Abstract
A series of Ru(II) complexes incorporating two 4,4′-bis(trifluoromethyl)-2,2′-bipyridine (4,4′-btfmb) coligands and thienyl-appended imidazo[4,5-f][1,10]phenanthroline (IP-nT) ligands was characterized and assessed for phototherapy effects toward cancer cells. The [Ru(4,4′-btfmb)2(IP-nT)]2+ scaffold has greater overall redox activity compared to Ru(II) polypyridyl complexes such as [Ru(bpy)3]2+. Ru-1T–Ru-4T have additional oxidations due to the nT group and additional reductions due to the 4,4′-btfmb ligands. Ru-2T–Ru-4T also exhibit nT-based reductions. Ru-4T exhibits two oxidations and eight reductions within the potential window of −3 to +1.5 V. The lowest-lying triplets (T1) for Ru-0T–2T are metal-to-ligand charge transfer (3MLCT) excited states with lifetimes around 1 μs, whereas T1 for Ru-3T–4T is longer-lived (~20–24 μs) and of significant intraligand charge-transfer (3ILCT) character. Phototoxicity toward melanoma cells (SK-MEL-28) increases with n, with Ru-4T having a visible EC50 value as low as 9 nM and PI as large as 12,000. Ru-3T and Ru-4T retain some of this activity in hypoxia, where Ru-4T has a visible EC50 as low as 35 nM and PI as high as 2,900. Activity over six biological replicates is consistent and within an order of magnitude. These results demonstrate the importance of lowest-lying 3ILCT states for phototoxicity and maintaining activity in hypoxia.
Keywords: Ruthenium polypyridyl complexes, photosensitizers (PS), photobiology, intraligand charge-transfer (ILCT), ligand-to-ligand charge transfer (LLCT), metal-to-ligand charge transfer (MLCT), intraligand (IL), melanoma, photodynamic therapy (PDT), hypoxia
Graphical Abstract

A series of Ru(II) polypyridyl complexes incorporating two 4,4′-bis(trifluoromethyl)-2,2′-bipyridine (4,4′-btfmb) coligands and a variable thienyl-appended imidazo[4,5-f][1,10]phenanthroline (IP-nT) ligand was characterized and assessed for phototherapy effects toward cancer cells. Ru-4T has a visible EC50 value as low as 9 nM and PI as large as 12,000.
1. INTRODUCTION
Cancer continues to be a leading cause of mortality worldwide.1 Despite substantial progress in available medical treatments,2–6 there remains a critical need for innovative therapeutic strategies and adjuvants to complement traditional methods like surgery, radiotherapy, and chemotherapy. In this regard, photodynamic therapy (PDT) is a possible avenue for targeted cancer therapy. It employs a nontoxic photosensitizer (PS), light, and molecular oxygen to produce cytotoxic reactive oxygen species (ROS), importantly singlet oxygen (1O2), that destroy tumors and tumor vasculature and can induce antitumor immunity.7–10 Phototoxic effects are restricted to oxygenated areas where light is delivered and PS has accumulated. The localized action of PDT results in fewer side effects and improved quality of life for patients.11,12 PDT can be enhanced by tumor-selective PSs and precise delivery of light. In addition, the light protocol can be optimized by altering the drug-to-light interval (DLI) and tuning the wavelength, fluence, and irradiance.
There is also a drive to develop light-responsive compounds that generate ROS at low oxygen tension or function via alternate mechanisms that complement the 1O2 pathway. This capacity would allow effective treatment of hypoxic tumors and the use of high light irradiance. Metal complexes, particularly Ru(II) polypyridyl systems, are of significant interest in this regard. The strategic combination of ligands and metals allows access to a range of excited-state configurations with unique photophysical characteristics. Approaches have included the photorelease of bulky ligands to expose phototoxic metals and/or ligands,13–21 photocaging of chemotherapeutics and enzyme inhibitors,18,22–43 photoredox reactions,44,45 and enhancing ROS production even under low oxygen conditions.17,20,46
Our research group has a strong focus on metal complexes as PSs for their diverse mechanisms of action. Their modular design and relatively straightforward synthesis allow for rapid alteration of their physicochemical, photophysical, and biological characteristics and aligns with our tumor-specific approach to PS development. We advocate that no singular ideal PS exists; rather, PS design should be tailored to the intended application. Our Ru(II) polypyridyl complex, TLD1433, is a prime example, currently in Phase II clinical trials for treating non-muscle invasive bladder cancer (NMIBC) with PDT (Chart 1a).11,47,48 It exhibits significant phototoxicity toward cancer cells while maintaining low dark toxicity. Clinically, it is activated using green light to prevent harm to underlying muscle tissue.
Chart 1.

(a) Structure of Racemic TLD1433 and (b) Structures of Racemic Complexes in This study.
To expand our understanding of oligothiophene-containing coordination complexes, we are examining variations in their central metals, coordinating ligands, thienyl groups, counterions, and coordination geometries.11,14,17,20,46,49–51 We aim to establish structure-activity relationships (SARs) for medicinally active complexes that reconcile their physicochemical, photophysical, electrochemical, and biological properties with the ideal phototherapy profile for a given clinical setting. In this study we introduce a series of Ru(II) PSs (Ru-nT), each featuring two 4,4′-bis(trifluoromethyl)-2,2′-bipyridine (4,4′-btfmb) coligands and a thienyl-containing imidazo[4,5-f][1,10]phenanthroline (IP-nT) ligand with n=1–4 thiophene rings. The members are compared to the parent [Ru(4,4′-btfmb)2(LL)]2+ complexes with LL=IP-0T, 1,10-phenanthroline (phen), or 4,4′-btfmb (Chart 1b). The syntheses and structural characterization of complexes Ru-3T and Ru-4T were previously published.52,53 In this study the photocytotoxic effects of [Ru(4,4′-btfmb)2(IP-nT)]2+ (n=1–4) and the parent complexes without nT groups are examined on melanoma cells under various light conditions and oxygen levels. We also report on their lipophilicities, steady-state absorption and emission characteristics, excited-state properties, electronic configurations, and electrochemical profiles.
2. MATERIALS AND METHODS
The complexes in this series were characterized by 1H NMR, high-performance liquid chromatography (HPLC), and ESI+ mass spectrometry (MS). They were evaluated for lipophilicity, ground and excited-state characteristics using absorption and emission spectroscopy, electrochemistry, and (photo)cytotoxicity. Further details regarding procedures and characterization data are available in the Supplementary Information (SI).
2.1. Instrumentation
A CEM Discover microwave reactor was used to perform microwave reactions. Flash column chromatography was carried out on the Teledyne ISCO EZ Prep UV model of CombiFlash® EZ Prep using SILICYCLE SiliaSep 25 g prepacked silica cartridges. Size-exclusion chromatography was performed using a gravity column packed with Sephadex® LH-20. The NMR spectra were collected on JEOL 500 MHz spectrometers (University of North Carolina at Greensboro, University of Texas at Arlington) operating at 500 MHz for 1H experiments, and on an Agilent 700 MHz Magnet spectrometer (The Joint School of Nanoscience and Nanoengineering at Greensboro) operating at 700 MHz for 1H experiments. The chemical shifts are reported in parts per million (ppm) and were referenced to the residual solvent peaks. High-resolution ESI+ mass spectra were obtained using a Thermo Fisher Scientific LTQ Orbitrap XL instrument (Triad Mass Spectrometry Laboratory at University of North Carolina at Greensboro) and Shimadzu IT-TOF instrument (Shimadzu Center for Advanced Analytical Chemistry at University of Texas at Arlington). HPLC analyses were carried out on an Agilent/Hewlett Packard 1100 series instrument in 100 μM solutions in methanol using a Hypersil GOLD C18 reversed-phase column with an A→B gradient (98% → 5% A; A=0.1% formic acid in H2O, B=0.1% formic acid in MeCN). Reported retention times are accurate to within ±0.1 min.
2.2. Synthesis and Characterization
We previously published the synthetic methods and structural characterization of Ru-3T and Ru-4T,52,53 and Ru(4,4′-btfmb)3]2+ has been reported by others.54–57 To the best of our knowledge, all other complexes presented in this study have not been reported. All complexes were prepared as racemic mixtures. Solvents and reagents were purchased from commercial sources and used without further purification. Water used for all biological experiments was deionized to a resistivity ≥ 18.2 MΩ using either a Barnstead or Milli-Q® filtration system. Methanol was purchased from Fisher Scientific (ACS grade for synthesis, HPLC grade for LC eluent, Optima grade for HPLC and MS sample preparation). Deuterated solvents for NMR were purchased from Cambridge Isotope Laboratories. Ruthenium(III) trichloride trihydrate was purchased from Ark Pharm and Acros Organics. Ru(4,4′-btfmb)2Cl2•2H2O58 and IP-based ligands59 were prepared according to adapted literature procedures. The synthesis of IP-based ligands follows that described below for IP-4T. [2,2′:5′,2″:5″,2‴-Quaterthiophene]-5-carbaldehyde (4T-CHO) was prepared as previously described.60,61 The final products are synthetically characterized in Figures S1–S21 via 1H NMR, 1H–1H COSY NMR, HPLC, and ESI+–MS. The Cl− salts of final complex products were obtained via anion metathesis on HCl-treated Amberlite IRA-410 resin with methanol as eluent and isolated in vacuo. Final complexes are a mixture of Δ/Λ isomers.
rac-[Ru(4,4′-btfmb)3](Cl)2 Ru(Cl)3·xH2O (58 mg, 0.2 mmol) and 4,4′-btfmb (175 mg, 0.6 mmol) was added to a microwave vessel containing argon-purged ethylene glycol (3 mL), and then the mixture was subject to microwave irradiation at 180 °C for 45 min with stirring. The resulting dark red solution was then transferred to a separatory funnel with deionized water (25 mL) and CH2Cl2 (25 mL). After gentle agitation, the CH2Cl2 was drained and the remaining aqueous layer was washed with CH2Cl2 (25 mL) until the CH2Cl2 layer was colorless. Then, CH2Cl2 (25 mL) and saturated aqueous KPF6 (5 mL) was added, and the mixture was shaken gently. The CH2Cl2 layer was drained and the product was further extracted from the aqueous layer with CH2Cl2 (25 mL) until the aqueous layer was colorless. The CH2Cl2 extracts were then combined and concentrated under reduced pressure. The crude product was then purified using silica gel flash column chromatography with a gradient of MeCN to 10% water in MeCN, followed by 7.5% water in MeCN with 0.5% KNO3. The product-containing fractions were then combined and concentrated under vacuum, then transferred to a separatory funnel with CH2Cl2 (25 mL), deionized water (25 mL), and saturated aqueous KPF6 (1 mL). The resulting mixture was gently agitated and the CH2Cl2 layer was drained. Additional CH2Cl2 (25 mL) was used to extract the remaining product until the aqueous layer was colorless. The CH2Cl2 layers were then combined and dried under vacuum. This was then converted to the corresponding Cl− salt in quantitative yield using Amberlite IRA-410 with MeOH as the eluent, then purifying further using Sephadex LH-20 with MeOH as the eluent, affording a dark red solid (50 mg, 20%). 1H NMR (400 MHz, MeOD-d3, ppm): δ 9.41 (d, J=1.9 Hz, 6H), 8.15 (d, J=5.9 Hz, 6H), 7.86 (dd, J=6.0, 1.8 Hz, 6H). HRMS (ESI+) m/z for [M-2Cl−]2+ calcd: 489.0169; found: 488.9529. HPLC retention time: 22.41 min (99.5% purity by peak area).
rac-[Ru(4,4′-btfmb)2(phen)](Cl)2 (Ru-phen) Ru(4,4′-btfmb)2Cl2·2H2O (91 mg, 0.12 mmol) and phen (22 mg, 0.1 mmol) were added to a microwave vessel containing argon-purged ethylene glycol (4 mL) and subjected to microwave irradiation at 180 °C for 15 min. The resulting dark red mixture was then isolated and purified in the same manner as [Ru(4,4′-btfmb)3](Cl)2, yielding the desired product as a dark red solid (43 mg, 45%). 1H NMR (400 MHz, MeOD-d3, ppm): δ 9.38 (d, J=14.1 Hz, 4H), 8.81 (d, J=8.0 Hz, 2H), 8.36 (s, 2H), 8.28 (d, J=5.9 Hz, 2H), 8.23 (d, J=5.3 Hz, 2H), 7.94–7.84 (m, 6H), 7.67 (dd, J=6.0, 1.8 Hz, 2H). HRMS (ESI+) m/z for [M-2Cl−]2+ calcd: 433.0295; found: 432.9739. HPLC retention time: 22.41 min (99.5% purity by peak area).
