Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

A sodium ion-selective photosensitizer: dibrominated F-BODIPY as a fluorescence imaging and therapeutic agent

Thomas Schwarze*a, Mazen Al Akramia, Julian Heinricha, Vinja Hergla, Eric Sperlicha, Alexandra Kellinga, Tobias Sprengerab, Nicolas Jahna, Tillmann Klamrotha and Nora Kulak*a
aInstitute of Chemistry, University of Potsdam, 14476 Potsdam, Germany. E-mail: schwarth@uni-potsdam.de; nora.kulak@uni-potsdam.de
bFaculty of Medicine, Health and Medical University (HMU), 14471 Potsdam, Germany

Received 18th August 2025 , Accepted 20th September 2025

First published on 22nd September 2025


Abstract

Herein, we report that the production of singlet oxygen (1O2) is exclusively regulated by sodium ions in aqueous solution by the use of a Na+-selective photosensitizer (PS), a 2,6-dibrominated F-BODIPY dye equipped with benzo-15-crown-5. The PS showed an enhanced fluorescence quantum yield (Φf) and an enhanced singlet oxygen quantum yield (ΦΔ) in the presence of Na+. A detailed theoretical study uncovered the underlying photophysical pathways which are responsible for both functional characteristics of the PS, therapeutic and Na+ imaging properties.


1. Introduction

Photodynamic therapy (PDT) is a non-invasive and very powerful method to kill cancer cells by singlet oxygen (1O2) generated through light and a photosensitizer (PS).1,2 Several PSs, mainly porphyrin derivatives are approved for the PDT treatment for different types of cancer such as skin, lung, bladder, and breast cancer.3 In malign breast cancer cells the pH value can be more acidic and the Na+ level is up to five times higher than in benign cells (raising from around 20 mM to over 100 mM Na+).4 A very powerful and non-invasive but costly technique to visualise Na+ in the human body is based on magnetic resonance imaging (MRI) of 23Na.5 A more cost-effective method to image Na+ in vivo is the use of fluorescence spectroscopy.6,7 For precise identification and targeted light irradiation of tumor tissue, a fluorescence imaging-guided PDT is very helpful.8,9 A further class of promising triplet PSs for PDT are based on boron-dipyrromethene (BODIPY) dyes,10–12 when for instance substituted in 2,6-position with heavy atoms such as iodine13–15 or bromine.16 Two decades ago, the group of Akkaya et al. reported on 2,6-dibromo-substituted F-BODIPYs as triplet PSs to efficiently produce 1O2.16 Further, O’Shea et al. published a while ago, that the 1O2 generation rate can be regulated by protons.17 There, a photoinduced electron transfer (PET) is blocked by protonation of an amine donor.17 Moreover, in a pioneering work Akkaya et al. showed that a PS consisting of 2,6-diiodo- and 3,5-dipyridylethenyl-substituted F-BODIPY equipped in meso-position with a benzo-15-crown-5 can modulate and enhance 1O2 production by both H+ and Na+ in acetonitrile (ACN).18 Meanwhile, some factors that control the 1O2 efficiency have been uncovered such as pH, light, hydrogen peroxide, nucleic acids, proteins etc.18–21

In a recent study, we reported on a benzo-15-crown-5-equipped F-BODIPY dye 1a (Fig. 1) for a reliable fluorescence detection of Na+ in the pH range from 3 to 10 by fluorescence enhancement caused by an off-switching of a PET by Na+ in aqueous solution.22 Herein, we now report on a detailed experimental and theoretical study of the regulation of 1O2 exclusively by Na+ and the fluorescence sensing of Na+ by a PS in ACN and aqueous solution. Our overriding goal is to design a PS which shows an enhanced 1O2 production as well as an enhanced fluorescence response only in malign, but not in benign tissue. As a trigger we selected the enhanced Na+ level in breast cancer cells. By fine tuning the Na+ complexing abilities of the Na+-responsive PS (dissociation constant Kd) we aimed to manipulate the 1O2 evolution and fluorescence response. We designed PS 1 to be both, a therapeutic and an imaging agent regulated by the enhanced Na+ level in tumor tissue. PS 1 is a combination of the photostable triplet PS 2,23 a 2,6-dibromo-substituted F-BODIPY dye, and the pH-stable and Na+-selective binding unit 3, benzo-15-crown-524 (Fig. 1).


image file: d5cp03172a-f1.tif
Fig. 1 Studied Na+-selective 2,6-dibrominated F-BODIPY PS 1 (left: molecular structure obtained from XRD) and reference compounds 1a, 2, 2a and 3. H atoms are omitted for clarity.

