Abstract
Mitochondrial uncouplers are actively sought as potential therapeutics. Here, we report the first successful synthesis of mitochondria-targeted derivatives of the highly potent uncoupler 3,5-ditert-butyl-4-hydroxybenzylidene-malononitrile (SF6847), bearing a cationic alkyl(triphenyl)phosphonium (TPP) group. As a key step of the synthesis, we used condensation of a ketophenol with malononitrile via the Knoevenagel reaction. SF-C5-TPP with a pentamethylene linker between SF6847 and TPP, stimulating respiration and collapsing membrane potential of rat liver mitochondria at submicromolar concentrations, proved to be the most effective uncoupler of the series. SF-C5-TPP showed pronounced protonophoric activity on a model planar bilayer lipid membrane. Importantly, SF-C5-TPP exhibited rather low toxicity in fibroblast cell culture, causing mitochondrial depolarization in cells at concentrations that only slightly affected cell viability. SF-C5-TPP was more effective in decreasing the mitochondrial membrane potential in the cell culture than SF6847, in contrast to the case of isolated mitochondria. Like other zwitterionic uncouplers, SF-C5-TPP inhibited the growth of Bacillus subtilis in the micromolar concentration range.
Introduction
Compounds causing marked respiratory stimulation were described 90 years ago1 and later called uncouplers2 because they break the coupling between the oxidation of organic compounds and ATP synthesis; in other words, they uncouple phosphorylation from oxidation. Even in the early years, it became clear that uncoupling could be beneficial under certain conditions.3 At present, mitochondrial uncouplers are considered promising in fighting various diseases.4−7 Yet the medicinal application of the majority of uncouplers is precluded by rather high toxicity.8 To this end, making their action more targeted seems to be rational. Several attempts have been made to create a mitochondria-targeted uncoupler by attaching a lipophilic cationic group, such as triphenylphosphonium (TPP), widely used for mitochondria targeting,9−13 to conventional uncouplers, such as 2,4-dinitrophenol (DNP)14 and carbonyl cyanide m-chlorophenyl hydrazone (CCCP).15 Unfortunately, these derivatives appeared to be weak uncouplers. By contrast, linking decyl-TPP to fluorescein yielded a rather effective mitochondrial uncoupler coined mitoFluo, albeit showing poor protonophoric activity on model lipid membranes.16 Of importance, mitoFluo exhibited neuro- and nephroprotective effects in rat models of traumatic brain and kidney injuries,17 which was obviously associated with the reduction of ROS generation by mitochondria, i.e., the antioxidant action of mitoFluo. Here, we report the synthesis of one of the most potent small-molecule mitochondrial uncouplers 3,5-ditert-butyl-4-hydroxybenzylidene-malononitrile (SF6847)18 appended to alkyl-TPP with varying carbon chain length, SF-Cn-TPP. Of note, SF6847, introduced in 197119 and extensively studied later,20−23 now still remains a promising scaffold for the design of new mitochondrial bioenergetics’ inhibitors.24 Importantly, the pentamethylene linker-containing derivative SF-C5-TPP not only stimulated respiration and decreased membrane potential of isolated rat liver mitochondria (RLM) at submicromolar concentrations, although less effectively than SF6847, but also elicited transmembrane proton current across planar bilayer lipid membranes. In fibroblast cell culture, SF-C5-TPP was more effective in mitochondrial depolarization than SF6847, in contrast to isolated RLM.
Results and Discussion
Synthesis of SF-Cn-TPP
SF-Cn-TPP conjugates were prepared in three steps starting from 2,6-ditert-butylphenol according to Scheme 1.
Scheme 1. Synthesis of SF-Cn-TPP Conjugates.
The Friedel–Crafts acylation of 2,6-ditert-butylphenol with chloroanhydrides of ω-bromoalkane carboxylic acids was carried out in the presence of titanium tetrachloride at −20 °C. The low temperature seems to be crucial for the successful implementation of this reaction because the dealkylation and/or migration of tert-butyl substituents occurred at 0 °C or above.25,26 The Knoevenagel condensation of 4-acyl-2,6-ditert-butylphenols with malononitrile appears to be a rather difficult task even in the case of the simplest representative of these ketophenols, such as 4-acetyl-2,6-ditert-butylphenol. In fact, the known methyl-modified SF6847 (TX-1122)27,28 was obtained by heating in benzene in the presence of ammonium acetate as a catalyst, but the yield was poor. We found that the previously described method,29 in which condensation occurred in hexamethyldisilazane (HMDS)/acetic acid mixture, allows the preparation of alkyl-containing derivatives of SF6847 in good yields (unpublished results). To our great surprise, the method of Knoevenagel condensation appeared to be useful also in the preparation of ω-bromoalkyl-modified SF6847 (SF-CnBr). This procedure implies that HMDS transforms in the presence of acetic acid to AcOSiMe3, which serves as a water removal agent, and AcONH4, which is a promoter. There were concerns that acetate might substitute for bromide in such harsh conditions. Indeed, the reaction of 4-(5′-bromopentanoyl)–2,6-ditert-butylphenol, I–C4Br (Figure S1), with malononitrile resulted in a mixture of three compounds (Scheme 2), and desirable SF-C4Br was isolated only in a poor yield of 23%.
Scheme 2. Reaction of I–C4Br with Malononitrile.

Surprisingly, if the bromoalkyl substituent contained five methylene groups or more, the reaction proceeded smoothly, and no analogous byproducts were observed. ω-Bromoalkyl-modified SF6847 (SF-CnBr) smoothly reacted with triphenylphosphine to give desirable SF-Cn-TPP conjugates in good to excellent yields. The 1H, 13C, and 31P NMR spectra and Liquid chromatography-mass spectrometry (LC-MS) data for the compounds are presented in Figures S1–S29.
Estimation of pKa of SF-C5-TPP
As expected from the literature data on the pKa of SF6847,21,22 the pKa value measured spectrophotometrically from the pH dependence of absorbance at a maximum of the SF-C5-TPP spectrum (Figure S30) was equal to 7.17 in the aqueous solution and 7.16 in the presence of DPhPC liposomes. Therefore, similar to the majority of conventional uncouplers,30 SF-C5-TPP is a weak acid with pKa close to physiological pH.
SF-Cn-TPP Stimulated Respiration of the Isolated Rat Liver Mitochondria
Figure 1 displays the effect of SF-Cn-TPP on RLM respiration upon the addition of ADP (100 μM). As it is seen from the time courses of oxygen consumption by RLM (a), SF-C5-TPP caused pronounced acceleration of RLM respiration at a concentration of 250 nM both before the addition of ADP and in state 4 (after ADP conversion into ATP). The comparison of the concentration dependences of the RLM respiration rate for a series of SF-Cn-TPP with n = 4, 5, 10 (b) shows that the pentamethylene linker-containing derivative was the most potent. Alkyl-TPP compounds lacking the 3,5-ditert-butyl-4-hydroxybenzylidene-malononitrile group, C4-TPP and C10-TPP, were of no effect at the submicromolar and micromolar concentrations (open circles and filled triangles), while SF6847 (open diamonds) was effective at nanomolar concentrations.