rac-[Ru(4,4′-btfmb)2(IP)](Cl)2 (Ru-0T) Ru(4,4′-btfmb)2Cl2·2H2O (91 mg, 0.12 mmol) and IP (22 mg, 0.1 mmol) were added to a microwave vessel containing argon-purged ethylene glycol (4 mL) and subjected to microwave irradiation at 180 °C for 15 min. The resulting dark red mixture was then isolated and purified in the same manner as [Ru(4,4′-btfmb)3](Cl)2, yielding the desired product as a dark red solid (81.3 mg, 51%). 1H NMR (400 MHz, MeOD-d3, ppm): δ 9.38 (dd, J=16.6, 1.9 Hz, 4H), 9.10 (d, J=8.4 Hz, 2H), 8.68 (s, 1H), 8.29 (d, J=5.9 Hz, 2H), 8.15 (dd, J=5.3, 1.3 Hz, 2H), 7.95 (d, J=6.0 Hz, 2H), 7.92 (dd, J=8.3, 5.3 Hz, 2H), 7.87 (dd, J=6.0, 1.9 Hz, 2H), 7.66 (dd, J=6.0, 1.9 Hz, 2H). HRMS (ESI+) m/z for [M-2Cl−]2+ calcd: 453.0326; found: 452.9735. HPLC retention time: 20.33 min (>98% purity by peak area).
rac-[Ru(4,4′-btfmb)2(IP-1T)](Cl)2 (Ru-1T). Ru(4,4′-btfmb)2Cl2·2H2O (91 mg, 0.12 mmol) and IP-1T (30 mg, 0.1 mmol) were added to a microwave vessel containing ethylene glycol (4 mL) and subjected to microwave irradiation at 180 °C for 15 min. The dark red solution was transferred to a separatory funnel with H2O (25 mL) and CH2Cl2 (25 mL). CH2Cl2 layer was used to wash the aqueous layer. CH2Cl2 (25 mL) and saturated aqueous KPF6 (5 mL) were used to extract the product from the aqueous layer. The product was then purified using silica gel flash column chromatography. The product-containing fractions were transferred to a separatory funnel with CH2Cl2 (25 mL), H2O (25 mL), and saturated aqueous KPF6 (1 mL) and the [Ru(4,4′-btfmb)2(IP-1T)](PF6)2 product was isolated via extraction. The PF6− salt was then converted to the corresponding Cl− salt in quantitative yield using Amberlite IRA-410 with MeOH as the eluent, followed by further purification using Sephadex LH-20 with MeOH as the eluent, yielding [Ru(4,4′-btfmb)2(IP-1T)](Cl)2 as a dark red solid (43 mg, 25%). 1H NMR (400 MHz, MeOD-d3, ppm): δ 9.40 (d, J=1.9 Hz, 2H), 9.36 (d, J=1.8 Hz, 2H), 9.19 (s, 2H), 8.29 (d, J=5.8 Hz, 2H), 8.12 (d, J=5.2 Hz, 2H), 8.01 (d, J=3.6 Hz, 1H), 7.97 (d, J=6.1 Hz, 2H), 7.91 (dd, J=8.3, 5.2 Hz, 2H), 7.88 (dd, J=6.0, 1.9 Hz, 2H), 7.74 (d, J=5.0 Hz, 1H), 7.68 (dd, J=6.1, 1.8 Hz, 2H), 7.31 (dd, J=5.0, 3.7 Hz, 1H). HRMS (ESI+) m/z for [M-2Cl]2+ calcd: 494.0265; found: 494.0261, m/z for [M-2Cl-H]+ calcd: 987.0456; found: 987.0474. HPLC retention time: 21.75 min (99% purity by peak area).
rac-[Ru(4,4′-btfmb)2(IP-2T)](Cl)2 (Ru-2T). Ru(4,4′-btfmb)2Cl2·2H2O (91 mg, 0.12 mmol) and IP-2T (38 mg, 0.1 mmol) were added to a microwave vessel containing argon-purged ethylene glycol (4 mL) and subjected to microwave irradiation at 180 °C for 15 min. The resulting dark red mixture was then isolated and purified in the same manner as Ru-1T, yielding the desired product as a dark solid (33 mg, 29%). 1H NMR (400 MHz, MeOD-d3, ppm): δ 9.40 (d, J=1.9 Hz, 2H), 9.37 (d, J=1.9 Hz, 2H), 9.17 (s, 2H), 8.29 (d, J=5.8 Hz, 2H), 8.13 (dd, J=5.2, 1.2 Hz, 2H), 7.99 (d, J=6.1 Hz, 2H), 7.94 (d, J=3.8 Hz, 1H), 7.91 (dd, J=8.3, 5.2 Hz, 2H), 7.88 (dd, J=5.9, 1.9 Hz, 2H), 7.71–7.67 (m, 2H), 7.47 (dd, J=5.1, 1.1 Hz, 1H), 7.43–7.38 (m, 2H), 7.13 (dd, J=5.1, 3.6 Hz, 1H). HRMS (ESI+) m/z for [M-2Cl−]2+ calcd: 535.0203. Found: 535.0201, m/z for [M-2Cl−-H]+ calcd: 1069.0333; found: 1069.0353. HPLC retention time: 22.86 min (95.2% purity by peak area).
rac-[Ru(4,4′-btfmb)2(IP-3T)](Cl)2 (Ru-3T). Ru(4,4′-btfmb)2Cl2·2H2O (151 mg, 0.2 mmol) and IP-3T (76 mg, 0.164 mmol) were added to a microwave vessel containing argon-purged ethylene glycol (4 mL) and subjected to microwave irradiation at 180 °C for 15 min. The resulting dark red mixture was then isolated and purified in the same manner as Ru-1T, yielding the desired product as a dark red solid (55 mg, 27%). 1H NMR (700 MHz, MeOD-d3, ppm): δ 9.40 (d, J=2.0 Hz, 2H), 9.37 (d, J=1.9 Hz, 2H), 9.26 (s, 1H), 9.11 (s, 1H), 8.29 (d, J=5.8 Hz, 2H), 8.14 (dd, J=5.2, 1.3 Hz, 2H), 7.99 (s, 2H), 7.95–7.90 (m, 3H), 7.88 (dd, J=6.0, 1.9 Hz, 2H), 7.69 (dd, J=6.1, 1.9 Hz, 2H), 7.40 (d, J=3.8 Hz, 1H), 7.35 (d, J=3.8 Hz, 1H), 7.31 (dd, J=3.5, 1.1 Hz, 1H), 7.23 (d, J=3.8 Hz, 1H), 7.09 (dd, J=5.1, 3.6 Hz, 1H). 13C NMR (700 MHz, MeOD-d3, ppm): δ 163.28, 163.08, 162.89, 159.60, 159.43, 155.07, 154.53, 150.34, 141.88, 140.97, 140.81, 140.77, 140.60, 139.10, 137.67, 135.97, 131.57, 129.82, 129.20, 126.94, 126.39, 125.84, 125.75, 125.40, 125.33, 125.23, 124.45, 124.34, 123.11, 123.02, 122.89, 122.79, 119.88. HRMS (ESI+) m/z for [M-2Cl−]2+ calcd: 576.0142; found: 576.0141, m/z for [M-2Cl−-H]+ calcd: 1151.0211; found: 1151.0232. HPLC retention time: 23.80 min (95.4% purity by peak area).
rac-[Ru(4,4′-btfmb)2(IP-4T)](Cl)2 (Ru-4T). Ru(4,4′-btfmb)2Cl2·2H2O (114 mg, 0.2 mmol) and IP-4T (90 mg, 0.164 mmol) were added to a microwave vessel containing argon-purged ethylene glycol (4 mL) and subjected to microwave irradiation at 180 °C for 15 min. The reaction mixture was transferred to a 100 mL beaker and diluted with ~30 mL H2O, then treated with 3 mL of saturated aqueous KPF6, and stirred for 5 minutes. At this time, a red precipitate formed and was collected using a Büchner filtration apparatus. The product was then purified following the same procedure as described for Ru-1T, yielding [Ru(4,4′-btfmb)2(IP-4T)](Cl)2 as a dark red solid (77 mg, 35%). 1H NMR (700 MHz, MeOD-d3, ppm): δ 9.40 (d, J=2.0 Hz, 2H), 9.37 (d, J=1.9 Hz, 2H), 9.26 (s, 1H), 9.11 (s, 1H), 8.29 (d, J=5.8 Hz, 2H), 8.14 (dd, J=5.2, 1.3 Hz, 2H), 7.99 (s, 2H), 7.95–7.90 (m, 3H), 7.88 (dd, J=6.0, 1.9 Hz, 2H), 7.69 (dd, J=6.1, 1.9 Hz, 2H), 7.40 (d, J=3.8 Hz, 1H), 7.35 (d, J=3.8 Hz, 1H), 7.31 (dd, J=3.5, 1.1 Hz, 1H), 7.23 (d, J=3.8 Hz, 1H), 7.09 (dd, J=5.1, 3.6 Hz, 1H). 13C NMR (175 MHz, MeOD-d3, ppm): δ 163.28, 163.08, 162.89, 159.60, 159.43, 155.07, 154.53, 150.34, 141.88, 140.97, 140.81, 140.77, 140.60, 139.10, 137.67, 135.97, 131.57, 129.82, 129.20, 126.94, 126.39, 125.84, 125.75, 125.40, 125.33, 125.23, 124.45, 124.34, 123.11, 123.02, 122.89, 122.79, 119.88. HRMS (ESI+) m/z: [M–2Cl]2+ calcd for C49H26F12N8RuS3: 576.0142; found: 576.0141. [M–2Cl–H]+ calcd for C49H25F12N8RuS3: 1151.0211; found: 1151.0232. HPLC retention time: 23.80 min (99% purity by peak area).
2.3. Computational Details
The complexes of this series were studied with a combination of DFT and TDDFT methods62 as implemented in the Gaussian 16 software63. This computational protocol has been used to describe a variety of metal complex PSs for PDT, including Ru(II) and Os(II) systems.17,46,49,51,64–71 The geometries of the lowest energy singlet and triplet excited states were optimized with the PBE0 exchange-correlation functional72 combined with the 6–31+G(d,p) basis set for all atoms except Ru, which was described with the Stuttgart-Dresden pseudopotential73. The M06 exchange correlation functional was used to compute the electronic excited states on top of the optimized geometries through the TDDFT formalism making use of the Tamm-Dancoff approximation (TDA).74 This level of theory has been employed recently to describe 3MLCT and 3ILCT/3LLCT excited states and to predict emission energies for related oligothiophene-based Ru(II) and Os(II) complexes, providing a more accurate description of the vertical triplet state energies with respect to the conventional TDDFT analysis (which tends to underestimate the energy gaps75). The calculations were performed in water within the framework of the integral equation formalism polarizable continuum model (IEFPCM),76–78 with the dielectric constant of ε = 80 and default Gaussian 16 parameters. The excited state topologies were analyzed by post-processing of the Gaussian 16 output in two ways: (i) determination of the natural transition orbitals (NTOs) with the Chemissian 4.67 software,79 and (ii) computation of the charge-transfer descriptors through fragment-based analyses performed with the TheoDORE 3.1.1 toolbox80.
2.4. Electrochemistry
Voltammetry was performed in dimethylformamide (DMF, Fisher HPLC grade) that had been dried and deoxygenated with an Inert PureSolv MD7 solvent purification system, with 100 mM tetrabutylammonium hexafluorophosphate (TBAPF6) (Fisher) as the supporting electrolyte, in a two-compartment low volume cell with the three-electrode configuration under argon. A 3 mm glassy carbon disc was used as the working electrode with a platinum wire counter electrode and a Ag/AgCl/4M KCl reference electrode. Ferrocene (Fc) was used as an internal standard. The complex solutions were approximately 4 mM for oxidation sweeps and 0.25 mM for reduction sweeps.
Measurements were conducted at room temperature using a WaveNow potentiostat (Pine Research Company) with Aftermath software. Cyclic differential-pulse voltammetry (CDPV) measurements used a sweep rate of 2 mV·s−1 with a modulation amplitude varying from 12.5 to 100 mV. For reversible processes, the formal redox potential E°′ was taken as the average of Epa (anodic peak potential) and Epc (cathodic peak potential). For quasi-reversible processes, only Epa or Epc is reported.
3. RESULTS AND DISCUSSION
3.1. Synthesis and Characterization
[Ru(4,4′-btfmb)3]+2 and Ru-nT were synthesized utilizing our established method for related Ru(II) 2,9-dimethyl-1,10-phenanthroline systems.17 The initially isolated PF6− salts of the complexes were purified using silica gel flash chromatography and then converted to their corresponding Cl− salts on Amberlite IRA-410. Then size-exclusion chromatography was carried out on Sephadex LH-20 to give final yields of approximately 60% for [Ru(4,4′-btfmb)3]+2, Ru-0T, Ru-1T, and Ru-3T, around 40% for Ru-2T, and close to 30% for Ru-4T. These complexes underwent thorough characterization by 1D and 2D 1H NMR spectroscopy (Figures 1, and S1–S7), with signal assignments for [Ru(4,4′-btfmb)3]2+ and Ru-0T–Ru-4T conducted using 1H–1H COSY NMR. The assignments aligned with those of our previously reported, related compounds.17,20,46 Additionally, these complexes were characterized by high-resolution ESI+ MS (Figures S8–S14). The complexes were ≥95% pure by HPLC (Figures S15–S21).
