Maryam
Taghizadeh Shool
a,
Hadi
Amiri Rudbari
*a,
Tania
Gil-Antón
b,
José V.
Cuevas-Vicario
b,
Begoña
García
b,
Natalia
Busto
*bc,
Nakisa
Moini
d and
Olivier
Blacque
e
aDepartment of Chemistry, University of Isfahan, Isfahan 81746-73441, Iran. E-mail: h.a.rudbari@sci.ui.ac.ir; hamiri1358@gmail.com
bDepartamento de Química, Facultad de Ciencias, Universidad de Burgos, Plaza Misael Bañuelos s/n, 09001, Burgos, Spain. E-mail: nbusto@ubu.es
cDepartamento de Ciencias de la Salud, Facultad de Ciencias de la Salud, Universidad de Burgos, Hospital Militar, Paseo de los Comendadores, s/n, 09001 Burgos, Spain
dDepartment of Chemistry, Faculty of Physics and Chemistry Alzahra University, P.O. Box 1993891176, Vanak Tehran, Iran
eDepartment of Chemistry, University of Zurich, Winterthurerstrasse 190, CH-8057, Zurich, Switzerland
First published on 20th April 2022
Ru(II) polypyridyl complexes are widely used in biological fields, due to their physico-chemical and photophysical properties. In this paper, a series of new chiral Ru(II) polypyridyl complexes (1–5) with the general formula {Δ/Λ-[Ru(bpy)2(X,Y-sal)]BF4} (bpy = 2,2′-bipyridyl; X,Y-sal = 5-bromosalicylaldehyde (1), 3,5-dibromosalicylaldehyde (2), 5-chlorosalicylaldehyde (3), 3,5-dichlorosalicylaldehyde (4) and 3-bromo-5-chlorosalicylaldehy (5)) were synthesized and characterized by elemental analysis, FT-IR, and 1H/13C NMR spectroscopy. Also, the structures of complexes 1 and 5 were determined by X-ray crystallography; these results showed that the central Ru atom adopts a distorted octahedral coordination sphere with two bpy and one halogen-substituted salicylaldehyde. DFT and TD-DFT calculations have been performed to explain the excited states of these complexes. The singlet states with higher oscillator strength are correlated with the absorption signals and are mainly described as 1MLCT from the ruthenium centre to the bpy ligands. The lowest triplet states (T1) are described as 3MLCT from the ruthenium center to the salicylaldehyde ligand. The theoretical results are in good agreement with the observed unstructured band at around 520 nm for complexes 2, 4 and 5. Biological studies on human cancer cells revealed that dihalogenated ligands endow the Ru(II) complexes with enhanced cytotoxicity compared to monohalogenated ligands. In addition, as far as the type of halogen is concerned, bromine is the halogen that provides the highest cytotoxicity to the synthesized complexes. All complexes induce cell cycle arrest in G0/G1 and apoptosis, but only complexes bearing Br are able to provoke an increase in intracellular ROS levels and mitochondrial dysfunction.
In recent decades, ruthenium complexes have become an attractive option for biological application due to their distinct features such as (1) existence stable different oxidation states under biological conditions, (2) less toxicity due to higher selectivity of cancer cells toward healthy cells, and (3) ability to mimic iron in binding biomolecules such as transferrin and albumin because these proteins play a key role in the transport of metallodrugs and their receptors are largely overexpressed on the surface of malignant cells.7–10 Despite the synthesis of a large number of ruthenium complexes with anticancer properties, only a few of them, like NAMI-A and KP1019 (Fig. 1) have been employed in human clinical studies,11–13 probably due to their poor water solubility, insufficient effectiveness or ungovernable interaction with serum proteins.14 As a result, the synthesis of new Ru-based compounds is still necessary in order to improve the physicochemical properties (water solubility) and anticancer activity of complexes.15–17 The reactivities of transition metal coordination compounds often become controlled by the environment around the coordination sphere. Hence, polypyridines with multiple covalently bonded pyridine groups exhibit unique photophysical and redox properties.18,19 Bipyridine analogues not only function as supporting ligands stabilizing metal complexes, but also are utilized as photosensitizers20 and phosphorescent materials.21
For the first time, two Ru(II) polypyridyl complexes, [Ru(phen)3](ClO4)2 and [Ru(bpy)3](ClO4)2 have been biologically investigated in the 1950s.10 Ru(II) polypyridyl complexes have been widely investigated in cellular imaging, chemotherapy and photodynamic therapy due to their unique photochemical and photophysical properties, which can in turn be controlled by suitable variations of the auxiliary and primary ligands around the Ru(II) metal center.22,23 The relationship between the number of heteroatoms involved in the supporting ligand and the reactivity of the complex has been reported in a ruthenium complex containing bipyridine analogues.24 The investigation of the electronic properties of cyclometalated ruthenium polypyridyl has continued to be active for many years.25–27