Figure 1.
Effect of SF-C5-TPP at various concentrations on the basal respiration rate and respiration rates in states 3 and 4 of the isolated RLM (a). Respiration rates are indicated as relative values in the course of each record. The maximal respiration rate for the control mitochondria could be attained by the addition of 40 μM 2,4-dinitrophenol (DNP). Incubation medium: RLM (0.5 mg protein/ml), MOPS 20 mM (pH 7.2), sucrose 250 mM, EGTA 1 mM, MgCl2 2 mM, KH2PO4 2 mM, succinate 5 mM, and rotenone 2 μM. The effect of SF-C5-TPP (filled circles), SF-C4-TPP (open triangles), SF-C10-TPP (gray squares), C10-TPP (open circles), C4-TPP (filled triangles), SF6847 (open diamonds), and CCCP (filled diamonds) on RLM respiration (b). Points and error bars correspond to mean ± SD obtained from at least 3 independent experiments.
SF-Cn-TPP Decreased Membrane Potential of the Isolated Rat Liver Mitochondria
In line with the data on mitochondrial respiration, SF-C5-TPP also exhibited the highest efficacy of the SF-Cn-TPP series in decreasing the RLM membrane potential, as measured by changes in safranine O absorbance (Figure 2). Thus, the compounds of the SF-Cn-TPP series proved to be mitochondrial uncouplers, with n = 5 being the optimal alkyl chain length for the uncoupling activity. This kind of chain length dependence (with an optimum) is in line with previous results on the protonophores derived from fluorescein,31 7-nitrobenz-2-oxa-1,3-diazole,32 rhodamine 19,33p-chlorophenol,34,35 and 7-hydroxycoumarin,36,37 and also corresponds to other studies of chain length impact on membrane permeability and biological activity of small molecules.38−40
Figure 2.
Effect of SF-C5-TPP (a, black line), SF-C4-TPP (a, red line), SF-C10-TPP (a, blue line), CCCP (b, black line), C4-TPP (b, red line), and C10-TPP (b, blue line) on membrane potential in isolated RLM, as measured by changes in safranine O (15 μM) absorbance. (a) Three successive additions of the compounds (0.25 μM each) and subsequent additions (0.50 μM each) are marked by arrows. The gray line shows the effect of sequential additions of SF6847, 5 nM each. (b). Sequential additions of the compounds, 2 μM each, are marked by arrows. The black line shows the effect of sequential additions of CCCP, 50 nM each. For experimental conditions, see the Experimental Section.
SF-C4-TPP Accumulated in the Isolated Rat Liver Mitochondria Upon Energization
Figure 3 depicts the effect of respiring mitochondria on the potential of a tetraphenylphosphonium (TPP+)-selective electrode observed in the presence of SF-C4-TPP or TPP+. The addition of RLM to the experimental cuvette brought about the uptake of SF-C4-TPP similar to that of TPP+ due to membrane potential generation resulting from the oxidation of endogenous substrates, which was then suppressed by the addition of the respiratory complex I inhibitor rotenone leading to SF-C4-TPP release. The addition of succinate as the respiratory substrate again caused the uptake of SF-C4-TPP by the mitochondria, which was followed by its release after the addition of the conventional uncoupler DNP, similar to previous observations with other zwitterionic uncouplers.16,41,42 With SF-C5-TPP, it was hardly possible to observe its energy-dependent accumulation in mitochondria because of its high uncoupling activity, leading to membrane potential collapse.
Figure 3.
Effect of respiring RLM on the voltage of a tetraphenylphosphonium-selective electrode in the presence of SF-C4-TPP (red curve) or TPP+ (black curve). The solution was MOPS 20 mM, KH2PO4 10 mM, sucrose 250 mM, MgCl2 1 mM, EGTA 0.2 mM, and bovine serum albumin 0.5 mg/mL, pH 7.4. Additions: rotenone 2 μM, succinate 2 mM, DNP 40 μM, antimycin A 1 μM. T = 23 °C. Concentration of mitochondrial protein, 1 mg/mL.
Direct Measurements of Proton Transport Induction by SF-Cn-TPP on the Planar Bilayer Lipid Membrane (BLM)
To monitor the SF-C5-TPP-induced transport of protons, we detected the electrical current across a BLM formed from soybean phosphatidylcholine (asolectin) under voltage-clamp conditions. The addition of 5 μM phloretin, known to cause a reduction of the membrane dipole potential, thus hindering anion translocation but promoting cation translocation,43−45 led to a marked increase in the SF-C5-TPP-mediated current (Figure 4a, red track) across the asolectin membrane. Therefore, translocation of the protonated cationic form of SF-C5-TPP across the membrane contributes significantly to the total current, similar to the case of other zwitterionic uncouplers.39 By contrast, phloretin caused a sharp drop in the current mediated by the classical anionic protonophore SF6847 (Figure 4a, black track).
Figure 4.
SF-C5-TPP induces an electrical current through BLM that is selective for protons. (a) Typical traces of the induction of the BLM current by SF-C5-TPP (0.5 μM, red curve) or SF6847 (0.1 μM, black curve) and the effect of phloretin (5 μM). BLM was made from a mixture of lipids from soybeans (asolectin). The solution contained Tris (10 mM), MES (10 mM), β-alanine (10 mM), and KCl (100 mM), pH 6.8. BLM voltage: 50 mV. (b) The current-voltage (I–V) curve for 0.5 μM SF-C5-TPP under symmetrical (closed black circles, pH cis and trans were 6.8) and asymmetrical (open red circles, pH cis was 6.8, pH trans was 7.8) conditions. Straight lines were drawn to guide the eye.
To evaluate the ion selectivity of SF-C5-TPP-mediated
current, I–V characteristics
were determined under asymmetrical conditions of different pH values
at the opposite sides of the BLM (pHcis = 7.8, pHtrans = 6.8, Figure 4b,
red open circles). The value of a zero-current voltage (Vzero) under these asymmetrical conditions was −51.1
mV, having a “minus” sign at the side with lower pH.
The Vzero value showed high proton selectivity
of the conductance (
at ΔpH = 1.0), as follows from the
Nernst equation. Based on three repeats, the average Vzero value was −43.4 ± 7.7 mV.
Depolarization of Mitochondria in Cultured Cells Incubated with SF-Cn-TPP
In further experiments, we measured the uncoupling efficacy of SF-Cn-TPP in the cell culture. Figure 5a displays concentration dependences of changes in the fluorescence intensity of the potential-sensitive dye TMRM, reflecting mitochondrial membrane depolarization in MRC5-SV40 cells incubated with SF-Cn-TPP for 60 min. It is seen that SF-C5-TPP provoked a reduction of mitochondrial membrane potential in cells even at a concentration of 0.4 μM. The comparison of the data in Figures 5a and 2 reveals a sharp difference in the relative efficacy of SF-C5-TPP and SF6847 in the cell culture with respect to isolated mitochondria, namely, in the cells, SF-C5-TPP was more effective in mitochondrial depolarization than SF6847, whereas in isolated RLM, the case was vice versa, which could be associated with alleviation of hampered SF6847 penetration into cells by attaching an alkyl-TPP. The enhanced depolarizing effect of the TPP-conjugated derivatives of SF6847 on cellular mitochondria was also observed with JC-1, another potentially sensitive fluorescent dye (Figure 5b).