Figure 1.

1H NMR spectra showing aromatic region for [Ru(4,4′-btfmb)3](Cl)2 and Ru-nT (n=0–4) in MeOD-d3 (Cl− salts; 298 K). All spectra were collected at 500 MHz, except for Ru-4T, which was collected at 700 MHz.
The lipophilicities of the chloride salts of complexes were determined by measuring their distribution between 1-octanol and 10 mM phosphate buffer (pH 7.4) following the “shake-flask” method as we described previously.49 The log Do/w values for this series span 4 orders of magnitude, with [Ru(4,4′-btfmb)3]2+ being the most hydrophilic at about −2 and Ru-4T most lipophilic near +2 (Figure 2). The three reference compounds lacking thiophene rings ([Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-0T) as well as Ru-1T show a preference for the aqueous buffer and accordingly have negative log Do/w values. In contrast, complexes with 2–4 thiophenes (Ru-2T, Ru-3T, Ru-4T) have positive log Do/w values and increase on going from n=2 to 4. The addition of the trifluoromethyl group qualitatively improved the overall aqueous solubility of the complexes with positive log Do/w values compared to analogous Ru(II) and Os(II) complexes with other coligands where 4T often leads to precipitation at the octanol/buffer interface.46,49
Figure 2.

Lipophilicity of [Ru(4,4′-btfmb)3](Cl)2, Ru-phen, and Ru-nT (n=0–4) in 1-octanol and phosphate buffer using the in-house “shake-flask” method.
3.2. Computation
Singlet states.
The singlet ground-state structures for [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-nT (n=0–4) optimized in water at the DFT/PBEO level of theory are shown in Figure 3. The associated geometric parameters (Table S3) confirm the octahedral arrangement of coordinating nitrogen atoms around the central Ru(II) ion and similar Ru-N bond distances for all of the compounds. The thiophene ring of Ru-1T and the first thiophene ring of Ru-nT (n=2–4) are coplanar with the coordinating IP ligand, but additional thienyl rings are conformationally more flexible. In the case of Ru-4T, the fourth thiophene ring is twisted as much as 18° with respect to the IP plane.
Figure 3.

Optimized geometries of [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-nT (n=0–4) in a water environment at the PBE0/6–31+G(d,p)/SDD level of theory. For the sake of clarity, the 4,4′-btfmb coligands are shown in grey for all the complexes except for [Ru(4,4′-btfmb)3]2+.
The nT chain length impacts the frontier molecular orbitals and the resulting electronic transitions. The systematic reduction of the highest occupied molecular orbital (HOMO)–lowest unoccupied molecular orbital (LUMO) gaps with longer thiophene chains is due to an increase in the HOMO energies as the orbital contribution from nT increases. The LUMOs are primarily bftmb-based (>70%) with similar energies across the series, but the nT contribution to the HOMO reaches 51% for Ru-3T and almost 66% for Ru-4T (Figures 4 and S22, Table S4). The red-shift of the lowest energy spin-allowed singlet–singlet electronic transitions with increasing n is accompanied by a change in the nature of these states toward mixed 1LLCT/1ILCT/1IL character for n>2 as shown by the natural transition orbital (NTO) analysis (Figures 5 and S23), consistent with other oligothiophene-bearing Ru(II) and Os(II) complexes.17,20,46,51,67 The NTOs are predominantly 1MLCT (Ru→4,4′-btfmb) for the complexes lacking thiophene rings and Ru-1T, with electronic transitions computed in the range of 438 to 459 nm (only slightly higher in energy than the experimental bands falling between 456 and 471 nm in Figure 3). The NTOs for Ru-2T are of mixed 1MLCT (53%)/1LLCT (26%) character (1LLCT is nT→4,4′-btfmb), and the lowest-energy transition is computed at 455 nm in close agreement with the experimental band at 462 nm. In contrast, the NTOs for Ru-3T and Ru-4T lack 1MLCT character and are mixed 1LLCT/1ILCT/1IL. Both complexes have lowest-energy transitions with significant 1LLCT (nT→IP) character (55% for Ru-3T and 45% for Ru-4T) but differ in the relative contributions of 1ILCT versus 1IL. The NTOs for Ru-3T have a larger contribution from 1IL (IP-based ππ*) compared to 1ILCT (nT-based CT). The opposite is true for Ru-4T, where the 1ILCT contribution is slightly more than that of 1IL. The computed transitions are 469 and 493 nm for Ru-3T and Ru-4T, respectively, which lie within the broadened spectral features in the experimental spectra.
Figure 4.

(a) Calculated HOMO, LUMO, HOMO-1 and LUMO+1 orbital energies. (b) Percent contribution of the nT, bftmb, IP and Ru fragments to the HOMO and LUMO orbitals, for Ru-nT (n=0–4) in the singlet ground state, at the M06/6–31+G(d,p)/SDD level of theory. Additional details can be found in the SI.
Figure 5.

Occupied and virtual NTOs of the computed lowest-energy singlet-singlet transitions in water (λ) with the predominant character indicated. Additional NTOs are reported in Figure S23.
Triplet states.
The lowest energy excited triplet state (T1) structures for [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-nT are similar to their corresponding ground-state structures except that those involving oligothiophenes adopt a completely planar chain conformation for maximum π-conjugation, with successive nT groups antiplanar with respect to one another. The geometrical parameters of these optimized T1 states are compiled in Table S3 alongside the data for S0. All complexes, with the exception of Ru-3T and Ru-4T, have T1 states that are 3MLCT (Ru→btfmb), as shown by the NTO analyses (Figures 6 and S24, Table 1) and supported by Mulliken spin densities (MSDs) close to 1 for the Ru(II) center (Table S3). In sharp contrast, T1 is localized primarily to the IP-nT ligand for Ru-3T and Ru-4T and its nature changes to mixed 3ILCT/3LLCT/3IL. The largest contributor to these mixed T1 states is 3ILCT (CT within the nT chain) and is slightly more for Ru-4T (51% vs 39%), whereas the 3IL contribution (nT-localized) is similar for both at ~20% and 3LLCT (CT from nT to IP) contributes more for Ru-3T (25%) and less for Ru-4T (16%). The 3MC states involving Ru(II) and 3IL states localized to 4,4′-btfmb or IP are higher in energy and do not contribute to the computed NTOs for the lower-energy excited states.
Figure 6.

(a) Computed T1 adiabatic energies for [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-nT (n=0–4). (b) Molecular fragments (left) defined to quantify the molecular topology of the T1 excited states and their character (right). The corresponding NTOs are reported in Figure S24.
Table 1.
Adiabatic labels (State) and adiabatic 3MLCT and 3ILCT energies (ΔEadia) for [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-nT (n=0–4) in a water environment at the TD-M06/6–31+G**/SDD level of theory. Additional details can be found in Table S5.
| Compound | 3MLCT | 3ILCT | ||
|---|---|---|---|---|
| State | ΔEadia(eV) | State | ΔEadia(eV) | |
| [Ru(4,4′-btfmb) 3 ] 2+ | T1 | 2.01 | / | / |
| Ru-phen | T1 | 1.91 | / | / |
| Ru-0T | T1 | 1.91 | / | / |
| Ru-1T | T1 | 1.91 | / | / |
| Ru-2T | T1 | 1.91 | / | / |
| Ru-3T | T2 | 1.99 | T1 | 1.54 |
| Ru-4T | T2 | 1.88 | T1 | 1.42 |
The computed adiabatic 3MLCT energies lie near 1.9–2.0 eV across the series, regardless of whether this state is T1 or T2 (Tables 1 and S5). These 3MLCT energies are lower (by ~0.2 to 0.3 eV) than those for related Ru(II) polypyridyl systems without substituted ligands due to the electron-withdrawing groups of the 4,4′-btfmb coligands in this series. These groups do not impact the 3ILCT/3LLCT/3IL energies, which are very similar to those computed for the related series with 1,10-phenanthroline (phen) coligands.51 However, the lower 3MLCT energies in the Ru(II) 4,4′-btfmb complexes result in an 3MLCT-based T1 for Ru-2T (rather than the IP-nT-based triplet that tends to be T1 in other related families). The mixed 3ILCT/3LLCT/3IL states that are T1 in the case of Ru-3T and Ru-4T have computed adiabatic energies of 1.54 and 1.42 eV, respectively. The drop in energy of the T1 state for Ru-3T and Ru-4T is directly related to its organic 3ILCT character and length of the nT chain as we have seen previously.17,51
3.3. Spectroscopy
3.3.1. Electronic Absorption
The UV-vis absorption spectra of the series are overlaid in Figure 7, and the corresponding molar extinction coefficients for various peak maxima are summarized in Table 2. The spectra display some of the general features of typical Ru(II) polypyridyl complexes81 but with some distinctions for the complexes bearing IP-nT ligands. The sharp peaks near 295 nm and shorter are due to π‒π* transitions involving the 4,4′-btfmb ligand,55 as well as phen and/or IP in the cases of Ru-phen and Ru-nT. The energies of these transitions are not affected by the length of the thiophene chain appended to the IP ligand. The broader and less intense peaks between 400 and 500 nm with a local maximum at 456 nm in [Ru(4,4′-btfmb)3]+2 are due to Ru2+(dπ)→L(π*) 1MLCT transitions involving mainly the 4,4′-btfmb ligands as the π* acceptor orbitals. The substitution of one 4,4′-btfmb ligand for phen or IP causes the lowest-energy 1MLCT transitions to red-shift by 10–20 nm. Assuming that the 4,4′-btfmb π* orbitals remain the acceptor orbitals for the lowest-energy 1MLCT transitions in all cases, the effect of phen or IP is primarily on the dπ orbital energy. This effect is most evident for Ru-phen, Ru-0T, and Ru-1T. For complexes containing the IP-nT ligand, additional transitions overlap the 1MLCT transitions. These contributions from the IP-nT ligands can be seen in the absorption spectra of the analogous uncomplexed IP-nT ligands and free oligothiophenes.82 The computed NTOs provide information on their contributions and characters in the Ru(II) complexes (Figures 5 and S23). The lowest-energy singlet–singlet transition is mixed 1MLCT/1LLCT for Ru-2T wherein 1LLCT involves nT→4,4′-btfmb CT, while these transitions are mixed 1LLCT/1IL/1ILCT and 1LLCT/1ILCT/1IL for Ru-3T and Ru-4T, respectively. In these cases, 1LLCT is the major contributor and involves nT→IP CT, 1ILCT involves nT→nT CT, and 1IL involves localized ππ* transitions within nT. These mixed transitions shift to longer wavelengths and increase in 1ILCT character with increasing n as we have observed previously in related compounds.11,17,46,47,51
Figure 7:

UV-vis spectra of [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-nT (n=0–4) normalized to the π‒π* peak near 295 nm
Table 2:
Molar Extinction Coefficients at Selected Wavelengths for the Complexes of This Study.
| cmpd | λabs (nm) (log (ε / M−1 cm−1)) |
|---|---|
| [Ru(4,4′-btfmb)3]+2 | 244 (4.34), 294 (4.84), 427 (4.01), 456 (4.12), |
| Ru-phen | 260 (4.69), 296 (4.80), 353 (4.05), 432 (4.09), 471 (4.12) |
| Ru-0T | 248 (4.68), 296 (4.84), 341 (4.13), 435 (4.12), 470 (4.13) |
| Ru-1T | 245 (4.57), 294 (4.92), 335 (4.51), 462 (4.20) |
| Ru-2T | 250 (4.68), 295 (4.89), 370 (4.69), 462 (4.35) |
| Ru-3T | 251 (4.58), 294 (4.76), 410 (4.57), 426 (4.55)a |
| Ru-4T | 246 (4.57), 297 (4.81), 351 (4.33), 441 (4.70), 468 (4.62)a |
The maximum of the longest wavelength singlet–singlet transition is obscured.
3.3.2. Singlet Oxygen Sensitization
The singlet oxygen quantum yields (ΦΔ) of the complexes were calculated by measuring the intensity of O2 phosphorescence (1Δg→3Σg) centered at 1260 nm against [Ru(bpy)3]2+ as the standard (ΦΔ=0.56).83 These are reported in Table 3. [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-0T are moderately efficient 1O2 generators (ΦΔ=0.47–0.64) and similar to [Ru(bpy)3]2+. However, the thiophene-containing complexes Ru-1T–Ru-4T exhibit larger differences (ΦΔ=0.13–0.66) and are less efficient than what has been observed for related compounds.11,47 The largest 1O2 quantum yield was measured for Ru-3T at 66%. While all the complexes show a wavelength-dependence for ΦΔ, Ru-4T exhibits a notable concentration dependence for the 1O2 quantum yield (Figure S25 and Table S6). These differences in ΦΔ appear to be unrelated to differences in other photophysical parameters such as emission wavelengths and triplet lifetimes (vide infra).