Nowadays, the most attractive Ru(II) polypyridyl complex is the TLD-1433 compound (Fig. 1), that has recently entered phase II clinical studies for the treatment of nonmuscle invasive bladder cancer.14,28
To explore the structures of ruthenium complexes with salicylaldehyde derivatives, C. Chen et al. have studied the coordination modes of salicylaldehyde derivatives in the Ru(II) nitrosyl and Ru(II) bis(2,2′-bipyridine) complexes, with the cationic ruthenium complex [Ru(bpy)2(κ2-O,O-salCl)](PF6) being similar to ClByRu(3).29 The antitumoral and antimicrobial biological activity of some ruthenium carbonyl derivatives of the bis-(salicylaldehyde)phenylenediimine Schiff base ligand have been studied. The data showed that the complexes have the capacity of inhibiting the metabolic growth of the investigated bacteria to different extents, which may indicate their broad-spectrum properties, especially for the bipyridine derivative.30
Lastly, it is well-known that natural products contain halogens in their structures. Therefore, halogenation should be an invaluable approach for the structural modification of natural products for drug development.31 Thus, halogen atoms are widely used as substituents in medicinal chemistry which enhance the bioactivity and bioavailability of drugs through attractive intermolecular interaction (halogen bonding) between an electrophilic site on a halogen and a nucleophilic site of the molecule namely the lone pair of heteroatoms like N,O and S in proteins.32–34 Also, the introduction of halogens into the phenyl ring decreases drug metabolism.35–37 Therefore, halogen bonding is a powerful tool to design more effective medicinal compounds for medicinal chemistry. For instance, it has been recently described that in a series of Pd(II) complexes with halogen-substituted Schiff bases and 2-picolylamine, the number and types of halogens influence not only the chirality but also their cytotoxicity towards breast cancer cells.38 In contrast, Hartinger et al. found no significant differences as a function of halogens in the anticancer activity of piano stool Ru(II) complexes bearing 8-hydroxyquinoline.39
Recently we synthesized a series of Cu(diimine)(X,Y-sal)(NO3) complexes, where the diimine is either bpy or phen, sal is salicylaldehyde, and X and Y are Cl, Br, I and H. The data set showed the potential of these bpy derivatives for further in vivo studies.40
On the basis of these promising results, we synthesized such kinds of compounds with different metal ions. From the first attempt, based on the above description of the anticancer activity of the ruthenium compounds, we concluded on the one side that the ruthenium atom is the best alternative for the copper atom and, on the other side, the properties of the Ru(II)-bpy, Ru(II)-salicylaldehyde base derivatives and the introduction of halogens into the phenyl ring have prompted us to report the synthesis, structural characterization, and antibacterial and anticancer activity of this series of novel chiral {Δ/Λ-[Ru(bpy)2(X,Y-sal)]BF4} complexes (1–5), where X,Y-sal is monoanionic halogenated salicylaldehyde (X = Cl, Br and H; Y = Cl and Br).
The IR spectra of the complexes exhibit common characteristic bands for CO (aldehyde) and B–F (BF4). The stretching frequency for C
O in compounds 1, 2, 3, 4 and 5 occurs at 1600.6, 1579.4, 1584.2, 1583.2 and 1578.4 cm−1, respectively.41 The main stretching frequency for the BF4 anion occurs at 1055.8, 1056.8, 1057.7, 1058.7 and 1054.8 cm−1 for compounds 1, 2, 3, 4 and 5, respectively.42
All complexes showed well-defined 1H/13C NMR spectra (Fig. S1†), permitting the unambiguous identification and assessment of purity. In the 1H NMR spectra of the complexes, the aldehyde proton (CHO) from the halogen-substituted salicylaldehyde ligand gives rise to signals at 8.92, 9.04, 8.93, 9.05 and 9.04 ppm for 1, 2, 3, 4 and 5, respectively. In the range of 6.5–8.7 ppm the signals of the aromatic protons from the halogen-substituted salicylaldehyde ligand appear to be overlapped with those from the bpy ligand. The main modification observed in the 1H NMR spectra of the complexes in relation to that of the free salicylaldehyde ligand is the absence of a resonance at ∼10.90 ppm assigned to the proton of the phenol oxygen, indicating its deprotonation.43
The 13C NMR spectra show 27 signals for all complexes. The peak observed at ∼188.00 ppm is ascribed to the aldehyde carbon atom. The existence of this peak in the spectra of the complexes supports the presence of the salicylaldehyde ligand in the structure of the Ru(II) complexes. The peaks in the range of 104.6–169.9 ppm are assigned to the aromatic protons.