Figure 5.
Effect of SF-C5-TPP, SF-C4-TPP, and SF6847 on mitochondrial membrane potential of MRC5-SV40 cells. (a) Changes in mitochondrial membrane potential after 60 min incubation of MRC5-SV40 cells with increasing concentrations of compounds, as assessed with the TMRM dye. (b) Changes in mitochondrial membrane potential after 60 min of incubation of MRC5-SV40 cells with increasing concentrations of the compounds, as assessed with the JC-1 dye. Results are shown as the mean of a minimum of three independent measurements with standard deviation (SD).
Influence of SF-Cn-TPP on Cell Viability
Resazurin-based measurements of cell survival in the presence of SF-Cn-TPP (Figure 6) showed rather low toxicity of these compounds toward MRC5-SV40 cells at submicromolar concentrations that caused a marked decrease in mitochondrial membrane potential (Figure 5a). This result is consistent with recent findings indicating that cell viability could not be correlated with mitochondrial bioenergetics.46
Figure 6.

Effect of 24 h incubation with increasing concentrations of the compounds on the survival of MRC5-SV40 cells. The mean values of a minimum of three independent measurements with the standard deviation (SD) are shown.
Antibacterial Activity of SF-Cn-TPP
Like other zwitterionic uncouplers,35,39,47 SF-Cn-TPP (n = 4, 5, 10) exhibited antibacterial activity with Gram-positive species. In particular, the growth of Bacillus subtilis was inhibited at micromolar concentrations of these compounds (Table 1).
Table 1. Minimal Inhibitory Concentrations (MICs) for B. subtilis (μM).
| substance | MIC |
|---|---|
| SF6847 | 0.2 |
| SF-C4-TPP | 2 |
| SF-C5-TPP | 2 |
| SF-C10-TPP | 2 |
| CCCP | 5 |
It should be noted that lipophilic TPP substituents, such as alkyl-TPP, ensure the targeting of compounds to respiring mitochondria due to their inside negative membrane potential.9 Alkyl-TPP cations per se are not protonophoric uncouplers because they cannot bind and dissociate protons; therefore, they lack the ability to carry protons selectively by themselves across membranes. By contrast, triphenylphosphonium ylides possess high protonophoric uncoupling activity due to the acidity of a methylene group linking phosphorus to an ester group in these compounds.39 On the other hand, several research groups use the term uncoupler for lipophilic TPP cations,48−50 because they are able to effectively depolarize mitochondria in cells. Obviously, the mechanism of their action is not the selective transport of protons through lipid membranes but can be the induction of nonspecific membrane permeability similar to the detergent action on membranes in general51−53 and mitochondria in particular.54−58 Another possibility is connected to the presence of fatty acids in the majority of cellular membranes. According to several studies58,59, lipophilic TPP cations can transport protons due to the formation of complexes with fatty acids.
SF6847, which was first reported as the most powerful fungicidal and acaricidal substance of the newly synthesized 3,5-dialkyl-4-hydroxybenzylidene-malononitrile derivatives,19 soon appeared to be the strongest mitochondrial uncoupler.18 Later, SF6847 also turned out to be one of the tyrphostins, a group of compounds that are able to inhibit protein tyrosine kinases, playing a key role in normal cell division and abnormal cell proliferation.60,61 Tyrphostins were proposed as useful agents for the treatment of cancer by inhibiting tumor cell proliferation.62,63 On the other hand, a series of mitochondria-targeted compounds, including TPP derivatives, exhibited pronounced anticancer activity and were coined mitocans.64,65 The presence of lipophilic cationic groups in the structure of mitocans ensured their selective accumulation in tumor cells due to the increased mitochondrial membrane potential compared to normal tissues.66 Combining the advantages of tyrphostins, mitocans, and protonophoric uncouplers, the TPP-linked derivatives of SF6847 may offer promising results both in treating malignant diseases and cancer chemoprevention.67
Experimental Section
Chemicals
Most chemicals, including diphytanoylphosphatidylcholine (DPhPC), were purchased from Sigma-Aldrich, St. Louis, MO.
NMR and LC-MS Measurements
1H, 13C, and 31P spectra were recorded on Bruker Avance III HD 500 operating at 500, 125, and 202 MHz, respectively. HC-HSQC and HC-HMBC were conducted on Bruker Avance III HD 500. LC-MS measurements were conducted by using an Acquity UPLC System (Waters, Milford, MA) and TQD (Waters, Milford, MA).
General Procedure for the Synthesis of Phosphonium Salts SF-Cn-TPP
A mixture of 1.3 equiv of triphenylphosphine and 1 equiv of SF-CnBr (Supplement) in acetonitrile was heated at 85 °C for 3–4 days. Then, the solution was evaporated in vacuo, and the residue was dissolved in a minimal amount of chloroform. A 3-fold excess of hexane was added, and the dark oily precipitate was isolated by centrifugation. The dissolution–reprecipitation procedure was repeated 3 more times. Finally, the target phosphonium salt was dried in vacuo.
(6,6-Dicyano-5-(4′-hydroxy-3′,5′-ditert-butylphenyl)hex-5-en-1-yl)triphenylphosphonium bromide (SF-C4-TPP): yellow powder, 77% yield. 1H NMR (500 MHz, CDCl3): δ 7.87–7.91 (m, 6H, Ph), 7.80–7.84 (m, 3H, Ph), 7.70–7.74 (m, 6H, Ph), 7.33 (s, 2H, CH), 5.81 (s, 1H, OH), 3.94–4.00 (m, 2H, CH2P), 3.05 (t, 3JHH = 7.2 Hz, 2H, CH2C(=C(CN)2)), 1.96–2.02 (m, 2H, CH2), 1.66–1.74 (m, 2H, CH2), 1.45 (s, 18H, C(CH3)3). 13C{1H} NMR (125 MHz, CDCl3): δ 179.66, 158.34, 136.67, 135.12 (4JCP = 2.7), 133.90 (2JCP = 10.0), 130.54 (3JCP = 12.71), 125.70, 125.00, 118.13 (1JCP = 86.3), 114.12, 113.69, 80.81, 35.66, 34.66 (C(CH3)3), 30.08 (C(CH3)3), 29.7 (2JCP = 16.6), 22.26 (1JCP = 50.6), 21.76 (3JCP = 3.7). 31P{1H} NMR (202 MHz, CDCl3): δ 24.75. LC-MS (ESI+):m/z calcd for C40H44N2OP ([M]+), 599.32, found 599.23.
