. J. L.; Inorg. Chim. Acta 2008, 361, 1569. Ravera et al.2525 Ravera, M.; Cassino, C.; Baracco, S.; Osella, D.; Eur. J. Inorg. Chem. 2006, 740. investigated the redox chemistry of [LH][trans-RuIIICl4L2] (L = imidazole or indazole) complexes in water at different pH values. For the indazole and imidazole complexes the E0' (RuIII/RuII) reduction potential of –0.160 and –0.460 V vs. Ag/AgCl was obtained at pH 4.0. The authors showed that despite the structural similarities, the electrochemically generated RuII species undergoes different ligand displacement reactions. Using nuclear magnetic resonance (NMR) experiments, the authors have also shown that one of the axial imidazole ligands is displaced by water in the imidazole complex [RuIICl4(Im)2]2– and a Cl/water substitution takes place in the indazole complex [RuIICl4(Ind)2]2–. The chemical reduction in the presence of glutathione (GSH) was also observed by the authors. Reisner et al.2323 Reisner, E.; Arion, V. B.; da Silva, M. F. C. G.; Lichtenecker, R.; Eichinger, A.; Kepler, B. K.; Kukushkin, V. Y.; Pombeiro, A. J. L.; Inorg. Chem. 2004, 43, 7083. investigated the electrochemical behavior of [trans-RuIIICl4L(DMSO)] and [trans-RuIIICl4L2] complexes (L = imidazole, 1,2,4-triazole, indazole) in dimethylformamide (DMF) and DMSO. For the NAMI-A and KP1019 complexes, E0' (RuIII/RuII) reduction potentials of –0.220 and –0.720 V vs. normal hydrogen electrode (NHE) was found, respectively, in DMF. The authors showed that the general equation proposed by Lever,2828 Lever, A. B. P.; Inorg. Chem. 1990, 29, 1271. which assumes an additive contribution of all the ligands to the redox potential of the complex, predicts very well the measured reduction potential of the RuIII complexes, and they were able to propose an equation to predict the reduction potential of single negatively charged complexes of RuIII in aqueous medium at pH 7. For the Na[RuIIICl4(DMSO)(Im)] complex, Alessio et al.1919 Alessio, E.; Balducci, G.; Lutman, A.; Mestroni, G.; Calligaris, M.; Attia, W. M.; Inorg. Chim. Acta 1993, 203, 205. obtained E0' (RuIII/RuII) reduction potentials of –0.001 V vs. saturated sodium chloride calomel electrode (SSCE). The aquation reaction revealed to be pH dependent. In the pH range of 7.1 to 9.0 the final aquated product was [RuIICl3(DMSO)(H2O)(Im)] and, at pH 6 the final detected product was [RuIICl4(DMSO)(H2O)], with exchange of the imidazole ligand by water. It has been shown that at physiological pH and 37 ºC (phosphate buffer, 0.9% NaCl), stepwise dissociation of two chloride ions from NAMI-A occurs within minutes.2929 Mestroni, G.; Alessio, E.; Sava, G.; Pacor, S.; Coluccia, M.; Boccarelli, A.; Met.-Based Drugs 1994, 1, 41.

These RuIII anticancer complexes have also been shown to be reduced by common biological reductants such as GSH and ascorbic acid. Hartmann et al.2020 Hartmann, M.; Lipponer, K.-G.; Kepler, B. K.; Inorg. Chim. Acta 1998, 267, 137. showed that, at pH 7 and 0.1 M phosphate buffer, (H2Im)[RuCl4(HIm)2] is promptly reduced by GSH. Schluga et al.2424 Schluga, P.; Hartinger, C. G.; Egger, A.; Reisner, E.; Galanski, M.; Jakupec, M. A.; Kepler, B. K.; Dalton Trans. 2006, 1796. showed that the binding of KP1019 toward the deoxyribonucleic acid (DNA)-modeling nucleotide guanosine monophosphate (GMP) is increased upon addition of GSH but not of ascorbic acid. That is, GSH reduces the RuIII complex and facilitates the binding to the nucleotide. However, the mechanism involved in the chemical reduction of these RuIII anticancer complexes is not yet fully understood. It is believed that an inner-sphere mechanism is involved.2121 Frasca, D. R.; Clarke, M. J.; J. Am. Chem. Soc. 1999, 121, 8523.,3030 Haukka, M.; Ahlgrén, M.; Pakkanen, T. A.; J. Chem. Soc., Dalton Trans. 1996, 1927.

All these results suggest that in the mechanism associated with these RuIII anticancer complexes, activation by reduction and exchange of chloride ligands by water molecules of the solvent (hydrolysis or aquation) are two important steps for the antitumor function of such metallodrugs. After being activated, distributed and located in cell compartments, the way the ruthenium metallodrugs kills the cancer cells is not well fully understood. In contrast with cisplatin which is well-known to interact with DNA, the primary interaction target(s) for the ruthenium metallodrugs is still an open question.

Based on the observations listed above, three events are connected with the biological activity of NAMI-A and KP1019: (i) reduction of the RuIII center; (ii) chloride exchange reaction with water molecules of the solvent medium; and (iii) interaction with the biological targets. These reactions can compete with each other, as schematically shown in Figure 2, for the NAMI-A compound. The compound can initially be reduced following pathway b, generating RuII complex 1b, and this reduced complex can then undergo ligand exchange reaction with water of the solvent, generating the monoaquo complex 2. Alternatively, the complex can initially undergo ligand displacement reaction, following pathway a, generating the monoaquo-RuIII species 1a, which is posteriorly reduced to the monoaquo-RuII species 2. Once the reduced monoaquo-RuII species is formed other competitive processes can take place, as for instance the successive displacement of chloride by water through pathway c or direct interaction with the biological target L, pathway d, generating the adduct 3 in which the target L displaces a water molecule. The sequence of the events or, the preferred pathways, will depend on the kinetics of these reactions and also on the reduction potential of the species, chloride concentration and thermodynamic stability of the complexes formed.

Figure 2
Schematic representation of the possible competitive biological reaction paths for the NAMI-A complex (1). In pathway (a) NAMI-A initially undergoes ligand displacement reaction generating the monoaquo-RuIII species 1a, which is posteriorly reduced to the monoaquo-RuII species 2. In pathway (b) the compound is initially reduced generating RuII complex 1b and undergoes ligand exchange reaction with water, generating the monoaquo complex 2. After complex 2 is generated it can be consecutively hydrolyzed through pathway (c) or interact with the biological targets L in pathway (d), generating the adduct 3.