1 | 5 | |
---|---|---|
Empirical formula | C54H42B2Br2F8N8O5Ru2 | C58H52B2Br2Cl2F8N8O7Ru2 |
Formula weight | 1418.53 | 1579.55 |
Temperature (K) | 298(2) | 160(1) |
Wavelength (Å) | 0.71073 | 1.54184 |
Crystal system | Triclinic | Triclinic |
Space group |
P![]() |
P![]() |
a (Å) | 10.508(2) | 11.7796(7) |
b (Å) | 12.625(3) | 12.0345(7) |
c (Å) | 13.078(3) | 12.9612(8) |
α (°) | 108.15(3) | 63.333(6) |
β (°) | 107.82(3) | 70.416(6) |
γ (°) | 94.36(3) | 76.745(5) |
Volume (Å3) | 1541.1(8) | 1540.17(19) |
Z/calculated density (g cm−3) | 1/1.529 | 1/1.703 |
Absorption coefficient (mm−1) | 1.861 | 7.006 |
F(000) | 702 | 786.0 |
θ range (°) | 2.655 to 29.329 (θ) | 7.92 to 149.008 (2θ) |
h; k; l ranges | −13,14; ±17; −18, 17 | ±14, −15, 12; −16, 15 |
Reflections collected/unique | 28![]() |
25![]() |
Refinement method | Full-matrix least-squares on F2 | Full-matrix least-squares on F2 |
Data/restraints/parameters | 8382/0/379 | 6270/195/416 |
Goodness-of-fit on F2 | 0.978 | 1.039 |
Final R indices [I > 2σ(I)] | R 1 = 0.0555/wR2 = 0.1611 | R 1 = 0.0583/wR2 = 0.1430 |
R indices (all data) | R 1 = 0.0836/wR2 = 0.1795 | R 1 = 0.0960/wR2 = 0.1635 |
Largest diff. peak and hole (e Å−3) | 1.091/−1.065 | 0.97/−0.93 |
CCDC number | 2150468 | 2108989 |
1 | 5 | |||
---|---|---|---|---|
Experimental XRD | Calculated DFT | Experimental XRD | Calculated DFT | |
Ru(1)–N(1) | 2.056(3) | 2.080 | 2.042(6) | 2.081 |
Ru(1)–N(2) | 2.027(4) | 2.055 | 2.031(6) | 2.053 |
Ru(1)–N(3) | 2.043(3) | 2.065 | 2.040(5) | 2.064 |
Ru(1)–N(4) | 2.042(4) | 2.081 | 2.045(6) | 2.077 |
Ru(1)–O(1) | 2.066(3) | 2.094 | 2.083(4) | 2.084 |
Ru(1)–O(2) | 2.080(3) | 2.103 | 2.054(5) | 2.105 |
N(1)–Ru(1)–N(4) | 177.08(15) | 175.92 | 175.2(2) | 176.20 |
N(1)–Ru(1)–O(1) | 85.94(14) | 87.49 | 88.54(19) | 86.82 |
N(1)–Ru(1)–O(2) | 95.05(13) | 95.79 | 95.0(2) | 95.50 |
N(1)–Ru(1)–N(2) | 79.33(15) | 79.05 | 79.7(2) | 79.13 |
N(3)–Ru(1)–N(2) | 89.21(15) | 91.69 | 88.8(2) | 93.05 |
N(2)–Ru(1)–N(4) | 97.97(15) | 97.78 | 95.9(2) | 97.51 |
N(2)–Ru(1)–O(1) | 88.17(14) | 88.95 | 91.87(18) | 87.77 |
N(2)–Ru(1)–O(2) | 174.26(12) | 174.78 | 173.8(2) | 174.16 |
N(3)–Ru(1)–N(1) | 99.66(14) | 98.53 | 98.4(2) | 99.37 |
N(3)–Ru(1)–N(4) | 79.10(15) | 78.90 | 79.5(2) | 78.95 |
N(3)–Ru(1)–O(1) | 173.27(13) | 173.96 | 173.1(2) | 173.80 |
N(3)–Ru(1)–O(2) | 92.93(13) | 89.88 | 89.0(2) | 90.08 |
N(4)–Ru(1)–O(1) | 95.12(15) | 95.05 | 93.5(2) | 94.85 |
N(4)–Ru(1)–O(2) | 87.67(14) | 87.41 | 89.3(2) | 87.93 |
O(1)–Ru(1)–O(2) | 90.26(13) | 90.01 | 91.10(17) | 89.66 |
Both complexes, 1 and 5, have similar structures and crystallize in the space group P, with one enantiomer of the complex occupying the asymmetric unit.