(7,7-Dicyano-6-(4′-hydroxy-3′,5′-ditert-butylphenyl)hept-6-en-1-yl)triphenylphosphonium bromide (SF-C5-TPP): yellow powder, 82% yield. 1H NMR (500 MHz, CDCl3): δ 7.85–7.89 (m, 6H, Ph), 7.79–7.82 (m, 3H, Ph), 7.69–7.73 (m, 6H, Ph), 7.38 (s, 2H, CH), 5.79 (s, 1H, OH), 3.89–3.95 (m, 2H, CH2P), 2.88 (t, 3JHH = 7.9 Hz, 2H, CH2C(=C(CN)2)), 1.82–1.86 (m, 2H, CH2), 1.67–1.75 (m, 2H, CH2), 1.51–1.59 (m, 2H, CH2), 1.46 (s, 18H, C(CH3)3). 13C{1H} NMR (125 MHz, CDCl3): δ 180.26, 158.19, 136.60, 135.01 (4JCP = 3.6), 133.77 (2JCP = 10.0), 130.52 (3JCP = 12.7), 125.66, 125.46, 118.30 (1JCP = 85.4), 114.10, 113.91, 80.66, 36.23, 34.65 (C(CH3)3), 30.07 (C(CH3)3), 29.63 (2JCP = 16.4), 28.66, 22.71 (1JCP = 50.5), 22.01 (3JCP = 4.0). 31P{1H} NMR (202 MHz, CDCl3): δ 24.46. LC-MS (ESI+):m/z calcd for C41H46N2OP ([M]+), 613.33, found 613.08.
(12,12-Dicyano-11-(4′-hydroxy-3′,5′-ditert-butylphenyl)dodec-11-en-1-yl)triphenylphosphonium bromide (SF-C10-TPP): dark-yellow powder, 94% yield. 1H NMR (500 MHz, CDCl3): δ 7.79–7.87 (m, 9H, Ph), 7.70–7.74 (m, 6H, Ph), 7.43 (s, 2H, CH), 5.80 (s, 1H, OH), 3.74–3.81 (m, 2H, CH2P), 2.91 (t, 3JHH = 7.8 Hz, 2H, CH2C(=C(CN)2)), 1.63 (bs 4H, 2CH2), 1.44–1.53 (m, 2H, CH2), 1.47 (s, 18H, C(CH3)3), 1.20–1.35 (m, 10H, 5CH2),. 13C{1H} NMR (125 MHz, CDCl3): δ 180.35, 157.99, 136.56, 134.99 (4JCP = 3.6), 133.72 (2JCP = 10.0), 130.50 (3JCP = 12.6), 125.72, 125.60, 118.44 (1JCP = 85.4), 114.28, 113.78, 80.66, 36.92, 34.67 (C(CH3)3), 30.37 (2JCP = 15.8), 30.10 (C(CH3)3), 29.21, 29.17, 29.13, 29.11, 29.05, 28.84, 22.78 (1JCP = 49.3), 22.68 (3JCP = 4.4). 31P{1H} NMR (202 MHz, CDCl3): δ 24.49. LC-MS (ESI+):m/z calcd for C46H56N2OP ([M]+), 683.41, found 682.95.
Measurements of Absorbance Spectra
Absorption spectra of SF-C5-TPP were recorded with a SPECORD 50 ultraviolet–visible (UV–vis) spectrophotometer (Analytik Jena, Jena, Germany).
Isolation of Mitochondria from Rat Liver
Rat liver mitochondria (RLM) were isolated by differential centrifugation68 in the medium containing 250 mM sucrose, 5 mM MOPS, 1 mM EGTA, and pH 7.4. The final washing was carried out in the medium, additionally supplemented with bovine serum albumin (0.1 mg/mL). Protein concentration was measured by using the Biuret method. Handling of animals and experimental procedures were performed according to the international guidelines for animal care and use and were adopted by the Ethics Committee of Belozersky Institute of Physico-Chemical Biology at the Lomonosov Moscow State University (protocol #3 on February 15, 2019).
Mitochondrial Respiration
To measure oxygen consumption (respiration) by RLM, we applied a standard polarographic technique based on a Clark-type oxygen electrode (Strathkelvin Instruments, U.K.) at 25 °C using 782 system software. The incubation medium contained sucrose (250 mM), MOPS (5 mM), and EGTA (1 mM) at pH 7.4. The mitochondrial protein content was 0.8 mg/mL. Oxygen consumption rates were measured in nmol/min·mg protein.
Membrane Potential (Δψ) Measurements in Isolated Mitochondria
Δψ was estimated using the safranine O dye.69 The difference in the absorbance at 555 and 523 nm (ΔA) was evaluated by using an Aminco DW-2000 spectrophotometer in dual-wavelength mode. The incubation medium contained sucrose 250 mM, MOPS 5 mM, KH2PO4 0.5 mM, EGTA 1 mM, rotenone 2 μM, succinate 5 mM, oligomycin 1 μg/mL, safranine O 15 μM, and pH 7.4. The mitochondrial protein concentration was about 0.6–0.9 mg of protein/ml, and the temperature was 25 °C.
Monitoring the Uptake of SF-C4-TPP by Energized Rat Liver Mitochondria with a TPP+-Selective Electrode
Accumulation of SF-C4-TPP in RLM as compared to that of tetraphenylphosphonium (TPP+) cations was monitored using a TPP+-selective electrode (NIKO-ANALIT, Moscow, Russia) immersed in a 1 mL chamber with a constantly stirred buffer solution.
Electrical Current Measurements on Planar Bilayers
A bilayer lipid membrane (BLM) was formed by the brush technique70 from a 2% decane solution of asolectin on a 0.8 mm aperture in a Teflon septum, separating the experimental chamber into two 3 mL compartments. The electrical current was measured with two Ag/AgCl electrodes placed into the solutions on the two sides of the BLM via agar bridges (ground electrode in the cis compartment) using a Keithley 428 amplifier (Cleveland, Ohio).
Flow Cytometry
Mitochondrial membrane potential changes in MRC5-SV40 cells were assessed by the fluorescence of tetramethylrhodamine methyl ester (TMRM) (Invitrogen) or JC-1 (Lumiprobe),71 using an Amnis FlowSight Imaging Flow Cytometer (Luminex Corporation, Seattle, WA). With 100 nM TMRM, the cells were stained for 15 min. Then, the cells were incubated with SF-Cn-TPP or SF6847 at various concentrations for 1 h. After this treatment, the cells were detached with trypsin/versene and washed from the medium with phosphate-buffered saline. The median fluorescence of the cells was analyzed with a flow cytometer (excitation 548 nm, emission 574 nm). With 1 μg/mL JC-1, the cells were incubated at 37 °C for 30 min, then washed from the medium, and then analyzed with the flow cytometer. Membrane potential was assessed by the ratio of the fluorescence in the red channel (exc. 514 nm, em. 590 nm), emitted by the so-called “J-aggregates” in highly energized mitochondria, and the fluorescence in the green channel (exc. 514 nm, em. 529 nm), emitted by JC-1 in de-energized mitochondria.
Cell Viability Study
Induced cell death in MRC5-SV40 cells was measured by fluorometric analysis with the CellTiterBlue reagent (Promega) in accordance with the manufacturer’s protocol. Each measurement was repeated a minimum of 3 times. The data are shown as the mean of at least 3 independent replicates with standard deviation (SD).