2. Results and discussion

A bromination at positions 2 and 6 of the F-BODIPY 1a22 with N-bromosuccinimide (NBS) yielded the novel PS 1 in a moderate yield of 47%.25 As references, the F-BODIPYs 2 and 2a without benzo-15-crown-5 moiety were synthesized as described.22,26 Benzo-15-crown-5 (3) is commercially available. The novel PS 1 was characterized by 1H and 13C NMR spectroscopy as well as electrospray ionization mass spectrometry.25 The molecular structure of 1 was confirmed by X-ray analysis (Fig. 1).25 Single crystals of 1 were obtained by slow solvent evaporation (ethyl acetate/hexane, v/v, 1/1). At first, we recorded UV/Vis absorption spectra of 1 and 2 in ACN (Fig. S2a). The absorption spectra of 1 and 2 are very similar, in the range from 350 nm to 550 nm, to each other. They show the most intense absorption band (S0 to S1 transition) with a local maximum (λmax) at about 525 nm (vibronic 0–0 state) with a shoulder at about 490 nm (vibronic 0–1 state).27 The molar extinction coefficients (ελ) at λmax for 1 (77[thin space (1/6-em)]000 M−1 cm−1) and 2 (75[thin space (1/6-em)]000 M−1 cm−1) in ACN are comparable to each other, suggesting that the phenylic substituent in meso-position of the F-BODIPY in 1 does not significantly extend the π-electron system of the F-BODIPY chromophore. As found in the molecular structure of 1 the phenyl ring is almost orthogonal to the planar F-BODIPY core (dihedral angle 86.1°, Fig. 1) which electronically decouples the F-BODIPY from the benzo-15-crown-5. Then, we recorded UV/Vis absorption spectra of 1 and 2 (cdye = 10−5 M and 10−6 M, respectively) in different ACN/water mixtures and found a good solubility of 1 and 2 up to a ACN/water mixture of 1/9 (v/v) (Fig. S2c–f), but 2 showed a blue shift of λmax when the water amount was increased (Fig. 2d and f).25 Thus, 2 is only an appropriate spectroscopic reference compound for 1 in ACN. Moreover, to ensure complete solubility of 1, we decided for further investigation to use as an aqueous solution an ACN/water mixture of 1/3 (v/v). Further, the fluorescence emission maxima of 1 and 2 (c = 10−6 M) were also very similar to each other in ACN (539 nm (1) and 540 (2)) (Fig. S6a), but their fluorescence quantum yields (Φf) differ from each other (Φf = 0.010 (1), Φf = 0.207 (2)).25 The low Φf value of 2 is caused by a heavy atom quenching effect which is typical for a triplet PS.28 Probably, in 1 an additional quenching process, such as in 1a (Φf = 0.25822 in ACN(1a)), a reductive PET from the benzo-15-crown-5 (electron donor) to the excited and decoupled 2,6-dibrominated F-BODIPY core (electron acceptor) occurs.29–31 Solvent effects on the Φf values for 1 are found because the reductive PET in 1 is more favorable in polar solvents (Table S3).25 The low Φf values of 1, 1a and 2 in polar solvents make them suitable candidates as PS to produce efficiently 1O2 in ACN and aqueous solution. Then we monitored the 1O2 production by recording the absorbance of 1,3-diphenylisobenzofuran (DPBF) as a singlet oxygen scavenger at 410 nm in ACN, aqueous solution (ACN/water, v/v, 1/3) and 1,4-dioxane/dimethyl sulfoxide (v/v, 99/1).25 The following singlet oxygen quantum yields (ΦΔ) were calculated: 0.199 ± 0.010 for 1, 0.239 ± 0.010 for 1a and 0.495 ± 0.092 for 2 in ACN, 0.527 ± 0.012 for 1, 0.048 ± 0.002 for 1a and 0.521 ± 0.014 for 2 in 1,4-dioxane/dimethyl sulfoxide (v/v, 99/1) as well as for 1 0.126 ± 0.003 in aqueous solution (ACN/water, v/v, 1/3). The triplet PS 2 generates more 1O2 than 1 and 1a in ACN and exhibits very similar ΦΔ values in both polar and non-polar solvents. The intersystem crossing (ISC) process in 2, caused by the heavy atom effect of the two bromine atoms, results in a well populated triplet state (T1), which is less dependent on the solvent polarity.25,32 The PS 1 exhibits a similar ΦΔ value in non-polar environments to that of 2, and shows a higher ΦΔ value compared to its behaviour in more polar solvents. In contrast, 1a displays the opposite trend: it has a higher ΦΔ value in polar solvents and a lower ΦΔ value in non-polar environments. In 1, two deactivation pathways from the S1 state to the T1 state are conceivable. Firstly, ISC, which is typical for heavy atom containing triplet PS,13,16,28 and secondly, a spin–orbit charge-transfer (SOCT)-ISC process, which predominates in heavy atom-free triplet PS,33–36 such as PET-based PS, where a charge-separated 1CT state is formed and stabilized in polar solvents.34 In general, the SOCT-ISC proceeds much faster than the ordinary ISC between π to π* states.37 For the heavy atom-free triplet PS 1a, we observed a higher ΦΔ value in polar solvents compared to the reference F-BODIPY 2a (ΦΔ = 0.09 in ACN).38 This enhancement is likely due to a SOCT-ISC process facilitated by the polar environment. Moreover, we observed similar ΦΔ values for 1 and 1a in ACN, indicating that in both PS, the SOCT-ISC is likely the predominant pathway from the 1CT state to the T1 state, in 1 the SOCT-ISC process likely predominates in polar solvents whereas conventional ISC is more dominant in non-polar environments.25
image file: d5cp03172a-f2.tif
Fig. 2 Fluorescence intensity (If) of 1 (c = 10−6 M, λex = 500 nm) in the presence of different Na+ concentrations (a) in ACN and (b) in ACN/water, (v/v, 1/3). Fluorescence quantum yields (Φf) (black) and singlet oxygen quantum yields (ΦΔ) (red) of 1 in the presence of different Na+ concentrations (c) in ACN and (d) in ACN/water (v/v, 1/3). Photographs under UV light (366 nm) of 1 (c = 10−6 M) in the presence of different Na+ concentrations (e) in ACN and (f) in ACN/water (v/v, 1/3).