There are some open questions related to the interaction of NAMI-A and KP1019 complexes with their final biological target, the last step shown in Figure 2. Some studies have suggested that NAMI-A acts on cell cycle blockade at the G2/M transition point, due to the accumulation of inactive Cdk1 proteins, which is responsible for reporting DNA molecule damage.3131 Bergamo, A.; Gaiddon, C.; Schellens, J. H. M.; Beijnen, J. H.; Sava, G.; J. Inorg. Biochem. 2012, 106, 90.,3232 Han Ang, W.; Dyson, P. J.; Eur. J. Inorg. Chem. 2006, 20, 4003. In addition, NAMI-A also causes inhibition of topoisomerase proteins, which are important enzymes in DNA replication and packaging processes.3333 Qian, C.; Wang, J. Q.; Song, C. L.; Wang, L. L.; Ji, L. N.; Chao, H.; Metallomics 2013, 5, 844. On the other hand, it is believed that the cytotoxic activity of KP1019 is based on the endoplasmic reticulum (ER), with stress-related effects rather than DNA damage.1717 Trondl, R.; Heffeter, P.; Kowol, C. R.; Jakupec, M. A.; Berger, W.; Kepler, B. K.; Chem. Sci. 2014, 5, 2925. KP1019 has been shown to cause oxidative stress and perturbation of ER functions in cancer cells. In addition, the 78 kDa glucose regulatory protein (GRP78) is reduced to a normal level.3434 Bierle, L. A.; Reich, K. L.; Taylor, B. E.; Blatt, E. B.; Middleton, S. M.; Burke, S. D.; Stultz, L. K.; Hanson, P. K.; Partridge, J. F.; Miller, M. E.; PLoS One 2015, 10, e0138085.,3535 Flocke, L. S.; Trondl, R.; Jakupec, M. A.; Keppler, B. K.; Invest. New Drugs 2016, 34, 261. Thus, apoptosis is induced by intrinsic mitochondrial pathway.3636 Smith, C. A.; Sutherland-Smith, A. J.; Kepler, B. K.; Kratz, F.; Baker, E. N.; J. Biol. Inorg. Chem. 1996, 1, 424.,3737 Clarke, M. J.; Zhu, F.; Frasca, D. R.; Chem. Rev. 1999, 99, 2511.

Therefore, despite the progress in the biochemical understanding of the biological activity of NAMI-A and KP1019, fundamental studies on the kinetic and thermodynamics of the competitive processes exemplified in Figure 2 can provide important information on the bioavailability and reactivity of these compounds. In this regard, in the present work, we employed the density functional theory (DFT) to investigate the electrochemistry of NAMI-A, KP1019 and their hydrolysis products, as well as their interactions with S-donor targets cysteine (Cys) and GSH, and N-donor target guanine nucleobase (at the coordination sites N3 (GN3) and N7 (GN7)). We believe that the study of the different structural characteristics, as well as the thermodynamic properties of RuII adducts formation with some strategic biological targets and DNA nucleobases in solution, can provide important information to understand the interaction of these compounds in the cellular environment.

Methodology

Full unconstrained geometry optimizations and frequency calculations of all species involved in this work were carried out at the DFT level of theory3838 Parr, R. G.; Yang, W.; Density Functional Theory of Atoms and Molecules; Oxford University Press: New York, 1989. with the Tao, Perdew, Staroverov, and Scuseria hybrid meta-generalized gradient approximation (TPSSh) exchange-correlation functional3939 Staroverov, V. N.; Scuseria, G. E.; Tao, J.; Perdew, J. P.; J. Chem. Phys. 2003, 119, 12129. as implemented in the ORCA program.4040 Neese, F.; Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 73. The Ahlrichs full electron def2-triple-zeta valence polarization (TZVP) basis set4141 Weigend, F.; Ahlrichs, R.; Phys. Chem. Chem. Phys. 2005, 7, 3297. was used for all atoms. Scalar relativistic effects were treated using the zeroth order regular approximation (ZORA) formalism.4242 van Lenthe, E.; Baerends, E. J.; Snijders, J. G.; J. Chem. Phys. 1993, 99, 4597.

43 van Lenthe, E.; Baerends, E. J.; Snijders, J. G.; J. Chem. Phys. 1994, 101, 9783.
-4444 van Lenthe, E.; van Leeuwen, R.; Baerends, E. J.; Snijders, J. G.; Int. J. Quantum Chem. 1996, 57, 281. To speed up the calculations the resolution of the identity4545 Neese, F.; J. Comput. Chem. 2003, 24, 1740. was used for the Coulomb part and the chain of sphere approach4646 Neese, F.; Wennmohs, F.; Hansen, A.; Becker, U.; Chem. Phys. 2009, 356, 98. for the exchange part of the Fock matrix employing the def2-TZVP/J auxiliary basis set.4747 Weigend, F.; Phys. Chem. Chem. Phys. 2006, 8, 1057. We have shown that this combination of exchange-correlation functional and basis set provides good structural and energetic results for a series of transition metal complexes4848 Ferreira, D. E. C.; De Almeida, W. B.; Neves, A.; Rocha, W. R.; Comput. Theor. Chem. 2012, 979, 89. and also to describe the electronic spectrum of ruthenium-amine compounds.4949 Chagas, M. A.; Rocha, W. R.; Chem. Phys. Lett. 2014, 612, 78. Calculations on the open-shell structures were carried out within the unrestricted Kohn-Sham formalism. For the RuIII species with d55 Murray, B. S.; Babak, M. V.; Hartinger, C. G.; Dyson, P. J.; Coord. Chem. Rev. 2016, 306, 86. configuration, only the S = 1 / 2 configuration was evaluated since it is the most stable spin state of RuIII complexes.5050 Chen, J.; Chen, L.; Liao, S.; Zheng, K.; Ji, L.; Dalton Trans. 2007, 3507.

51 Vargiu, A. V.; Robertazzi, A.; Magistrato, A.; Ruggerone, P.; Carloni, P.; J. Phys. Chem. B 2008, 112, 4401.
-5252 Rodrigues, G. L. S.; Rocha, W. R.; J. Phys. Chem. B 2016, 120, 11821. The optimized Cartesian coordinates of all species studied in this work can be found in the Supplementary Information (SI) section.

The reduction potential of the complexes was calculated using equation 1, as shown by Rulíšek.5353 Rulíšek, L.; J. Phys. Chem. C 2013, 117, 16871.