Growth of Bacteria
The standard laboratory strain of B. subtilis subs. subtilis Cohn 1872, specifically strain PY79, was utilized in this study. Bacterial cells from overnight cultures were diluted in a fresh Luria–Bertani broth. Subsequently, 200 μL of bacterial cell cultures containing 5 × 105 cells/mL were dispensed into 96-well plates (Eppendorf AG, Hamburg, Germany). The tested compounds were introduced into the wells to attain the desired concentrations. The cells were allowed to grow for 21 h at 30 °C. Optical densities at 620 nm were recorded by using a Thermo Scientific Multiskan FC plate reader equipped with an incubator (Thermo Fisher Scientific). All experiments were conducted with a minimum of three replicates.
Acknowledgments
This work was financially supported by the Russian Science Foundation (project No. 21-14-00062).
Glossary
Abbreviations
- SF-C5-TPP
7,7-dicyano-6-(4′-hydroxy-3′,5′-ditert-butylphenyl)hept-6-en-1-yltriphenylphosphonium bromide
- SF-C4-TPP
6,6-dicyano-5-((4′-hydroxy-3′,5′-ditert-butylphenyl)hex-5-en-1-yl)triphenylphosphonium bromide
- SF-C10-TPP
12,12-dicyano-11-((4′-hydroxy-3′,5′-ditert-butylphenyl)dodec-11-en-1-yl)triphenylphosphonium bromide
- SF6847
3,5-ditert-butyl-4-hydroxybenzylidene-malononitrile
- DNP
2,4-dinitrophenol
- CCCP
carbonyl cyanide m-chlorophenyl hydrazone
- DPhPC
diphytanoylphosphatidylcholine
- RLM
rat liver mitochondria
- BLM
bilayer lipid membrane
- ΔΨ
mitochondrial membrane potential
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.3c08621.
Includes the synthesis procedures, the NMR spectra, the MS spectra, and the pKa estimation data (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- Cutting W. C.; Tainter M. L. Actions of dinitrophenol. Exp. Biol. Med. 1932, 29, 1268–1269. 10.3181/00379727-29-6315. [DOI] [Google Scholar]
- Loomis W. F.; Lipmann F. Reversible inhibition of the coupling between phosphorylation and oxidation. J. Biol. Chem. 1948, 173, 807–808. 10.1016/S0021-9258(18)57455-X. [DOI] [PubMed] [Google Scholar]
- Tainter M. L.; Stockton A. B.; Cutting W. C. Dinitrophenol in the treatment of obesity: final report. J. Am. Med. Assoc. 1935, 105, 332–337. 10.1001/jama.1935.02760310006002. [DOI] [Google Scholar]
- Goedeke L.; Shulman G. I. Therapeutic potential of mitochondrial uncouplers for the treatment of metabolic associated fatty liver disease and NASH. Mol. Metab. 2021, 46, 101178 10.1016/j.molmet.2021.101178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shrestha R.; Johnson E.; Byrne F. L. Exploring the therapeutic potential of mitochondrial uncouplers in cancer. Mol. Metab. 2021, 51, 101222 10.1016/j.molmet.2021.101222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zorov D. B.; Andrianova N. V.; Babenko V. A.; Pevzner I. B.; Popkov V. A.; Zorov S. D.; Zorova L. D.; Plotnikov E. Y.; Sukhikh G. T.; Silachev D. N. Neuroprotective potential of mild uncoupling in mitochondria. Pros and Cons. Brain Sci. 2021, 11, 1050 10.3390/brainsci11081050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kotova E. A.; Antonenko Y. N. Fifty years of research on protonophores: mitochondrial uncoupling as a basis for therapeutic action. Acta Nat. 2022, 14, 4–13. 10.32607/actanaturae.11610. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grundlingh J.; Dargan P. I.; El-Zanfaly M.; Wood D. M. 2,4-dinitrophenol (DNP): a weight loss agent with significant acute toxicity and risk of death. J. Med. Toxicol. 2011, 7, 205–212. 10.1007/s13181-011-0162-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murphy M. P.; Smith R. A. Targeting antioxidants to mitochondria by conjugation to lipophilic cations. Annu. Rev. Pharmacol. Toxicol. 2007, 47, 629–656. 10.1146/annurev.pharmtox.47.120505.105110. [DOI] [PubMed] [Google Scholar]
- Wen R.; Banik B.; Pathak R. K.; Kumar A.; Kolishetti N.; Dhar S. Nanotechnology inspired tools for mitochondrial dysfunction related diseases. Adv. Drug Delivery Rev. 2016, 99, 52–69. 10.1016/j.addr.2015.12.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zielonka J.; Joseph J.; Sikora A.; Hardy M.; Ouari O.; Vasquez-Vivar J.; Cheng G.; Lopez M.; Kalyanaraman B. Mitochondria-targeted triphenylphosphonium-based compounds: syntheses, mechanisms of action, and therapeutic and diagnostic applications. Chem. Rev. 2017, 117, 10043–10120. 10.1021/acs.chemrev.7b00042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Demine S.; Renard P.; Arnould T. Mitochondrial Uncoupling: A Key Controller of Biological Processes in Physiology and Diseases. Cells 2019, 8, 795 10.3390/cells8080795. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang X.; Sarkar S.; Ashokan A.; Surnar B.; Kolishetti N.; Dhar S. All-in-one Pyruvate Dehydrogenase Kinase Inhibitor for Tracking, Targeting, and Enhanced Efficacy. Bioconjugate Chem. 2023, 34, 1122–1129. 10.1021/acs.bioconjchem.3c00152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blaikie F. H.; Brown S. E.; Samuelsson L. M.; Brand M. D.; Smith R. A.; Murphy M. P. Targeting dinitrophenol to mitochondria: limitations to the development of a self-limiting mitochondrial protonophore. Biosci. Rep. 2006, 26, 231–243. 10.1007/s10540-006-9018-8. [DOI] [PubMed] [Google Scholar]
- Iaubasarova I. R.; Khailova L. S.; Firsov A. M.; Grivennikova V. G.; Kirsanov R. S.; Korshunova G. A.; Kotova E. A.; Antonenko Y. N. The mitochondria-targeted derivative of the classical uncoupler of oxidative phosphorylation carbonyl cyanide m-chlorophenylhydrazone is an effective mitochondrial recoupler. PLoS One 2020, 15, e0244499 10.1371/journal.pone.0244499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Denisov S. S.; Kotova E. A.; Plotnikov E. Y.; Tikhonov A. A.; Zorov D. B.; Korshunova G. A.; Antonenko Y. N. A mitochondria-targeted protonophoric uncoupler derived from fluorescein. Chem. Commun. 2014, 50, 15366–15369. 10.1039/C4CC04996A. [DOI] [PubMed] [Google Scholar]
