Gennaro
Sanità‡
a,
Maria Laura
Alfieri‡
b,
Barbara
Carrese
c,
Serena
Damian
b,
Vincenza
Mele
c,
Gaetano
Calì
d,
Brigida
Silvestri
e,
Sebastiano
Marra
e,
Susan
Mohammadi
f,
Giuseppina
Luciani
g,
Paola
Manini
*b and
Annalisa
Lamberti
*c
aInstitute of Applied Sciences and Intelligent Systems Unit of Naples, National Research Council, Naples, Italy
bDepartment of Chemical Sciences, University of Naples Federico II, Naples, Italy. E-mail: pmanini@unina.it
cDepartment of Molecular Medicine and Medical Biotechnology, University of Naples Federico II, Naples, Italy. E-mail: annalisa.lamberti@unina.it
dInstitute of Endocrinology and Experimental Oncology, National Research Council, Naples, Italy
eDepartment of Civil, Architectural and Environmental Engineering, University of Naples Federico II, Naples, Italy
fScuola Superiore Meridionale, Naples, Italy
gDepartment of Chemical, Materials and Industrial Production Engineering, University of Naples Federico II, Naples, Italy
First published on 25th October 2024
In this work, we report on the synthesis and properties of a new sensitizer for photodynamic therapy applications, constituted by a ruthenium(II) complex (1) featuring a ligand inspired from natural isoquinoline alkaloids. The spectroscopic analysis revealed that 1 is characterized by an intense red emission (λem = 620 nm, Φ = 0.17) when excited at 550 nm, a low energy radiation warranting for a safe therapeutic approach. The phototoxicity of 1 on human breast cancer (Hs578T) and melanoma (A375) cell lines was assessed after irradiation using a LED lamp (525 nm, total fluence 10 J cm−2). In vitro biological assays indicated that the cytotoxicity of 1 was significantly enhanced by light reaching IC50 values below the micromolar threshold. The cell damage induced by 1 proved to be strictly connected with the overproduction of reactive oxygen species (ROS) responsible for mitochondrial dysfunction leading to the activation of caspases and then to apoptosis, and for DNA photocleavage leading to cell cycle arrest.
Known photosensitizers currently used for clinical applications are based on tetrapyrrolic moieties, including for example porphyrins, chlorins and phthalocyanines.18–21 However, these compounds share similar drawbacks including complications in the synthesis and purification, low water solubility, weak photostability, and lack of tumor tissue selectivity aside from poor tissue penetration of shorter-wavelength visible light, which limit their potential use.22–25 Consequently, different attempts have been made over the years to improve the effectiveness of the PDT treatment. Among others, the use of transition metal complexes26–30 and more specifically of ruthenium(II) complexes featuring polypyridyl ligands31–38 was found to be a valuable alternative for PDT owing to the outstanding photophysical properties of ruthenium(II) complexes.31,34,39,40
The great potential of ruthenium(II) complexes relies also on: 1) the multiple mechanisms through which they can enter cells, such as passive diffusion, active transport and endocytosis; 2) the different cellular targets they can address, such as cell nuclei, mitochondria or lysosomes; 3) the way they can damage and kill cancer cells, i.e. via DNA intercalation, protein interaction and ROS production; 4) the low systemic toxicity and selective antimetastatic properties.41–44 Overall, this evidence prompted the design of different kinds of ruthenium(II) complexes to look for the best performing PS for PDT application. This kind of investigation is possible considering that the photophysical and chemical properties of ruthenium(II)-based complexes (e.g., charge, solubility, ligand conformations and metal- and ligand-based redox potentials) can be finely tuned through the proper selection of the ligand–metal combination and designing appropriate geometries for specific interactions with biological targets, which thus makes them attractive for photobiological applications.45–48