In the cationic part of these complexes, the deprotonated aldehyde coordinates to the Ru(II) atom through the phenol-O and aldehyde-O atoms, forming a virtually planar six-membered chelate ring [maximum deviation from the least-squares plane = 0.053 Å (1) and 0.040 (5)], and two bidentate bpy co-ligands through their nitrogen atoms.44
The ruthenium atoms in both structures adopt a slightly distorted octahedral coordination geometry (Fig. 2). The Ru–Nbpy bond lengths are in the range of 2.027(4)–2.056(3) Å and 2.031(6)–2.045(6) Å for 1 and 5, respectively. The Ru–Oaldehyde bond lengths are 2.080(3) and 2.054(5) Å for 1 and 5, respectively, while the Ru–Ophenol bond lengths are 2.066(3) and 2.083(4) Å for 1 and 5, respectively. The Ru–N bond trans to the Ru–Ophenol bond (for 1: 2.043(3) Å; for 5: 2.040(5) Å) is longer than the Ru–N bond trans to the Ru–Oaldehyde bond (for 1: 2.027(4) Å; for 5: 2.031(6) Å). These results are consistent with the stronger trans influence of the Ophenol atom compared to that of the Oaldehyde atom.45
Because of the geometrical arrangement of the bpy chelating ligands and centrosymmetric space group (P for both complexes), the configuration at the Ru(II) metal center is Δ or Λ. Therefore, two enantiomers are possible such as Δ and Λ.46
The most noticeable distortion of the ideal octahedral geometry corresponds to the N–Ru–N bond angles, formed by the chelating bpy ligands, which are near 80° for both complexes (Table 2). These angles are shorter than ideal 90° found in a regular octahedron due to the geometrical requirements of the chelate rings formed by the bpy ligands.46
The similarity of both structures can be confirmed in the best way by overall conformation. As shown in Fig. 3, the dihedral angles between the three coordinating planes (N(1)–Ru(1)–N(2), N(3)–Ru(1)–N(4) and O(1)–Ru(1)–O(2)) in the two structures are slightly different. The dihedral angles are 93.956, 87.374 and 92.182° for complex 1 and 91.654, 89.335 and 91.172° for complex 5, respectively for planes O(1)–Ru(1)–O(2)/N(1)–Ru(1)–N(2), O(1)–Ru(1)–O(2)/N(3)–Ru(1)–N(4) and N(1)–Ru(1)–N(2)/N(3)–Ru(1)–N(4).