(1) E 0 V = 27 . 21 G ox a . u . G red a . u . E abs 0 NHE V

where Gox and Gred are the Gibbs free energies of the oxidized and reduced forms, respectively. E0abs (NHE) is the absolute potential of the normal hydrogen electrode of 4.218 V, recommended by Isse and Gennaro.5454 Isse, A. A.; Gennaro, A.; J. Phys. Chem. B 2010, 114, 7894. Gibbs free energy of the species was obtained using equation 2:

(2) G = E elec nuc + Δ G solv + G therm

where ΔGsolv is the free energy of solvation, Gtherm is the thermal contribution, obtained under the harmonic oscillator and rigid rotor approach and Eelec-nuc is the total electronic and nuclear energy of the species. The ΔGsolv was obtained using the solvation model based on density (SMD) of Truhlar and co-workers5555 Marenich, A. V.; Cramer, C. J.; Truhlar, D. G.; J. Phys. Chem. B 2009, 113, 6378. as implemented in the ORCA program,4040 Neese, F.; Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 73.,5555 Marenich, A. V.; Cramer, C. J.; Truhlar, D. G.; J. Phys. Chem. B 2009, 113, 6378. in which the electrostatic contribution is obtained using the conductor-like screening model (COSMO) of Klamt and Schuurmann.5656 Klamt, A.; Schuurmann, G.; J. Chem. Soc., Perkin Trans. 2 1993, 799. We have shown that this combination of methods/basis set provides good reduction potentials for ruthenium complexes.5252 Rodrigues, G. L. S.; Rocha, W. R.; J. Phys. Chem. B 2016, 120, 11821.

The Gibbs free energy for the interactions involving the RuII-monoaquo species and the biomolecules (GN3 and GN7 sites, Cys and GSH) were computed using the thermodynamic cycle exemplified in Figure 3 (for the KP1019 monohydrated compound), at the TPSSh/def2-TZVP level of theory.

Figure 3
Example of the thermodynamic cycle used to compute the Gibbs free energy for the interaction of the RuII-monoaquo complexes with the biological molecules L (GN3 and GN7 sites, Cys and GSH). ΔGg, ΔGsolv and ΔGsol are the Gibbs free energy of the reaction in gas phase, of solvation and of the reaction in solution, respectively. q is zero when L is the protonated form of cysteine and glutathione (CysH, GSH) and the DNA nucleobases, and –1 when L is the anionic form of cysteine and glutathione (Cys, GS). When the protonated forms of Cys and GSH are used, the final products for these interactions contain H3O+ instead of H2O.

Solvation free energies were computed as described before and the thermal corrections for the zero point energy at 298.15 K were obtained through harmonic frequency calculations at the TPSSh/def2-TZVP. The overall free energy in solution (ΔGsol) was then computed using equation 3:

(3) Δ G sol = Δ G g + Δ Δ G solv

where ΔGg is the free energy in gas phase and Δ(ΔGsolv) is the difference in free energy of solvation of the species involved. All calculations were performed using the ORCA program package.4040 Neese, F.; Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 73.

Results and Discussion

Molecular structure, reduction potentials and hydrolysis reactions

The TPSSh/def2-TZVP optimized structures of NAMI-A and KP1019 are shown in Figure 4. The complexes show octahedral geometry around the metallic center and the computed structural parameters are in excellent agreement with the experimental findings1919 Alessio, E.; Balducci, G.; Lutman, A.; Mestroni, G.; Calligaris, M.; Attia, W. M.; Inorg. Chim. Acta 1993, 203, 205.,5757 Peti, W.; Pieper, T.; Sommer, M.; Keppler, B. K.; Giester, G.; Eur. J. Inorg. Chem. 1999, 1999, 1551. with deviations in the order of 0.01 Å in the bond lengths and 0.01º in the bond angles.

Figure 4
TPSSh/def2-TZVP optimized structures of (a) NAMI-A and (b) KP1019. Bond distances are given in Å.

It is worth noting that the KP1019 structure is stabilized by intramolecular hydrogen bonds of NH···Cl type, in accordance with the determined X-ray crystal structure.2323 Reisner, E.; Arion, V. B.; da Silva, M. F. C. G.; Lichtenecker, R.; Eichinger, A.; Kepler, B. K.; Kukushkin, V. Y.; Pombeiro, A. J. L.; Inorg. Chem. 2004, 43, 7083. When a rotation of 45º around the Ru-indazole bond is performed, the new generated structure is ca. 10 kcal mol-1 higher in energy. Therefore, the intramolecular hydrogen bonds have an important effect of stabilizing the KP1019 structure around 10 kcal mol-1. Both NAMI-A and KP1019 are paramagnetic species and the calculations showed that the excess a spin densities is localized almost exclusively at the dz2 orbital of the ruthenium center with ra = 0.900 for NAMI-A and rα = 0.852 for KP1019.

The calculated one electron reduction potentials of the NAMI-A, KP1019 and their respective products from the chloride ligand exchange with water are summarized in Table 1. As it can be seen, while the reduction potential of the KP1019 complex is unresponsive relative to chloride exchange, the reduction potential of NAMI-A increases 0.132 V after the first hydrolysis, and 0.361 and 0.443 V after the second hydrolysis in cis and trans positions, respectively. After the third hydrolysis, the reduction potential increases 0.795 V.

Table 1
Computed reduction potential for the NAMI-A, KP1019 and their hydrolysis products

Electrochemical studies conducted by Brindell et al.5858 Brindell, M.; Piotrowska, D.; Shoukry, A. A.; Stochel, G.; van Eldik, R.; J. Biol. Inorg. Chem. 2007, 12, 809. have shown that the hydrolysis products of NAMI-A have redox potentials in the range of 0.187-0.597 V vs. NHE. The wave recorded at E0 = 0.597 V was assigned to a mixture of [RuCl3(H2O)(DMSO)(Im)] and [RuCl2(H2O)2(DMSO)(Im)]+, corresponding to an increase of 0.299 V when compared with the value of E0 = 0.298 V measured for NAMI-A. An increase of the 0.12 V in the reduction potential of the monoaquo complex was also observed.2222 Ravera, M.; Baracco, S.; Cassino, C.; Zanello, P.; Osella, D.; Dalton Trans. 2004, 2347.,2323 Reisner, E.; Arion, V. B.; da Silva, M. F. C. G.; Lichtenecker, R.; Eichinger, A.; Kepler, B. K.; Kukushkin, V. Y.; Pombeiro, A. J. L.; Inorg. Chem. 2004, 43, 7083. For the KP1019 complex, it is also observed an increase in the reduction potential after the first hydrolysis. Ravera et al.2525 Ravera, M.; Cassino, C.; Baracco, S.; Osella, D.; Eur. J. Inorg. Chem. 2006, 740. measured the reduction potential of [RuIII(H2O)Cl3(Ind)2] at –0.072 V vs. Ag/AgCl, which corresponds to an increase of 0.088 V compared with the KP1019 reduction potential. Therefore, the trend observed in our computed values for NAMI-A and its hydrolysis products are in line with the experimental findings and shows that it is easier to reduce the hydrolyzed products than the original compound. However, for the KP1019, our results show that the reduction potential decreases after the first hydrolysis, increases very little in the second hydrolysis and returns almost to its original value after the third hydrolysis. Therefore, the complexes show distinct electrochemical behavior, which may partially explain their biological diversity.