In this connection, we recently explored the possibility of taking advantage of biologically relevant ligands to build bio-inspired transition metal complexes.49 By pursuing this strategy, the PS may be easily recognized by the cellular environment and can accumulate in the tissues. Previous studies have demonstrated, for example, the high and selective cytotoxicity of ruthenium(II) complexes experiencing the presence of β-carboline ligands. By taking advantage of the presence of a heterocyclic platform common to many natural and synthetic alkaloids,50,51 the β-carboline ligand can camouflage the metal and foster the diffusion of the ruthenium(II) complex into the cellular target via specific receptors. By this way, high and selective cytotoxicity and antitumor activity against various cancer cells through multiple mechanisms (i.e. interfering with DNA synthesis, inhibiting DNA topoisomerases I and II) have been reached.52–54
By pursuing a similar approach, herein we report on the synthesis and characterization of the photo-physical properties of a ruthenium(II) complex (1) featuring a ligand inspired from natural isoquinoline alkaloids (Fig. 1). On the basis of literature data reporting on the use of PDT as an excellent alternative in the treatment and diagnosis of breast cancer and melanoma compared to the conventional surgery, chemotherapy and radiotherapy,55,56 we have assayed for the first time the light-promoted enhanced cytotoxicity of 1 on the human breast carcinoma cell line (Hs578T) and human melanoma cell line (A375) to assess the potential use of the complex as an efficient PS.
The identity of both the amide 2 and the ligand 3 was confirmed by NMR spectroscopy and MALDI mass spectrometry (Fig. S1–S3†).
The complex 1 was synthesized according to a procedure reported in the literature.54 A suspension of cis-Ru(bpy)2Cl2 in a mixture of ethanol and water (1:
1 (v/v)) was treated under an argon atmosphere with the ligand 3 and triethylamine under reflux conditions (Scheme 2). After 3 h, the reaction mixture was cooled down and treated with a water solution of NH4PF6. The formation of a dark-red solid was observed; this latter was collected by centrifugation to give the complex 1 in good yield (74%).
Mono- and bidimensional NMR spectroscopy (Fig. S5–S10†) was carried out to confirm the structure of the complex 1.
In particular, the 1H and 1H,1H COSY spectra displayed a series of signals both in the aromatic and aliphatic proton regions. Four multiplets at 3.46, 2.82, 2.73 and 2.41 ppm were ascribable to the methylene protons of the isoquinoline unit along with the four singlets at 3.93 and 3.91 ppm, relative to the proton of the methoxyl group, and at 7.56 and 6.85 ppm due to the aromatic protons. The spectrum was completed by the signals of two 2,2′-bipyridine units along with the signals of the protons of the phenyl residue. The 13C, 1H,13C HSQC and 1H,13C HMBC spectra allowed for the complete assignment of all the resonances, definitely supporting the structural identification of the complex 1. In good agreement were the data obtained from the MALDI mass spectrometry analysis, revealing the presence of the peaks of the two ionic portions in the positive (m/z 680, [Ru(bpy)23]+) and negative (m/z 145, [PF6]−) ion modes (Fig. S12 and S13†). The complex 1 exhibited good solubility in PBS at the micromolar concentration, unlike the parent compound Ru(bpy)3(PF6)2, suggesting its potential use for biological assays and therapeutic purposes.