![]() | ||
Fig. 3 A view of the structural overlap of cationic part of 1 and 5, having an RMSD deviation of 0.024 Å. Hydrogen atoms are omitted for clarity. |
Fig. 4 displays the energy levels and the isosurface contour plots of the selected frontier molecular orbitals for complex 1. The electronic structure of complexes 2–5 is very similar to the one calculated for complex 1 (see Fig. S3–S6†). In all of them the HOMO–LUMO gap is ranging between 3.10 and 3.13 eV (see Fig. S2†). In compound 1 (as a representative example) the HOMO is contributed by the orbitals of the Ruthenium atom (50.7%) and the salicylaldehyde ligand (39.3%) while the LUMO and LUMO+1 are mainly spread over the bpy ligands (see Table S1†) in a similar manner as it has been described for similar complexes47 or related complexes of ruthenium with bpy ligands and a chelating oxygen donor ligand.48,49
TD-DFT calculations have been performed to explore the nature of the low-lying singlet and triplet states with the geometries of the ground state. Tables 3 and S2† summarize the calculated excited states. For complex 1, the absorption in the experimental spectra (Fig. 5A) appeared at 492 nm is assigned to the singlet excited state S8 (445.5 nm) and it is mainly a double transition from the HOMO−2 to the LUMO and to the LUMO+1 with a calculated oscillator strength of 0.1034 corresponding to a Metal-to-Ligand Charge Transfer (1MLCT) from the ruthenium center to the bpy ligands. Lower energy singlet excited states displayed very low values of the oscillator strength. In the same complex, the signal appeared at 363 nm is assigned to the singlet excited state S20 (343.7 nm, with a calculated oscillator strength of 0.0705), which is mainly a double transition from the HOMO−2 to the LUMO+3 and from the HOMO−1 to the LUMO+5. Both HOMO−2 and HOMO−1 show a high participation of atomic orbitals of the ruthenium center, and the LUMO+3 and the LUMO+5 are centered on the bpy ligands, therefore these transitions can be described as a Metal-to-Ligand charge Transfer (1MLCT). Similar results can be observed with complexes 2–5 as both the experimental absorption spectra and the calculated electronic structure of all of them are very close.
![]() | ||
Fig. 5 (A) Absorption spectra of 20 μM of all complexes in DMSO. (B) Fluorescence spectra of all complexes (120 μM) in water at λex = 403 nm. |
Complex | Estate | Energy (eV) | λ (nm) | f.osc. | Monoexcitations | Nature | Description |
---|---|---|---|---|---|---|---|
a Vertical excitation energies (E), dominant monoexcitations with contributions (within parentheses) of >15%, the nature of the electronic transition, and the description of the excited state are summarized. | |||||||
1 | S1 | 2.174 | 570.3 | 0.0067 | HOMO → LUMO (82) | dπ(Ru) + πsal → π*bpy | 1MLCT/1LLCT |
S2 | 2.177 | 569.6 | 0.0041 | HOMO → LUMO+1 (74) | dπ(Ru) + πsal → π*bpy | 1MLCT/1LLCT | |
S3 | 2.215 | 559.8 | 0.0168 | HOMO−1 → LUMO (81) | dπ(Ru) → π*bpy | 1MLCT | |
S 8 | 2.782 | 445.5 | 0.1034 | HOMO−2 → LUMO (56) | dπ(Ru) → π*bpy | 1MLCT | |
HOMO−2 → LUMO + 1 (27) | dπ(Ru) → π*bpy | 1MLCT | |||||
S 20 | 3.607 | 343.7 | 0.0704 | HOMO−2 → LUMO + 3 (55) | dπ(Ru) → π*bpy | 1MLCT | |
HOMO−1 → LUMO + 5 (27) | dπ(Ru) → π*bpy | 1MLCT | |||||
T1 | 1.784 | 695.0 | HOMO−1 → LUMO + 2 (93) | dπ(Ru) → π*sal | 3MLCT | ||
T2 | 1.796 | 690.2 | HOMO−1 → LUMO (54) | dπ(Ru) → π*bpy | 3MLCT | ||
HOMO → LUMO (27) | dπ(Ru) + πsal → π*bpy | 3MLCT/3LLCT | |||||
T3 | 1.842 | 673.0 | HOMO → LUMO + 1 (77) | dπ(Ru) + πsal → π*bpy | 3MLCT/3LLCT | ||
T4 | 1.953 | 634.7 | HOMO → LUMO + 2 (73) | dπ(Ru) + πsal → π*sal | 3MLCT/3LC | ||
T5 | 1.989 | 623.5 | HOMO−1 → LUMO (27) | dπ(Ru) → π*bpy | 3MLCT | ||
HOMO−1 → LUMO + 1 (38) | dπ(Ru) → π*bpy | 3MLCT | |||||
HOMO → LUMO (29) | dπ(Ru) + πsal → π*bpy | 3MLCT/3LLCT |
The fluorescence emission spectra of all compounds are reported in Fig. 5B. The emission spectra of these complexes display important differences. Complexes bearing two halogen atoms on the salicylic ring (2, 4 and 5) feature a non-structured emission band at about 520 nm, while complexes with only one halogen atom (1 and 3) do not show any band in the same region. The first five calculated triplet excited states are reported in Tables 3 and S2.