For both complexes, the lowest unoccupied molecular orbitals (LUMO), which can be directly correlated with the electron accepting capability of the complex, are located at the metallic center, as can be seen in Figures S1 and S2 (SI section). The next orbitals LUMO + 1 are within 0.2-1.0 eV higher in energy and, therefore, the reduction process takes place effectively at the metallic center. In the KP1019 complex there is an indazole σ-donor ligand that contributes to increase the electronic density at the metallic center, making it more difficult to reduce. For the NAMI-A the π-acceptor DMSO ligand withdraws electron density from the metal, stabilizes the RuII species and facilitates the reduction. The increase of the reduction potential with subsequent hydrolysis is in qualitative agreement with the experimental results, showing an increase of the reduction potential for each chloride exchanged by water.5858 Brindell, M.; Piotrowska, D.; Shoukry, A. A.; Stochel, G.; van Eldik, R.; J. Biol. Inorg. Chem. 2007, 12, 809.,5959 Brindell, M.; Stawoska, I.; Supel, J.; Skoczowski, A.; Stochel, G.; van Eldik, R.; J. Biol. Inorg. Chem. 2008, 13, 909.

Our TPSSh/def2-TZVP/SMD computed reduction potentials are in good agreement with the available experimental values. The variation observed may be related to description of the solvent effects as well as spin-orbit effects, as was pointed out by other authors for the reduction potential of some ruthenium-containing compounds.6060 Srnec, M.; Chalupsky, J.; Fojta, M.; Zendlova , L.; Havran, L.; Hocek, M.; Kyvala, M.; Rulíšek, L.; J. Am. Chem. Soc. 2008, 130, 10947.,6161 Rulíšek, L.; J. Phys. Chem. C 2013, 117, 16871. In some cases, quantitative agreement is obtained only when explicit solvent molecules are employed. However, we have recently shown that the procedure used in this work produces excellent results for the reduction potential of ruthenium complexes with amine ligands.5252 Rodrigues, G. L. S.; Rocha, W. R.; J. Phys. Chem. B 2016, 120, 11821.

The TPSSh/def2-TZVP optimized structures for the products of the hydrolysis reactions of NAMI-A, in the +3 and +2 oxidation states, are shown in Figures 5 and 6, respectively, which highlight the main structural parameters around the metallic center. The optimized structures for the products of the hydrolysis reactions of KP1019, in the +3 and +2 oxidation states, are shown in Figures S3 and S4 (SI section), respectively. The thermodynamics associated with the chloride exchange reaction by water was evaluated considering the ruthenium atom in the +3 and +2 formal oxidation states, in order to determine the preferential path for the hydrolysis reaction (pathways a or b, Figure 2). The computed TPSSh/def2-TZVP Gibbs free energies of reaction for the hydrolysis reactions at 298.15 K are quoted in Table 2.

Figure 5
TPSSh/def2-TZVP optimized structures for the NAMI-A complex in the +3 formal oxidation state (NAMIIII) and its mono- (NAMIIII-(H2O)), di- (NAMIIII-(H2O)2-cis, NAMIIII-(H2O)2-trans) and tri- (NAMIIII-(H2O)3) water substituted derivatives. Bond distances are given in Å.
Figure 6
TPSSh/def2-TZVP optimized structures of the reduced form for of NAMI-A complex in the +2 formal oxidation state (NAMIII) and its mono- (NAMIII-(H2O)), di- (NAMIII-(H2O)2-cis, NAMIII-(H2O)2-trans) and tri- (NAMIII-(H2O)3) water substituted derivatives. Bond distances are given in Å.

Table 2
Computed Gibbs free energy for the ligand displacement reactions involving water and chloride ligands of NAMI-A and KP1019 complexes in their normal and reduced forms

For both NAMI-A and KP1019 in the +2 and +3 oxidation state, the chloride exchange reaction by water is highly favorable and the Gibbs free energy varies almost linearly with the number of water molecules. For instance, as can be seen in Table 2, the free energy for the second and third hydrolysis of NAMI-A is 1.95 and 2.80 times the free energy for the first hydrolysis, respectively, for the +3 oxidation state, and 2.00 and 2.96 for the +2 oxidation state. It is worth noting that the free energy for the first hydrolysis of NAMI-A is almost of the same magnitude for the +2 (–60.4 kcal mol-1) and +3 (–57.4 kcal mol-1) oxidation states, which is also true for the KP1019. However, the hydrolysis with the RuII is always more favorable than the respective RuIII complexes.

Interestingly, the thermodynamic results reported here correlate well with the Becke, three-parameter, Lee-Yang-Parr (B3LYP)/6-31G(d,p) kinetic data of Vargiu et al.,6262 Vargiu, A. V.; Robertazzi, A.; Magistrato, A.; Ruggerone, P.; Carloni, P.; J. Phys. Chem. B 2008, 112, 4401. who showed that the activation energy for the first hydrolysis decreases about 3 kcal mol-1 with the reduction of NAMI-A. For the second hydrolysis this decrease is even more pronounced, ranging from ca. 7 kcal mol-1 for the substitution in cis and 9 kcal mol-1 for the substitution in trans position.6262 Vargiu, A. V.; Robertazzi, A.; Magistrato, A.; Ruggerone, P.; Carloni, P.; J. Phys. Chem. B 2008, 112, 4401. These results show that the hydrolysis of the reduced complexes is thermodynamically more favorable and kinetically more accessible than the corresponding RuIII complexes. On the other hand, when the same analysis is made for KP1019, we observe that the reduction does not influence the overall Gibbs free energy of the reaction, since the magnitude of the reaction free energy is almost the same for the RuII and RuIII complexes. This fact may aid in explaining the different mechanism of action of both drugs.