UV-vis λmax, nm (log![]() |
UV-vis λmax,d nm (log![]() |
PL λem, nm (λecc, nm) | Φ%e (λem, nm) | PL λem,d,e nm (λecc, nm) | |
---|---|---|---|---|---|
a Determined in diluted solutions (1 × 10−5 M). b Determined in CH2Cl2. c Determined in ACN. d Determined in PBS solution (pH 7.4). e Determined relatively to fluorescein (Φ = 0.9 in a 0.1 M solution of NaOH). | |||||
1 | 294 (4.87), 352 (4.05), 430 (3.96), 504 (4.07), 558 (3.97), 606 (sh) | 290, 341, 439, 482, 550 (sh) | 620 (550) | 0.17 | 665 (550) |
Ru(bpy)3(PF6)2c | 288 (4.98), 390 (sh), 426 (sh), 451 (4.17) | Low solubility | 615 (451) | 0.09 | Low solubility |
The UV-vis spectrum of 1 (Fig. 2) displayed the typical profile of ruthenium(II) complexes with polypyridyl ligands (see data from Ru(bpy)3(PF6)2). In detail, it has been possible to note the presence of: a) an intense maximum (logε > 4 M−1 cm−1) below 300 nm assigned to the spin-allowed π–π* ligand centred (LC) transitions (see the UV-vis spectrum of 3 for analogies); b) a series of intense maxima (log
ε ≈ 4 M−1 cm−1) in the range of 300–500 nm assigned to metal-to-ligand charge transfer (1MLCT) transitions; c) weak and broad absorption maxima at longer wavelengths (>500 nm) ascribable to spin forbidden metal-to-ligand charge transfer (3MLCT) transitions. The absorption maxima of 1 proved to be red shifted with respect to those from Ru(bpy)3(PF6)2 as a consequence of the reduction of the HOMO–LUMO gap due to the higher electron-releasing properties of the C^N ligand via σ-donation from the carbanion with respect to 2,2′-bipyridine (N^N).57
![]() | ||
Fig. 2 UV-vis (blue trace) and emission (red trace) spectra of 1 in CH2Cl2 (solid line) and in PBS at pH 7.4 (dotted line). |
The complex 1 exhibited a red emission with a maximum set at 620 nm and an emission quantum yield of Φ = 0.17 (relative to fluorescein) comparable to those reported in the literature for similar complexes and higher than the parent compound Ru(bpy)3(PF6)2.
The photophysical properties of 1 have been investigated also in phosphate buffer saline (PBS) at pH 7.4; the aim was to assess if any change in the optical properties of 1 may occur when incubated in the medium for cell culture biological assays and to evaluate the stability of the complex with time.
As shown in Fig. 2, the absorption profile of 1 showed a slight reduction of the intensity of some maxima, such as the one at 558 nm, and a significant red shift of the emission maximum (from 620 to 665 nm). Under these conditions, the complex 1 also proved to be stable over 24 h incubation time. These data have been taken into account for the design of the biological assays. A spectrophotometric approach was also pursued to evaluate the hydrophobic/hydrophilic character of the ruthenium complex. The logP value measured for 1 proved to be −2.06 suggesting a moderate hydrophilic character.
![]() | ||
Fig. 3 Schematic representation of the experimental procedure followed to assess the phototoxicity of the ruthenium complex 1 on Hs578T or A375 cell lines. |
The irradiation was performed by using an Arkeo 96 led well unit provided by Cicci Research s.r.l., a portable device constituted by an array of 96 LEDs with 12 independent lines, fitting with the 96 well plates used for the biological tests. All the irradiation sessions were carried out by using a total fluence of 10 J cm−2.
After the incubation, the metabolic functionalities of the cells were evaluated through a combined multiparametric approach based on the following assays/analyses: 1) the CellTiter-GLO® and Live Cell Explorer® assays to evaluate the vitality of cells; 2) the flow cytometry and confocal microscopy analyses to assess the cellular uptake of the complex; 3) the propidium iodide assay to evaluate the onset of apoptosis; 4) the Caspase-Glo® assay to estimate the activity of caspases; 5) the western blot analysis to evaluate the expression of the anti-apoptotic protein Bcl-xL and the pro-apoptotic protein Bak; 6) the ROS-Glo® H2O2 assay to measure the amount of intracellular ROS; 7) the agarose gel electrophoresis analysis to check for the ability of 1 to induce the DNA photocleavage.
Both assays were carried out on Hs578T and A375 cancer cell lines (Fig. 4) and on a mammary breast fibrocystic disease cell line (MCF10a), selected as the non-tumoral control cell line (Fig. S14†).