† In complexes 1 and 3, the triplet T1 lies 0.39 eV and 0.37 eV respectively lower than the corresponding singlet S1. This difference in the energy is bigger in the complexes featuring a second halogen atom on the salicylaldehyde ligand, with differences around 0.50 eV in these complexes. In complexes with only one halogen atom on the salicylaldehyde ligand (1 and 3) the triplets T1 and T3 are closer in energy (separated by 0.058 eV and 0.059 eV respectively) than those in the complexes with two halogen atoms in the salicylaldehyde ligand (with differences in the energy of 0.146 eV, 0.157 eV and 0.152 eV for 2, 4 and 5 respectively). In all cases, the first five triplet excited states are results of metal-to-ligand charge transfer from the ruthenium center to the bpy or to the salicylaldehyde ligands, along with some ligand–ligand charge transfer and ligand centered character. The excited states that display this ligand–ligand or ligand centered character are those in which there is an important participation of the HOMO as this molecular orbital is composed mainly of orbitals belonging to the ruthenium atom and to the salicylaldehyde ligand (see Table S2†). The lowest energy triplet state, T1, is described for all complexes as a Metal-to-Ligand Charge Transfer (3MLCT) from the ruthenium center to the salicylaldehyde ligand with the only one exception of complex 3 in which the same transition corresponds to the calculated exited state T3, with T1 for complex 3 being a combination of transitions from the HOMO to the LUMO and to the LUMO+1 (see Table S2†) that can be described as a 3MLCT from the ruthenium to the bpy ligands. The strong component of 3MLCT in the calculated triplet excited states is in good accordance with the observed unstructured band at around 520 nm observed for complexes 2, 4 and 5.
The geometries of the lowest triplet states T1 and T2 of complexes 1–5 were optimized using the spin-unrestricted DFT approach. After this geometry relaxation, the differences in the energy of each T1 and T2 state with their related S0 are calculated (adiabatic energy differences) and gathered in Table 4. In all cases, the calculated energy values are underestimated. The optimized triplet excited state of the lowest energy displays a spin-density distribution that supports the description of the TD-DFT calculations. For complexes 1, 2, 4 and 5, the spin-density distribution of the optimized T1 state is spread mainly over the salicylaldehyde ligand and the ruthenium center with values of spin densities for complex 1, as a representative example, of 0.77e for Ru and 1.20e for the salicylaldehyde ligand, in good agreement with the TD-DFT calculated T1 excited state (see Fig. 6 and Fig. S7† and Table 4 and Table S3†). Similarly, the second triplet state optimized is mainly spread over the bpy ligand and the ruthenium atom (0.84e for Ru and 1.01 for bpy ligand in complex 1 as example), in good agreement with the TD-DFT calculated T2 excited state. In the case of complex 3, the optimized T1 state displays an analogous spin-density distribution to the one described for the other complexes, but this state is represented by the excited state T3 in the TD-DFT calculations, as said above. As can be seen in Table 4 the difference in the energy between T1 and T3 for complex 3 is only 0.059 eV, and probably the apparent disorder of the triplet states in this complex arises from the fact that the triplet states calculated by TD-DFT are obtained over the geometry of the singlet ground state and the geometry of the triplet states is slightly modified.
![]() | ||
Fig. 6 Spin-density contours (0.0008 a.u.) calculated for fully relaxed T1 (left) and T2 (right) states of complexes 1 (up) and 3 (down). |
1 | 2 | 3 | 4 | 5 | |
---|---|---|---|---|---|
Emission (experimental) | — | 2.38; 522 | — | 2.39; 518 | 2.37; 524 |
Adiabatic T1–S0 | 1.68; 738 | 1.63; 761 | 1.67; 742 | 1.60; 775 | 1.58; 784 |
Adiabatic T2–S0 | 1.82; 680 | 1.85; 669 | 1.81; 686 | 1.83; 676 | 1.84; 672 |
To confirm this hypothesis, the evolution of complexes 1–5 in solution was monitored using NMR 1H spectroscopy. Initially, the stability of the complexes in DMSO was verified as no changes in the compounds were observed for 120 hours at room temperature (see Fig. S12A–E†). The influence of the presence of water in these solutions was studied by adding a small amount of deuterated water to the samples. As shown in Fig. S13A–E,† the complexes bearing two halogen atoms on the salicylaldehyde ligand (2, 4 and 5) undergo slight decomposition because we find the appearance of the signal of the free ligand at about 10 ppm. This decomposition process seems to be much slower in the complexes with only one halogen on the salicylaldehyde ligand (1 and 3), and can be attributed to the different steric hindrance of the halogen atom in comparison with the hydrogen atom.