The Gibbs free energy involved in the reduction of NAMI-A and KP1019 is computed as –98.0 and –110.0 kcal mol-1, respectively. Therefore, these results suggest that both complexes have thermodynamic preference to be reduced first, prior to their hydrolysis, thus following pathway b (Figure 2). In fact, solvolysis of the electrochemically generated RuII species has been observed for both KP1019 and NAMI-A33 Clarke, M. J.; Coord. Chem. Rev. 2003, 236, 209.,2222 Ravera, M.; Baracco, S.; Cassino, C.; Zanello, P.; Osella, D.; Dalton Trans. 2004, 2347.,2323 Reisner, E.; Arion, V. B.; da Silva, M. F. C. G.; Lichtenecker, R.; Eichinger, A.; Kepler, B. K.; Kukushkin, V. Y.; Pombeiro, A. J. L.; Inorg. Chem. 2004, 43, 7083.,2727 Reisner, E.; Arion, V. B.; Kepler, B. K.; Pombeiro, A. J. L.; Inorg. Chim. Acta 2008, 361, 1569. and, in general, the water/chloride exchange reaction is accelerated upon reduction of the RuIII species. However, in the biological medium, it is not straightforward to attribute a preferred pathway for the mechanism of action, since several processes can alter the reduction potential of the drug and, therefore, alter the rate of the solvolysis process. For instance, in the biological medium the prodrugs can be reduced by well-known biological reductants such as reduced nicotinamide adenine dinucleotide (NADH), ascorbate, tocopherol (vitamin E), Cys and GSH, for which the reduction potential at physiological pH is around –0.3 to –0.5 V.6363 Hamer, M.; Suarez, S. A.; Neuman, N. I.; Alvarez, L.; Muñoz, M.; Marti, M. A.; Inorg. Chem. 2015, 54, 9342.,6464 Suarez, A. A.; Neuman, N. I.; Muñoz, M.; Álvarez, L.; Bikiel, D. E.; Brondino, C. D.; Ivanović-Burmazović, I.; Miljkovic, J. L.; Filipovic, M. R.; Martí, M. A.; Doctorovich, F.; J. Am. Chem. Soc. 2015, 137, 4720. However, the reaction mechanism involved in the reduction of these complexes is not known and, therefore, in the time scale required for the biological reduction of these prodrugs, solvolysis can takes place. NMR experiments have shown that, under buffered conditions, KP1019 is reduced within 3.5 h in the presence of GSH and within min in the presence of ascorbic acid.2424 Schluga, P.; Hartinger, C. G.; Egger, A.; Reisner, E.; Galanski, M.; Jakupec, M. A.; Kepler, B. K.; Dalton Trans. 2006, 1796. It has been suggested that the transport and delivery of the RuIII drugs occur through the interaction with serum proteins such as transferrin and albumin,1717 Trondl, R.; Heffeter, P.; Kowol, C. R.; Jakupec, M. A.; Berger, W.; Kepler, B. K.; Chem. Sci. 2014, 5, 2925. and ruthenium complexes are not reduced effectively while bound to these proteins.2727 Reisner, E.; Arion, V. B.; Kepler, B. K.; Pombeiro, A. J. L.; Inorg. Chim. Acta 2008, 361, 1569. Therefore, aquation can take place prior to reduction and alter the redox properties.

The results reported in Tables 1 and 2 also support the hypothesis that the reduced form of the complexes studied here, as well as their hydrolysis products, are relevant for their antitumor activity, since their formation is thermodynamically favorable and their reduction potentials are biologically accessible. For both complexes, the thermodynamics of the hydrolysis processes are extremely favorable and the Gibbs free energy calculations for the successive hydrolysis reactions revealed that all chloride ligands could be favorably replaced by water. Interestingly, this evidence is in line with the recent work of Keppler and co-workers,6565 Bijelic, A.; Theiner, S.; Keppler, B. K.; Rompel, A.; J. Med. Chem. 2016, 59, 5894. who reported the X-ray structure of KP1019 bound to human serum albumin (HSA). The authors showed that KP1019 lost all of the original ligands when crystallized with the protein. An analogous behavior was also observed for the NAMI-A complex and NAMI-A-type complexes.6666 Levina, A.; Lay, P. A.; Inorg. Chem. Front. 2014, 1, 44.,6767 Casini, A.; Temperini, C.; Gabbiani, C.; Supuran, C. T.; Messori, L.; ChemMedChem 2010, 5, 1989.

Interactions with biological targets

After being activated, distributed and located in cell compartments, the drug kills the cancer cell through interactions with different cellular components or with DNA. In fact, the primary target for these interactions is still unknown. Wang et al.6868 Wang, F.; Xu, J.; Wu, K.; Weidt, S. K.; Mackay, C. L.; Langridge-Smith, P. R. R.; Sadler, P. J.; Dalton Trans. 2013, 42, 3188. have shown that Ru-based complexes follow different reaction mechanisms than cisplatin in the competitive reaction between S-donor ligands such as GSH and N-donor ligands. In this mechanism, the reaction between the ruthenium complex and GSH leads to the formation of a ruthenium thiolate compound, which can then undergo oxidation by oxygen to produce an original sulfenate complex. The oxidation of coordinated GSH in the thiolate complex seems to provide an easy route for GSH exchange and Ru-N7 binding formation. This may provide a unique mechanism of action and explain the lack of cross-resistance with platinum drugs, which is a potential clinical advantage. Other cytotoxic activities from NAMI-A and KP1019 may be attributed to their interactions with DNA nucleobases generally to the N7 site of guanine.1717 Trondl, R.; Heffeter, P.; Kowol, C. R.; Jakupec, M. A.; Berger, W.; Kepler, B. K.; Chem. Sci. 2014, 5, 2925.,2424 Schluga, P.; Hartinger, C. G.; Egger, A.; Reisner, E.; Galanski, M.; Jakupec, M. A.; Kepler, B. K.; Dalton Trans. 2006, 1796. It is worth mentioning that interaction of ruthenium compounds with nucleic acid and derivatives generates compounds with chemical and photophysical properties with important applications in nanotechnologies, medical diagnosis and therapies, and have recently been reviewed by Flamme et al.6969 Flamme, M.; Clarke, E.; Gasser, G.; Hollenstein, M.; Molecules 2018, 23, 1515.

Since our results suggested that both NAMI-A, KP1019 and their hydrolyzed products are very prone to be reduced, we evaluated the affinity of the NAMI-A and KP1019 for different intracellular components, though the calculation of the thermodynamics of the interaction of the reduced and mono-hydrated forms of NAMI-A and KP1019 with different S-donors (GSH and Cys) and N-donors (GN3 and GN7 sites) ligands.