Further, we recorded UV/Vis absorption spectra of 1 (c = 10−5 M) in the presence of Na+ in ACN and in an ACN/water mixture of 1/3 (v/v) (Fig. S3a and b). The absorption at 540 nm (λmax) is nearly unaffected by Na+. The complexation of Na+ within the benzo-15-crown-5 in 1 can be observed by an enhanced blue-shift of the π → π* transition from around 280 nm to 270 nm, (Fig. S3a and b) which is typical for cation complexation of benzo-crown ethers.39 Then, we measured the influence of Na+ on the fluorescence intensity (If), Φf and ΦΔ of 1 in ACN and aqueous solution (ACN/water, v/v, 1/3). The If of 1 is enhanced with increasing Na+ concentrations in ACN and aqueous solution (ACN/water, v/v, 1/3) (Fig. 2a and b). The relative course of both titration curves (λem = 540 nm, Fig. S7b and d) is similar but the maximum FE is reached at different Na+ concentrations, in ACN at 5 mM and in aqueous solution (ACN/water, v/v, 1/3) at 2 M, respectively. The fluorescence enhancement factor (FEF) induced by Na+ in ACN is 11.6 ± 0.1 at 5 mM Na+ and in ACN/water (v/v, 1/3) is 9.1 ± 0.5 at 2 M Na+, respectively. We also observed an enhancement of the Φf values of 1 in the presence of different Na+ concentrations in ACN and aqueous solutions (ACN/water, v/v, 1/3) (Fig. 2c, d and Tables S4, S5). Here, we observed the highest Φf value for 1 at 5 mM Na+ in ACN (Φf = 0.132 ± 0.004) and in aqueous solution (ACN/water, v/v, 1/3) at 2000 mM Na+ (Φf = 0.108 ± 0.016). Probably, the FE is caused by blocking the PET process in 1 by Na+, as also found for 1a + Na+.22 Na+ raises the oxidation potential of the PET electron donor benzo-15-crown-5 in ACN and aqueous solution.40 Therefore, the reductive PET process in 1 + Na+ becomes more unlikely as expressed by the Rehm–Weller equation.30 Moreover, we also determined an enhanced ΦΔ value for 1 in the presence of different Na+ concentrations in both ACN and aqueous solution (ACN/water, v/v, 1/3), (Fig. 2c, d and Tables S1, S2). We also observed the highest ΦΔ value for 1 at 5 mM Na+ in ACN (ΦΔ = 0.367 ± 0.007) and in aqueous solution (ACN/water, v/v, 1/3) at 2000 mM Na+ (ΦΔ = 0.185 ± 0.006). Moreover, we determined for 1a + 5 mM Na+ a ΦΔ value of 0.137 ± 0.006 in ACN which is close to the ΦΔ value of 0.09 of 2a in ACN.38 Overall, we observed for 1 an enhancement of If, Φf and ΦΔ by Na+ and for 1a an enhancement of If and Φf but a reduction of ΦΔ by Na+ in polar solvents. Further, we calculated the limit of detection (LOD) from the fluorescence titration data of 1 + Na+ (LOD = 3σ/m) in ACN and aqueous solution (ACN/water, v/v, 1/3).25 The PS 1 shows a lower sensitivity towards Na+ in ACN with a LOD of (9.45 ± 0.6) μM as in aqueous solution (ACN/water, v/v, 1/3) (11.5 ± 1.1) mM, respectively (Fig. S9a and b). We also found a good linear relationship between the fluorescence intensity of 1 + Na+ in ACN and aqueous solution (ACN/water, v/v, 1/3) (from 0 mM to 0.14 mM Na+, R2 = 0.9966 (ACN), from 0 mM to 100 mM Na+, R2 = 0.9992 (ACN/water, v/v, 1/3), Fig. S9a and b) at 540 nm, respectively. More importantly, we calculated from the fluorescence intensity changes of 1 + Na+ their dissociation constants (Kd) in ACN and in aqueous solution (ACN/water, v/v, 1/3) resulting in Kd values of (0.16 ± 0.02) mM and (209 ± 5) mM, respectively.25 The latter Kd value of 1 + Na+ in aqueous solution is biologically relevant, since it is close to the Na+ level in malign breast cancer cells.4 The Kd value of 1 + Na+ is significantly lower in ACN than in aqueous solution caused by the fact that a solvent like ACN that does not coordinate strongly with Na+ and a complexation of Na+ within the benzo-15-crown-5 is less hampered. In addition to it, the slopes of the plots for 1 + Na+ (log[thin space (1/6-em)](cNa+) vs. log[thin space (1/6-em)][(IfIf[thin space (1/6-em)]min)/(If[thin space (1/6-em)]maxIf)]) in ACN and aqueous solution (ACN/water, v/v, 1/3) were nearly 1 (Fig. S8a and b),25 suggesting a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 binding ratio between Na+ and 1. Moreover, to elucidate the binding stoichiometry between 1 with NaClO4 in solution, we carried out 1H NMR experiments in CD3CN (Fig. S12).25 Thus, a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 binding stoichiometry of 1 with NaClO4 was confirmed by a Job's plot analysis (Fig. S13).25 We observed a downfield shift of the benzo-15-crown-5 protons until one equivalent NaClO4 in the 1H NMR spectra of 1 (Fig. S12) assuming that Na+ is coordinated within the benzo-15-crown-5 in 1.

