Next Article in Journal
Pathological Mechanisms of Bortezomib-Induced Peripheral Neuropathy
Previous Article in Journal
PEG-BHD1028 Peptide Regulates Insulin Resistance and Fatty Acid β-Oxidation, and Mitochondrial Biogenesis by Binding to Two Heterogeneous Binding Sites of Adiponectin Receptors, AdipoR1 and AdipoR2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fluorescence Quenching Studies on the Interactions between Chosen Fluoroquinolones and Selected Stable TEMPO and PROXYL Nitroxides

Faculty of Chemistry, University of Gdańsk, Wita Stwosza 63, 80-308 Gdańsk, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(2), 885; https://doi.org/10.3390/ijms22020885
Submission received: 20 November 2020 / Revised: 7 January 2021 / Accepted: 13 January 2021 / Published: 17 January 2021
(This article belongs to the Section Physical Chemistry and Chemical Physics)

Abstract

:
The influence of the stable 2,2,6,6-tetramethylpiperidinyl-N-oxyl (TEMPO) nitroxide and its six C4-substituted derivatives, as well as two C3-substituted analogues of 2,2,5,5-tetramethylpyrrolidynyl-N-oxyl (PROXYL) nitroxide on the chosen fluoroquinolone antibiotics (marbofloxacin, ciprofloxacin, danofloxacin, norfloxacin, enrofloxacin, levofloxacin and ofloxacin), has been examined in aqueous solutions by UV absorption as well as steady-state and time-resolved fluorescence spectroscopies. The mechanism of fluorescence quenching has been specified and proved to be purely dynamic (collisional) for all the studied systems, which was additionally confirmed by temperature dependence experiments. Moreover, the selected quenching parameters—that is, Stern–Volmer quenching constants and bimolecular quenching rate constants—have been determined and explained. The possibility of electron transfer was ruled out, and the quenching was found to be diffusion-limited, being a result of the increase in non-radiative processes. Furthermore, as the chosen nitroxides affected the fluorescence of fluoroquinolone antibiotics in different ways, an influence of the structure and the type of substituents in the molecules of both fluoroquinolones and stable radicals on the quenching efficiency has been determined and discussed. Finally, the impact of the solvent’s polarity on the values of bimolecular quenching rate constants has been explained. The significance of the project comes from many applications of nitroxides in chemistry, biology and industry.

1. Introduction

The compounds 2,2,6,6-tetramethylpiperidinyl-N-oxyl (TEMPO), 2,2,5,5-tetramethylpyrrolidynyl-N-oxyl (PROXYL) and their derivatives—because of a large resonance and steric effects—are very stable [1]. These membrane-permeable nitroxides exhibit efficient protection of cells and tissues from damages related to an overproduction of free radicals (among others nitrogen dioxide [2]), which is due to the oxidative [3] and nitrative stress conditions [4]. This stems from the fact that they are able to catalyze the superoxide dismutation (mimic activity) [5], as well as detoxify carbon-centered [6], hydroxyl [7] and peroxyl radicals [8]. Furthermore, profluorescent nitroxides have been previously reported as probes for the detection of free radicals, as well as damage mediated by these species, for example acute pancreatitis and chronic splanchnic ischemia [9]. A variety of TEMPO and PROXYL analogues is also used as—among others—spin labels [10,11], radical scavengers [12], lipid oxidants [13], polymerization inhibitors [14], candidates for development as antiarrhythmic preventive drugs [15,16] or catalysts for chemical oxidation of alcohols [17]. The studies performed with the participation of these stable paramagnetic intermediates are known to result in some important phenomena, such as the quenching of photoexcited molecules [18] or formation of singlet oxygen [19]. Moreover, as the magnetic properties in the ground state are different before and after the photoexcitation, they seem to be attractive for controlling magnetic properties by light [20].
The fluoroquinolone antibiotics are a class of fully synthetic, broad-spectrum antibacterial agents with a unique mechanism of action and wide clinical use, structurally related to nalidixic acid and oxolinic acid [21]. Their activities result from the inhibition of various kinds of bacterial enzymes (among others DNA gyrase and topoisomerase IV) involved in the control of DNA synthesis and are vital for chromosome function, replication and transcription [22,23]. Fluoroquinolones are generally highly effective against both aerobic Gram (−) and many Gram (+) bacteria [24] and, thus, very helpful in the treatment of a wide range of infections, including urinary, bone-joint, enteral, soft tissue and respiratory tract infections, typhoid fever, prostatitis, blood poisoning, sinusitis, gonorrhea and bacterial gastroenteritis [25]. On the other hand, there is great concern for the fluoroquinolones that are widely used and at the same time are not readily biodegradable by microorganisms [26]. Different generations of fluoroquinolone antibiotics have been already developed to enlarge the spectrum of their activity and to solve problems of their resistance [27]. Since it has been proven that various types of fluoroquinolones—depending on substituents in the structure—exhibit quite different antibacterial responses and stability, it became very important to study the physical and chemical properties of these molecules in various environments [28]. Although the mechanism of action of all fluoroquinolones is very similar, there are many significant differences in their antimicrobial spectrum of activity, pharmacokinetic characteristics, safety profiles, transport characteristics and complexing properties—mainly because of a diversity of functional groups in antibiotics molecules [29]. All these prove that a precise understanding of the interaction mechanisms between antibiotics and other compounds is essential and of great importance.
In the following study, spectroscopic techniques, namely UV absorption and steady-state fluorescence spectroscopy, supported by time-resolved fluorescence lifetimes measurements, were used to investigate the interactions between different fluoroquinolones and selected TEMPO and PROXYL derivatives in aqueous solutions. Since all nitroxides were found to decrease the fluorescence intensity of the chosen antibiotics, while individual fluoroquinolones are known to exhibit particular properties in relation to their chemical structures [30], the main goal of the studies was to understand the mechanism of quenching involved in these systems as well as to determine whether the structure of both the fluorophore and the quencher has an influence on the observed fluorescence quenching efficiency.