The TPSSh/def2-TZVP optimized Cartesian coordinates of the free ligands, the protonated form of cysteine and glutathione (CysH and GSH) and guanine, can be found in the SI section. For the GSH, the optimization started with the most stable conformation found at physiological pH, taken from molecular dynamics conformational sampling studies.7070 Lampela, O.; Juffer, A. H.; Rauk, A.; J. Phys. Chem. A 2003, 107, 9208.,7171 Zhang, R.; Wu, W.; Luo, S.; J. Solution Chem. 2011, 40, 1784. Figure 7 shows the possible adducts that can be formed by the interaction of the mono-aquo species of the reduced forms of NAMI-A ([Ru(Cl)3(H2O)(Im)(DMSO)], NAMIII-H2O) and KP1019 ([Ru(Cl)3(H2O)(In)2], KP1019II-H2O) with the biological molecules. The optimized structures of the adducts are presented in Figures 8 and 9, which highlight the main structural parameters around the metallic center. As can be seen, the complexes exhibit an octahedral arrangement of the ligands, with the angles around the metal deviating from the optimal values for the octahedral geometry.

Figure 7
Schematic representation of the reactions between the RuII-monoaquo complexes of (a) NAMI-A and (b) KP1019 with the GN3 and GN7 sites, CysH and GSH.
Figure 8
TPSSh/def2-TZVP optimized structures of the adducts formed by the reaction of the mono-aquo species of NAMI-A (NAMIII-(H2O)) with the GN3 and GN7 sites, CysH and GSH. A1, A2, A3, A4 and A5 are the structures obtained from the reaction with CysH, GSH in circular conformation, GN3 and GN7 sites and GSH in extended conformation, respectively. Bond distances are given in Å.
Figure 9
TPSSh/def2-TZVP optimized structures of the adducts formed by the reaction of the mono-aquo species of KP1019 (KP1019II-(H2O)) with the GN3 and GN7 sites, CysH and GSH. B1, B2, B3, B4, B5 are the structures obtained from the reaction with CysH, GSH in circular conformation, GN3 and GN7 sites and GSH in extended conformation, respectively.

For each of the reactions showed in Figure 7, we have used a thermodynamic cycle to compute the Gibbs free energy for the ligand displacement reaction in solution, as the one exemplified in Figure 3. The final Gibbs free energy for the interaction in aqueous solution was then computed using equation 3 and the results are quoted in Table 3. As is shown in equation 3, the overall Gibbs free energy of the reaction is composed of two major components: (i) the Gibbs free energy of the reaction in gas phase (ΔGg); and (ii) the variation of the solvation free energies of the species.

Table 3
Computed Gibbs free energy for the reaction of the RuII-monoaquo complexes (NAMIII-H2O and KP1019II-H2O) with the GN3 and GN7 sites, CysH and GSH

The results quoted in Table 3 show that the interaction of NAMIII-H2O and KP1019II-H2O complexes with the GN3 site is unfavorable. The interaction with guanine at the GN7 site is only slightly favorable in solution, with DGsol = –0.4 kcal mol-1 for NAMIII-H2O and –4.8 kcal mol-1 for the interaction with KP1019II-H2O. It is interesting to note that the intereaction of both complexes with the depronated form of GSH and Cys, generating the complexes [RuCl3(DMSO)(Im)(Cys)]2–, [RuCl3(In)2(Cys)]2–, [RuCl3(DMSO)(Im)(GS)]2–, [RuCl3(In)2(GS)]2–, with overall –2 charge, is highly unfavorable to happen in solution. This may be attributed to the unfavorable interaction between two negatively charged species. On the other hand, the coordination of GSH in the protonated form is slightly favorable for NAMIII-H2O and KP1019II-H2O with solvation free energies of –0.7 and –2.6 kcal mol-1, respectively, for the water displacement reaction. Also, displacement of water by the protonated cysteine is also favorable for both complexes. Since it is well known that both NAMI-A and KP1019 interact with GSH under buffered condition,2424 Schluga, P.; Hartinger, C. G.; Egger, A.; Reisner, E.; Galanski, M.; Jakupec, M. A.; Kepler, B. K.; Dalton Trans. 2006, 1796. these results suggest, therefore, that the S-donor ligands may interact with the metallodrug in the protonated form, which leads to negative Gibbs free energy of reaction. For both complexes, the interaction with the GN7 site and GSH are of the same magnitude, meaning that these ligands can compete for binding to the metallodrug. And, in the case of KP1019, interaction with Cys is also of the same magnitude. The unfavorable (or only slightly favorable) free energy values for the interaction of the complexes with GN3 and GN7 sites suggest that the principal mechanism of action of NAMI-A and KP1019 may not involve the direct interaction with DNA. Indeed, it has been shown that the cytotoxic activity of KP1019 is based on ER stress-related effects while NAMI-A interacts with some proteins responsible for DNA replication.3131 Bergamo, A.; Gaiddon, C.; Schellens, J. H. M.; Beijnen, J. H.; Sava, G.; J. Inorg. Biochem. 2012, 106, 90.

32 Han Ang, W.; Dyson, P. J.; Eur. J. Inorg. Chem. 2006, 20, 4003.

33 Qian, C.; Wang, J. Q.; Song, C. L.; Wang, L. L.; Ji, L. N.; Chao, H.; Metallomics 2013, 5, 844.

34 Bierle, L. A.; Reich, K. L.; Taylor, B. E.; Blatt, E. B.; Middleton, S. M.; Burke, S. D.; Stultz, L. K.; Hanson, P. K.; Partridge, J. F.; Miller, M. E.; PLoS One 2015, 10, e0138085.
-3535 Flocke, L. S.; Trondl, R.; Jakupec, M. A.; Keppler, B. K.; Invest. New Drugs 2016, 34, 261.

Conclusions

In this work, we employed the DFT at the TPSSh/def2-TZVP level, in combination with the SMD continuum solvation model, to investigate the electrochemistry of NAMI-A, KP1019 and their hydrolysis products, as well as the thermodynamics of their interactions with S-donor (CysH and GSH) and N-donor (GN3 and GN7 coordination sites) biological molecules. Our results show that the compounds exhibit different electrochemical behavior upon hydrolysis. The reduction potential of NAMI-A is sensitive to the degree of hydrolysis, increasing with the number of chloride ligands replaced by water. On the other hand, the reduction potential of KP1019 does not vary with the hydrolysis and remains almost constant. Our results showed that thermodynamically, the NAMI-A and KP1019 complexes have a thermodynamic preference to be reduced first and to undergo hydrolysis after, since the free energy involved in the reduction process of the complexes is more negative. For both complexes, the interaction with the GN7 site and GSH are of the same magnitude (ΔGsol ca. –0.6 kcal mol-1) meaning that these ligands can compete for binding to the metallodrug. And, in the case of KP1019, interaction with CysH is also of the same magnitude. The unfavorable (or only slightly favorable) free energy values for the interaction of the complexes with guanine at the N3 and N7 sites suggest that the principal mechanism of action of NAMI-A and KP1019 may not involve the direct interaction with DNA. The calculations showed that the interaction of the complexes with GSH and Cys is only favorable in the neutral or protonated forms, and this is an important information to propose a reaction mechanism for the reduction reaction of these metallodrugs by thiol-like reductants. It is also worth noting that, according to the calculations, once the first hydrolysis takes place the complexes have thermodynamic preference to undergo further hydrolysis than to interact with the S-donor or N-donor ligands studied in this work, and this is somehow in line with recent experimental findings.