EPR experiments were carried out with 2,2,6,6-tetramethylpiperidine (TEMP) as a 1O2 specific spin-trap agent.25 It was added to 1 and 1 + 5 mM NaClO4, and a strong EPR signal of 2,2,6,6-tetramethylpiperidinyloxyl (TEMPO) was observed after light irradiation in ACN (Fig. S30). We found for 1 + 5 mM NaClO4 a two times higher intensity of the TEMPO signal at 336.58 mT than for 1 indicating that in the presence of Na+ more 1O2 is produced.

We further investigated the influence of varying aqueous pH values on the fluorescence performance of 1.25 1 shows very stable invariant fluorescence emission signals in the pH value range from 3.04 to 10.04 (Fig. S11a). Moreover, we observed for 1 in ACN and aqueous solution (ACN/water, v/v, 1/3) over a time period of 360 min a relatively photostable fluorescence signal at 540 nm (Fig. S11b) meaning that the photobleaching of 1 is negligible. To verify selectivity of 1 for Na+ towards other important biological cations such as Li+, K+, NH4+, Mg2+, Ca2+, Mn2+, Fe3+, Cu2+ and Zn2+, we measured the fluorescence intensities in the presence of these cations at their respective concentrations that are biologically relevant in aqueous solution (ACN/water, v/v, 1/3).25 The fluorescence performance of 1 is only slightly impacted (Fig. S10), showing that 1 is a Na+-selective fluorescent imaging tool.