2. Results and Discussion

For all the studied fluoroquinolone antibiotics, their UV absorption and fluorescence emission spectra were recorded under the action of the chosen nine nitroxide radicals. As similar observations were made for all 63 combinations, in the following paper, the spectra are presented, and their changes are explained on the example of danofloxacin and 4-hydroxy-TEMPO—the system in case of which the most effective fluorescence quenching was observed. Figure S1A in Supplementary Materials shows UV absorption spectra of the aqueous solution of danofloxacin (10 µM) in the presence of increasing amounts of 4-hydroxy-TEMPO, while Figure S1B presents the spectra of pure 4-hydroxy-TEMPO at analogous concentrations in aqueous solutions. From the inspection of these spectra, it can be clearly observed that in the studied range of concentrations, 4-hydroxy-TEMPO exhibits no influence on the UV spectrum of danofloxacin. When the concentration of 4-hydroxy-TEMPO increases, the band with a maximum at 243 nm (attributed to the presence of the nitroxide) gradually appears with no impact on the bands attributed to the antibiotic (namely 282 and 344 nm). Figure S2 in Supplementary Materials shows the fluorescence emission spectra of the excited at 340 nm pure danofloxacin and one spectrum in the presence of increasing concentrations of 4-hydroxy-TEMPO. The addition of the nitroxide immediately affects solely the fluorescence intensity of the fluoroquinolone with no changes in the shape of its spectrum and the shift of the band with maximum at 440 nm, which suggests no change in hydrophobicity of the microenvironment of the potential drug binding region [31]. All these findings (from both UV absorption and fluorescence emission measurements) prove that strictly physical interactions between danofloxacin and 4-hydroxy-TEMPO occur, with no option for a complex (also emissive exciplex) formation or photochemistry involved.
Figure 1 shows the Stern–Volmer plots obtained for the steady-state fluorescence quenching of the studied fluoroquinolones by various TEMPO and PROXYL derivatives in aqueous solutions. In the chosen range of nitroxides’ concentrations, there are no deviations from the observed linearity—the fluorescence quenching shows very clearly a linear dependence such as the one represented by the Stern–Volmer equation. Since it has been previously proven that 4-hydroxy-TEMPO impacts (shortens) on a fluorescence lifetime of selected fluoroquinolones [32], it indicates that dynamic, collisional fluorescence quenching occurred. Consequently, Table 1 presents the newly determined values of Stern–Volmer dynamic quenching constants (KD) along with linear correlation coefficients (R2) for Stern–Volmer plots and bimolecular quenching rate constants (kq) obtained for the studied systems. The latter ones were calculated based on determined values of fluorescence lifetimes, which were found to be in good agreement with literature data [33,34]. The postulated mechanism of fluorescence quenching corresponds well with the results obtained from UV absorption experiments, since collisions among the fluorophore and quencher molecules would not change the absorption spectrum. Additionally, the above finding has been confirmed by experiments at various temperatures. From the inspection of Figure S3 in Supplementary Materials, it can be observed that the increase in temperature results in greater values of Stern–Volmer constants (higher slopes). It is characteristic for dynamic quenching (the higher the temperature, the faster diffusion and, thus, more collisions)—contrary to static quenching, where the increase in the temperature favors dissociation of weakly bound complexes (higher temperature is likely to reduce the stability of the complexes, resulting in lower static quenching constants) [35]. The fact that stable nitroxides quench the fluorescence of the studied fluoroquinolones through the collisions is quite interesting, since a vast majority of scientific data reports the static quenching of these fluorophores by various quenchers [31,36,37,38,39,40].