The assays were performed after incubation for 24 h with different amounts of a PBS solution of 1 under dark conditions or after irradiation at 525 nm.
The results reported in Fig. 4 and in Fig. S16† showed that the complex 1 exhibited a dose-dependent toxicity that proved to be significantly enhanced by irradiation. As a matter of fact, the CellTiter-GLO® assay revealed that the half-maximal inhibitory concentrations (IC50) of 1, set at 1.1 and 1.4 μM under dark conditions in the case of the Hs578T and A375 cell lines, drop down at 0.5 and 0.6 μM after irradiation (Fig. 4A and B). These data, also supported by the results from the Live Cell Explorer assay (Fig. 4C and D), indicated that the cytotoxic effect of the ruthenium complex 1 is enhanced by light reaching IC50 values below the micromolar threshold, quite lower with respect to cisplatin (IC50 = 30.3 and 4.9 μM in the case of Hs578T58 and A375 (ref. 59) cell lines, respectively) and other similar ruthenium complexes.
Similar results were also observed in the case of the MCF10a cell line with IC50 = 1.2 μM under dark conditions and IC50 = 0.4 μM after irradiation (Fig. S14 and S15†). These data indicated that the cytotoxic effect of 1 was not selective toward tumoral cells, suggesting that the application/accumulation of the photosensitizer on the damaged tissues was necessary before irradiation.
Finally, control experiments carried out in the absence of the photosensitizer revealed that the irradiation step did not cause loss in cell viability, in agreement with Lifshits.60 Overall, these preliminary data suggested the potential use of 1 as a photosensitizer for local therapeutic applications.
As detailed in the next sections, to delineate the mechanism of action of 1, other biological assays have been carried out.
The results showed a concentration dependent uptake in both cell lines already after 3 h of incubation, while no increase was observed after 6 h of incubation. In particular, in Hs578T cells mean fluorescence intensity (MFI) increases of about 5 and 12 were observed by using 1 μM and 10 μM, respectively, whereas in A375 cells an MFI increase of about 10 and 30 was detected by using 1.5 μM and 15 μM, respectively (Table 2). The results suggest a quick cellular uptake and a very good cell membrane permeability of the complex 1.
Time (h) | Hs578T | A375 | ||
---|---|---|---|---|
MFI (a.u.) | MFI (a.u.) | |||
3 | Control | 2.56 ± 0.12 | Control | 0.85 ± 0.04 |
3 | 1 (1 μM) | 14.16 ± 0.87 | 1 (1.5 μM) | 9.2 ± 0.01 |
3 | 1 (10 μM) | 35.38 ± 4.15 | 1 (15 μM) | 29.4 ± 1.37 |
6 | Control | 2.85 ± 0.33 | Control | 0.63 ± 0.07 |
6 | 1 (1 μM) | 14.82 ± 0.95 | 1 (1.5 μM) | 7.24 ± 0.34 |
6 | 1 (10 μM) | 35.33 ± 2.1 | 1 (15 μM) | 16.7 ± 1.56 |
For confocal microscopy analysis, both cell lines were incubated with the complex 1. After 6 h of incubation, the images revealed the presence of the complex in the perinuclear area confirming that the cellular uptake occurred (Fig. 5). Furthermore, in good agreement with the data obtained by flow cytometry analysis, the red fluorescence signal from 1 proved to be more intense in the case of Hs578T cells than in A375 cells.