Moreover, we tested the DNA cleavage activity of 1 (c = 200 μM) with plasmid DNA at pH 7.4 with or without green light irradiation in the absence or presence of NaCl (Fig. 3) or NaClO4 (Fig. S28). Degradation of supercoiled DNA (form I) to open-circular/nicked (form II) and linear DNA (form III) was monitored via gel electrophoresis.25 Under green light irradiation, we observed DNA cleavage by 1 (lane j) forming 69% DNA form II (single-strand breaks) and even 1% form III (double-strand breaks). When the sample was not irradiated, no cleavage activity of 1 was observed (lane d about 40% form II). Surprisingly, the cleavage activity is not enhanced by NaCl (lanes k and l, Fig. 3) or NaClO4 (Fig. S28). Probably, the stabilisation of the negatively charged DNA double helix (phosphate backbone) by Na+ due to electrostatic interactions41 results in a lower DNA cleavage activity of 1.

Further, we crystallized 1 with NaClO4 in a molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1 from a chloroform/acetonitrile (v/v, 3/1) mixture to get more insights on the binding characteristics of Na+ within 1. X-ray analysis provided the molecular structure of the Na+ complex [Na(1)(ClO4)] (Fig. 4). Na+ is mainly coordinated by the five oxygen atoms of the benzo-15-crown-5 in 1 and shows a good fit-in-size into the cavity (Fig. 4a). Notably, the two symmetry-equivalent bridging perchlorate anions are disordered which influences the total number of coordination bonds of the Na+ (Fig. S23 and S24). We also found an electronic decoupling of the F-BODIPY from the Na+ complexed benzo-15-crown-5 unit because in the molecular structure of [Na(1)(ClO4)] the phenyl ring is almost orthogonal to the planar F-BODIPY core (dihedral angle 82.2°, Fig. 4b).


image file: d5cp03172a-f3.tif
Fig. 3 (a) Nuclease activity towards plasmid DNA pBR322 (0.025 μg μL−1) of PS 1 (c = 200 μM) in Tris buffer (5 mM, pH 7.4) w/wo NaCl (25 or 100 mM). Samples in lanes g–l were incubated under irradiation by green light for 50 min whereas samples in lanes a-f were not irradiated. Lane a: DNA ladder (form I, II and III), lane b + g: DNA reference, lanes c + i: 100 mM NaCl, lanes d + j: 1, lanes e + k: 1 + 25 mM NaCl, lanes f + l: 1 + 100 mM NaCl, lane h: 25 mM NaCl. (b) Visualization of the extent of DNA cleavage in percent with standard deviation as error bars.

image file: d5cp03172a-f4.tif
Fig. 4 Molecular structure of [Na(1)ClO4] with a space-filling model of Na (crystal radius42 regarding coordination number). (a) Front view and (b) side view. H atoms are omitted for clarity.

Complementary to the experiments, we performed (time-dependent) density functional theory [(TD-)DFT] and singlet/triplet spin–orbit coupling (SOC) calculations of 1, a dibromine-free F-BODIPY dye 1a and 2 at the B3LYP/def2-TZVP level of theory43–45 in ORCA 6.0.25,46 The bright S1 state of 1 in Fig. 5c is given by a local transition on the BODIPY part from MODye to the LUMO, while the optically dark 1CT state shows strong charge transfer character from MOCT to the LUMO. We find that the addition of Na+ leads to an energetic stabilization of the MOCT, while the two BODIPY-localized MOs remain mostly unaffected as shown in Fig. 5d due to the greater spatial distance to the crown ether part of 1. The excitation energy of the 1CT state is thus increased relative to the bright S1 state after Na+ complexation in agreement with the reported experimental findings.25 Furthermore, bromination leads to a one order of magnitude increase in the computed singlet/triplet SOCs of 1 compared to 1a due to the heavy-atom effect of the bromine atoms.25

Overall, the fluorescence quenching observed for 1 is likely due to a reductive PET process. In polar solvents, a 1CT state is formed and stabilized, and its conversion to T1 state via a SOCT-ISC mechanism is probable. The resulting T1 state subsequently generates a moderate amount of 1O2 in polar solvents. Furthermore, we assume that Na+ interrupts the reductive PET process in 1 in polar solvents, leading to an increase in the energy of the 1CT state. As a result, population of the T1 state via ISC, facilitated by the heavy atoms (bromine), becomes more favourable and efficient, thereby restoring both fluorescence and 1O2 generation of the dibrominated F-BODIPY core (Φf and ΦΔ values of 2 in polar media). As a result we observed for 1 + Na+ higher Φf and ΦΔ values compared to 1 without Na+ (Fig. 5a and b). The latter can lead to degradation of DNA under irradiation (Fig. 3a and b). Moreover, we found for the dibromine-free F-BODIPY dye 1a in the presence of Na+ also an enhanced Φf value but a reduced ΦΔ value in polar solvents. The presence of Na+ blocks the reductive PET process in 1a, resulting in an elevation of the 1CT energy level. This effectively restores the fluorescence of the dibromine-free F-BODIPY core, where intersystem crossing (ISC) is considered highly unlikely (Fig. S34a and b). To the best of our knowledge, this is the first report that only a metal ion, here Na+, regulates 1O2 evolution. The enhanced 1O2 production by Na+ can be useful to selectively kill malign cancer cells after irradiation when the Kd value of the Na+-selective PS fits to the Na+ levels in the cancer cells.


image file: d5cp03172a-f5.tif
Fig. 5 Jablonski diagram of the postulated mechanism of the photosensitized production of 1O2 in the PS 1 in polar solvents (a) without Na+ and (b) with Na+. (c) Molecular orbitals (MOs) of 1 corresponding to the S1 and 1CT excited states (see further explanations in the text). (d) Relative MO energy level changes of 1 due to the addition of Na+.

3. Conclusions

In summary, we synthesized the novel and Na+ selective PS 1 consisting of a benzo-15-crown-5 and a dibrominated F-BODIPY dye shows a fluorescence signal which is photostable and invariant to a wide pH value range from 3.04 to 10.04. Further, 1 is a fluorescent tool with high Na+ selectivity and Na+ sensitivity, fast Na+ response and the Na+ induced fluorescence enhancement is even recognizable with the naked eye after irradiation with UV light. Moreover, we observed higher Φf and ΦΔ values for 1 in the presence of Na+ in polar solvents. The Kd value of 1 + Na+ is (209 ± 5) mM in aqueous solution and fits better to the Na+ level in malign cancer cells (around 100 mM Na+) than to benign cells (around 20 mM Na+).4 PS 1 is a suitable therapeutic as well as a Na+ imaging agent. 1 could be a useful PS for cancer therapy because 1 could image tumorous tissue, and targeted light irradiation would selectively kill cancer cells through 1O2 generation. Currently, we are designing PSs applicable in PDT for a deeper tissue penetration by extending the π-system in position 2 and 5 of the F-BODIPY core to shift absorbance to the near-infrared (NIR) region.47

Author contributions

Thomas Schwarze: conceptualization, methodology, investigation, formal analysis, writing – original draft; Mazen Al Akrami: investigation, data curation; Julian Heinrich: formal analysis, data curation; Vinja Hergl: investigation; Alexandra Kelling: formal analysis; Eric Sperlich: formal analysis, visualisation, methodology; Tobias Sprenger: formal analysis; Nicolas Jahn: formal analysis, visualisation; Tillmann Klamroth: supervision; Nora Kulak: funding acquisition, supervision, writing – review & editing.

Conflicts of interest

The authors declare no conflict of interest.

Data availability

The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: synthesis, data from NMR, EPR, UV/Vis and fluorescence spectroscopy, cyclic voltammetry, single crystal X-ray diffraction, DNA cleavage experiments and DFT calculations. See DOI: https://doi.org/10.1039/d5cp03172a.

CCDC 2457233 and 2477061 contain the supplementary crystallographic data for this paper.48a,b

Acknowledgements

The authors thank Matthias Hartlieb for providing access to a PhotoCube reactor (ThalesNano). The German Research Foundation (DFG) is acknowledged for funding within SFB 1636 (Project ID 510943930) for establishing EPR experiments under irradiation.