As a final note, it is important to mention that, despite the agreements with some experimental evidences, our study was focused on the thermodynamics of the process, to discuss the preferential interactions. However, the kinetics of the processes studied in this work will certainly play an important role in determining the preferential pathway for these competitive interactions.

  • Supplementary Information
    Supplementary information (plot of the frontier molecular orbitals for KP1019, NAMI-A and their hydrolysis products, figures of the optimized structures of the KP1019 complexes and TPSSh/def2-TZVP optimized Cartesian coordinates of all species studied in this work) is available free of charge at http://jbcs.sbq.org.br as PDF file.

Acknowledgments

The authors would like to thank Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq; INCT-Catálise) and Fundação de Amparo à Pesquisa do Estado de Minas Gerais (FAPEMIG) for the financial support and research grants.

References

  • 1
    Rosenberg, B.; VanCamp, L.; Trosko, J. E.; Mansour, H. V.; Nature 1969, 222, 385.
  • 2
    Rosenberg, B.; Van Camp, L.; Krigas, T.; Nature 1965, 205, 698.
  • 3
    Clarke, M. J.; Coord. Chem. Rev 2003, 236, 209.
  • 4
    Mjos, K. D.; Orvig, C.; Chem. Rev 2014, 114, 4540.
  • 5
    Murray, B. S.; Babak, M. V.; Hartinger, C. G.; Dyson, P. J.; Coord. Chem. Rev 2016, 306, 86.
  • 6
    Johnstone, T. C.; Suntharalingam, K.; Lippard, S. J.; Chem. Rev 2016, 116, 3436.
  • 7
    Pasini, A.; Zunino, F.; Angew. Chem., Int. Ed. Engl. 1987, 26, 615.
  • 8
    Boulikas, T.; Vougiouka, M.; Oncol. Rep. 2003, 10, 1663.
  • 9
    Smith, N. A.; Sadler, P. J.; Philos. Trans. R. Soc., A 2013, 371, 20120519.
  • 10
    Sava, G.; Pacor, S.; Mestroni, G.; Alessio, E.; Clin. Exp. Metastasis 1992, 10, 273.
  • 11
    Hartinger, C. G.; Jakupec, M. A.; Zorbas-Seifried, S.; Groessl, M.; Egger, A.; Berger, W.; Zorbas, H.; Dyson, P. J.; Keppler, B. K.; Chem. Biodiversity 2008, 5, 2140.
  • 12
    Sava, G.; Bergamo, A.; Zorzet, S.; Gava, B.; Casarsa, C.; Cocchietto, M.; Furlani, A.; Scarcia, V.; Serli, B.; Iengo, E.; Alessio, E.; Mestroni, G.; Eur. J. Cancer 2002, 38, 427.
  • 13
    Leijen, S.; Burgers, S. A.; Baas, P.; Pluim, D.; Tibben, M.; van Werkhoven, E.; Alessio, E.; Sava, G.; Beijnen, J. H.; Schellens, J. H.; Invest. New Drugs 2015, 33, 201.
  • 14
    Sava, G.; Zorzet, S.; Turrin, C.; Francesca, V.; Soranzo, M. R.; Zabucchi, G.; Cocchietto, M.; Bergamo, A.; DiGiovine, S.; Pezzoni, G.; Sartor, L.; Garbisa, S.; Clin. Cancer Res. 2003, 9, 1898.
  • 15
    Gava, B.; Zorzet, S.; Spessotto, P.; Cocchietto, M.; Sava, G.; J. Pharmacol. Exp. Ther 2006, 317, 284.
  • 16
    Alessio, E.; Eur. J. Inorg. Chem 2017, 2017, 1549.
  • 17
    Trondl, R.; Heffeter, P.; Kowol, C. R.; Jakupec, M. A.; Berger, W.; Kepler, B. K.; Chem. Sci 2014, 5, 2925.
  • 18
    Costa, G.; Balducci, G.; Alessio, E.; Tavagnacco, C.; Mestroni, G.; J. Electroanal. Chem 1990, 296, 57.
  • 19
    Alessio, E.; Balducci, G.; Lutman, A.; Mestroni, G.; Calligaris, M.; Attia, W. M.; Inorg. Chim. Acta 1993, 203, 205.
  • 20
    Hartmann, M.; Lipponer, K.-G.; Kepler, B. K.; Inorg. Chim. Acta 1998, 267, 137.
  • 21
    Frasca, D. R.; Clarke, M. J.; J. Am. Chem. Soc 1999, 121, 8523.
  • 22
    Ravera, M.; Baracco, S.; Cassino, C.; Zanello, P.; Osella, D.; Dalton Trans 2004, 2347.
  • 23
    Reisner, E.; Arion, V. B.; da Silva, M. F. C. G.; Lichtenecker, R.; Eichinger, A.; Kepler, B. K.; Kukushkin, V. Y.; Pombeiro, A. J. L.; Inorg. Chem 2004, 43, 7083.
  • 24
    Schluga, P.; Hartinger, C. G.; Egger, A.; Reisner, E.; Galanski, M.; Jakupec, M. A.; Kepler, B. K.; Dalton Trans 2006, 1796.
  • 25
    Ravera, M.; Cassino, C.; Baracco, S.; Osella, D.; Eur. J. Inorg. Chem 2006, 740.
  • 26
    Cebrián-Losantos, B.; Reisner, E.; Kowol, C. R.; Roller, A.; Shova, S.; Arion, V. B.; Kepler, B. K.; Inorg. Chem 2008, 47, 6513.
  • 27
    Reisner, E.; Arion, V. B.; Kepler, B. K.; Pombeiro, A. J. L.; Inorg. Chim. Acta 2008, 361, 1569.
  • 28
    Lever, A. B. P.; Inorg. Chem 1990, 29, 1271.
  • 29
    Mestroni, G.; Alessio, E.; Sava, G.; Pacor, S.; Coluccia, M.; Boccarelli, A.; Met.-Based Drugs 1994, 1, 41.
  • 30
    Haukka, M.; Ahlgrén, M.; Pakkanen, T. A.; J. Chem. Soc., Dalton Trans 1996, 1927.
  • 31