- Antonenko Y. N.; Denisov S. S.; Silachev D. N.; Khailova L. S.; Jankauskas S. S.; Rokitskaya T. I.; Danilina T. I.; Kotova E. A.; Korshunova G. A.; Plotnikov E. Y.; Zorov D. B. A long-linker conjugate of fluorescein and triphenylphosphonium as mitochondria-targeted uncoupler and fluorescent neuro- and nephroprotector. Biochim. Biophys. Acta, Gen. Subj. 2016, 1860, 2463–2473. 10.1016/j.bbagen.2016.07.014. [DOI] [PubMed] [Google Scholar]
- Muraoka S.; Terada H. 3,5-Di-tert-butyl-4-hydroxybenzylidenemalononitrile, a new powerful uncoupler of respiratory-chain phosphorylation. Biochim. Biophys. Acta, Bioenerg. 1972, 275, 271–275. 10.1016/0005-2728(72)90047-3. [DOI] [PubMed] [Google Scholar]
- Horiuchi F.; Fujimoto K.; Ozaki T.; Nishizawa Y. 3, 5-Dialkyl-4-hydroxybenzylidenemalononitrile; A New Group of Fungicidal and Acaricidal Agents. Agric. Biol. Chem. 1971, 35, 2003–2007. 10.1271/bbb1961.35.2003. [DOI] [Google Scholar]
- Terada H. Some biochemical and physicochemical properties of the potent uncoupler SF 6847 (3,5-di-tert-butyl-4-hydroxybenzylidenemalononitrile). Biochim. Biophys. Acta, Bioenerg. 1975, 387, 519–532. 10.1016/0005-2728(75)90090-0. [DOI] [PubMed] [Google Scholar]
- Terada H.; Kumazawa N.; Ju-Ichi M.; Yoshikawa K. Molecular basis of the protonophoric and uncoupling activities of the potent uncoupler SF-6847 ((3,5-di-tert-butyl-4-hydroxybenzylidene)malononitrile) and derivatives. Regulation of their electronic structures by restricted intramolecular rotation. Biochim. Biophys. Acta, Bioenerg. 1984, 767, 192–199. 10.1016/0005-2728(84)90187-7. [DOI] [PubMed] [Google Scholar]
- Miyoshi H.; Nishioka T.; Fujita T. Quantitative relationship between protonophoric and uncoupling activities of analogs of SF6847 (2,6-di-t-butyl-4-(2′,2′-dicyanovinyl)phenol). Biochim. Biophys. Acta, Bioenerg. 1987, 891, 293–299. 10.1016/0005-2728(87)90224-6. [DOI] [PubMed] [Google Scholar]
- Miyoshi H.; Guo Z.; Fujita T. R.; Iwamura H. R. Acidity and Dynamic Structure of the Potent Uncoupler SF6847 (2,6-Di-t-butyl-4-(2,2-dicyanovinyl)phenol) and Its Derivatives. Bull. Chem. Soc. Jpn. 1993, 66, 269–274. 10.1246/bcsj.66.269. [DOI] [Google Scholar]
- Uno S.; Kimura H.; Murai M.; Miyoshi H. Exploring the quinone/inhibitor-binding pocket in mitochondrial respiratory complex I by chemical biology approaches. J. Biol. Chem. 2019, 294, 679–696. 10.1074/jbc.RA118.006056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Desai J.; Misra A. N.; Nair K. B. Synthesis and Characterization of Homologue of Irganox 1076 - Some Novel Observations. Synth. Commun. 2003, 33, 199–205. 10.1081/SCC-120015701. [DOI] [Google Scholar]
- Shcherbakova I. V.; Ukhin L. Yu.; Komissarov V. N.; Kuznetsov E. V.; Polyakov A. V.; Yanovskii A. I.; Struchkov Yu. T. 2-Benzopyrilium salts. 30. Synthesis and properties of 2-benzopyrilium salts with the 2,6-di-t-butylphenyl substituent in position 3. Chem. Heterocycl. Compd. 1987, 23, 824–829. 10.1007/BF00473448. [DOI] [Google Scholar]
- Hansen B. S.; Tullin S.; Hansen T. K.; Colding-Joergensen M.. Safe Chemical Uncouplers for the Treatment of Obesity. WO Patent WO2004041256A2, 2004.
- Hori H.; Fujii T.; Kubo A.; Pan N.; Sako S.; Tada C.; Matsubara T. KIH-201: A New Enhancer of Chemiluminescence in the Luminol-Hydrogen Peroxide-Peroxidase System Characterized by A High-Performance Luminometer. Anal. Lett. 1994, 27, 1109–1122. 10.1080/00032719408000282. [DOI] [Google Scholar]
- Barnes D. M.; Haight A. R.; Hameury T.; McLaughlin M. A.; Mei J.; Tedrow J. S.; Toma J. D. R. New conditions for the synthesis of thiophenes via the Knoevenagel/Gewald reaction sequence. Application to the synthesis of a multitargeted kinase inhibitor. Tetrahedron 2006, 62, 11311–11319. 10.1016/j.tet.2006.07.008. [DOI] [Google Scholar]
- Terada H. The interaction of highly active uncouplers with mitochondria. Biochim. Biophys. Acta, Rev. Bioenerg. 1981, 639, 225–242. 10.1016/0304-4173(81)90011-2. [DOI] [PubMed] [Google Scholar]
- Shchepinova M. M.; Denisov S. S.; Kotova E. A.; Khailova L. S.; Knorre D. A.; Korshunova G. A.; Tashlitsky V. N.; Severin F. F.; Antonenko Y. N. Dodecyl and octyl esters of fluorescein as protonophores and uncouplers of oxidative phosphorylation in mitochondria at submicromolar concentrations. Biochim. Biophys. Acta, Bioenerg. 2014, 1837, 149–158. 10.1016/j.bbabio.2013.09.011. [DOI] [PubMed] [Google Scholar]
- Denisov S. S.; Kotova E. A.; Khailova L. S.; Korshunova G. A.; Antonenko Y. N. Tuning the hydrophobicity overcomes unfavorable deprotonation making octylamino-substituted 7-nitrobenz-2-oxa-1,3-diazole (n-octylamino-NBD) a protonophore and uncoupler of oxidative phosphorylation in mitochondria. Bioelectrochemistry 2014, 98, 30–38. 10.1016/j.bioelechem.2014.02.002. [DOI] [PubMed] [Google Scholar]
- Khailova L. S.; Silachev D. N.; Rokitskaya T. I.; Avetisyan A. V.; Lyamzaev K. G.; Severina I. I.; Il’yasova T. M.; Gulyaev M. V.; Dedukhova V. I.; Trendeleva T. A.; Plotnikov E. Y.; Zvyagilskaya R. A.; Chernyak B. V.; Zorov D. B.; Antonenko Y. N.; Skulachev V. P. A short-chain alkyl derivative of Rhodamine 19 acts as a mild uncoupler of mitochondria and a neuroprotector. Biochim. Biophys. Acta, Bioenerg. 2014, 1837, 1739–1747. 10.1016/j.bbabio.2014.07.006. [DOI] [PubMed] [Google Scholar]