![]() | ||
Fig. 5 Representative confocal microscopy images of Hs578T cells (A) and A375 cells (B). Cell nuclei were stained with Hoechst 33342, complex 1 is visible as a red color. Scale bars: 5 μm. |
To clarify the mechanism of cell apoptosis induced by 1, the activities of caspase-9 and caspase-3/7 were examined by using specific luminogenic substrates (Caspase-Glo® assay). Caspase-9 is the initiator caspase related to the intrinsic or mitochondrial pathway of apoptosis. When activated, caspase-9 cleaves and activates downstream effector caspase-3 and -7, thus resulting in apoptosis.61 To evaluate caspase activation, Hs578T and A375 cell lines were incubated with 1 and analyzed both under dark conditions and after 4.5 h post irradiation (this incubation time was chosen taking into account that the activation of caspases in the apoptotic pathway is an early event).62–64 Under all tested conditions, both cell lines showed an increase of the caspase activity in a dose-dependent manner after irradiation, compared to dark conditions. The caspase 3/7 activity increased after irradiation, with an enhancement of about 2 and 2.8 times in Hs578T cells (Fig. 7A) with the concentration of 1 set at 0.5 and 1 μM, respectively, and of about 2.2 and 3.2 times in A375 cells (Fig. 7B) with the concentration of 1 set at 1 and 1.5 μM, respectively.
![]() | ||
Fig. 7 Caspases 3/7 and 9 activities. Caspase-GLO® 3/7 assay in Hs578T cells (A) and in A375 cells (B); Caspase-GLO® 9 assay in Hs578T cells (C) and in A375 cells (D). *** P ≤ 0.001. |
The caspase-9 activity proved to be significantly enhanced in Hs578T cells (Fig. 7C), with an increase of about 7.2 and 8.4 times after administration of 1 (0.5 and 1 μM, respectively), and to a lesser extent in A375 cells (Fig. 7D), with an increase of about 2.2 and 2.3 times after incubation of 1 (1 and 1.5 μM, respectively). In both cases, no increase of caspase activity was observed when the incubation was carried out under dark conditions. These data suggest that an intrinsic apoptosis pathway could be photoactivated in both tested cell lines after incubation with the ruthenium complex.
The Bcl-2 proteins are involved in the regulation of apoptosis through the control of mitochondrial membrane permeability and the release of cytochrome c and/or Smac/Diablo. They include anti-apoptotic proteins, such as Mcl-1, Bcl-2, and Bcl-xL, and pro-apoptotic proteins, including Bax, Bad, Bak, Bid, and Bim.65
Therefore, the effects of 1 on the expression levels of the anti-apoptotic protein Bcl-xL and the pro-apoptotic protein Bak were evaluated by performing a western blot analysis in Hs578T and A375 cell lines treated with the ruthenium complex and irradiated at a wavelength of 525 nm. The results indicated that 1 is able to inhibit the expression of the Bcl-xL protein and stimulate the expression of the Bak protein, with an overall reduction of the Bcl-xL/Bak ratio (Fig. 8).
ROS are normally produced in normal metabolic living cells. At low concentrations, these species have an important role in cell signaling, while at high concentrations, they can become toxic due to the interaction with proteins, DNA, lipids, and other biological molecules affecting their functioning. To investigate the possible implication of 1 in the formation of ROS, the ruthenium complex has been administered to Hs578T and A375 cell lines and the amount of intracellular ROS was measured after light irradiation by the ROS-Glo® H2O2 assay.
The data reported in Fig. 9 show an increase of the ROS production after irradiation in both cell lines. The effect was more evident in the Hs578T cell line with a ROS production up to 1.8 times higher with respect to the control experiments, whereas in the case of the A375 cell line, the ROS production was 1.3 times higher compared to the control. In both cases, the results suggest that ROS-mediated pathways can trigger apoptosis.
![]() | ||
Fig. 9 ROS levels determined in Hs578T (A) and in A375 (B) cell lines after incubation with 1 and light irradiation. *** P ≤ 0.001. |
As shown in Fig. 10, the complex 1 does not induce significant DNA cleavage under dark conditions, whereas after irradiation a dose-dependent reduction of the supercoiled form is visible. This result indicates that the DNA cleavage is determined by photoinduced processes mediated by 1 and supports the possibility of using 1 as a photosensitizer to induce cellular toxicity by DNA damage.71
![]() | ||
Fig. 10 Photocleavage of pEGFP-N3 plasmid DNA induced by 1 under dark and light irradiation conditions. |
The investigation of the photophysical properties of the complex allowed us to assess that, both in dichloromethane and in PBS, low energetic radiation (a green light, λ = 525 nm) was necessary to promote the sensitization of the complex.