Notes and references

  1. Y. Allamyradov, J. ben Yosef, B. Annamuradov, M. Ateyeh, C. Street, H. Whipple and A. O. Er, Photochemistry, 2024, 4, 434 Search PubMed.
  2. J. F. Algorri, M. Ochoa, P. Roldán-Varona and J. M. López-Higuera, Cancers, 2021, 13, 4447 CrossRef.
  3. A.-G. Niculescu and A. M. Grumezescu, Appl. Sci., 2021, 11, 3626 CrossRef.
  4. R. Ouwerkerk, M. A. Jacobs, K. J. Macura, A. C. Wolff, V. Stearns, S. D. Mezban, N. F. Khouri, D. A. Bluemke and P. A. Bottomley, Breast Cancer Res. Treat., 2007, 106, 151 CrossRef PubMed.
  5. L. O. Poku, M. Phil, Y. Cheng, K. Wang and X. Sun, J. Magn. Reson. Imaging, 2021, 53, 995 CrossRef PubMed.
  6. J. M. Dubach, E. Lim, N. Zhang, K. P. Francis and H. Clark, Integr. Biol., 2011, 3, 142 CrossRef PubMed.
  7. O. Iamshanova, P. Mariot, V. Lehenkyi and N. Prevarskaya, Eur. Biophys. J., 2016, 45, 765 CrossRef PubMed.
  8. J. P. Celli, B. Q. Spring, I. Rizvi, C. L. Evans, K. S. Samkoe, S. Verma, B. W. Pogue and T. Hasan, Chem. Rev., 2010, 110, 2795 CrossRef PubMed.
  9. H. Luo and S. Gao, J. Control. Release, 2023, 362, 425 CrossRef PubMed.
  10. A. Kamkaew, S. H. Lim, H. B. Lee, L. V. Kiew, L. Y. Chung and K. Burgess, Chem. Soc. Rev., 2013, 42, 77 RSC.
  11. Z. Mao, J. H. Kim, J. Lee, H. Xiong, F. Zhang and J. S. Kim, Coord. Chem. Rev., 2023, 476, 214908 CrossRef.
  12. E. Bassan, A. Gualandi, P. G. Cozzi and P. Ceroni, Chem. Sci., 2021, 12, 6607 RSC.
  13. T. Yogo, Y. Urano, Y. Ishitsuka, F. Maniwa and T. Nagano, J. Am. Chem. Soc., 2005, 127, 12162 CrossRef CAS PubMed.
  14. J. Piskorz, W. Porolnik, M. Kucinska, J. Dlugaszewska, M. Murias and J. Mielcarek, ChemMedChem, 2021, 16, 399 CrossRef CAS PubMed.
  15. P. Rybczynski, A. Smolarkiewicz-Wyczachowski, M. Ziegler-Borowska, D. Kedziera, J. Piskorz, S. Bocian and A. Kaczmarek-Kedziera, Int. J. Mol. Sci., 2021, 22, 6735 CrossRef CAS.
  16. S. Atilgan, Z. Ekmekci, A. L. Dogan, D. Guc and E. U. Akkaya, Chem. Commun., 2006, 4398 RSC.
  17. S. O. McDonnell, M. J. Hall, L. T. Allen, A. Byrne, W. M. Gallagher and D. F. O’Shea, J. Am. Chem. Soc., 2005, 127, 16360 CrossRef CAS.
  18. S. Ozlem and E. U. Akkaya, J. Am. Chem. Soc., 2009, 131, 48 CrossRef CAS.
  19. S. Callaghan and M. O. Senge, Photochem. Photobiol. Sci., 2018, 17, 1490 CrossRef CAS.
  20. W. Wu, X. Shao, J. Zhao and M. Wu, Adv. Sci., 2017, 4, 1700113 CrossRef PubMed.
  21. N. Kwon, H. Weng, M. A. Rajora and G. Zheng, Angew. Chem., Int. Ed., 2025, 64, e202423348 Search PubMed.
  22. T. Sprenger, T. Schwarze, H. Müller, E. Sperlich, A. Kelling, H.-J. Holdt, J. Paul, V. Martos Riaño and M. Nazaré, ChemPhotoChem, 2023, 7, e202200270 CrossRef CAS.
  23. E. N. Nuraneeva, E. V. Antina, G. B. Guseva, M. B. Berezin and A. I. Vyugin, Inorg. Chim. Acta, 2018, 482, 800 CrossRef CAS.
  24. R. M. Izatt, R. E. Terry, D. P. Nelson, Y. Chan, D. J. Eatough, J. S. Bradshaw, L. D. Hansen and J. J. Christensen, J. Am. Chem. Soc., 1976, 98, 7626 CrossRef CAS.
  25. For detailed experimental and theoretical data see SI.
  26. E. N. Nuraneeva, G. B. Guseva, E. V. Antina, O. A. Lodochnikova, D. R. Islamov and L. E. Nikitina, Dyes Pigm., 2022, 201, 110202 CrossRef CAS.
  27. W. Hu, X.-F. Zhang, X. Lu, S. Lan, D. Tian, T. Li, L. Wang, S. Zhao, M. Feng, J. Zhang and J. Lumin, J. Lumin., 2018, 194, 185 CrossRef CAS.
  28. Y. P. Rey, D. G. Abradelo, N. Santschi, C. A. Strassert and R. Gilmour, Eur. J. Org. Chem., 2017, 2170 CrossRef CAS.
  29. For 1, a reductive PET process from the benzo-15-crown-5 to the dibrominated F-BODIPY core has a negative ΔGPET value of −0.03 eV according to the Rehm-Weller equation ΔGPET = EoxEred − ΔE00 − ΔGion-pair.27 The oxidation potential (Eox) of benzo-15-crown-5 is 1.00 V and the reduction potential (Ered) of 2 is −1.20 V vs. Ag/AgNO3 as the reference electrode in ACN.23 In ACN, the ΔE00 value of 2 (533 nm) is about 2.33 eV (transition energy between the vibrationally relaxed ground and excited state of the fluorophore).23 The attractive energy, ΔGion-pair is assumed to be −0.1 V.28.
  30. D. Rehm and A. Weller, Isr. J. Chem., 1970, 8, 259 CrossRef CAS.
  31. A. P. de Silva, H. Q. N. Gunaratne, J.-L. Habib-Jiwan, C. P. McCoy, T. E. Rice and J.-P. Soumillion, Angew. Chem., 1995, 107, 1889 CrossRef.
  32. W. Hu, R. Zhang, X.-F. Zhang, J. Liu and L. Luo, Spectrochim. Acta, Part A, 2022, 272, 120965 CrossRef CAS PubMed.
  33. X. Zhang, Z. Wang, Y. Hou, Y. Yan, J. Zhao and B. Dick, J. Mater. Chem. C, 2021, 9, 11944 RSC.
  34. M. A. Filatov, Org. Biomol. Chem., 2020, 18, 10 RSC.
  35. P. P. Chebotaev, A. A. Buglak, A. Sheehan and M. A. Filatov, Phys. Chem. Chem. Phys., 2024, 26, 25131 RSC.
  36. A. A. Buglak, M. O. Senge, A. Charisiadis, A. Sheehan and M. A. Filatov, Chem. – Eur. J., 2021, 27, 9934 CrossRef CAS.
  37. J. T. Buck, A. M. Boudreau, A. DeCarmine, R. W. Wilson, J. Hampsey and T. Mani, Chem, 2019, 5, 138 CAS.
  38. M. A. Filatov, T. Mikulchyk, M. Hodee, M. Dvoracek, V. N. K. Mamillapalli, A. Sheehan, C. Newman, S. M. Borisov, D. Escudero and I. Naydenova, J. Mater. Chem. C, 2025, 13, 6993 RSC.
  39. C. J. Pederson, J. Am. Chem. Soc., 1967, 89, 7017 CrossRef.
  40. S. Kenmoku, Y. Urano, K. Kanda, H. Kojima, K. Kikuchi and T. Nagano, Tetrahedron, 2004, 60, 11067 CrossRef CAS.
  41. Z.-J. Tan and S.-J. Chen, Biophys. J., 2006, 90, 1175 CrossRef CAS PubMed.
  42. R. D. Shannon, Acta Cryst., 1976, A32, 751 CrossRef CAS.
  43. A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS.
  44. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785 CrossRef CAS PubMed.
  45. F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2005, 7, 3297–3305 RSC.
  46. F. Neese, WIREs Comput. Mol. Sci., 2022, 12, e1606 CrossRef.
  47. W. Porolnik, W. Szczolko, T. Koczorowski, M. Falkowski, E. Wieczorek-Szweda and J. Piskorz, Spectrochim. Acta, Part A, 2024, 314, 124188 CrossRef CAS PubMed.
  48. (a) CCDC 2457233: Experimental Crystal Structure Determination, 2025 DOI:10.5517/ccdc.csd.cc2ngylp; (b) CCDC 2477061: Experimental Crystal Structure Determination, 2025 DOI:10.5517/ccdc.csd.cc2p4l6n.

This journal is © the Owner Societies 2025
Click here to see how this site uses Cookies. View our privacy policy here.