Table 3:
Spectroscopic data for compounds [Ru(4,4′-btfmb)3](PF6)2, Ru-phen, and Ru-0T–Ru-4T as PF6− salts. Excitation wavelengths are indicated in parenthesis. Emission lifetimes were measured following a <5 ns 355 nm laser pulse.
| cmpd | RT emission | 77 K emission | ΦΔ (λex / nm) | τTA / μs) | ||||
|---|---|---|---|---|---|---|---|---|
| λem. (λex) / nm | Φem | τem / μs | λem. (λex) / nm | Φem,77K | λex=355 nm | λex=532 nm | ||
| [Ru(4,4′-btfmb)3]2+ | 636 (458) | 1.6×10−1 | 1.5 | 604,653 (460) | 2.5×10−1 | 0.47 (462) | 1.5 | 1.5 |
| Ru-phen | 660 (470) | 9.0×10−2 | 0.89 | 619, 668 (472) | 2.4×10−1 | 0.64 (470) | 0.89–0.94 | 0.77–0.92 |
| Ru-0T | 662 (470) | 9.1×10−2 | 0.83 | 619, 669 (470) | 1.9×10−1 | 0.50 (469) | 0.78–0.86 | 0.78–0.87 |
| Ru-1T | 661 (460) | 1.8×10−2 | 0.79 | 620, 670 (469) | 2.0×10−1 | 0.13 (462) | 0.81–0.93 | 0.71–0.87 |
| Ru-2T | 659 (457) | 4.1×10−3 | 0.62 | 634, 687 (463) | 4.3×10−2 | 0.28 (466) | 0.61–0.72 | 0.63–0.71 |
| Ru-3T | 664 (462) | 1.3×10−3 | 0.79 | 623, 673 (470) | 9.7×10−3 | 0.66 (470) | 20–21 | 22–24 |
| Ru-4T | 650 (435) | v. wk. | 0.64 | 625, 672 (468) | 1.4×10−3 | 0.40 (467) | 19–20 | 20–21 |
3.3.3. Emission
All of the complexes in the series exhibited a broad, featureless red emission band at room temperature (Figure 8a and Table 3). This emission was centered around 636 nm (τem=1.5 μs) for the parent [Ru(4,4′-btfmb)3]2+ complex in MeCN at room temperature (RT) and shifted to shorter wavelengths with vibronic intervals of around 1350 cm−1 at 77 K (Figure 4b).54–57 Such behavior is consistent with emissive 3MLCT states in Ru(II) complexes with polypyridyl ligands.84 The thermally induced Stokes shift (ΔES) of around 830 cm−1 is slightly smaller than the related model complex [Ru(bpy)3]2+ (ΔES=1127 cm−1)85 but in agreement with the assignment.86 Similar to what was observed for the 1MLCT absorption bands, the introduction of a phen or an IP ligand results in bathochromic shifts of up to 30 nm but otherwise similar spectra and lifetimes (τem=0.8–0.9 μs). The 3MLCT emission energies and lifetimes (τem=0.6–0.8 μs) also do not depend on the number of thiophenes in the IP-nT ligand, suggesting that the π* acceptor orbitals in the Ru(dπ)→L(π*) transitions involve the 4,4′-btfmb coligands. The emission quantum yield (Φem) ranges from 16% for [Ru(4,4′-btfmb)3]2+ to 9% for the complexes without thienyl groups. The value of Φem drops to around 2% for Ru-1T, and additional thienyl rings further decrease the emission output to <1%. Although Ru-4T produces weak but detectable emission, a value for Φem was not calculated due to the extremely low signal-to-noise ratio. The values for Φem increase up to 10-fold at 77 K but Ru-4T is still only about 0.14%. While the energy of the emissive 3MLCT state does not appear to change across the series, the emission quantum yields decrease with increasing n and implicate additional excited-state deactivation pathways. The computed NTOs indicate that the nature of T1 changes substantially with the introduction of three or four thienyl groups. The lowest-energy triplet states are of 3MLCT character for all of the complexes except Ru-3T and Ru-4T, where T1 becomes mixed 3ILCT/3LLCT/3IL (with 3ILCT involving nT→nT CT, 3LLCT involving nT→IP CT, and 1IL based on localized ππ* transitions within n). These organic triplets involving the IP-nT ligand are nonemissive, and thus the very weak emission arises from the 3MLCT state that is T2 for Ru-3T and Ru-4T and of similar energy to the other complexes in the series.
Figure 8:

Normalized emission spectra [Ru(4,4′-btfmb)3]2+ and the Ru-nT series as PF6− salts at RT and at 77K. The RT emission was measured in MeCN degassed by freeze-pump-thaw (5 cycles). The 77K emission was measured in a 4:1 EtOH:MeOH glass. Excitation wavelengths are noted in parentheses.
3.3.4. Transient Absorption and Excited State Pathways
The triplet excited states were investigated using nanosecond transient absorption (TA) spectroscopy with excitation from a 355 nm (Figure 9) or 532 nm laser (Figure S26) of ≤5 ns pulse width. The responses with the two different excitation wavelengths are similar, suggesting similar excited-state dynamics. The differential excited-state absorption (ESA) spectra were collected using solutions of the PF6− salts of the compounds in degassed (5x freeze-pump-thaw) MeCN. Early time slices are presented in Figures 9 and S26, and the full set of relaxation spectra over different time points are collected in Figures S27 and S28. Transient lifetimes were measured at the ESA maxima and bleach minima and are listed in Table 3.
Figure 9:

Transient absorption (TA) spectra of [Ru(4,4′-btfmb)3](PF6)2, Ru-phen, and Ru-nT as PF6− salts in degassed MeCN at RT (λex=355 nm) integrated over the indicated time slice following the excitation pulse. The profiles for λex=532 nm are similar (Figure S26). The horizontal dashed lines indicate ΔO.D.=0 for each compound differentiated by color.
The ESA profiles of [Ru(4,4′-btfmb)3]2+, Ru-phen, Ru-0T, and Ru-1T are similar, consisting primarily of a strong 1MLCT ground-state bleach near 450 nm superimposed with new ESA characteristic of the 3MLCT state. The stronger ESA near 375 nm involves 4,4′-bftmb− transitions, and the extremely weak and broad absorption past 525 nm is due to 4,4′-btfmb− or LMCT transitions involving Ru(II). The TA lifetimes match the emission lifetimes and support lowest-lying 3MLCT states for these complexes as predicted from the NTO analyses.
The ESA profile of Ru-2T exhibits features consistent with a triplet excited-state localized to the IP-nT ligand, as we have previously reported.11,17,50,51,87–89 This broad and rather intense ESA near 450–700 nm is superimposed on the bleach with its minimum near 370 nm due to a strong ground-state absorption contributed by the IP-2T ligand (Figure 9). While Ru-2T has the signature of an nT-based triplet, its TA lifetime matches the emissive 3MLCT lifetime at all wavelengths. This observation and the NTO analyses suggest that the 3MLCT state computed as T1 and the ligand-localized triplet are close in energy and decay with a common lifetime.90 In this regard, Ru-2T behaves differently from some its close analogs containing the IP-2T ligand, where T1 is of mixed 3LLCT/3ILCT character with a prolonged lifetime,51 due to the lower 3MLCT energies of Ru(II) polypyridyl complexes incorporating 4,4′-btfmb coligands.
The ESA spectra of Ru-3T and Ru-4T are characteristic of oligothiophene-based 3ILCT states, with longer TA lifetimes. The excited state relaxation pathways are proposed in Scheme 1. Ru-3T shows a bleach corresponding to the ground-state ππ* transition near 410 nm and a broad ESA with its maximum near 610 nm. For Ru-4T these were shifted to around 430 nm and 655 nm, respectively. These ESA features are similar to those of the free IP-3T and IP-4T ligands.46,49 The decays are monoexponential, with τ=~20 μs for both complexes, indicating that the long-lived 3ILCT is decoupled from the shorter-lived emissive 3MLCT with computed T2 energies near 1.9–2.0 eV. This is further supported by the very low emission quantum yields for Ru-3T and Ru-4T and computed T1 energies of 1.54 and 1.42 eV, respectively, with significant 3ILCT character.
Scheme 1.

Jablonski Diagram Illustrating the Excited-State Relaxation Pathways of Complexes Ru-3T and Ru-4T. Energies are not to scale.
3.4. Electrochemistry
The electrochemistry of Ru(II) polypyridyl complexes is typified by single electron processes involving one-electron oxidation of the metal center and three sequential reductions on each of the ligands.91 Oxidation of the Ru2+ center (+0.98 V versus Fc, MeCN) tends to be electrochemically reversible, and the ensuing low-spin 4d5 complex is chemically stable. In complexes like [Ru(bpy)3]2+, the first reduction (−1.72 V versus Fc, MeCN) involves the lowest-energy ligand π* orbital. Since the low spin 4d6 configuration is thereby unaffected, the complex remains substitutionally inert, and the process is also reversible. The added electron is localized on one ligand, and thus [Ru(bpy)3]2+ exhibits three sequential one-electron reductions under straightforward electrochemical conditions. In a potential window widened by low temperature cyclic voltammetry, [Ru(bpy)3]2+ has been shown to participate in a total of six one-electron reductions.92,93
There are additional redox processes for some members of the present series due to the presence of the electrochemically active oligothiophene94 unit in complexes Ru-2T–Ru-4T. The formal redox potentials of the series were measured by cyclic differential pulse voltammetry (CDPV) to enhance the signal, with ferrocene (Fc) as an internal reference (E1/2 (Fc/Fc+)=0.380 V vs SCE95). The potentials are compiled in Table 4 and compared graphically in Figure 10 with tentative assignments. The CDPV plots for oxidation and reduction are shown in Figures S29 and S30, respectively.
Table 4.
Formal redox potentials for the hexafluorophosphate salts of the complexes measured using CDPV at approximately 1.0 mM in MeCN containing TBAPF6. The potentials are referenced in volts (V) against ferrocene as the internal standard. The working and reference electrodes were glassy carbon and Ag/AgCl/4M KCl, respectively. Overlapping waves were deconvoluted mathematically (error approximately ±0.02 V).
| Compound | nT2−→nT3− | nT−→nT2− | nT→nT− | LL2−→LL12− | LL1−→LL12− | LL3→LL3− | LL2→LL2− | LL1→LL1− | Ru2+→Ru3+ | nT→nT+ |
|---|---|---|---|---|---|---|---|---|---|---|
| [Ru(4,4′-btfmb)3]2+ | −2.75 | −2.21 | −1.61 | −1.39 | −1.24 | +1.34 | ||||
| Ru-phen | −2.70 | −2.30 | −2.01 | −1.49 | −1.29 | +1.18 | ||||
| Ru-0T | −2.81 | −2.29 | −2.14 | −1.54 | −1.33 | +1.18 | ||||
| Ru-1T | −2.75 | −2.30 | −2.12 | −1.51 | −1.32 | +1.22 | +1.02 | |||
| Ru-2T | −2.81 | −2.71 | −2.29 | −2.12 | −1.51 | −1.31 | +1.26 | +0.89 | ||
| Ru-3T | −2.96 | −2.54 | −2.86 | −2.32 | −2.15 | −1.54 | −1.32 | +1.29 | +0.69 | |
| Ru-4T | −2.93 | −2.55 | −2.23 | −2.83 | −2.33 | −2.13 | −1.52 | −1.32 | +1.27 | +0.58 |
Figure 10:

Redox potentials (vs. Fc) of [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-nT as PF6− salts in degassed MeCN at RT and their tentative assignments.
The metal oxidation of [Ru(4,4′-btfmb)3]2+ and the three one-electron reductions agree well with published data55 and are around 0.3–0.5 V more positive than the corresponding processes in [Ru(bpy)3]2+, a consequence of the electron-withdrawing nature of the 4,4′-btfmb ligands. The shift of the Ru2+→ Ru3+ oxidation is slightly attenuated when one 4,4′-btfmb ligand is replaced with phen or IP-nT. The oxidation of nT becomes more favorable with increasing n, consistent with the behavior of free oligothiophenes.96 The oxidation potential of the thienyl group ranges from +1.02 V for Ru-1T to +0.58 for Ru-4T and is less positive than the metal center in all cases, indicating that nT is more easily oxidized than Ru(II) regardless of the number of thiophenes. The oxidation of nT does shift the Ru2+→ Ru3+ oxidation slightly more positive, consistent with a decrease in electron density on the metal. For Ru-3T and Ru-4T there is at least one additional peak between those for nT and Ru2+ oxidation that is of lesser intensity and solvent-dependent (Figure S29). This peak, not listed in Table 4, is thought to arise from the oxidation of the σ-dimer radical cation ((Ru-nT)2+• → (Ru-nT)2 + 2H+ + e−) given the known abilities of certain oligothiophenes to undergo electrochemical σ-dimerization through their α positions. However, this requires further investigation to confirm.
[Ru(4,4′-btfmb)3]2+ and the other complexes without thiophenes (Ru-phen, Ru-0T) as well as Ru-1T exhibit five sequential reductions spanning −1.24 to −2.81 V (Figure 10). For [Ru(4,4′-btfmb)3]2+, this involves sequential one-electron reductions on each of the three 4,4′-btfmb ligands followed by second reductions on two of those ligands (within the experimental potential window). We believe this study is the first to report the second reductions of the two 4,4′-btfmb ligands occurring between −2.2 and −2.9 V for [Ru(4,4′-btfmb)3]2+.
When one of the 4,4′-btfmb ligands is replaced by phen, IP-0T, or IP-1T, the third reduction involves phen or IP and is less favorable by around 0.4–0.5 V (owing to their lack of electron-withdrawing −CF3 substituents). Ru-2T–Ru-4T exhibit similar reductions involving the 4,4′-btfmb coligands and IP, but they accommodate additional reductions on the oligothienyl groups. For Ru-2T, the 2T group accepts only one electron and this sixth reduction is the least favorable and occurs near −2.81 V. The 3T group of Ru-3T can be doubly reduced (−2.54 and −2.96 V), where the first nT reduction is easier than the last 4,4′-btfmb reduction. Continuing this trend, the 4T group of Ru-4T accommodates three electrons (−2.23, −2.55, and −2.93 V), with the first nT reduction occurring near that of IP at −2.13 V. The effect of the thiophene chain length is dramatic. The nT reduction potential shifts positive by more than 0.5 V on going from two to four thiophenes, and the Ru-4T complex can accommodate at least eight extra electrons in its ground state at room temperature.
3.5. In Vitro Photobiological Activity
The compounds in this series were assessed for their cytotoxicity in the absence of light (dark) as well as their light-triggered cytotoxicity against human skin melanoma cells (SK-MEL-28) and lung carcinoma cells (A549) cultured as 2D monolayers (Figure 11). The experimental details can be found in our previously published work11,17,20 as well as in the SI. The spectral output of the light sources is shown in Figure S31, their overlap with the absorption profiles of the complexes is shown in Figure S32, and the estimated absorbed photon flux is compiled in Table S7. The results are summarized in Figure 11 and Tables S8, S9.
Figure 11.

In vitro cytotoxicity and photocytotoxicity log (EC50 ± SEM) values (left) and PI values (right) obtained from dose–response curves in the A549 cell line (a) and SK-MEL-28 melanoma cell line (b) with [Ru(4,4′-btfmb)3](Cl)2 and Ru-phen–Ru-4T. Treatments include dark (black circles) or light delivered at a fluence of 100 J cm−2 and irradiance of ~20 mW cm−2. The light wavelengths were broadband visible (400–700 nm, blue squares), 523 nm (green inverted triangles), or 633 nm (red triangles). Data collected under normoxic (~18.5% O2) and hypoxic (1% O2) conditions is represented with closed symbols and open symbols, respectively.
3.5.1. Normoxia.
For each photobiological assay in normoxia, SK-MEL-28 or A549 cells were seeded into two sets of 384-well plates: one for dark cytotoxicity evaluation and the other for photocytotoxicity assessment. After allowing the cells to adhere for 3–5 hours at 37°C, cells were dosed with the compounds (1 nM to 300 μM). The light plates were exposed to light treatments after a 13–22 h DLI, while the dark plates remained in the incubator. The light treatments were delivered at a fluence of 100 J cm−2 emitted from broadband-visible (400–700 nm, 21 mW cm−2) or monochromatic (±2.5 nm) green (523 nm, 18 mW cm−2) or red (633 nm, 18 mW cm−2) LEDs.
After light treatment, the plates were allowed to incubate in normoxia at 37°C for 24 h. Cell viability was then assessed indirectly using a resazurin-based cell viability assay. Sigmoidal fits of the dose-response curves were used to calculate the effective concentrations required to reduce cell viability by 50% (EC50 values) for both treatment conditions. The amplified cytotoxic effects upon light exposure, known as phototherapeutic indices (PIs), were calculated as ratios of dark EC50 to light EC50 values.
The complexes were mostly nontoxic to both cell lines in the dark. [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-0T exhibited dark EC50 values that exceeded the highest doses used in the assay (>300 μM), indicating a lack of toxicity. They were also inactive against both cell lines under any light condition. Ru-1T–Ru-3T were considered nontoxic toward both cell lines, with dark EC50 values >170 μM. Ru-4T had the lowest dark EC50 values, which were still relatively high at 100 μM and 120 μM in SK-MEL-28 and A549 cells, respectively.
SKMEL28 cells.
Increasing the number of thiophene rings (n=1–4 thienyl groups) increased the potency with visible light, ranging from 22 μM (PI=11) for Ru-1T to as low as 10 nM (PI=10,000) for the most potent compound Ru-4T in SK-MEL-28. Appending two thiophenes (Ru-2T) improved the potency 36-fold, shifting the EC50 values to sub-micromolar values near 0.61 μM (PI=290). About 100-fold enhancement in photocytotoxicity was seen moving to three thiophene rings (Ru-3T; SK-MEL-28: EC50=0.21 μM, PI=920), and over 2000-fold improvement occurred with four thiophenes (Ru-4T; SK-MEL-28: EC50=10 nM, PI=10,000).
The activity of the series was mostly diminished with longer-wavelength green or red light. With green light, Ru-1T lost most of its activity (EC50=103 μM, PI=2). The activity Ru-2T dropped by a factor of 3 but was still single-digit micromolar (EC50=1.9 μM, PI=91), and Ru-4T dropped by 10-fold but remained sub-micromolar (EC50=0.12 μM, PI=815). Ru-3T, on the other hand, maintained its activity (EC50=0.27 μM, PI=731). With red light, Ru-3T (EC50=7.02 μM; PI=28) and Ru-4T (EC50=2.74 μM; PI=37) showed modest activity. This outcome aligns with the anticipated behavior of compounds that exhibit minimal absorption of red light.97
A549 cells.
The A549 cell line proved more resistant to the light-triggered compounds than SK-MEL-28 under similar conditions, but the overall trend in activity remained the same. Ru-1T (visible EC50=29 μM, PI=7) was the least active thienyl complex, followed by Ru-2T (visible EC50=1.37 μM, PI=127), Ru-3T (visible EC50=0.70 μM, PI=280), and finally Ru-4T (EC50=0.077 μM, PI=1500). Treatments with red and green light similarly attenuated the overall activity of the series.
3.5.2. Hypoxia.
The assays under hypoxic conditions mirrored the procedure followed for normoxia, with a notable exception: post-cell adhesion, the plates—both for dark and light conditions—were transferred to a hypoxia chamber with 1% O2 atmosphere for a duration of 2–3 h prior to introducing the compounds. After a DLI of 18 h in the hypoxia chamber, the concentration of dissolved O2 was verified in the assay wells using an immersive optical probe before sealing the plates to be light treated with transparent qPCR film that has a low gas permeability. Light was delivered outside the hypoxia chamber for approximately 1.5 h, while the dark plates were kept inside an incubator. The films were removed from the light plates at the end of the illumination period. Both dark and light plates were then incubated in normoxia (37°C, 5% CO2, with relative humidity above 90%) for 20–23 h, after which time the cell viability was assessed.
Hypoxic conditions broadly attenuated the activity of the series in both cell lines, where A549 cells were completely resistant to all treatments and SK-MEL-28 cells were moderately sensitive under certain conditions. Regardless of light-treatment, [Ru(4,4′-btfmb)3](Cl)2 and compounds Ru-0T–Ru-2T exhibited no photocytotoxic effect under hypoxic conditions toward SK-MEL-28 cells. On the other hand, Ru-3T showed an EC50 value of ~1 μM with visible and green light treatments. With red light, Ru-4T was less active (EC50=26 μM, PI=4), but otherwise the series was inactive. The dramatic decline in the activity of Ru-3T and Ru-4T, coupled with the complete inactivity of the other compounds in the series under hypoxic conditions, suggests that oxygen plays a pivotal role in their photocytotoxic mechanisms. This observation solidifies the hypothesis that the primary driver behind the observed photocytotoxicity of these compounds in normoxic conditions stems from oxygen-dependent photophysical processes.
3.5.3. Biological replicates.
Figure 11 and Table S9 show representative activity against SK-MEL-28 for one biological replicate based on the mean values from three technical replicates with minimal standard deviation. Since a greater degree of variation is expected over biological replicates, we assessed the activity of the compounds (Ru-3T and Ru-4T) of highest activity over five biological replicates (each carried out in triplicate) with SK-MEL-28 cells, as illustrated in Figure S34 and detailed in Tables S10–S13. Repeat 0 corresponds to the data in Figure 11 and Table S9. The subsequent biological replicates are labeled Repeats 1–5. SK-MEL-28 cells were selected for these studies because this is the cell line we have historically used to rank potency across all compounds made in our laboratory.17
Ru-3T and Ru-4T were nontoxic over all biological replicates in both normoxia and hypoxia without light activation, with mean EC50 values of 197 μM for Ru-3T and 100 μM for Ru-4T. The visible EC50 values for Ru-3T in normoxia were in the range of 70–700 nM (mean=440 nM), and their visible PIs spanned 305–2700 (PIavg=1400). The EC50 values for Ru-4T under the same conditions ranged from 9–59 nM with a mean of 17 nM (PI=1900–12000, mean=6500). The results from all 6 replicates varied less an order of magnitude.
Using green light, the EC50 values for Ru-3T fluctuated between 0.22 and 0.66 μM, with PIs between 300 and 1000. The mean values were determined to be 0.44 μM and 548, respectively. For Ru-4T, the results were slightly more variable, with green EC50 values ranging from 54 to 320 nM, and the corresponding PIs falling in the 800 to 2000 range. Even though Ru-4T was generally more active than Ru-3T, the difference in their activities was only around 3-fold.
Under red light illumination in normoxia, the EC50 values for Ru-3T and Ru-4T were reproducibly in the single digit micromolar regime. Ru-3T exhibited red EC50 values between 2 and 7 μM (mean=4.6 μM) and PIs ranging from 23 to 101 (PIavg=52). The activity of Ru-4T was slightly more consistent than Ru-3T, with red EC50 values falling between 1.3 and 3.3 μM (mean=2.2 μM) with PIs between 21 and 91 (PIavg=53). The responses with red light for Ru-3T and Ru-4T is notable as both complexes exhibit vanishingly low molar extinction coefficients at 633 nm. This characteristic has been reported previously,97 and tends to require lowest-lying 3ππ* triplets with prolonged lifetimes such as those observed for Ru-3T and Ru-4T.
The activity of Ru-3T in hypoxia was mostly consistent but repeats 3 (with red light) and 4 (with vis and green light) were outliers: repeat 3 displayed unusually high activity compared to the other replicates, and no activity at all was observed in repeat 4. In some of the replicates, the visible and green activity of Ru-3T improved when compared to our initial measurements, but all values were still within one order of magnitude. With visible light, the EC50 value was 1.2 μM (PI=170) in repeat 0 but improved to a range of 300–870 nM (PI=230–700, PIavg=300). Similarly, the initial green EC50 value for Ru-3T was 1.1 μM (PI=180), which lowered to 0.3–1.0 μM (PI=200–700, PIavg=270) in subsequent assays. The observation of two distinct outliers, and the fact that the activity of Ru-3T was actually greater in subsequent biological replicates, underscores the importance of performing biological replicates when evaluating photobiological efficacy.
The activity of Ru-4T varied slightly more in hypoxia than it did in normoxia, but all EC50 values were still within roughly one order of magnitude. Using visible light treatment, repeats 0 and 2–5 fell between 0.13–0.54 μM (PI=200–400), but repeat 1 presented a slightly lower EC50 of 35 nM (PI=2900). With green and red light treatments, Ru-4T proved to be consistently more active than what was originally observed in repeat 0, with initial EC50 values of 1.0 μM (PI=90) and 26 μM (PI=4), respectively, which improved to a range of 200–900 nM (PI=100–500, PIavg=210) and 3–9 μM (PI=7–30, PIavg=21) during replicates 1–5. The average hypoxic red activity (<10 μM) of Ru-4T is particularly notable for this class of PS, and is comparable to the “ubertoxin” ML19C01 (which presents sub-nanomolar normoxic phototoxicity).17,20
In conclusion, across six biological replicates, Ru-4T has consistently displayed superior activity compared to Ru-3T, particularly when activated with visible light under normoxic conditions. Under these conditions, Ru-4T had an average EC50 of 24 nM (PIavg=6500). Notably, the values obtained in all biological replicates were within one order of magnitude, making Ru-4T more consistent than ML19C01, which varies by up to 6 orders of magnitude.17,20 This marked improvement in consistency may be related to the improved aqueous solubility observed in this family of complexes, and the effects of fluorination on biological reproducibility is being investigated further.