- Rokitskaya T. I.; Terekhova N. V.; Khailova L. S.; Kotova E. A.; Plotnikov E. Y.; Zorov D. B.; Tatarinov D. A.; Antonenko Y. N. Zwitterionic protonophore derived from 2-(2-hydroxyaryl)alkenylphosphonium as an uncoupler of oxidative phosphorylation. Bioconjugate Chem. 2019, 30, 2435–2443. 10.1021/acs.bioconjchem.9b00516. [DOI] [PubMed] [Google Scholar]
- Terekhova N. V.; Khailova L. S.; Rokitskaya T. I.; Nazarov P. A.; Islamov D. R.; Usachev K. S.; Tatarinov D. A.; Mironov V. F.; Kotova E. A.; Antonenko Y. N. Trialkyl(vinyl)phosphonium Chlorophenol Derivatives as Potent Mitochondrial Uncouplers and Antibacterial Agents. ACS Omega 2021, 6, 20676–20685. 10.1021/acsomega.1c02909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krasnov V. S.; Kirsanov R. S.; Khailova L. S.; Firsov A. M.; Nazarov P. A.; Tashlitsky V. N.; Korshunova G. A.; Kotova E. A.; Antonenko Y. N. Alkyl esters of umbelliferone-4-acetic acid as protonophores in bilayer lipid membranes and ALDH2-dependent soft uncouplers in rat liver mitochondria. Bioelectrochemistry 2022, 145, 108081 10.1016/j.bioelechem.2022.108081. [DOI] [PubMed] [Google Scholar]
- Krasnov V. S.; Kirsanov R. S.; Khailova L. S.; Popova L. B.; Lyamzaev K. G.; Firsov A. M.; Korshunova G. A.; Kotova E. A.; Antonenko Y. N. Alkyl esters of 7-hydroxycoumarin-3-carboxylic acid as potent tissue-specific uncouplers of oxidative phosphorylation: Involvement of ATP/ADP translocase in mitochondrial uncoupling. Arch. Biochem. Biophys. 2022, 728, 109366 10.1016/j.abb.2022.109366. [DOI] [PubMed] [Google Scholar]
- Morstein J.; Capecchi A.; Hinnah K.; Park B.; Petit-Jacques J.; Van Lehn R. C.; Reymond J. L.; Trauner D. Medium-chain lipid conjugation facilitates cell-permeability and bioactivity. J. Am. Chem. Soc. 2022, 144, 18532–18544. 10.1021/jacs.2c07833. [DOI] [PubMed] [Google Scholar]
- Kirsanov R. S.; Khailova L. S.; Rokitskaya T. I.; Iaubasarova I. R.; Nazarov P. A.; Panteleeva A. A.; Lyamzaev K. G.; Popova L. B.; Korshunova G. A.; Kotova E. A.; Antonenko Y. N. Ester-stabilized phosphorus ylides as protonophores on bilayer lipid membranes, mitochondria and chloroplasts. Bioelectrochemistry 2023, 150, 108369 10.1016/j.bioelechem.2023.108369. [DOI] [PubMed] [Google Scholar]
- Pashirova T. N.; Bogdanov A. V.; Zaripova I. F.; Burilova E. A.; Vandyukov A. E.; Sapunova A. S.; Vandyukova I. I.; Voloshina A. D.; Mironov V. F.; Zakharova L. Y. Tunable amphiphilic p-systems based on isatin derivatives containing a quaternary ammonium moiety: the role of alkyl chain length in biological activity. J. Mol. Liq. 2019, 290, 111220 10.1016/j.molliq.2019.111220. [DOI] [Google Scholar]
- Biasutto L.; Sassi N.; Mattarei A.; Marotta E.; Cattelan P.; Toninello A.; Garbisa S.; Zoratti M.; Paradisi C. Impact of mitochondriotropic quercetin derivatives on mitochondria. Biochim. Biophys. Acta, Bioenerg. 2010, 1797, 189–196. 10.1016/j.bbabio.2009.10.001. [DOI] [PubMed] [Google Scholar]
- Iaubasarova I. R.; Khailova L. S.; Nazarov P. A.; Rokitskaya T. I.; Silachev D. N.; Danilina T. I.; Plotnikov E. Y.; Denisov S. S.; Kirsanov R. S.; Korshunova G. A.; Kotova E. A.; Zorov D. B.; Antonenko Y. N. Linking 7-Nitrobenzo-2-oxa-1,3-diazole (NBD) to Triphenylphosphonium Yields Mitochondria-Targeted Protonophore and Antibacterial Agent. Biochemistry (Moscow) 2020, 85, 1578–1590. 10.1134/S000629792012010X. [DOI] [PubMed] [Google Scholar]
- Andersen O. S.; Finkelstein A.; Katz I.; Cass A. Effect of phloretin on the permeability of thin lipid membranes. J. Gen. Physiol. 1976, 67, 749–771. 10.1085/jgp.67.6.749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Franklin J. C.; Cafiso D. S. Internal electrostatic potentials in bilayers - measuring and controlling dipole potentials in lipid vesicles. Biophys. J. 1993, 65, 289–299. 10.1016/S0006-3495(93)81051-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pohl P.; Rokitskaya T. I.; Pohl E. E.; Saparov S. M. Permeation of phloretin across bilayer lipid membranes monitored by dipole potential and microelectrode measurements. Biochim. Biophys. Acta, Biomembr. 1997, 1323, 163–172. 10.1016/S0005-2736(96)00185-X. [DOI] [PubMed] [Google Scholar]
- Doczi J.; Karnok N.; Bui D.; Azarov V.; Pallag G.; Nazarian S.; Czumbel B.; Seyfried T. N.; Chinopoulos C. Viability of MCF cells is not correlated with mitochondrial bioenergetics. Sci. Rep. 2023, 13, 10822 10.1038/s41598-023-37677-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nazarov P. A.; Kirsanov R. S.; Denisov S. S.; Khailova L. S.; Karakozova M. V.; Lyamzaev K. G.; Korshunova G. A.; Lukyanov K. A.; Kotova E. A.; Antonenko Y. N. Fluorescein derivatives as antibacterial agents acting via membrane depolarization. Biomolecules 2020, 10, 309 10.3390/biom10020309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Childress E. S.; Alexopoulos S. J.; Hoehn K. L.; Santos W. L. Small molecule mitochondrial uncouplers and their therapeutic potential. J. Med. Chem. 2018, 61, 4641–4655. 10.1021/acs.jmedchem.7b01182. [DOI] [PubMed] [Google Scholar]
- Kulkarni C. A.; Fink B. D.; Gibbs B. E.; Chheda P. R.; Wu M.; Sivitz W. I.; Kerns R. J. A novel triphenylphosphonium carrier to target mitochondria without uncoupling oxidative phosphorylation. J. Med. Chem. 2021, 64, 662–676. 10.1021/acs.jmedchem.0c01671. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zinovkina L. A.; Makievskaya C. I.; Galkin I. I.; Zinovkin R. A. Mitochondria-targeted Uncouplers Decrease Inflammatory Reactions in Endothelial Cells by Enhancing Methylation of the ICAM1 Gene Promoter. Curr. Mol. Pharmacol. 2024, 17, e150823219723 10.2174/1874467217666230815142556. [DOI] [PubMed] [Google Scholar]