This represents an important requisite for a safer approach to the photodynamic treatment. A set of biological tests was carried out on two selected human cell lines, breast cancer (Hs578T) and melanoma (A375), to assess the phototoxicity of the ruthenium complex 1 and to investigate the mechanisms of cellular damage. What emerged is that 1 is cytotoxic for both cell lines, in a more effective way toward Hs578T, and that the cytotoxicity is enhanced by light reaching IC50 values below the micromolar threshold, quite lower with respect to cisplatin and other similar ruthenium complexes. Light irradiation proved to be fundamental in causing the overproduction of ROS, a key event triggering a sequence of processes leading to cell death (Fig. 11). The increase in the levels of the proapoptotic protein Bak and the decrease in the levels of the antiapoptotic protein Bcl-xL were also evident after the photoactivation of 1, resulting in the activation of caspase-3/7/9 and the onset of apoptosis. Finally, experiments carried out on pEGFP-N3 plasmid DNA showed the ability of 1 to promote the DNA photocleavage.
Overall, these results pointed out that 1 can explicate its phototoxicity via a combined action that leads to the impairment of the main cellular activities and finally to death, thus opening a way to its possible use in the treatment of cancer.
UV-visible and emission spectra have been recorded on Jasco V-560 and Jasco FP-750 instruments. Quantum efficiencies (Φ) have been calculated using fluorescein (Φ = 0.9 in a 0.1 M solution of NaOH) as a reference. 1H and 13C NMR spectra have been registered on a Bruker DRX (400 MHz) instrument. 1H,1H COSY, 1H,13C HSQC, 1H,13C HMBC and NOESY experiments have been run at 400.1 MHz using standard pulse programs. MALDI mass spectra have been recorded on an AB Sciex TOF/TOF 5800 instrument using 2,5-dihydroxybenzoic acid as the matrix. The spectra represent the sum of 15000 laser pulses from randomly chosen spots per sample position.
Elemental analysis has been performed on a Thermo Scientific Flash Smart V CHNS/CHNS instrument.
1H NMR (400 MHz, CDCl3) δ ppm: 7.83 (dd, J = 8.2, 2.3 Hz, 2H), 7.71 (t, J = 8.2 Hz, 1H), 7.59 (t, J = 8.2 Hz, 2H), 6.56 (bs, 1H), 4.31 (s, 3H, –OCH3), 4.28 (s, 3H, –OCH3), 4.12 (dt, J = 6.4, 6.0 Hz, 2H), 3.32 (t, J = 6.4 Hz, 2H).
A solution of the amide 2 (969 mg, 3.4 mmol) in 11.7 mL of a 3:
11 (v/v) mixture of ethanol/dichloromethane was treated under stirring with POCl3 (3.28 mL, 35 mmol) and kept under reflux. After 5 h, petroleum ether was added (8 mL), and the mixture was kept under reflux. After 12 h, the reaction mixture was filtered and the solid was rinsed with water and treated with an aqueous solution of K2CO3 until pH 10 was reached. Then, the mixture was extracted with chloroform and the organic layers were dried with anhydrous sodium sulphate, filtered, and evaporated under reduced pressure. The crude residue was purified by liquid chromatography on silica gel using dichloromethane/methanol 95
:
5 (v/v) as an eluent to give the pure 1-phenyl-6,7-dimethoxy-3,4-dihydroisoquinoline 3 (885 mg, 95%).