    Bergamo, A.; Gaiddon, C.; Schellens, J. H. M.; Beijnen, J. H.; Sava, G.; J. Inorg. Biochem 2012, 106, 90.
  • 32
    Han Ang, W.; Dyson, P. J.; Eur. J. Inorg. Chem 2006, 20, 4003.
  • 33
    Qian, C.; Wang, J. Q.; Song, C. L.; Wang, L. L.; Ji, L. N.; Chao, H.; Metallomics 2013, 5, 844.
  • 34
    Bierle, L. A.; Reich, K. L.; Taylor, B. E.; Blatt, E. B.; Middleton, S. M.; Burke, S. D.; Stultz, L. K.; Hanson, P. K.; Partridge, J. F.; Miller, M. E.; PLoS One 2015, 10, e0138085.
  • 35
    Flocke, L. S.; Trondl, R.; Jakupec, M. A.; Keppler, B. K.; Invest. New Drugs 2016, 34, 261.
  • 36
    Smith, C. A.; Sutherland-Smith, A. J.; Kepler, B. K.; Kratz, F.; Baker, E. N.; J. Biol. Inorg. Chem 1996, 1, 424.
  • 37
    Clarke, M. J.; Zhu, F.; Frasca, D. R.; Chem. Rev 1999, 99, 2511.
  • 38
    Parr, R. G.; Yang, W.; Density Functional Theory of Atoms and Molecules; Oxford University Press: New York, 1989.
  • 39
    Staroverov, V. N.; Scuseria, G. E.; Tao, J.; Perdew, J. P.; J. Chem. Phys 2003, 119, 12129.
  • 40
    Neese, F.; Wiley Interdiscip. Rev.: Comput. Mol. Sci 2012, 2, 73.
  • 41
    Weigend, F.; Ahlrichs, R.; Phys. Chem. Chem. Phys 2005, 7, 3297.
  • 42
    van Lenthe, E.; Baerends, E. J.; Snijders, J. G.; J. Chem. Phys 1993, 99, 4597.
  • 43
    van Lenthe, E.; Baerends, E. J.; Snijders, J. G.; J. Chem. Phys 1994, 101, 9783.
  • 44
    van Lenthe, E.; van Leeuwen, R.; Baerends, E. J.; Snijders, J. G.; Int. J. Quantum Chem 1996, 57, 281.
  • 45
    Neese, F.; J. Comput. Chem 2003, 24, 1740.
  • 46
    Neese, F.; Wennmohs, F.; Hansen, A.; Becker, U.; Chem. Phys 2009, 356, 98.
  • 47
    Weigend, F.; Phys. Chem. Chem. Phys 2006, 8, 1057.
  • 48
    Ferreira, D. E. C.; De Almeida, W. B.; Neves, A.; Rocha, W. R.; Comput. Theor. Chem 2012, 979, 89.
  • 49
    Chagas, M. A.; Rocha, W. R.; Chem. Phys Lett 2014, 612, 78.
  • 50
    Chen, J.; Chen, L.; Liao, S.; Zheng, K.; Ji, L.; Dalton Trans 2007, 3507.
  • 51
    Vargiu, A. V.; Robertazzi, A.; Magistrato, A.; Ruggerone, P.; Carloni, P.; J. Phys. Chem. B 2008, 112, 4401.
  • 52
    Rodrigues, G. L. S.; Rocha, W. R.; J. Phys. Chem. B 2016, 120, 11821.
  • 53
    Rulíšek, L.; J. Phys. Chem. C 2013, 117, 16871.
  • 54
    Isse, A. A.; Gennaro, A.; J. Phys. Chem. B 2010, 114, 7894.
  • 55
    Marenich, A. V.; Cramer, C. J.; Truhlar, D. G.; J. Phys. Chem. B 2009, 113, 6378.
  • 56
    Klamt, A.; Schuurmann, G.; J. Chem. Soc., Perkin Trans. 2 1993, 799.
  • 57
    Peti, W.; Pieper, T.; Sommer, M.; Keppler, B. K.; Giester, G.; Eur. J. Inorg. Chem 1999, 1999, 1551.
  • 58
    Brindell, M.; Piotrowska, D.; Shoukry, A. A.; Stochel, G.; van Eldik, R.; J. Biol. Inorg. Chem 2007, 12, 809.
  • 59
    Brindell, M.; Stawoska, I.; Supel, J.; Skoczowski, A.; Stochel, G.; van Eldik, R.; J. Biol. Inorg. Chem 2008, 13, 909.
  • 60
    Srnec, M.; Chalupsky, J.; Fojta, M.; Zendlova , L.; Havran, L.; Hocek, M.; Kyvala, M.; Rulíšek, L.; J. Am. Chem. Soc 2008, 130, 10947.
  • 61
    Rulíšek, L.; J. Phys. Chem. C 2013, 117, 16871.
  • 62
    Vargiu, A. V.; Robertazzi, A.; Magistrato, A.; Ruggerone, P.; Carloni, P.; J. Phys. Chem. B 2008, 112, 4401.
  • 63
    Hamer, M.; Suarez, S. A.; Neuman, N. I.; Alvarez, L.; Muñoz, M.; Marti, M. A.; Inorg. Chem 2015, 54, 9342.
  • 64
    Suarez, A. A.; Neuman, N. I.; Muñoz, M.; Álvarez, L.; Bikiel, D. E.; Brondino, C. D.; Ivanović-Burmazović, I.; Miljkovic, J. L.; Filipovic, M. R.; Martí, M. A.; Doctorovich, F.; J. Am. Chem. Soc 2015, 137, 4720.
  • 65
    Bijelic, A.; Theiner, S.; Keppler, B. K.; Rompel, A.; J. Med. Chem 2016, 59, 5894.
  • 66
    Levina, A.; Lay, P. A.; Inorg. Chem. Front 2014, 1, 44.
  • 67
    Casini, A.; Temperini, C.; Gabbiani, C.; Supuran, C. T.; Messori, L.; ChemMedChem 2010, 5, 1989.
  • 68
    Wang, F.; Xu, J.; Wu, K.; Weidt, S. K.; Mackay, C. L.; Langridge-Smith, P. R. R.; Sadler, P. J.; Dalton Trans 2013, 42, 3188.
  • 69
    Flamme, M.; Clarke, E.; Gasser, G.; Hollenstein, M.; Molecules 2018, 23, 1515.
  • 70
    Lampela, O.; Juffer, A. H.; Rauk, A.; J. Phys. Chem. A 2003, 107, 9208.
  • 71
    Zhang, R.; Wu, W.; Luo, S.; J. Solution Chem 2011, 40, 1784.

Publication Dates

  • Publication in this collection
    Mar 2019

History

  • Received
    05 Aug 2018
  • Accepted
    11 Oct 2018
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br