- Prasad M.; Moulik S. P.; MacDonald A.; Palepu R. Self-Aggregation of Alkyl (C10-, C12-, C14-, and C16-) Triphenyl Phosphonium Bromides and their 1:1 Molar Mixtures in Aqueous Medium: A Thermodynamic Study. J. Phys. Chem. B 2004, 108, 355–362. 10.1021/jp036358+. [DOI] [Google Scholar]
- Zakharova L. Ya.; Kaupova G. I.; Gabdrakhmanov D. R.; Gaynanova G. A.; Ermakova E. A.; Mukhitov A. R.; Galkina I. V.; Cheresiz S. V.; Pokrovsky A. G.; Skvortsova P. V.; Gogolev Y. V.; Zuev Y. F. Alkyl triphenylphosphonium surfactants as nucleic acid carriers: complexation efficacy toward DNA decamers, interaction with lipid bilayers and cytotoxicity studies. Phys. Chem. Chem. Phys. 2019, 21, 16706–16717. 10.1039/C9CP02384D. [DOI] [PubMed] [Google Scholar]
- Sokolov S.; Zyrina A.; Akimov S.; Knorre D.; Severin F. Toxic Effects of Penetrating Cations. Membranes 2023, 13, 841 10.3390/membranes13100841. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bragadin M.; Dell’Antone P. Mitochondrial bioenergetics as affected by cationic detergents. Arch. Environ. Contam. Toxicol. 1996, 30, 280–284. 10.1007/BF00215809. [DOI] [PubMed] [Google Scholar]
- Khailova L. S.; Nazarov P. A.; Sumbatyan N. V.; Korshunova G. A.; Rokitskaya T. I.; Dedukhova V. I.; Antonenko Y. N.; Skulachev V. P. Uncoupling and Toxic Action of Alkyltriphenylphosphonium Cations on Mitochondria and the Bacterium Bacillus subtilis as a Function of Alkyl Chain Length. Biochemistry 2015, 80, 1589–1597. 10.1134/S000629791512007X. [DOI] [PubMed] [Google Scholar]
- Trnka J.; Elkalaf M.; Andel M. Lipophilic triphenylphosphonium cations inhibit mitochondrial electron transport chain and induce mitochondrial proton leak. PLoS One 2015, 10, e0121837 10.1371/journal.pone.0121837. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kafkova A.; Tilokani L.; Trcka F.; Sramkova V.; Vancova M.; Bily T.; Nebesarova J.; Prudent J.; Trnka J. Selective and reversible disruption of mitochondrial inner membrane protein complexes by lipophilic cations. Mitochondrion 2023, 68, 60–71. 10.1016/j.mito.2022.11.006. [DOI] [PubMed] [Google Scholar]
- Rokitskaya T. I.; Khailova L. S.; Korshunova G. A.; Antonenko Y. N. Efficiency of mitochondrial uncoupling by modified butyltriphenylphosphonium cations and fatty acids correlates with lipophilicity of cations: Protonophoric vs leakage mechanisms. Biochim. Biophys. Acta, Biomembr. 2023, 1865, 184183 10.1016/j.bbamem.2023.184183. [DOI] [PubMed] [Google Scholar]
- Severin F. F.; Severina I. I.; Antonenko Y. N.; Rokitskaya T. I.; Cherepanov D. A.; Mokhova E. N.; Vyssokikh M. Y.; Pustovidko A. V.; Markova O. V.; Yaguzhinsky L. S.; Korshunova G. A.; Sumbatyan N. V.; Skulachev M. V.; Skulachev V. P. Penetrating cation/fatty acid anion pair as a mitochondria-targeted protonophore. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 663–668. 10.1073/pnas.0910216107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gazit A.; Yaish P.; Gilon C.; Levitzki A. Tyrphostins I: synthesis and biological activity of protein tyrosine kinase inhibitors. J. Med. Chem. 1989, 32, 2344–2352. 10.1021/jm00130a020. [DOI] [PubMed] [Google Scholar]
- Levitzki A.; Gilon C. Tyrphostins as molecular tools and potential antiproliferative drugs. Trends Pharmacol. Sci. 1991, 12, 171–174. 10.1016/0165-6147(91)90538-4. [DOI] [PubMed] [Google Scholar]
- Gillespie J.; Dye J. F.; Schachter M.; Guillou P. J. Inhibition of pancreatic cancer cell growth in vitro by the tyrphostin group of tyrosine kinase inhibitors. Br. J. Cancer 1993, 68, 1122–1126. 10.1038/bjc.1993.491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Burger A. M.; Kaur G.; Alley M. C.; Supko J. G.; Malspeis L.; Grever M. R.; Sausville E. A. Tyrphostin AG17, [(3,5-Di-tert-butyl-4-hydroxybenzylidene)-malononitrile], Inhibits Cell Growth by Disrupting Mitochondria. Cancer Res. 1995, 55, 2794–2799. [PubMed] [Google Scholar]
- Biasutto L.; Dong L.-F.; Zoratti M.; Neuzil J. Mitochondrially targeted anti-cancer agents. Mitochondrion 2010, 10, 670–681. 10.1016/j.mito.2010.06.004. [DOI] [PubMed] [Google Scholar]
- Dong L.-F.; Gopalan V.; Holland O.; Neuzil J. Mitocans Revisited: Mitochondrial Targeting as Efficient Anti-Cancer Therapy. Int. J. Mol. Sci. 2020, 21, 7941 10.3390/ijms21217941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nadakavukaren K. K.; Nadakavukaren J. J.; Chen L. B. Increased rhodamine 123 uptake by carcinoma cells. Cancer Res. 1985, 45, 6093–6099. [PubMed] [Google Scholar]
- Huang M.; Myers C. R.; Wang Y.; You M. Mitochondria as a Novel Target for Cancer Chemoprevention: Emergence of Mitochondrial-targeting Agents. Cancer Prev. Res. 2021, 14, 285–306. 10.1158/1940-6207.CAPR-20-0425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Johnson D.; Lardy H.. Isolation of Liver or Kidney Mitochondria. In Methods in Enzymology; Elsevier, 1967; Vol. 10, pp 94–96. [Google Scholar]
- Åkerman K. E.; Wikstrom M. K. Safranine as a probe of the mitochondrial membrane potential. FEBS Lett. 1976, 68, 191–197. 10.1016/0014-5793(76)80434-6. [DOI] [PubMed] [Google Scholar]
- Mueller P.; Rudin D. O.; Tien H. T.; Wescott W. C. Methods for the formation of single bimolecular lipid membranes in aqueous solution. J. Phys. Chem. A 1963, 67, 534–535. 10.1021/j100796a529. [DOI] [Google Scholar]
- Gerencser A. A.; Chinopoulos C.; Birket M. J.; Jastroch M.; Vitelli C.; Nicholls D. G.; Brand M. D. Quantitative measurement of mitochondrial membrane potential in cultured cells: calcium-induced de- and hyperpolarization of neuronal mitochondria. J. Physiol. 2012, 590, 2845–2871. 10.1113/jphysiol.2012.228387. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.