4. CONCLUSIONS
Herein we synthesized a family of Ru(II) polypyridyl complexes featuring two 4,4′-btfmb coligands and the IP-nT ligand (n=1–4) and compared these systems to the reference compounds [Ru(4,4′-btfmb)3]2+, Ru-phen, and Ru-0T. This study stands as part of our larger initiative to examine the impact of structural variations on the physicochemical, photophysical, electrochemical, and (photo)biological properties of oligothienyl-containing metal complexes.
The complexes lacking thienyl groups and Ru-1T are relatively hydrophilic, while the lipophilicities of Ru-2T–Ru-4T increase with n. The number of thienyl groups in the Ru-nT complexes also have a striking influence on the ground-state absorption and TA spectra with transitions involving the nT groups (1LLCT and 1ILCT) shifting to longer wavelengths and increasing in 1ILCT character with increasing n. Based on their TA profiles, Ru-3T and Ru-4T have lowest-lying 3ILCT states, while the complexes lacking nT groups and Ru-1T as well as Ru-2T have 3MLCT states as T1. The TA lifetimes corroborate this assignment with Ru-3T and Ru-4T having triplet lifetimes on the order of 20–24 μs and ESA spectra consistent with 3ILCT states. The other complexes have triplet lifetimes and TA spectra consistent with 3MLCT states near 1 μs. The exception is Ru-2T, which has the signature of an nT-localized triplet but with a lifetime characteristic of the 3MLCT state. All of the complexes have emissive 3MLCT states, but with emission quantum yields that decrease with increasing n. The 1O2 quantum yields increase from 13 to 66% from one to three thiophenes but decrease to 40% at four. The parent complexes lacking thienyl groups generate 1O2 better than the Ru-nT complexes, with the exception that Ru-3T is slightly more efficient than Ru-phen.
Compared to Ru(II) polypyridyl complexes such as [Ru(bpy)3]2+, the parent [Ru(4,4′-btfmb)3]2+ complex is oxidized less readily but is reduced more easily owing to the electron-withdrawing nature of the −CF3-substituted bipyridyl coligands. All of the complexes undergo at least five reductions over the potential window investigated, whereas [Ru(bpy)3]2+ undergoes only three. For Ru-1T–Ru-4T, the additional oxidation due to the thienyl group (nT0/+) occurs before Ru2 oxidation. Ru-2T–Ru-4T underwent additional reductions involving their nT groups: one for Ru-2T, two for Ru-3T, and three for Ru-4T. The most redox-active of the series, Ru-4T, undergoes two oxidations and eight reductions over the potential window investigated.
Only complexes with thiophenes are phototoxic toward melanoma cells (SK-MEL-28), with potency increasing with n. Ru-4T demonstrated EC50 values in the low-nanomolar regime and PIs around 104 in normoxia under visible light. Notably, all values fell within one order of magnitude of each other over six biological replicates. Both Ru-3T and Ru-4T retain some of this activity in hypoxia, where Ru-4T has a visible EC50 as low as 35 nM and PI as high as 2,900. These results highlight the significance of lowest-lying 3ILCT states for enhancing phototoxicity toward cancer cells and maintaining activity in hypoxia. This family illustrates that Ru(II) coordinated to IP-3T or IP-4T ligands results in an excellent pharmacophore that can tolerate a variety of changes to the polypyridyl ancillary ligands. In this case, the redox activity of the system is changed substantially by the 4,4′-btfmb coligands (compared to our previously published analogs), and the complexes with three or four thiophene rings were still of high potency. The fluorinated polypyridyl ligands are of ongoing interest given the important role that fluorinated groups play in medicinal chemistry. These findings highlight Ru-4T as an excellent candidate for further investigations.
Supplementary Material
Acknowledgements
S.A.M. and C.G.C. thank the National Cancer Institute (NCI) of the National Institutes of Health (NIH) (Award R01CA222227) and the National Science Foundation (NSF) (Award NSF 2102459) for support. The content in this work is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. S.A.M. also thanks Dr. Daniel Todd as UNCG’s Triad Mass Spectrometry Facility manager and his assistants Jennifer Simpson and Diane Wallace. S.A.M. likewise thanks Dr. Franklin Moy (UNCG) and Dr. Brian Edwards (UTA) for their experimental support and instrument maintenance as NMR facility managers. A.F.M. thanks the grant PID2021-127554NA-I00 funded by the Spanish Ministry of Science and Innovation (MCIN/AEI/10.13039/501100011033) and by “ERDF A way of making Europe.” M.E.A. acknowledges the CINECA award under the ISCRA initiative (HyPS4DAT project) for the availability of high-performance computing resources.
Footnotes
The Supporting Information is available free of charge via the Internet at https://www.acs.org. Synthetic characterization (1D and 2D NMR, HPLC, and HRMS); computational details and additional results; spectroscopic characterization (emission and TA); electrochemical characterization; and (photo)biological data. (PDF)
S.A.M. has a potential research conflict of interest due to a financial interest with Theralase Technologies, Inc. and PhotoDynamic, Inc. A management plan has been created to preserve objectivity in research in accordance with UTA policy.
REFERENCES
- (1).Sung H; Ferlay J; Siegel RL; Laversanne M; Soerjomataram I; Jemal A; Bray F Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA A Cancer J Clin 2021, 71 (3), 209–249. 10.3322/caac.21660. [DOI] [PubMed] [Google Scholar]
- (2).Mellman I; Coukos G; Dranoff G Cancer Immunotherapy Comes of Age. Nature 2011, 480 (7378), 480–489. 10.1038/nature10673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).Marabelle A; Tselikas L; De Baere T; Houot R Intratumoral Immunotherapy: Using the Tumor as the Remedy. Annals of Oncology 2017, 28, xii33–xii43. 10.1093/annonc/mdx683. [DOI] [PubMed] [Google Scholar]
- (4).Meric-Bernstam F; Larkin J; Tabernero J; Bonini C Enhancing Anti-Tumour Efficacy with Immunotherapy Combinations. The Lancet 2021, 397 (10278), 1010–1022. 10.1016/S0140-6736(20)32598-8. [DOI] [PubMed] [Google Scholar]
- (5).Sawyers C Targeted Cancer Therapy. Nature 2004, 432 (7015), 294–297. 10.1038/nature03095. [DOI] [PubMed] [Google Scholar]
- (6).Min H-Y; Lee H-Y Molecular Targeted Therapy for Anticancer Treatment. Exp Mol Med 2022, 54 (10), 1670–1694. 10.1038/s12276-022-00864-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Mroz P; Hashmi JT; Huang Y-Y; Lange N; Hamblin MR Stimulation of Anti-Tumor Immunity by Photodynamic Therapy. Expert Rev. Clin. Immunol 2011, 7 (1), 75–91. 10.1586/eci.10.81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Anzengruber F; Avci P; de Freitas LF; Hamblin MR T-Cell Mediated Anti-Tumor Immunity after Photodynamic Therapy: Why Does It Not Always Work and How Can We Improve It? Photochem. Photobiol. Sci 2015, 14 (8), 1492–1509. 10.1039/C4PP00455H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Gollnick SO; Brackett CM Enhancement of Anti-Tumor Immunity by Photodynamic Therapy. Immunologic Research 2010, 46 (1–3), 216–226. 10.1007/s12026-009-8119-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).Gollnick SO Photodynamic Therapy and Antitumor Immunity. J Natl Compr Canc Netw 2012, 10 Suppl 2, S40–43. 10.6004/jnccn.2012.0173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Monro S; Colón KL; Yin H; Roque J; Konda P; Gujar S; Thummel RP; Lilge L; Cameron CG; McFarland SA Transition Metal Complexes and Photodynamic Therapy from a Tumor-Centered Approach: Challenges, Opportunities, and Highlights from the Development of TLD1433. Chem. Rev 2019, 119 (2), 797–828. 10.1021/acs.chemrev.8b00211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (12).McFarland SA; Mandel A; Dumoulin-White R; Gasser G Metal-Based Photosensitizers for Photodynamic Therapy: The Future of Multimodal Oncology? Curr. Opin. Chem. Biol 2020, 56, 23–27. 10.1016/j.cbpa.2019.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (13).Howerton BS; Heidary DK; Glazer EC Strained Ruthenium Complexes Are Potent Light-Activated Anticancer Agents. Journal of the American Chemical Society 2012, 134 (20), 8324–8327. 10.1021/ja3009677. [DOI] [PubMed] [Google Scholar]
- (14).Sainuddin T; Pinto M; Yin H; Hetu M; Colpitts J; McFarland SA Strained Ruthenium Metal–Organic Dyads as Photocisplatin Agents with Dual Action. J. Inorg. Biochem 2016, 158, 45–54. 10.1016/j.jinorgbio.2016.01.009. [DOI] [PubMed] [Google Scholar]
- (15).Roque J; Havrylyuk D; Barrett PC; Sainuddin T; McCain J; Colón K; Sparks WT; Bradner E; Monro S; Heidary D; Cameron CG; Glazer EC; McFarland SA Strained, Photoejecting Ru(II) Complexes That Are Cytotoxic Under Hypoxic Conditions. Photochem. Photobiol 2020, 96 (2), 327–339. 10.1111/php.13174. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (16).Cole HD; Roque JA; Lifshits LM; Hodges R; Barrett PC; Havrylyuk D; Heidary D; Ramasamy E; Cameron CG; Glazer EC; McFarland SA Fine-Feature Modifications to Strained Ruthenium Complexes Radically Alter Their Hypoxic Anticancer Activity. Photochem & Photobiology 2022, 98 (1), 73–84. 10.1111/php.13395. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (17).Roque III JA; Cole HD; Barrett PC; Lifshits LM; Hodges RO; Kim S; Deep G; Francés-Monerris A; Alberto ME; Cameron CG; McFarland SA Intraligand Excited States Turn a Ruthenium Oligothiophene Complex into a Light-Triggered Ubertoxin with Anticancer Effects in Extreme Hypoxia. J. Am. Chem. Soc 2022, 144 (18), 8317–8336. 10.1021/jacs.2c02475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).Cuello-Garibo J-A; Meijer MS; Bonnet S To Cage or to Be Caged? The Cytotoxic Species in Ruthenium-Based Photoactivated Chemotherapy Is Not Always the Metal. Chem. Commun 2017, 53 (50), 6768–6771. 10.1039/C7CC03469E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (19).Azar DF; Audi H; Farhat S; El-Sibai M; Abi-Habib RJ; Khnayzer RS Phototoxicity of Strained Ru(ii) Complexes: Is It the Metal Complex or the Dissociating Ligand? Dalton Trans 2017, 46 (35), 11529–11532. 10.1039/C7DT02255G. [DOI] [PubMed] [Google Scholar]
- (20).Cole HD; Roque JA; Shi G; Lifshits LM; Ramasamy E; Barrett PC; Hodges RO; Cameron CG; McFarland SA Anticancer Agent with Inexplicable Potency in Extreme Hypoxia: Characterizing a Light-Triggered Ruthenium Ubertoxin. J. Am. Chem. Soc 2022, 144 (22), 9543–9547. 10.1021/jacs.1c09010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (21).Sun Y; Heidary DK; Zhang Z; Richards CI; Glazer EC Bacterial Cytological Profiling Reveals the Mechanism of Action of Anticancer Metal Complexes. Mol. Pharm 2018, 15 (8), 3404–3416. 10.1021/acs.molpharmaceut.8b00407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Loftus LM; White JK; Albani BA; Kohler L; Kodanko JJ; Thummel RP; Dunbar KR; Turro C New RuII Complex for Dual Activity: Photoinduced Ligand Release and 1O2 Production. Chem. Eur. J 2016, 22 (11), 3704–3708. 10.1002/chem.201504800. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (23).Toupin NP; Nadella S; Steinke SJ; Turro C; Kodanko JJ Dual-Action Ru(II) Complexes with Bulky π-Expansive Ligands: Phototoxicity without DNA Intercalation. Inorganic Chemistry 2020, 59 (6), 3919–3933. 10.1021/acs.inorgchem.9b03585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (24).Lameijer LN; Ernst D; Hopkins SL; Meijer MS; Askes SHC; Le Dévédec SE; Bonnet S A Red-Light-Activated Ruthenium-Caged NAMPT Inhibitor Remains Phototoxic in Hypoxic Cancer Cells. Angew. Chem. Int. Ed 2017, 56 (38), 11549–11553. 10.1002/anie.201703890. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (25).Albani BA; Durr CB; Turro C Selective Photoinduced Ligand Exchange in a New Tris–Heteroleptic Ru(II) Complex. The Journal of Physical Chemistry A 2013, 117 (50), 13885–13892. 10.1021/jp4085684. [DOI] [PubMed] [Google Scholar]
- (26).Knoll JD; Albani BA; Durr CB; Turro C Unusually Efficient Pyridine Photodissociation from Ru(II) Complexes with Sterically Bulky Bidentate Ancillary Ligands. The Journal of Physical Chemistry A 2014, 118 (45), 10603–10610. 10.1021/jp5057732. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (27).Arora K; White JK; Sharma R; Mazumder S; Martin PD; Schlegel HB; Turro C; Kodanko JJ Effects of Methyl Substitution in Ruthenium Tris(2-Pyridylmethyl)Amine Photocaging Groups for Nitriles. Inorganic Chemistry 2016, 55 (14), 6968–6979. 10.1021/acs.inorgchem.6b00650. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (28).Huisman M; White JK; Lewalski VG; Podgorski I; Turro C; Kodanko JJ Caging the Uncageable: Using Metal Complex Release for Photochemical Control over Irreversible Inhibition. Chemical Communications 2016, 52 (85), 12590–12593. 10.1039/C6CC07083C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Li A; Yadav R; White JK; Herroon MK; Callahan BP; Podgorski I; Turro C; Scott EE; Kodanko JJ Illuminating Cytochrome P450 Binding: Ru(ii)-Caged Inhibitors of CYP17A1. Chem. Commun 2017, 53 (26), 3673–3676. 10.1039/C7CC01459G. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (30).Arora K; Herroon M; Al-Afyouni MH; Toupin NP; Rohrabaugh TN; Loftus LM; Podgorski I; Turro C; Kodanko JJ Catch and Release Photosensitizers: Combining Dual-Action Ruthenium Complexes with Protease Inactivation for Targeting Invasive Cancers. J. Am. Chem. Soc 2018, 140 (43), 14367–14380. 10.1021/jacs.8b08853. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Li A; Turro C; Kodanko JJ Ru(II) Polypyridyl Complexes Derived from Tetradentate Ancillary Ligands for Effective Photocaging. Accounts of Chemical Research 2018, 51 (6), 1415–1421. 10.1021/acs.accounts.8b00066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (32).Li A; Turro C; Kodanko JJ Ru(ii) Polypyridyl Complexes as Photocages for Bioactive Compounds Containing Nitriles and Aromatic Heterocycles. Chem. Commun 2018, 54 (11), 1280–1290. 10.1039/C7CC09000E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (33).Nisbett K; Tu Y-J; Turro C; Kodanko JJ; Schlegel HB DFT Investigation of Ligand Photodissociation in [RuII(Tpy)(Bpy)(Py)]2+ and [RuII(Tpy)(Me2bpy)(Py)]2+ Complexes. Inorganic Chemistry 2018, 57 (1), 231–240. 10.1021/acs.inorgchem.7b02398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (34).Rohrabaugh TN; Rohrabaugh AM; Kodanko JJ; White JK; Turro C Photoactivation of Imatinib–Antibody Conjugate Using Low-Energy Visible Light from Ru(ii)-Polypyridyl Cages. Chem. Commun 2018, 54 (41), 5193–5196. 10.1039/C8CC01348A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (35).Bahreman A; Limburg B; Siegler MA; Bouwman E; Bonnet S Spontaneous Formation in the Dark, and Visible Light-Induced Cleavage, of a Ru–S Bond in Water: A Thermodynamic and Kinetic Study. Inorganic Chemistry 2013, 52 (16), 9456–9469. 10.1021/ic401105v. [DOI] [PubMed] [Google Scholar]
- (36).Bahreman A; Rabe M; Kros A; Bruylants G; Bonnet S Binding of a Ruthenium Complex to a Thioether Ligand Embedded in a Negatively Charged Lipid Bilayer: A Two-Step Mechanism. Chemistry - A European Journal 2014, 20 (24), 7429–7438. 10.1002/chem.201400377. [DOI] [PubMed] [Google Scholar]
- (37).Göttle AJ; Alary F; Boggio-Pasqua M; Dixon IM; Heully J-L; Bahreman A; Askes SHC; Bonnet S Pivotal Role of a Pentacoordinate 3MC State on the Photocleavage Efficiency of a Thioether Ligand in Ruthenium(II) Complexes: A Theoretical Mechanistic Study. Inorganic Chemistry 2016, 55 (9), 4448–4456. 10.1021/acs.inorgchem.6b00268. [DOI] [PubMed] [Google Scholar]
- (38).Cuello-Garibo J-A; Pérez-Gallent E; van der Boon L; Siegler MA; Bonnet S Influence of the Steric Bulk and Solvent on the Photoreactivity of Ruthenium Polypyridyl Complexes Coordinated to L-Proline. Inorganic Chemistry 2017, 56 (9), 4818–4828. 10.1021/acs.inorgchem.6b02794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (39).Sun W; Wen Y; Thiramanas R; Chen M; Han J; Gong N; Wagner M; Jiang S; Meijer MS; Bonnet S; Butt H-J; Mailänder V; Liang X-J; Wu S Red-Light-Controlled Release of Drug-Ru Complex Conjugates from Metallopolymer Micelles for Phototherapy in Hypoxic Tumor Environments. Adv. Funct. Mater 2018, 28 (39), 1804227. 10.1002/adfm.201804227. [DOI] [Google Scholar]
- (40).Meijer MS; Talens VS; Hilbers MF; Kieltyka RE; Brouwer AM; Natile MM; Bonnet S NIR-Light-Driven Generation of Reactive Oxygen Species Using Ru(II)-Decorated Lipid-Encapsulated Upconverting Nanoparticles. Langmuir 2019, 35 (37), 12079–12090. 10.1021/acs.langmuir.9b01318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (41).Havrylyuk D; Hachey AC; Fenton A; Heidary DK; Glazer EC Ru(II) Photocages Enable Precise Control over Enzyme Activity with Red Light. Nat Commun 2022, 13 (1), 3636. 10.1038/s41467-022-31269-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (42).Ryan RT; Havrylyuk D; Stevens KC; Moore LH; Parkin S; Blackburn JS; Heidary DK; Selegue JP; Glazer EC Biological Investigations of Ru(II) Complexes with Diverse β-Diketone Ligands. Eur. J. Inorg. Chem 2021, 2021 (35), 3611–3621. 10.1002/ejic.202100468. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (43).Wachter E; Heidary DK; Howerton BS; Parkin S; Glazer EC Light-Activated Ruthenium Complexes Photobind DNA and Are Cytotoxic in the Photodynamic Therapy Window. Chem. Commun 2012, 48 (77), 9649. 10.1039/c2cc33359g. [DOI] [PubMed] [Google Scholar]
- (44).Huang H; Banerjee S; Qiu K; Zhang P; Blacque O; Malcomson T; Paterson MJ; Clarkson GJ; Staniforth M; Stavros VG; Gasser G; Chao H; Sadler PJ Targeted Photoredox Catalysis in Cancer Cells. Nat. Chem 2019, 11 (11), 1041–1048. 10.1038/s41557-019-0328-4. [DOI] [PubMed] [Google Scholar]
- (45).Baptista MS; Cadet J; Di Mascio P; Ghogare AA; Greer A; Hamblin MR; Lorente C; Nunez SC; Ribeiro MS; Thomas AH; Vignoni M; Yoshimura TM Type I and Type II Photosensitized Oxidation Reactions: Guidelines and Mechanistic Pathways. Photochemistry and Photobiology 2017, 93 (4), 912–919. 10.1111/php.12716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (46).Roque III JA; Barrett PC; Cole HD; Lifshits LM; Shi G; Monro S; von Dohlen D; Kim S; Russo N; Deep G; Cameron CG; Alberto ME; McFarland SA Breaking the Barrier: An Osmium Photosensitizer with Unprecedented Hypoxic Phototoxicity for Real World Photodynamic Therapy. Chem. Sci 2020, 11, 9784–9806. 10.1039/D0SC03008B. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (47).Shi G; Monro S; Hennigar R; Colpitts J; Fong J; Kasimova K; Yin H; DeCoste R; Spencer C; Chamberlain L; Mandel A; Lilge L; McFarland SA Ru(II) Dyads Derived from α-Oligothiophenes: A New Class of Potent and Versatile Photosensitizers for PDT. Coord. Chem. Rev 2015, 282–283, 127–138. 10.1016/j.ccr.2014.04.012. [DOI] [Google Scholar]
- (48).Kulkarni GS; Lilge L; Nesbitt M; Dumoulin-White RJ; Mandel A; Jewett MAS A Phase 1b Clinical Study of Intravesical Photodynamic Therapy in Patients with Bacillus Calmette-Guérin–Unresponsive Non–Muscle-Invasive Bladder Cancer. European Urology Open Science 2022, 41, 105–111. 10.1016/j.euros.2022.04.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Roque JA; Barrett PC; Cole HD; Lifshits LM; Bradner E; Shi G; von Dohlen D; Kim S; Russo N; Deep G; Cameron CG; Alberto ME; McFarland SA Os(II) Oligothienyl Complexes as a Hypoxia-Active Photosensitizer Class for Photodynamic Therapy. Inorg. Chem 2020, 59 (22), 16341–16360. 10.1021/acs.inorgchem.0c02137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Lifshits LM; Roque JA; Cole HD; Thummel RP; Cameron CG; McFarland SA NIR-Absorbing Ru II Complexes Containing α-Oligothiophenes for Applications in Photodynamic Therapy. ChemBioChem 2020, 21, 3594–3607. 10.1002/cbic.202000419. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (51).Cole HD; Vali A; Roque JA; Shi G; Kaur G; Hodges RO; Francés-Monerris A; Alberto ME; Cameron CG; McFarland SA Ru(II) Phenanthroline-Based Oligothienyl Complexes as Phototherapy Agents. Inorg. Chem 2023, 62 (51), 21181–21200. 10.1021/acs.inorgchem.3c03216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (52).Armstrong DW; Yu J; Cole HD; McFarland Sherri. A.; Nafie J Chiral Resolution and Absolute Configuration Determination of New Metal-Based Photodynamic Therapy Antitumor Agents. J. Pharm. Biom. Anal 2021, 204, 114233. 10.1016/j.jpba.2021.114233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (53).Handlovic TT; Wahab MF; Cole HD; Alatrash N; Ramasamy E; MacDonnell FM; McFarland SA; Armstrong DW Insights into Enantioselective Separations of Ionic Metal Complexes by Sub/Supercritical Fluid Chromatography. Analytica Chimica Acta 2022, 1228, 340156. 10.1016/j.aca.2022.340156. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (54).Farney EP; Chapman SJ; Swords WB; Torelli MD; Hamers RJ; Yoon TP Discovery and Elucidation of Counteranion Dependence in Photoredox Catalysis. J. Am. Chem. Soc 2019, 141 (15), 6385–6391. 10.1021/jacs.9b01885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (55).Furue M; Maruyama K; Oguni T; Naiki M; Kamachi M Trifluoromethyl-Substituted 2,2’-Bipyridine Ligands. Synthetic Control of Excited-State Properties of Ruthenium(II) Tris-Chelate Complexes. Inorg. Chem 1992, 31 (18), 3792–3795. 10.1021/ic00044a022. [DOI] [Google Scholar]
- (56).Maurer AB; Piechota EJ; Meyer GJ Excited-State Dipole Moments of Homoleptic [Ru(Bpy′)3]2+ Complexes Measured by Stark Spectroscopy. J. Phys. Chem. A 2019, 123 (41), 8745–8754. 10.1021/acs.jpca.9b05874. [DOI] [PubMed] [Google Scholar]
- (57).Nomrowski J; Wenger OS Photoinduced PCET in Ruthenium–Phenol Systems: Thermodynamic Equivalence of Uni- and Bidirectional Reactions. Inorg. Chem 2015, 54 (7), 3680–3687. 10.1021/acs.inorgchem.5b00318. [DOI] [PubMed] [Google Scholar]
- (58).Sullivan B; Salmon D; Meyer T Mixed Phosphine 2,2’-Bipyridine Complexes of Ruthenium 1978, 17, 3334–3341. [Google Scholar]