1H NMR (400 MHz, CDCl3) δ ppm: 7.61 (m, 2H), 7.40 (m, 3H), 6.78 (m, 2H), 3.82 (s, 3H, –OCH3), 3.80 (t, J = 7.28 Hz, 2H), 3.73 (s, 3H, –OCH3), 2.75 (t, J = 7.28 Hz, 2H). 13C NMR (100 MHz, CDCl3) δ ppm: 167.0, 151.3, 147.2, 138.5, 132.8, 129.6, 128.9, 128.2, 121.3, 111.9, 110.4, 56.2, 56.1, 47.2, 26.0. ESI+-MS: m/z 268.
1H NMR (400 MHz, CD2Cl2) δ ppm: 8.4–8.2 (m, 3H), 8.21 (d, J = 8.0 Hz, 2H), 8.02 (d J = 8.0 Hz, 2H), 7.94 (t, J = 8.0 Hz, 1H), 7.84–7.77 (m, 4H), 7.60 (d, J = 8.0 Hz, 1H), 7.55 (s, 1H), 7.4–7.3 (m, 2H), 7.3–7.2 (m, 2H), 6.9–6.8 (m, 2H), 6.81 (s, 1H), 6.51 (m, 1H), 3.92 (s, 3H, –OCH3), 3.90 (s, 3H, –OCH3), 3.46 (m, 1H), 2.8–2.7 (m, 2H), 2.37 (m, 1H). 13C NMR (100 MHz, CD2Cl2) δ ppm: 157.6, 156.6 (2C), 154.8 (2C), 151.3 (2C), 149.5, 148.7 (2C), 147.3, 136.1 (2C), 134.6, 133.6, 133.3, 131.6 (2C), 127.9, 126.8, 126.3, 126.2 (2C), 125.3, 122.8, 122.7 (4C), 121.5 (2C), 111.4, 110.5, 56.3, 55.9, 50.3, 28.1. MALDI-MS: m/z 680 ([M–PF6]+). Anal. calc. for C37H32F6N5O2PRu: C 53.88%, H 3.91%, N 8.49%, found: C 53.12%, H 3.88%, N 8.40%.
For the Live Cell Explorer assay, Hs578T and A375 were seeded into 96-well microplates at a density of 5 × 103 cells per well for Hs578T and 2.5 × 103 for A375 and, after 24 h, were incubated with 1 at two selected concentrations: 0.5 and 1 μM in the case of Hs578T and at 1 and 1.5 μM in the case of A375. The analysis was carried out by using the same experimental conditions described for the CellTiter-GLO® assay. Cell viability was assessed by fluorescence microscopy.
For confocal microscopy, cells (5 × 103 per coverslip) were plated on 10 mm glass coverslips placed on the bottom of a 24-well plate, allowed to attach for 24 h under normal cell culture conditions and then incubated with complex 1 at a concentration of 10 μM (Hs578T) and 15 μM (A375) for 6 h at 37 °C. The cells were washed with PBS, fixed in 2% formaldehyde for 10 min, and washed 3 times with PBS. The cell nuclei were then stained with Hoechst 33342 (Invitrogen, Carlsbad, CA, United States). The cells were then spotted on microscope slides and analyzed on an inverted and motorized microscope (Axio Observer Z.1) equipped with a 63×/1.4 Plan-Apochromat objective. The attached laser-scanning unit (LSM 700 4× pigtailed laser 405–488–555–639; Zeiss, Jena, Germany) enabled confocal imaging. For excitation, 405 and 555 nm lasers were used. Fluorescence emission was revealed using a main dichroic beam splitter and a variable secondary dichroic beam splitter. The signal in the red channel was acquired by means of a 640 long pass filter. Double staining fluorescence images were acquired separately using ZEN 2012 software in the red and blue channels at a sampling of 1024 × 1024 pixels, with the confocal pinhole set to one Airy unit and then saved in TIFF format.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4md00600c |
‡ G. S. and M. L. A. contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |