Ricardo G.
Teixeira†
a,
Lívia
Stenico
b,
Xavier
Fontrodona
c,
Isabel
Romero
c,
Radosław
Starosta
ad,
Maria João
Moreno
be,
Ana Isabel
Tomaz
a,
Lígia C.
Gomes-da-Silva
*be and
Andreia
Valente
*a
aCentro de Química Estrutural, Institute of Molecular Sciences, Departamento de Química e Bioquímica, Faculdade de Ciências, Universidade de Lisboa, Campo Grande, 1749-016 Lisboa, Portugal. E-mail: amvalente@ciencias.ulisboa.pt; ligia.silva@uc.pt
bCoimbra Chemistry Center – Institute of Molecular Sciences (CQC-IMS), University of Coimbra, 3004-535 Coimbra, Portugal
cDepartament de Química and Serveis Tècnics de Recerca, Universitat de Girona, C/M. Aurèlia Campmany, 69, E-17003 Girona, Spain
dFaculty of Chemistry, University of Wrocław, F. Joliot-Curie 14, 50-383, Wrocław, Poland
eDepartment of Chemistry, University of Coimbra, 3004-535 Coimbra, Portugal
First published on 29th October 2025
Cancer continues to pose a significant global health burden, prompting ongoing exploration of innovative therapeutic strategies. Ruthenium-based complexes have emerged as promising alternatives to platinum drugs due to their generally favorable pharmacological profiles. In this work, we report the synthesis and characterization of a novel series of fluorescent Ru(II)–cyclopentadienyl organometallic complexes of general formula [Ru(η5-C5H5)(NN)(Ph2P-CH2-pip-NBD)][PF6] (1–5), where NN represents a bipyridine or phenanthroline-based ligand and Ph2P-CH2-pip-NBD is a 4-nitrobenzo-2-oxa-1,3-diazole (NBD)-derived fluorescent phosphane conjugate. Structural characterization confirmed their piano-stool geometry via NMR, FTIR, UV-Vis, fluorescence spectroscopy, and X-ray crystallography (for 1, 3 and 5). The complexes exhibit notable stability in both organic and aqueous media. Cytotoxicity screening across three cancer cell lines (4T1 murine epithelial breast cancer, CT26 murine colon carcinoma, U2OS human osteosarcoma) and one non-cancerous line (3T3 murine embryonic fibroblasts) revealed that complexes 1, 3, and 5 display potent anticancer activity, particularly against U2OS. Fluorescence-based uptake and confocal microscopy demonstrated efficient internalization, primarily through caveolin-mediated endocytosis, and preferential accumulation in lysosomes. Enhanced fluorescence in acidic environments and co-localization with lysosomal markers confirm lysosomal tropism, highlighting the dual role of the NBD fluorophore for traceability and subcellular targeting. Additional mechanistic studies revealed that complexes 1, 3, and 5 induce oxidative stress and trigger apoptosis, suggesting that ROS generation contributes to their cytotoxic activity. These findings establish this class of compounds as promising lysosome-targeting agents.
Over time, ruthenium complexes have consistently shown encouraging results as the next prominent class of metal-based anticancer therapeutics.7 Until now, special interest in the preparation of ruthenium-based compounds has been supported by their excellent antitumor properties and good pharmacological profile both in vitro and in vivo,8–11 including their cytotoxicity in platinum-resistant cells12 and ability to reduce the formation of metastasis.13,14 However, and despite their promising anticancer properties, only three Ru(III) compounds (namely NAMI-A14 ([ImH][trans-Ru(Cl)4(Im)(DMSO)]; Im = imidazole and DMSO = dimethylsulfoxide), KP101915 and its soluble salt NKP133916 (KP1019 = [IndH][trans-RuCl4(Ind)2], Ind = indazole; NKP1339 = [Na][trans-RuCl4(Ind)2])), together with a Ru(II) photosensitizer compound (TLD1433, [Ru(Me2bipy)2(IP-3T)][Cl]2 (Me2bipy = 4,4′-dimethyl-2,2′-bipyridine; IP = imidazo[4,5 f][1,10]phenanthroline; 3T = α-terthienyl17)) have entered clinical trials.
Other classes of ruthenium anticancer agents, mostly based on the ruthenium–arene motif, have been developed. The Ru(II) arene scaffold has proven to be a very useful structure for the development of anticancer drugs due to its stability, solubility, and lipophilicity.18 Ru(II) arene complexes have an octahedral geometry, in which the arene ligand occupies three coordination positions and other ligands can bind to the three remaining coordination sites, giving a three-legged piano-stool type of structure. The combination of the ruthenium ion with arene ligands confers more hydrophobic character to the metal-based scaffold, promoting passive diffusion across the cell membrane and enhancing their cellular accumulation.19 In addition, the nature of the arene ligands assists in the cell uptake and the interaction with potential targets.20
In this frame, our research group has developed a library of organometallic compounds by exploring the chemistry of the ruthenium-(η5-cyclopentadienyl) (‘RuCp’) scaffold within the therapeutic context. We have focused on the general structure [Ru(Cp)(NN)(P)]+ (Cp = η5-cyclopentadienyl or Cp-derivatized; NN = N,N-heteroaromatic bidentate ligand (e.g., 2,2′-bipiridine or its derivatives); P = triphenylphosphane or its derivatives), which has shown remarkable anticancer activity against several human cancer cell lines, especially against resistant and more aggressive cancer cells.21 Yet, up to now, we have not explored the incorporation of a fluorescent label for traceability purposes. In this context, the 4-nitrobenzo-2-oxa-1,3-diazole (NBD) fluorophore emerges as a promising candidate, as it has already been employed in the derivatization of various therapeutic agents, including platinum-based complexes,22,23 as well as other small molecules. These studies demonstrate that NBD can be used to monitor cellular uptake, efflux, and intracellular distribution without significantly altering the biological activity of the parent compound. Notably, some of us have previously reported the synthesis of a phosphane derivative, 7-nitro-4-(4-(diphenylphosphinomethyl)piperazin-1-yl)-2,1,3-benzoxadiazole (Ph2P-CH2-pip-NBD), which has shown potential as a molecular probe for imaging living cells.24 This compound could conveniently replace triphenylphosphane in our complexes as exemplified by Florindo et al., who described half-sandwich ‘RuCp’ complexes bearing NBD-derived 2-pyridine and phosphane ligands with antimalarial activity.25 Beyond enabling traceability, this substitution facilitates lysosome targeting due to the preferential accumulation of such probes in acidic organelles. Moreover, targeting the endosomal/lysosomal pathway has emerged as a key target in cancer treatment strategies by promoting lysosomal membrane permeabilization, disrupting autophagy, and exploiting the acidic environment to enhance drug activation.26 Lysosomes contain a variety of hydrolases capable of degrading both intra- and extracellular components. Upon release into the cytosol, these enzymes can contribute to the breakdown of apoptotic proteins, ultimately triggering the mitochondrial apoptotic pathway and/or other cell death modalities.27 Such a mechanism of action has been observed for certain lysosome-targeting Ru(III) coordination compounds,28 as well as for some half-sandwich Ru(II)(η6-arene) complexes,29–33 with confocal microscopy playing a key role in confirming organelle-specific localization. However, no such reports exist for half-sandwich Ru(II)(η5-Cp) complexes, likely due to the absence of suitable conjugate structures – and thus the lack of intrinsic fluorescence – required for effective imaging.
In this study, we present the synthesis of a novel family of fluorescent ‘RuCp’ organometallic complexes with the general formula [Ru(η5-C5H5)(NN)(Ph2P-CH2-pip-NBD)][PF6] (1–5, Fig. 1) where Ph2P-CH2-pip-NBD is 7-nitro-4-(4-(diphenylphosphinomethyl-piperazin-1-yl)-2,1,3-benzoxadiazole) and NN represents a N,N-donor (bipyridine or phenanthroline). Ligands of type 4,4′-R-2,2′-bipyridine (R = CH3, OCH3, CH2OH, CH2Biotin) were selected for this study as an extension of our previous work, which indicated the possible role of some of them in surmounting multidrug resistance (MDR).34–37
![]() | ||
| Fig. 1 Chemical structures of compounds 1–5, including the numbering scheme for spectral assignments. All complexes were isolated as hexafluorophosphate salts. | ||
Solid state infrared spectra (KBr pellets) of the new complexes presented the characteristic stretching bands of the cyclopentadienyl (νC–H ∼3050 cm−1) and the phenyl aromatic rings (νC–H and νC
C at ∼3050–3100 cm−1 and ∼1430 cm−1, respectively) and the typical vibrational bands for PF6− anion (stretching νPF6 at 843 cm−1 and symmetric bending δPF6 at 558 cm−1, respectively). The successful introduction of the bidentate diimines and Ph2P-CH2-pip-NBD ligands into the coordination sphere of the metal was confirmed by the complete disappearance of the stretching band of the acetonitrile ligands (νC
N ∼2280 cm−1) present at the starting material [Ru(η5-C5H5)(NCCH3)3][PF6] and is fully supported by NMR data as well (Fig. S1–S15 of the SI). The overall analysis of the 1H and 31P{1H} NMR spectra of complexes 1–5 indicate the displacement of the three acetonitrile ligands in [Ru(η5-C5H5)(NCCH3)3]+ by the chosen N,N-heteroaromatic and phosphane co-ligands. In all cases, the resonances expected for the cyclopentadienyl, alkyl diphenylphosphane, and bipyridine/phenanthroline protons are observed in a 1
:
1
:
1 ratio, confirming the stoichiometry of the ligands in the final complex. The assignment of all resonances observed in 1H,13C{1H}-APT NMR (atom numbering is shown in Fig. 1) was supported by analysis of 2D NMR experiments (1H–1H COSY, 1H–13C HMQC and HMBC).
In short, the 1H NMR spectra of complexes 1–5 show discrete resonances attributed to the cyclopentadienyl ring in the typical range for cationic ‘RuCp’ compounds (∼δ 5.00–5.30 ppm in acetone-d6). Upon coordination of the N,N-bidentate, all the aromatic protons show a significant deshielding, especially the protons next to the coordinated nitrogen (Δδ ≈ 0.7 ppm), and a shielding on the H3 protons of the 2,2′-bipyridine derivative (ΔH3 ≈ −0.5 ppm). This proves the successful σ-coordination of the diimine ligand to the ruthenium center.34–37 The 31P{1H} NMR spectrum of complexes 1–5 is characterized by the presence of a sharp resonance at δ ∼45 ppm (acetone-d6), which is a significant deshielding of the phosphorous nuclei resonance is observed (Δδ up to 72 ppm), indicating the successful coordination of the alkyl-diphenylphosphane to the metal center. Detailed 13C{1H}-APT NMR data are described in the experimental section and follow the effects discussed with the 1H NMR results.
Complexes 1–5 are soluble in organic solvents such as dichloromethane, methanol, acetone, acetonitrile, and dimethylsulfoxide, and insoluble in water, n-hexane, and diethyl ether. The optical absorption and emission spectra of all compounds were recorded using up to 10−3 to 10−6 M solutions in acetonitrile and dimethylsulfoxide at room temperature. Fig. S16 (SI) shows the electronic absorption spectra recorded in acetonitrile for complex 1 as a representative example. All complexes showed absorption bands in the UV region (λ = 200–325 nm) that are ascribed to electronic transitions occurring in the organometallic fragment ({[Ru(η5-C5H5)(Ph2P-CH2-pip-NBD)]+}) and the coordinated N,N-ligand. In addition to these bands, this series of organometallics shows a common feature related to the presence of the fluorophore in their structure: the usual metal-to-ligand charge transfer (MLCT) band, frequently observed for this type of organometallic complexes (at λ ∼450 nm), is masked by the absorption of the NBD-derived probe. The fluorescent unit on the organometallic structure, which aims to probe the cellular uptake and distribution of the compounds, was further studied by fluorescence spectroscopy. The luminescence quantum yields (Table S1 of the SI) of the Ph2P-CH2-pip-NBD coordinated complexes (ΦF1: 0.3%, 2: 0.1%, 3: 0.2%, 4: 0.2%, 5: 0.2%) were found to be significantly lower than that of the Ph2P-CH2-pip-NBD free ligand (ΦF: 1.0%). Fig. S17 (SI) shows the absorption and emission spectra of all (non-protonated) complexes in dimethylsulfoxide.
| 1 | 3 | 5 | |
|---|---|---|---|
| Bond lengths (Å) | |||
| Ru1–P (21or 23) | 2.300(2) | 2.3031(13) | 2.3098(9) |
| Ru1–N7 | 2.090(7) | 2.080(4) | 2.098(3) |
| Ru1–N18 | 2.101(6) | 2.095(4) | 2.089(3) |
| Ru1–Cp (centroid) | 1.842(4) | 1.829(2) | 1.826(2) |
| Angles (°) | |||
| N18–Ru1–N7 | 76.5(3) | 76.04(16) | 78.06(10) |
| N18–Ru1–P (21or 23) | 89.49(17) | 88.39(11) | 88.50(7) |
| N7–Ru1–P (21or 23) | 87.73(19) | 89.17(11) | 88.32(7) |
| Cp(centroid)–Ru1–N18 | 127.6 (2) | 128.84(15) | 128.40(9) |
| Cp(centroid)–Ru1–N7 | 133.2(3) | 131.15(15) | 130.73(9) |
| Cp(centroid)–Ru1–P (21 or 23) | 124.98(15) | 126.82(9) | 126.95(6) |
1 and 3 crystallize in the triclinic system (centrosymmetric space group P
) whereas 5 crystallize in the monoclinic system (space group P1 21/c1). All the compounds displayed racemic crystals, with the presence of the two enantiomers. Each asymmetric unit includes a cationic organometallic ‘RuCp’ complex and a hexafluorophosphate anion. The asymmetric unit for 1 contains one acetone molecule. All complexes feature a six-coordinate Ru(II) center with a pseudo-octahedral piano-stool geometry. The ruthenium ion is coordinated by an η5-cyclopentadienyl ring in a π-bonded fashion, one disubstituted bipyridyl ligand (1, –CH3; and 3, –OCH3) or a phenanthroline ligand (5), and the remaining coordination site is occupied by the phosphine ligand Ph2P-CH2-pip-NBD.
The distances between the ruthenium centers and the centroids of the cyclopentadienyl rings range from 1.826(2) to 1.842(4) Å, consistent with values reported for similar compounds in the literature.35,37 A similar trend is observed for the Ru–P bond lengths, which fall within the range of 2.300(2)–2.3098(9) Å. Notably, for the compounds containing bipyridyl ligands (1, 3), the Ru–N bond distances remain largely unaffected by the nature of the substituents on the bipyridyl ring. However, in compound 5, which features a phenanthroline ligand, a slight increase in Ru–N bond lengths is observed compared to the previous two complexes (see Table 1). All compounds exhibit weak intramolecular hydrogen bonding interactions between the nitrogen atoms of the piperazine substituent (adjacent to the phosphorus atoms) and two hydrogen atoms from the cyclopentadienyl rings.
These values range from 2.737(15) to 3.127(11) Å, and an example of this type of interaction is illustrated in Fig. S18 of the SI.
The bond angles between donor atoms around the Ru center are comparable to those reported for other cyclopentadienyl-based complexes in the literature.36 In general, the Cp(centroid)–Ru–N angles are larger than the Cp(centroid)–Ru–P angles, likely due to the steric constraints imposed by the substituents on the phosphine ligand. For all complexes, the N–Ru–N angles reflect the geometric constraints inherent to the bipyridyl and phenanthroline ligands.
The packing structures of the compounds are shown in Fig. S19–S21 of the SI. Interestingly, the unit cells of all the described compounds exhibit a centre of inversion, which plays a significant role in determining the molecular packing. In compound 1, the packing along the a-axis reveals molecular chains in which all the molecules adopt the same orientation along the chain. However, alternating orientation is observed between adjacent chains (Fig. S19 of the SI). This intra-chain arrangement appears to be influenced by intermolecular hydrogen bonding, both between neighbouring molecules and between molecules and PF6− anions, which, along with solvent molecules, are located between the chains. These intermolecular interactions involve hydrogen atoms from phenyl and methylene groups of the phosphane substituents forming hydrogen bonds with oxygen atoms from the NO2 group of neighbouring molecules (H22D–O40 Å = 2.555(11) Å; H46–O40 Å = 2.613(12) Å). Additional hydrogen bonds are observed between PF6− counterions and hydrogens from the phenyl and bipyridyl rings of two adjacent molecules of two different chains (H50–F3S = 2.546(12) Å; H16A–F4S = 2.558(12) Å) (Fig. S19B of the SI). A similar intra-chain arrangement is observed in the packing of compound 3 along the a-axis (Fig. S20 of the SI), consistent with the presence of analogous intermolecular interactions in both complexes, as seen in compound 1. The molecular packing of 5 along the a-, b-, and c-axes is shown in Fig. S21 of the SI.
The stability of the cationic complexes was also evaluated in aqueous cell culture medium DMEM without phenol red. Complexes 1–5 were first solubilized in DMSO and then diluted to 5% DMSO with DMEM. Changes in shape or intensity are negligible over 24 h (absorbance variations below 10%, Fig. S23 of the SI). Overall, this set of compounds is robust regarding ligand exchange, maintaining the original structure over time, which is a highly desirable feature in this context.
Cells were incubated with complexes 1–5, Ph2P-CH2-pip-NBD and cisplatin (CDDP), and cell viability was assessed after a 48 h period, and their cytotoxicity (IC50 values) was determined using the Alamar blue assay. The cytotoxicity analysis revealed distinct variations in the toxicity of the tested compounds across the different cell lines. Complex 1 exhibited the highest cytotoxic effects in all cell lines tested, with low IC50 values, particularly in 4T1 (7.42 ± 0.63 μM) and U2OS (3.73 ± 0.28 μM). Complexes 2 and 5 also demonstrated significant toxicity, particularly in U2OS (IC50 = 4.57 ± 0.46 and IC50 = 6.94 ± 1.70 μM, respectively). 3T3 fibroblast also demonstrated moderate sensitivity to these compounds, with IC50 values for complexes 1 and 3 (IC50 = 7.58 ± 0.16 μM and IC50 = 13.00 ± 0.67 μM, respectively) comparable to those observed in cancer cells, thus suggesting potential off-target effects. On the other hand, 2 and 4 did not display significant toxicity, with IC50 values exceeding 100 μM in all tested cell lines (Table 2 and Fig. S24–S27 of the SI). This indicates that the incorporation of hydroxyl or biotin substituents on the bipyridine ligand has a detrimental impact on cytotoxic activity within this compound family. A similar trend has been previously reported for related systems.35,37
| 4T1 | CT26 | U2OS | 3T3 | |
|---|---|---|---|---|
| 1 | 7.42 ± 0.63 | 23.21 ± 0.17 | 3.73 ± 0.28 | 7.58 ± 0.16 |
| 2 | >100 | >100 | >100 | >100 |
| 3 | 11.14 ± 2.08 | >100 | 4.57 ± 0.46 | 13.00 ± 0.67 |
| 4 | >100 | >100 | >100 | >100 |
| 5 | 12.14 ± 1.23 | >100 | 6.94 ± 1.70 | 19.59 ± 3.73 |
| Ph2P-CH2-pip-NBD | 43.06 ± 3.7 | 15.61 ± 0.56 | 48.83 ± 3.29 | 14.49 ± 0.79 |
| CDDP | 8.68 ± 0.90 | 15.46 ± 4.96 | 16.86 ± 5.07 | 9.94 ± 0.38 |
The most unexpected observation was the response of CT26 cells to the organometallic complexes, as they exhibited markedly higher resistance, with IC50 values exceeding 100 μM for all tested compounds, except for complex 1, which showed a comparatively high IC50 of 23.21 ± 0.17 μM relative to its effects on other cancer cell lines. These observations may be explained by the high mutational load of CT26 cells.38 Overall, 1, 3, and 5 demonstrated significantly higher cytotoxicity than the Ph2P-CH2-pip-NBD fluorescent ligand (except for CT26 cells), highlighting these candidates as the most promising compounds. Additionally, these compounds showed similar cytotoxicity to cisplatin (with the exception of CT26 cells), being particularly more cytotoxic in the human osteosarcoma U2OS cell line.
Cells were incubated with complexes 1–5 at concentrations of 6.25 μM and 12.5 μM for 25 h and, after a washing step, immediately analyzed by flow cytometry (Fig. 3A and B). The gating and fluorescence histograms obtaining by flow cytometry are shown in Fig. S28–S35 of the SI. Complexes 1, 3, and 5 exhibited the highest cell uptake across all cell lines, whereas complexes 2 and 4 demonstrated the lowest levels of internalization. This trend in cell uptake correlates well with the compounds’ cytotoxicity. Notably, U2OS, the most sensitive cell line to complexes 1, 3, and 5 (cell uptake could not be recorded for complexes 1 and 3 at 12.5 μM due to high toxicity), showed the highest internalization of these complexes. However, it should be noted that the correlation between cell uptake and cytotoxicity was not observed in CT26 cells. In these cells, cell uptake was similar to that in other cell lines, but no significant toxicity (except for complex 1) was detected. Overall, a strong correlation between cell uptake and cytotoxicity was observed in 4T1, U2OS, and 3T3 cells, but not in CT26 cells, suggesting that these cells may possess more robust defense mechanisms against cell death mediated by the complexes. To confirm that all complexes were internalized rather than merely associated with the cell membrane, fluorescent microscopy was used to image cells incubated with the compounds for 6 h. The acquired images reveal that the compounds localize within the cells, accumulating in cytoplasmic dots and with some dispersed signal throughout the cytoplasm (Fig. 3C).
Given the observed intracellular accumulation of the fluorescent complexes as distinct punctate dots, the most likely internalization mechanism is endocytosis – a process by which the cell membrane engulfs the compound, forming small vesicles (endosomes) that transport it into the cell. In order to confirm endocytosis as the primary internalization route, we evaluated the cellular uptake of the compounds in the presence of inhibitors of specific endocytic pathways, namely, chloroquine (a clathrin-mediated endocytosis inhibitor) and genistein (a caveolin-mediated endocytosis inhibitor). Flow cytometry and fluorescence microscopy experiments demonstrated that co-incubation of complexes 1–5 with chloroquine induced a slight, though not statistically significant, increase in the uptake of the compounds. This modest increase in intracellular accumulation may be explained by chloroquine's interference with the maturation of early endosomes into late endosomes and their subsequent fusion with lysosomes. By disrupting this trafficking pathway, chloroquine may lead to the accumulation of endosomes, thereby reducing compound degradation and contributing to increased intracellular retention (Fig. 4A and C). In contrast, genistein significantly inhibited their internalization, with a more pronounced effect observed for 1, 3, and Ph2P-CH2-pip-NBD. These results suggest that complexes 1–5 are primarily internalized via endocytosis, namely caveolin-mediated endocytosis (Fig. 4B and C).
These findings are consistent with the lack of correlation between the partition coefficients (KP) for lipid membranes of POPC:POPS (LUVs diameter 117.5 ± 1.1 nm; zeta potential – 10.3 ± 0.8 mV) and cell uptake. KP values were determined for the Ph2P-CH2-pip-NBD free ligand (KP = 3.47 × 102, log
KP = 2.54 ± 0.23), complex 1 (KP = 5.80 × 102, log
KP = 2.76 ± 0.26), one of the compounds with higher uptake and cytotoxicity, and complex 2 (KP = 2.64 × 103, log
KP = 3.42 ± 0.41), which exhibited one of the lowest uptake and cytotoxicity levels. The Ph2P-CH2-pip-NBD and complex 1 showed lower KP values, indicating a weaker affinity for lipid membranes, while complex 2 had the highest KP, indicating greater membrane affinity (Fig. S36 of the SI). However, these values did not correlate with the observed cellular internalization trends, as complex 1 was more efficiently internalized than complex 2 (Fig. 3). This discrepancy can be explained by the fact that KP is expected to correlate with internalization via passive diffusion (which involves direct passage through the phospholipid bilayer), but not via endocytosis (which involves vesicular uptake). In fact, the results observed indicate an inverse correlation between the compound toxicity and its membrane affinity. The higher KP observed for complex 2 in comparison with complex 1 is also unexpected given that the only structural difference is the presence of two CH2OH groups in complex 2 instead of two CH3 groups in complex 1. It was also observed that while the fluorescence quantum yield in DMSO is only slightly higher for complex 1 (Table S1), the fluorescence increase observed upon association with the lipid membranes is 5-fold higher for this complex. This suggests that complex 2 interacts mostly with the polar lipid headgroups and thus, its higher value of KP is not expected to lead to a higher passive permeability.39–41 Those results further reinforce that cell uptake of these complexes occurs mostly by endocytosis.
![]() | ||
| Fig. 5 Ru(II)–cyclopentadienyl complexes (in green) mainly accumulate in lysosomes (in red). Representative images of 4T1 cells taken 6 h after incubation with complexes 1–5 and Ph2P-CH2-pip-NBD. | ||
Although the fluorescence quantum yields of the complexes were very low in DMSO (0.001 to 0.0027, Table S1 of the SI), relatively strong intracellular fluorescence was observed in our experiments using flow cytometry and fluorescence microscopy (Fig. 3–5). This was particularly evident when the compounds were localized in endosomes and/or lysosomes, which are acidic compartments. These findings suggest that, like the free Ph2P-CH2-pip-NBD ligand, complexes 1–5 may exhibit enhanced fluorescence at acidic pH.24 Indeed, a kinetic study conducted at different pH values (2, 6, and 7.4) revealed a significant increase in the fluorescence of 1 and 2 at pH 2, but not at pH 6 or 7.4. The fluorescence of Ph2P-CH2-pip-NBD was much higher at pH 2, even at small times after incubation at this pH, with a more moderate and slower increase at pH 6 (Fig. S37 of the SI). A very strong fluorescence increase was observed for complex 1 when at pH 2, occurring over 7 hours and stabilizing at the fluorescence intensity observed for Ph2P-CH2-pip-NBD, while no variation was observed at pH 6 or 7.4. A similar behavior was observed for complex 2, although with slower kinetics. The enhanced fluorescence at acidic pH may be due to the protonation of the amino groups in the complexes. Additionally, one cannot neglect the possibility of Ph2P-CH2-pip-NBD release from the ruthenium complex within the lysosomes, leading to increased fluorescence, as the quantum yield of Ph2P-CH2-pip-NBD is 5 to 10 times higher than that of the Ru complexes.
Yield: 93%. 1HNMR [(CD3)2CO, Me4Si] δ (ppm): 9.22 (d, 2H, 3JHH = 6,
6bipy), 8.43 (d, 1H, 3JHH = 9,
5′PPh2-CH2-pip-NBD), 7.90 (s, 2H,
3bipy), 7.37 (m, 2H, 3JHH = 7,
paraPPh2-CH2-pip-NBD), 7.21 (m, 6H,
metaPPh2-CH2-pip-NBD +
5bipy), 7.09 (t, 4H, 3JHH = 8,
orthoPPh2-CH2-pip-NBD), 6.57 (d, 1H, 3JHH = 9,
6′PPh2-CH2-pip-NBD), 5.11 (s, 5H, C5
5), 4.00 (m, 4H,
3PPh2-CH2-pip-NBD +
5PPh2-CH2-pip-NBD), 3.79 (s, 2H, PPh2-C
2-pip-NBD), 2.62 (t, 4H, 3JHH = 5,
2PPh2-CH2-pip-NBD +
6PPh2-CH2-pip-NBD), 2.45 (s, 6H, C
3bipy). APT-13C{1H} NMR [(CD3)2CO] δ (ppm): 156.3 (
6bipy), 156.0 (
2bipy), 149.0, (
4bipy), 146.1 (
1′Ph2P-CH2-pip-NBD), 146.0 + 145.8 (
2′Ph2P-CH2-pip-NBD +
3′Ph2P-CH2-pip-NBD), 136.3 (
5′Ph2P-CH2-pip-NBD), 133.5 (d, 2JCP = 11,
HorthoPh2P-CH2-pip-NBD), 131.8 (d, 1JCP = 38,
ipsoPh2P-CH2-pip-NBD), 130.6 (d, 4JCP = 2,
HparaPh2P-CH2-pip-NBD), 129.1 (d, 3JCP = 10,
HmetaPh2P-CH2-pip-NBD), 127.1 (
5bipy), 124.7 (
3bipy), 104.2 (
6′Ph2P-CH2-pip-NBD), 77.6 (d, 2JCP = 2,
5H5), 60.9 (d, 1JCP = 38,
1Ph2P-CH2-pip-NBD), 55.7 (d, 3JCP = 9,
2Ph2P-CH2-pip-NBD and
6Ph2P-CH2-pip-NBD), 50.1 (
3Ph2P-CH2-pip-NBD and
5Ph2P-CH2-pip-NBD), 20.8 (
H3). 31P{1H} NMR [(CD3)2CO] δ (ppm): 45.5 (s, Ph2
-CH2-pip-NBD), −144.3 (sept, 1JPF = 708,
F6). FTIR [KBr, cm−1]: 3053 (νC–H aromatic rings), 2921–2852 (νC–H alkanes), 1617 and 1436 (νC
C aromatic rings), 843 and 557 (νPF6 counterion). UV-Vis [NCCH3, λmax/nm (ε × 103/M−1 cm−1)]: 238 (sh), 290 (4.3), 340 (2.9), 481 (5.8). Elemental analysis calc. for C40H39F6N7O3P2Ru (942.79): C, 50.96, H, 4.17; N, 10.40. Found: C 51.1; H, 4.2; N, 10.2. ESI-MS: [1-PF6]+ calc. for [C40H39N7O3PRu]+: 797.83. Found: 797.98.
6bipy), 8.43 (d, 1H, 3JHH = 9,
5′Ph2P-CH2-pip-NBD), 7.98 (s, 2H,
3bipy), 7.37 (m, 4H, 3JHH = 7, HparaPPh2-CH2-pip-NBD +
5bipy), 7.21 (m, 4H,
metaPPh2-CH2-pip-NBD), 7.09 (t, 4H, 3JHH = 8, HorthoPPh2-CH2-pip-NBD), 6.57 (d, 1H, 3JHH = 9,
6′Ph2P-CH2-pip-NBD), 5.15 (s, 5H, C5
5), 4.77 (m, 6H, C
2OHbipy), 4.01 (m, 4H,
3Ph2P-CH2-pip-NBD +
5Ph2P-CH2-pip-NBD), 3.78 (s, 2H, Ph2P-C
2-pip-NBD), 2.62 (t, 4H, 3JHH = 5,
2Ph2P-CH2-pip-NBD +
6Ph2P-CH2-pip-NBD). APT-13C{1H} NMR [(CD3)2CO] δ (ppm): 156.4 (d, 3JCP = 2,
6bipy), 156.1 (
4bipy), 153.4 (
6bipy), 146.1 (
1′Ph2P-CH2-pip-NBD), 146.0 + 145.8 (
2′Ph2P-CH2-pip-NBD +
3′Ph2P-CH2-pip-NBD), 136.3 (
5′Ph2P-CH2-pip-NBD), 133.4 (d, 2JCP = 11,
HorthoPh2P-CH2-pip-NBD), 131.6 (d, 1JCP = 38,
ipsoPh2P-CH2-pip-NBD), 130.7 (d, 4JCP = 2,
HparaPh2P-CH2-pip-NBD), 129.1 (d, 3JCP = 10,
HmetaPh2P-CH2-pip-NBD), 123.5 (
5bipy), 120.9 (
3bipy), 104.2 (
6′Ph2P-CH2-pip-NBD), 77.7 (d, 2JCP = 2, C5H5), 62.5 (
H2bipy), 59.8 (d, 1JCP = 38, C1Ph2P-CH2-pip-NBD), 55.7 (d, 3JCP = 9, C2Ph2P-CH2-pip-NBD and C6Ph2P-CH2-pip-NBD), 50.1 (C3Ph2P-CH2-pip-NBD and C5Ph2P-CH2-pip-NBD). 31P{1H} NMR [(CD3)2CO] δ (ppm): 45.5 ppm (s, PPh2-CH2-pip-NBD), −144.3 (sept, 1JPF = 708, PF6). FTIR [KBr, cm−1]: 3444 (νO–H), 3078–3058 (νC–H aromatic rings), 2925–2854 (νC–H alkanes), 1612 and 1436 (νC
C aromatic rings), 1332 (νN–O), 1132 (νC–O), 847 and 558 (νPF6 counterion). UV-Vis [NCCH3, λmax/nm (ε × 103/M−1 cm−1)]: 238 (sh), 290 (4.1), 339 (2.6), 480 (5.1). Elemental analysis calc. for C40H39F6N7O5P2Ru (974.79): C, 49.29, H, 4.03; N, 10.06. Found: C, 49.5; H, 4.2; N, 10.0. ESI-MS: [2-PF6]+ calc. for [C40H39N7O5PRu]+: 829.83. Found: 829.98.
1H NMR [(CD3)2CO, Me4Si] δ (ppm): 9.13 (d, 2H, 3JHH = 6,
6bipy), 8.43 (d, 1H, 3JHH = 9,
5′Ph2P-CH2-pip-NBD), 7.62 (s, 2H,
3bipy), 7.39 (m, 2H, 3JHH = 7,
paraPh2P-CH2-pip-NBD), 7.23(m,4H,
metaPh2P-CH2-pip-NBD), 7.13 (t, 4H, 3JHH = 8,
orthoPh2P-CH2-pip-NBD), 7.00 (dd, 2H,
5bipy), 6.57 (d, 1H, 3JHH = 9,
6′Ph2P-CH2-pip-NBD), 5.05 (s, 5H, C5
5), 4.01 (m, 4H,
3Ph2P-CH2-pip-NBD + H5Ph2P-CH2-pip-NBD), 3.97 (s, 6H, OC
3bipy), 3.78 (s, 2H, Ph2P-C
2-pip-NBD), 2.62 (t, 4H, 3JHH = 5,
2Ph2P-CH2-pip-NBD +
6Ph2P-CH2-pip-NBD).
APT-13C{1H} NMR [(CD3)2CO] δ (ppm): 166.2 (
2bipy), 156.8 (
4bipy), 156.7 (d, 3JCP = 2,
6bipy), 145.3 (
1′Ph2P-CH2-pip-NBD), 145.1 + 145.0 (
2′Ph2P-CH2-pip-NBD +
3′Ph2P-CH2-pip-NBD), 135.4 (
5′Ph2P-CH2-pip-NBD), 132.7 (d, 2JCP = 11,
HorthoPh2P-CH2-pip-NBD), 131.4 (d, 1JCP = 38,
ipsoPh2P-CH2-pip-NBD), 129.7 (d, 4JCP = 2,
HparaPh2P-CH2-pip-NBD), 128.3 (d, 3JCP = 10,
HmetaPh2P-CH2-pip-NBD), 112.5 (
5bipy), 109.2 (
3bipy), 103.3 (
6′Ph2P-CH2-pip-NBD), 76.0 (d, 2JCP = 2,
5H5), 60.9 (d, 1JCP = 38,
1Ph2P-CH2-pip-NBD), 59.0 (d, 1JCP = 38,
1Ph2P-CH2-pip-NBD), 56.0 (O
H3), 54.9 (d, 3JCP = 9,
2Ph2P-CH2-pip-NBD and
6Ph2P-CH2-pip-NBD), 49.2 (
3Ph2P-CH2-pip-NBD and
5Ph2P-CH2-pip-NBD). 31P{1H} NMR [(CD3)2CO] δ (ppm): 46.2 ppm (s, Ph2
-CH2-pip-NBD), −144.3 (sept, 1JPF = 708,
F6). FTIR [KBr, cm−1]: 3086–3052 (νC–H aromatic rings), 2939–2805 (νC–H alkanes), 1614 and 1436 (νC
C aromatic rings), 1129 (νC–O), 842 and 557 (νPF6 counterion). UV-Vis [NCCH3, λmax/nm (ε × 103/M−1 cm−1)]: 236 (sh), 291 (4.7), 340 (2.7), 480 (5.6). Elemental analysis calc. for C40H39F6N7O5P2Ru (974.79): C, 49.29, H, 4.03; N, 10.06. Found: C, 49.4; H, 4.1; N, 10.1. ESI-MS: [3-PF6]+ calc. for [C40H39N7O5PRu]+: 829.83. Found: 829.96.
6bipy), 8.43 (d, 1H, 3JHH = 9,
5′Ph2P-CH2-pip-NBD), 8.02 (s, 2H,
3bipy), 7.40 (m, 4H,
paraPh2P-CH2-pip-NBD +
5bipy), 7.22 (t, 4H, 3JHH = 7,
metaPh2P-CH2-pip-NBD), 7.08 (t, 4H, 3JHH = 8,
orthoPh2P-CH2-pip-NBD), 6.58 (d, 1H, 3JHH = 9,
6′Ph2P-CH2-pip-NBD), 6.06 (d, 2H, N
Biotin), 5.87 (br d, 2H, N
Biotin), 5.26 (m, 4H, C
2Biotin), 5.19 (s, 5H, C5
5), 4.49 (m, 2H, SCH2C
Biotin), 4.32 (m, 2H, CHBiotin), 4.01 (m, 4H,
3Ph2P-CH2-pip-NBD +
5Ph2P-CH2-pip-NBD), 3.82 (s, 2H, Ph2P-C
2-pip-NBD), 3.22 (2H, S-C
Biotin), 2.63 (t, 4H, 3JHH = 5,
2Ph2P-CH2-pip-NBD +
6Ph2P-CH2-pip-NBD), 2.50 (m, 4H, SC
2Biotin), 1.78–1.49 (m, 12H, C
2C
2C
2C
2Biotin). APT-13C{1H} NMR [(CD3)2CO] δ (ppm): 173.6 (CO, bipy(CH2Biotin)2 ester), 164.2 (CO, bipy(CH2Biotin)2 urea), 156.9 (br,
6bipy), 156.0 (
2bipy), 147.1 (
4bipy), 146.1 (
1′Ph2P-CH2-pip-NBD), 145.9 + 145.8 (
2′Ph2P-CH2-pip-NBD +
3′Ph2P-CH2-pip-NBD), 136.5 (
5′Ph2P-CH2-pip-NBD), 133.3 (d, 2JCP = 11,
HorthoPh2P-CH2-pip-NBD), 131.2 (d, 1JCP = 38,
ipsoPh2P-CH2-pip-NBD), 130.7 (d, 4JCP = 2,
HparaPh2P-CH2-pip-NBD), 129.2 (d, 3JCP = 10,
HmetaPh2P-CH2-pip-NBD), 124.5 (
5bipy), 122.3 (
3bipy), 104.2 (
6′Ph2P-CH2-pip-NBD), 77.9 (d, 2JCP = 2,
5H5), 64.1 (
H2bipy), 62.4 (
HBiotin), 60.7 (S
H2-CHBiotin), 59.7 (d, 1JCP = 38,
1Ph2P-CH2-pip-NBD), 56.5 (SCHBiotin), 59.7 (d, 1JCP = 38,
1Ph2P-CH2-pip-NBD), 55.7 (d, 3JCP = 9,
2Ph2P-CH2-pip-NBD and
6Ph2P-CH2-pip-NBD), 50.0 (
3Ph2P-CH2-pip-NBD and
5Ph2P-CH2-pip-NBD), 40.9 (S
H2Biotin), 34.1 + 29.2* + 25 (
H2
H2
H2
H2Biotin)* under the solvent peak. 31P{1H} NMR [(CD3)2CO] δ (ppm): 45.31 ppm (s, Ph2
-CH2-pip-NBD), −144.3 (sept, 1JPF = 708,
F6). FTIR [KBr, cm−1]: 3388–3263 (νN–H), 3077 (νC–H aromatic rings), 2928–2859 (νC–H alkanes), 1732–1698 (νC
O urea, ester), 1434 (νC
C aromatic rings), 1440 (νC
C aromatic rings), 843 and 557 (νPF6 counterion). UV-Vis [NCCH3, λ/nm (ε × 103/M−1 cm−1)]: 240 (sh), 290 (4.2), 340 (1.9), 480 (5.1). Elemental analysis calc. for C60H67F6N11O9P2RuS2 (1427.38): C, 50.49, H, 4.73; N, 10.79; S, 4.49. Found: C, 50.3; H, 4.6; N, 10.6; S, 4.1. ESI MS: [4-PF6]+ calc. for [C60H67N11O9PRuS2]+: 1282.42. Found: 1281.99.
2phen), 8.47 (d, 2H, 3JHH = 6,
4phen), 8.43 (d, 1H, 3JHH = 9,
5′Ph2P-CH2-pip-NBD), 7.92 (s, 2H,
6phen), 7.77 (m, 2H,
3phen), 7.16 (t, 2H, 3JHH = 7,
paraPPh2-CH2-pip-NBD), 6.97 (t, 4H, 3JHH = 7,
metaPPh2-CH2-pip-NBD), 6.90 (t, 4H, 3JHH = 8,
orthoPh2P-CH2-pip-NBD), 6.57 (d, 1H, 3JHH = 9,
6′Ph2P-CH2-pip-NBD), 5.27 (s, 5H, C5
5), 3.99 (m, 4H,
3Ph2P-CH2-pip-NBD +
5Ph2P-CH2-pip-NBD), 3.87 (s, 2H, PPh2-C
2-pip-NBD), 2.62 (t, 4H, 3JHH = 5,
2Ph2P-CH2-pip-NBD +
6Ph2P-CH2-pip-NBD). APT-13C{1H} NMR [(CD3)2CO] δ (ppm): 156.8 (
2phen), 147.6 (
5phen), 145.8 (
1′Ph2P-CH2-pip-NBD), 145.6 + 145.5 (
2′Ph2P-CH2-pip-NBD +
3′Ph2P-CH2-pip-NBD), 136.0 (
5′PPh2-CH2-pip-NBD), 135.2 (
4phen), 132.5 (d, 2JCP = 11,
HorthoPh2P-CH2-pip-NBD), 130.8 (d, 1JCP = 38, CipsoPh2P-CH2-pip-NBD), 130.1 (d, 4JCP = 2,
HparaPh2P-CH2-pip-NBD), 128.3 (d, 3JCP = 10,
HmetaPh2P-CH2-pip-NBD), 127.6 (
6phen), 125.1 (
3phen), 103.8 (
6′Ph2P-CH2-pip-NBD), 77.2 (d, 2JCP = 2,
5H5), 59.29 (d, 1JCP = 38,
1Ph2P-CH2-pip-NBD), 55.4 (d, 3JCP = 6,
2Ph2P-CH2-pip-NBD and
6Ph2P-CH2-pip-NBD), 49.7 (
3Ph2P-CH2-pip-NBD and
5Ph2P-CH2-pip-NBD). 31P{1H} NMR [(CD3)2CO] δ (ppm): 46.1 (s, Ph2
-CH2-pip-NBD), −144.2 (sept, 1JPF = 708,
F6). FTIR [KBr, cm−1]: 3077 (νC–H aromatic rings), 2928–2859 (νC–H alkanes), 1434 (νC
C aromatic rings), 840 and 558 (νPF6 counterion). UV-Vis [NCCH3, λ/nm (ε × 103/M−1 cm−1)]: 239 (sh), 290 (4.3), 339 (2.9), 482 (5.6). Elemental analysis calc. for C40H35F6N7O3P2Ru (938.76): C, 51.18, H, 3.76; N, 10.44. Found: C, 51.2; H, 3.8; N, 10.7. ESI-MS: [5-PF6]+ calc. for [C40H35N7O3PRu]+: 794.16. Found: 794.11.
432 reflections to a maximum θ angle of 26.57° (0.79 Å resolution), of which 18
801 were independent (average redundancy 9.810, completeness = 99.0%, Rint = 11.95%, Rsig = 6.63%) and 13
670 (72.71%) were greater than 2σ(F2). The final cell constants of a = 11.8493(15) Å, b = 19.432(3) Å, c = 20.703(3) Å, α = 78.293(4)°, β = 77.765(5)°, γ = 84.304(5)°, volume = 4553.9(10) Å3, are based upon the refinement of the XYZ-centroids of 9881 reflections above 20 σ(I) with 5.350° < 2θ < 52.33°. The ratio of minimum to maximum apparent transmission was 0.950. The calculated minimum and maximum transmission coefficients (based on crystal size) are 0.9220 and 0.9860.
The structure was solved and refined using the space group P
, with Z = 2 for the formula unit, C93.50H105F12N14O10.50P4Ru2. The final anisotropic full-matrix least-squares refinement on F2 with 1099 variables converged at R1 = 10.17%, for the observed data and wR2 = 26.16% for all data. The goodness-of-fit was 1.054. The largest peak in the final difference electron density synthesis was 2.007 e− Å−3 and the largest hole was −1.957 e−Å−3 with an RMS deviation of 0.178 e− Å−3. On the basis of the final model, the calculated density was 1.566 g cm−3 and F(000), 2208e−.
The crystal lattice contained a large amount of diffuse electron density that could not be appropriately modelled. The SQUEEZE routine within PLATON was employed to resolve this problem, resulting in void electrons count per cell of 215 that was assigned to seven solvent acetone molecules. The collected data was of moderate quality, primarily due to the crystal's mosaicity. This resulted in a mixture of well-defined diffraction spots and others that were significantly broadened and poorly resolved, so the R factors are slightly higher than expected.
238 reflections to a maximum θ angle of 28.43° (0.75 Å resolution), of which 23
334 were independent (average redundancy 7.553, completeness = 99.3%, Rint = 6.22%, Rsig = 4.33%) and 17
730 (75.98%) were greater than 2σ(F2). The final cell constants of a = 11.8831(7) Å, b = 19.9377(11) Å, c = 20.9074(12) Å, α = 70.752(2)°, β = 85.908(2)°, γ = 88.956(2)°, volume = 4664.5(5) Å3, are based upon the refinement of the XYZ-centroids of 9848 reflections above 20 σ(I) with 5.559° < 2θ < 56.03°. The ratio of minimum to maximum apparent transmission was 0.908. The calculated minimum and maximum transmission coefficients (based on crystal size) are 0.8710 and 0.9320.
The structure was solved and refined using the space group P
, with Z = 4 for the formula unit, C42H43Cl4F6N7O5P2Ru. The final anisotropic full-matrix least-squares refinement on F2 with 1294 variables converged at R1 = 7.15%, for the observed data and wR2 = 17.81% for all data. The goodness-of-fit was 1.136. The largest peak in the final difference electron density synthesis was 1.760 e− Å−3 and the largest hole was −1.669 e− Å−3 with an RMS deviation of 0.126 e− Å−3. On the basis of the final model, the calculated density was 1.630 g cm−3 and F(000), 2320e−. The Fluorines of one of the PF6− anions was disordered and modelled with the PART command with occupancies close to 50% over two sites. One of the solvent CH2Cl2 molecules was disordered and modelled with the PART command with occupancies of 60% and 40% over two sites. The phenyl ring C30-C35 was disordered and modelled with the PART command with occupancies of 70% and 30% over two sites. The phenyl ring C24a-C29a was disordered and modelled with the PART command with occupancies of 60% and 40% over two sites. One of the CH2-Piperazine moiety, atoms c36A, N37A, C38A, C39A, N40A, C41A and C42A was disordered and modelled with the PART command with occupancies of 50% over two sites.
107 reflections to a maximum θ angle of 27.55° (0.77 Å resolution), of which 8902 were independent (average redundancy 4.730, completeness = 99.3%, Rint = 7.96%, Rsig = 6.28%) and 6595 (74.08%) were greater than 2σ(F2). The final cell constants of a = 18.1160(10) Å, b = 13.1312(8) Å, c = 16.5005(11) Å, β = 98.267(2)°, volume = 3884.4(4) Å3, are based upon the refinement of the XYZ-centroids of 9918 reflections above 20 σ(I) with 5.176° < 2θ < 54.68°. The ratio of minimum to maximum apparent transmission was 0.744. The calculated minimum and maximum transmission coefficients (based on crystal size) are 0.8050 and 0.9400.
The structure was solved and refined using the space group P1 21/c1, with Z = 1 for the formula unit, C160H142F24N28O13P8Ru4. The final anisotropic full-matrix least-squares refinement on F2 with 539 variables converged at R1 = 4.81%, for the observed data and wR2 = 9.10% for all data. The goodness-of-fit was 1.077. The largest peak in the final difference electron density synthesis was 0.75 e− Å−3 and the largest hole was −0.850 e− Å−3 with an RMS deviation of 0.103 e− Å−3. On the basis of the final model, the calculated density was 1.613 g cm−3 and F(000), 1914e−.
In both protocols, at the indicated time points, cells were washed once with PBS, detached with trypsin-EDTA and centrifuged for 5 minutes at 2000 rpm. Cells were then resuspended in PBS and immediately analysed by flow cytometry (Agilent, Novocyte® 3000). The fluorescence associated to the Ru(II) compounds was measured upon excitation with the 488 nm laser, and detection was performed using the 530/30 nm filter. An average of 15
000 events were collected, and the data are presented as the mean fluorescence normalized to the mean fluorescence of untreated cells.
000 cells per well) were seeded in an 8-well iBidi plate. After 24 hours, complexes 1–5, and Ph2P-CH2-pip-NBD, were added to the cells at concentrations of 9 μM, followed by an incubation period of 6 hours. Afterwards, 150 nM Lysotracker Red (Life Technologies) was added to the cells and let for 2.5 hours. Cells were then washed twice with PBS and images were acquired.
In a parallel experiment, cells were incubated for 30 minutes with chloroquine (100 μM) or genistein (200 μM), followed by the addition of complexes 1, 3, and 4 at a final concentration of 9 μM.
In both experiments, imaging was performed using a Carl Zeiss LSM 980 confocal microscope with excitation lasers set to 488 and 561 nm for complexes 1–5 and Lysotracker Red, respectively.
:
1, dissolved in an azeotropic chloroform/methanol solution (87
:
13 v/v), with previous equilibration of the mixture for 30 minutes at room temperature. Large multilamellar vesicles (MLVs) were prepared by hydration of the lipid film with PBS (10 mM sodium phosphate at pH = 7.4 with 150 mM NaCl) to a total lipid concentration of 10 mM, followed by 3 cycles of freeze/taw (in liquid nitrogen and at 30 °C, respectively). For the preparation of large unilamelar vesicles (LUVs), the MLVs were first subject to 3 cycles of extrusion through two-staked polycarbonate 100 nm pore-size filters (Nucleopore) using a water-jacketed extruder (Lipex Biomembranes, Vancouver, British Columbia, Canada) with freeze/taw between the extrusions, followed by 10 additional extrusion steps.47 The size and surface potential of the LUVs was characterized by dynamic light scattering and zeta potential measurements (Zetasizer Nano ZS, Malvern, UK).
The partition coefficients of the compounds were characterized by addition of the compounds dissolved in dimethylsulfoxide to LUVs solutions pre-warmed at 37 °C and 10 different lipid concentrations, with a final concentration of the compounds of 5 or 15 μM and 1% DMSO. The solutions were allowed to equilibrate for 5 minutes at 37 °C, and their fluorescence at λexc = 440 nm/λem = 590 nm was measured using a plate reader (SpectraMax iD5, Molecular Devices, Berkshire, UK). Blank LUVs solutions containing 1% DMSO were also prepared and the signal subtracted from that of the test solutions. Titrations with a maximum lipid concentration of 5 mM were performed with 100 μL solution in 96 well black/clear bottom plates (ThermoFisher Scientific Inc., Walthem, MA, USA), while for a maximum lipid concentration of 10 mM a spectradrop micro-volume microplate with a light path of 0.5 mm (Molecular Devices, Berkshire, UK) was used to decrease possible artifacts due to high light scattering from the LUVs. The partition coefficient (KP) was obtained from the best fit of eqn (1) for the variation of the fluorescence intensity (I) with the lipid concentration ([L]), which assumes simple partition, additivity in the fluorescence from the compound in the aqueous phase (IW) and associated with the LUVs (IM), and considering a molar lipid volume (
L) equal to 0.76 dm3 mol−1.48
![]() | (1) |
000 cells per well) were seeded in 24-well plates. After 24 hours, compounds 1, 3, and 5 (15 and 25 μM) were added to the cells and incubated for 6 hours. Then, cells were stained with 5 μM of CellROX® Deep Red reagent (Thermo Fisher Scientific, Inc.) for 30 minutes, washed three times with PBS, detached using trypsin–EDTA, and centrifuged for 5 minutes at 2000 rpm. Afterwards, the cells were resuspended in PBS and immediately analyzed by flow cytometry (Agilent, NovoCyte® 3000). The fluorescence associated with the CellROX® red reagent was measured upon excitation with the 640 nm laser, and detection was performed using the 675/30 nm filter. An average of 12
000 events were collected, and the data are presented as the mean fluorescence normalized to the mean fluorescence of untreated cells.
000 cells per well) were seeded in 24-well plates. After 24 hours, compounds 1, 3, and 5 (15 and 25 μM) were added to the cells. Morphological changes were analyzed after 6, 24, and 48 hours of incubation using optical and fluorescence microscopy. For this, 1 μL of Hoechst 33342 (10 mg mL−1, Thermo Fisher Scientific) was added to each well 20 minutes prior to imaging. Subsequently, cell images were immediately acquired using a ZEISS Axiovert 5 microscope equipped with an HXP 120 V fluorescence system and a 20× objective.
CCDC 2476799 (1), 2476798 (3) and 2476797 (5) contain the supplementary crystallographic data for this paper.49a–c
Footnote |
| † Current address: Centro de Química Estrutural, Institute of Molecular Sciences, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais 1, 1049–001 Lisboa, Portugal; and Associação do Instituto Superior Técnico para a Investigação e Desenvolvimento (IST-ID), Avenida António José de Almeida 12, 1000-043 Lisboa Portugal. |
| This journal is © The Royal Society of Chemistry 2026 |