IC50 (μM) | SI = IC50, Hek293/IC50, A2780 | ||||
---|---|---|---|---|---|
A549 | SW480 | A2780 | Hek293 | ||
CDDP | 3.5 ± 0.6 | 5.1 ± 0.6 | 4.0 ± 0.6 | 2.0 ± 0.3 | 0.5 |
ClSal | >50 | >50 | >50 | >50 | — |
BrSal | >50 | >50 | >50 | >50 | — |
Cl 2 Sal | >50 | >50 | >50 | >50 | — |
BrClSal | >50 | >50 | >50 | >50 | — |
Br 2 Sal | >50 | >50 | >50 | >50 | — |
ByRu | >50 | >50 | >50 | >50 | — |
1 | 6.5 ± 0.9 | 2.5 ± 0.5 | 2.7 ± 0.3 | 8.4 ± 0.4 | 3.1 |
2 | 1.3 ± 0.3 | 1.5 ± 0.2 | 0.8 ± 0.2 | 2.0 ± 0.1 | 2.5 |
3 | 7.9 ± 0.8 | 5.5 ± 0.5 | 3.3 ± 0.6 | 11.0 ± 1.0 | 3.3 |
4 | 2.8 ± 0.4 | 1.7 ± 0.3 | 0.7 ± 0.2 | 4.3 ± 0.6 | 6.1 |
5 | 2.3 ± 0.2 | 1.5 ± 0.2 | 0.8 ± 0.1 | 1.6 ± 0.5 | 2 |
Neither the ligands nor the metallic fragment (ByRu) are cytotoxic compounds. Interestingly, these Ru(II) complexes are less cytotoxic than previously studied Cu(bpy)(X-sal)(NO3) (X = Cl, Br, I or H).40 The cytotoxic potential of the Ru(II) complexes is influenced by both the number and the types of halogens in the salicylaldehyde ligand. Regarding the number of halogens, two exhibited higher cytotoxicity than one, and regarding the types of halogens, Br rendered the complex more cytotoxic than Cl. Interestingly, these results differ from those obtained for the Cu family where the monohalogenated complexes and those complexes bearing Cl as the halogen are the most cytotoxic derivatives.
On the other hand, if the selectivity index (SI) is considered, 4 is the most promising complex of the series since it displays the highest selectivity towards ovarian cancer cells (SI = 6.1). It seems that the increase in the cytotoxicity achieved with additional Br (in the position X of Scheme 1) is associated with a decrease in the selectivity of the Ru(II) complexes. Then, the cellular uptake of the Ru(II) metal complexes was studied by means of ICP-MS experiments. The collected results (Fig. 7) show that all the complexes are more internalized with cisplatin being the halogen key for the cellular accumulation since the complexes bearing Cl are less internalized than the Br-complexes.
![]() | ||
Fig. 7 Cellular accumulation of the Ru(II) complexes in A549 cells treated with 2 μM of the studied complexes during 24 h. CDDP is included for comparison purposes. |
In order to shed some light on the mechanism of action of these Ru(II) complexes, images of the A549 cells treated with 10 μM of the Ru(II) complexes at different incubation times were recorded. After 17 h of treatment, important morphological changes compatible with apoptosis such as cell shrinkage, cytoplasmic vacuolization, membrane blebbing, and apoptotic body formation were observed (Fig. S14†).
Once apoptosis induction is confirmed, we evaluate the cell cycle distribution of A549 cells treated with the Ru(II) complexes at their respective IC50 values by flow cytometry. An increase in the G0/G1 population along with a decrease in the percentage of cells in the S phase is observed due to the treatment with all the Ru(II) derivatives (Fig. 9). As for halogenation, it can be observed that Br induces a greater increase in the percentage of cells in G0/G1 than Cl. In addition, there is a reduction in the G2/M population for all the complexes except for monohalogenated 3.
Since both apoptosis and cell cycle arrest at G0/G1 may be a consequence of an increase in reactive oxygen species (ROS) levels, the ability of these Ru(II) complexes to induce ROS production was investigated by fluorescence measurements with the probe 2′-7′-dichlorofluorescein diacetate (DCFH-DA).51 DCFH-DA gets into the cells by passive diffusion where it is hydrolyzed by esterases. Then, it is oxidized by ROS to yield fluorescent dichlorofluorescein. The fluorescence of the cells treated with the half maximal inhibitory concentration of the Ru(II) complexes was collected after 4 h of treatment. The variation in the emission intensity with respect to the untreated cells (corrected by the number of cells) is plotted in Fig. 10A and reflects the intracellular ROS levels. The halogen seems to influence ROS production since the presence of Br is essential for ROS generation as complexes without Br are not able to produce ROS. In addition, among the complexes bearing Br, the monohalogenated complex 1 is the least efficient as the ROS generator suggesting that the position of Br is a key issue for their biological activity.
Mitochondrial membrane potential (MMP) is a key indicator of the mitochondrial bioenergetic state since it is the driving force for ATP production. On the other hand, mitochondrial activity is a prime source of endogenous ROS production and mitochondrial dysfunction may be responsible for the observed increase in ROS levels.52 Therefore, the effect of the Ru(II) complexes on the MMP was evaluated by means of the tetramethyl rhodamine methyl ester (TMRM) probe. The complexes bearing Br in the position X of Scheme 1 suffer mitochondrial depolarization since a substantial decrease in their MMP in comparison with untreated cells is observed (Fig. 10B).
The stability studies in solution have revealed that these complexes undergo decomposition with the release of the salicylaldehyde ligand being faster in dihalogenated than in monohalogenated complexes. In addition, it has been observed that halogenation is an important factor in the cytotoxicity of these complexes and in their mechanism of action. Indeed, dihalogenated complexes exhibit higher cytotoxicity than monohalogenated complexes and the type of halogen plays an important role. On the one side, complexes bearing chloride are more selective towards cancer cells than complexes bearing bromide. On the other side, bromide as an halogen is better than chloride in terms of half maximal inhibitory concentration in cancer cells. In fact, the presence of halogen is a decisive issue for their anticancer activity since complexes with Br are more internalized than complexes with Cl in cancer cells. Moreover, complexes bearing Br in the X position are not only the most cytotoxic complexes but also the most efficient derivatives in ROS generation along with an enhanced mitochondrion depolarization. In fact, oxidative stress and mitochondrial dysfunction seem to be the mechanism of action for this series of Ru(II) complexes.
Single-crystal X-ray diffraction data for 5 were collected at 160(1) K on a Rigaku OD XtaLAB Synergy, Dualflex, Pilatus 200 K diffractometer using a single wavelength X-ray source (Cu Kα radiation: l = 1.54184 Å) from a micro-focus sealed X-ray tube and an Oxford liquid-nitrogen Cryostream cooler. The selected suitable single crystal was mounted using polybutene oil on a flexible loop fixed on a goniometer head and immediately transferred to the diffractometer. Pre-experiments, data collection, data reduction and analytical absorption correction61 were performed with the program suite CrysAlisPro.62 Using Olex2,63 the structure was solved with the SHELXT64 small molecule structure solution program and refined with the SHELXL2018/3 program package65 by full-matrix least-squares minimization on F2. PLATON66 was used to check the result of the X-ray analysis. A solvent mask67 was used in Olex2 for structure 5 to take into account the residual electron density attributed to disordered solvent molecules of ethanol. Although they are not present in the final model, the formula moiety and the formula sum include the atoms of those molecules (two solvent molecules per cell) leading to many alerts in the checkCIF report.
For both structures, all hydrogen atoms were added at the ideal positions and constrained to ride on their parent atoms. Crystallographic data are listed in Table 1. Selected bond distances and angles are summarized in Table 2. More details concerning both crystal structures and their refinements can be found in the corresponding CIF files (ESI†).
For NMR 1H measurements, 5 mg of the compound was dissolved in 0.5 ml of DMSO-d6. The NMR 1H spectra were regularly registered in order to test the variations of the spectra. The influence of the presence of water was studied adding 0.1 ml of D2O to the samples and then, the evolution of the samples with time was monitored registering the NMR 1H spectra.
In all the experiments, Cisplatin (CDDP) was included as positive control for comparison purposes.
We are indebted to M. Mansilla (PCT of the Universidad de Burgos) for the technical support.
This research has made use of the high-performance computing resources of the Castilla y León Supercomputing Center (SCAYLE, https://www.scayle.es), financed by FEDER (Fondo Europeo de Desarrollo Regional).
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2150468 and 2108989. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d2dt00401a |
This journal is © The Royal Society of Chemistry 2022 |