- (59).Wang Z Comprehensive Organic Name Reactions and Reagents; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2010. 10.1002/9780470638859. [DOI] [Google Scholar]
- (60).Ghosh G; Colón KL; Fuller A; Sainuddin T; Bradner E; McCain J; Monro SMA; Yin H; Hetu MW; Cameron CG; McFarland SA Cyclometalated Ruthenium(II) Complexes Derived from α-Oligothiophenes as Highly Selective Cytotoxic or Photocytotoxic Agents. Inorg. Chem 2018, 57 (13), 7694–7712. 10.1021/acs.inorgchem.8b00689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (61).Chen R; Yang X; Tian H; Wang X; Hagfeldt A; Sun L Effect of Tetrahydroquinoline Dyes Structure on the Performance of Organic Dye-Sensitized Solar Cells. Chem. Mater 2007, 19 (16), 4007–4015. 10.1021/cm070617g. [DOI] [Google Scholar]
- (62).Casida ME Time–Dependent Density Functional Response Theory of Molecular Systems: Theory, Computational Methods, and Functionals. In Recent developments and applications of modern density functional theory; Seminario JM, Ed.; Theoretical and computational chemistry; Elsevier: Amsterdam; Netherlands, 1996; pp 155–192. [Google Scholar]
- (63).Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Petersson GA; Nakatsuji H; Li X; Caricato M; Marenich AV; Bloino J; Janesko BG; Gomperts R; Mennucci B; Hratchian HP; Ortiz JV; Izmaylov AF; Sonnenberg JL; Williams-Young D; Ding F; Lipparini F; Egidi F; Goings J; Peng B; Petrone A; Henderson T; Ranasinghe D; Zakrzewski VG; Gao J; Rega N; Zheng G; Liang W; Hada M; Ehara M; Toyota K; Fukuda R; Hasegawa J; Ishida M; Nakajima T; Honda Y; Kitao O; Nakai H; Vreven T; Throssell K; Montgomery JA Jr.; Peralta JE; Ogliaro F; Bearpark MJ; Heyd JJ; Brothers EN; Kudin KN; Staroverov VN; Keith TA; Kobayashi R; Normand J; Raghavachari K; Rendell AP; Burant JC; Iyengar SS; Tomasi J; Cossi M; Millam JM; Klene M; Adamo C; Cammi R; Ochterski JW; Martin RL; Morokuma K; Farkas O; Foresman JB; Fox DJ Gaussian 16 Revision C.01; 2016. [Google Scholar]
- (64).Ponte F; Alberto ME; De Simone BC; Russo N; Sicilia E Photophysical Exploration of Dual-Approach PtII–BODIPY Conjugates: Theoretical Insights. Inorg. Chem 2019, 58 (15), 9882–9889. 10.1021/acs.inorgchem.9b01002. [DOI] [PubMed] [Google Scholar]
- (65).Alberto ME; Francés-Monerris A A Multiscale Free Energy Method Reveals an Unprecedented Photoactivation of a Bimetallic Os(ii)–Pt(ii) Dual Anticancer Agent. Phys. Chem. Chem. Phys 2022, 24 (32), 19584–19594. 10.1039/D2CP02128E. [DOI] [PubMed] [Google Scholar]
- (66).Alberto ME; Russo N; Adamo C Synergistic Effects of Metals in a Promising RuII–PtII Assembly for a Combined Anticancer Approach: Theoretical Exploration of the Photophysical Properties. Chemistry - A European Journal 2016, 22 (27), 9162–9168. 10.1002/chem.201601089. [DOI] [PubMed] [Google Scholar]
- (67).Alberto ME; Pirillo J; Russo N; Adamo C Theoretical Exploration of Type I/Type II Dual Photoreactivity of Promising Ru(II) Dyads for PDT Approach. Inorg. Chem 2016, 55 (21), 11185–11192. 10.1021/acs.inorgchem.6b01782. [DOI] [PubMed] [Google Scholar]
- (68).Bertini L; Alberto ME; Arrigoni F; Vertemara J; Fantucci P; Bruschi M; Zampella G; De Gioia L On the Photochemistry of Fe2(Edt)(CO)4(PMe3)2, a [FeFe]-Hydrogenase Model: A DFT/TDDFT Investigation. Int J Quantum Chem 2018, 118 (9), e25537. 10.1002/qua.25537. [DOI] [Google Scholar]
- (69).Alberto ME; Adamo C Synergistic Effects in PtII-Porphyrinoid Dyes as Candidates for a Dual-Action Anticancer Therapy: A Theoretical Exploration. Chemistry - A European Journal 2017, 23 (60), 15124–15132. 10.1002/chem.201702876. [DOI] [PubMed] [Google Scholar]
- (70).Francés-Monerris A; Magra K; Darari M; Cebrián C; Beley M; Domenichini E; Haacke S; Pastore M; Assfeld X; Gros PC; Monari A Synthesis and Computational Study of a Pyridylcarbene Fe(II) Complex: Unexpected Effects of Fac/ Mer Isomerism in Metal-to-Ligand Triplet Potential Energy Surfaces. Inorg. Chem 2018, 57 (16), 10431–10441. 10.1021/acs.inorgchem.8b01695. [DOI] [PubMed] [Google Scholar]
- (71).Alberto ME; Mazzone G; Regina C; Russo N; Sicilia E Theoretical Exploration of the Photophysical Properties of Two-Component RuII–Porphyrin Dyes as Promising Assemblies for a Combined Antitumor Effect. Dalton Trans 2020, 49 (36), 12653–12661. 10.1039/D0DT02197K. [DOI] [PubMed] [Google Scholar]
- (72).Adamo C; Barone V Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model. The Journal of Chemical Physics 1999, 110 (13), 6158–6170. 10.1063/1.478522. [DOI] [Google Scholar]
- (73).Andrae D; Häußermann U; Dolg M; Stoll H; Preuß H Energy-Adjusted Ab Initio Pseudopotentials for the Second and Third Row Transition Elements. Theoret. Chim. Acta 1990, 77 (2), 123–141. 10.1007/BF01114537. [DOI] [Google Scholar]
- (74).Hirata S; Head-Gordon M Time-Dependent Density Functional Theory within the Tamm–Dancoff Approximation. Chemical Physics Letters 1999, 314 (3–4), 291–299. 10.1016/S0009-2614(99)01149-5. [DOI] [Google Scholar]
- (75).Peach MJG; Williamson MJ; Tozer DJ Influence of Triplet Instabilities in TDDFT. J. Chem. Theory Comput 2011, 7 (11), 3578–3585. 10.1021/ct200651r. [DOI] [PubMed] [Google Scholar]
- (76).Cossi M; Barone V Solvent Effect on Vertical Electronic Transitions by the Polarizable Continuum Model. The Journal of Chemical Physics 2000, 112 (5), 2427–2435. 10.1063/1.480808. [DOI] [Google Scholar]
- (77).Tomasi J; Mennucci B; Cammi R Quantum Mechanical Continuum Solvation Models. Chemical Reviews 2005, 105 (8), 2999–3094. 10.1021/cr9904009. [DOI] [PubMed] [Google Scholar]
- (78).Cossi M; Rega N; Scalmani G; Barone V Energies, Structures, and Electronic Properties of Molecules in Solution with the C-PCM Solvation Model. J. Comput. Chem 2003, 24 (6), 669–681. 10.1002/jcc.10189. [DOI] [PubMed] [Google Scholar]
- (79).Skripnikov L. Chemissan. http://www.chemissian.com.
- (80).Plasser F TheoDORE: A Toolbox for a Detailed and Automated Analysis of Electronic Excited State Computations. The Journal of Chemical Physics 2020, 152 (8), 084108. 10.1063/1.5143076. [DOI] [PubMed] [Google Scholar]
- (81).Juris A; Campagna S; Balzani V; Gremaud G; Von Zelewsky A Absorption Spectra, Luminescence Properties, and Electrochemical Behavior of Tris-Heteroleptic Ruthenium(II) Polypyridine Complexes. Inorg. Chem 1988, 27 (20), 3652–3655. 10.1021/ic00293a043. [DOI] [Google Scholar]
- (82).Becker RS; Seixas de Melo J; Maçanita AL; Elisei F Comprehensive Evaluation of the Absorption, Photophysical, Energy Transfer, Structural, and Theoretical Properties of α-Oligothiophenes with One to Seven Rings. J. Phys. Chem 1996, 100, 18683–18695. 10.1021/jp960852e. [DOI] [Google Scholar]
- (83).DeRosa M Photosensitized Singlet Oxygen and Its Applications. Coordination Chemistry Reviews 2002, 233–234, 351–371. 10.1016/S0010-8545(02)00034-6. [DOI] [Google Scholar]
- (84).Hissler M; Connick WB; Geiger DK; McGarrah JE; Lipa D; Lachicotte RJ; Eisenberg R Platinum Diimine Bis(Acetylide) Complexes: Synthesis, Characterization, and Luminescence Properties. Inorg. Chem 2000, 39 (3), 447–457. 10.1021/ic991250n. [DOI] [PubMed] [Google Scholar]
- (85).Lincoln R; Kohler L; Monro S; Yin H; Stephenson M; Zong R; Chouai A; Dorsey C; Hennigar R; Thummel RP; McFarland SA Exploitation of Long-Lived 3IL Excited States for Metal–Organic Photodynamic Therapy: Verification in a Metastatic Melanoma Model. J. Am. Chem. Soc 2013, 135 (45), 17161–17175. 10.1021/ja408426z. [DOI] [PubMed] [Google Scholar]
- (86).Goze C; Kozlov DV; Tyson DS; Ziessel R; Castellano FN Synthesis and Photophysics of Ruthenium(Ii) Complexes with Multiple Pyrenylethynylene Subunits. New Journal of Chemistry 2003, 27 (12), 1679. 10.1039/b307327k. [DOI] [Google Scholar]
- (87).Monro S; Cameron CG; Zhu X; Colón KL; Yin H; Sainuddin T; Hetu M; Pinto M; Fuller A; Bennett L; Roque J; Sun W; McFarland SA Synthesis, Characterization and Photobiological Studies of Ru(II) Dyads Derived from α-Oligothiophene Derivatives of 1,10-Phenanthroline. Photochem. Photobiol 2019, 95 (1), 267–279. 10.1111/php.13012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (88).Chettri A; Schneider KRA; Cole HD; Roque JA; Cameron CG; McFarland SA; Dietzek B String-Attached Oligothiophene Substituents Determine the Fate of Excited States in Ruthenium Complexes for Photodynamic Therapy. J. Phys. Chem. A 2021, 125 (32), 6985–6994. 10.1021/acs.jpca.1c04900. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (89).Chettri A; Roque JA; Schneider KRA; Cole HD; Cameron CG; McFarland SA; Dietzek B It Takes Three to Tango: The Length of the Oligothiophene Chain Determines the Nature of the Long-Lived Excited State and the Resulting Photocytotoxicity of a Ruthenium(II) Photodrug. ChemPhotoChem 2021, 5 (5), 421–425. 10.1002/cptc.202000283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (90).McClenaghan ND; Leydet Y; Maubert B; Indelli MT; Campagna S Excited-State Equilibration: A Process Leading to Long-Lived Metal-to-Ligand Charge Transfer Luminescence in Supramolecular Systems. Coordination Chemistry Reviews 2005, 249 (13–14), 1336–1350. 10.1016/j.ccr.2004.12.017. [DOI] [Google Scholar]
- (91).Juris A; Balzani V; Barigelletti F; Campagna S; Belser P; von Zelewsky A Ru(II) Polypyridine Complexes: Photophysics, Photochemistry, Eletrochemistry, and Chemiluminescence. Coordination Chemistry Reviews 1988, 84, 85–277. 10.1016/0010-8545(88)80032-8. [DOI] [Google Scholar]
- (92).Ohsawa Y; DeArmond MK; Hanck KW; Morris DE; Whitten DG; Neveux PE Spatially Isolated Redox Orbitals: Evidence from Low-Temperature Voltammetry. J. Am. Chem. Soc 1983, 105 (21), 6522–6524. 10.1021/ja00359a045. [DOI] [Google Scholar]
- (93).Ohsawa Y; Hanck KW; DeArmond MK A Systematic Electrochemical and Spectroscopic Study of Mixed-Ligand Ruthenium(II) 2,2′-Bipyridine Complexes [Ru(Bpy)3-nLn]2+ (N=0,1,2 and 3). Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 1984, 175 (1–2), 229–240. 10.1016/S0022-0728(84)80358-7. [DOI] [Google Scholar]
- (94).Roncali J Conjugated Poly(Thiophenes): Synthesis, Functionalization, and Applications. Chemical Reviews 1992, 92 (4), 711–738. 10.1021/cr00012a009. [DOI] [Google Scholar]
- (95).Pavlishchuk VV; Addison AW Conversion Constants for Redox Potentials Measured versus Different Reference Electrodes in Acetonitrile Solutions at 25°C. Inorganica Chimica Acta 2000, 298 (1), 97–102. 10.1016/S0020-1693(99)00407-7. [DOI] [Google Scholar]
- (96).Diaz AF; Crowley J; Bargon J; Gardini GP; Torrance JB Electrooxidation of Aromatic Oligomers and Conducting Polymers. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 1981, 121, 355–361. 10.1016/S0022-0728(81)80592-X. [DOI] [Google Scholar]
- (97).Yin H; Stephenson M; Gibson J; Sampson E; Shi G; Sainuddin T; Monro S; McFarland SA In Vitro Multiwavelength PDT with 3IL States: Teaching Old Molecules New Tricks. Inorganic Chemistry 2014, 53 (9), 4548–4559. 10.1021/ic5002368. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