To gain a deeper insight into the mechanism involved, it should be noted that there is no obvious correlation between the observed fluorescence quenching efficiency (expressed by the values of bimolecular quenching rate constants) and ionization energies of the studied quenchers [41] and antibiotics [32]. Thus, the mechanism of electron transfer can be excluded. Furthermore, it is hard to observe a relationship between kq values and nitroxides’ molecular radii—probably because of comparable sizes of the chosen radicals and fluoroquinolones. Generally, it is more than likely that various factors simultaneously have an influence on the reported decrease in fluorescence. Being amphoteric compounds, fluoroquinolone antibiotics may occur in various forms related to the values of dissociation constants pKa—at pH near neutral they exist mainly as zwitterionic species [42]. Since at the same conditions nitroxide radicals may occur as neutral molecules as well as both negatively and positively charged ions, electrostatic interactions should not be ruled out. It has been confirmed by our previous studies, which revealed that better packing and access of a quencher to the electrically polarized molecule of a fluorophore increases the efficiency of fluorescence quenching [43]. Additionally, bimolecular quenching rate constants are of the same order of magnitude as diffusion rate constants obtained for studies performed in water [35]—thus, it is highly probable that the bimolecular process is diffusion-limited, caused by the increase in rate constants of non-radiative transitions, such as an internal conversion or an intersystem crossing. Interestingly, in the case of all studied fluoroquinolones, kq values obtained for 3-carboxy-PROXYL are slightly higher than for 4-carboxy-TEMPO. Although the reasons for the observed difference between five-membered and six-membered analogues used as quenchers may come from the electron distribution, they are not clear, and evidently, there is a scope for further investigations. By looking at the results obtained for 3-carbamoyl-PROXYL in water and the same quencher in methanol, it can be concluded that the greater value of dielectric constant of the solvent (the more polar solvent), the lower efficiency of the fluorescence quenching of all fluoroquinolones. With a high probability, it is a consequence of two effects, one associated with preferential solvation and related changes in intermolecular interactions and the second one connected with the changes in environmental viscosity.

3. Materials and Methods

All reagents, that is, the ones chosen for the studies of fluoroquinolone antibiotics (marbofloxacin, ciprofloxacin, danofloxacin, norfloxacin, enrofloxacin, levofloxacin and ofloxacin), TEMPO (2,2,6,6-tetramethylpiperidinyl-N-oxyl) and its C4-substituted derivatives (4-hydroxy-TEMPO, 4-methoxy-TEMPO, 4-carboxy-TEMPO, 4-acetamido-TEMPO, 4-oxo-TEMPO and TEMPO methacrylate) as well as two C3-substituted analogues of PROXYL (2,2,5,5-tetramethylpyrrolidynyl-N-oxyl), namely 3-carboxy-PROXYL and 3-carbamoyl-PROXYL, were purchased from Sigma Aldrich (Poland). All reagents were of purity no less than 97% and were used without further purification. The double-distilled water with a conductivity less than 0.18 μS∙cm−1 as well as methanol (purity >99.8%, water <0.1%) were used as solvents. The stock solutions of fluoroquinolones and nitroxides were prepared just before use by dissolving an appropriate amount of the substance in water (with the help of sonication if necessary; in one case, i.e., 3-carbamoyl-PROXYL, additionally in methanol). To avoid self-quenching or inner filter effects, the solutions of all antibiotics were prepared keeping the constant concentration (10 µM). In all fluorometric measurements, the optical density of the solutions did not exceed 0.1. The concentration of the stock solutions of TEMPO and PROXYL radicals was equal to 0.1 M. The molecular structures of the studied compounds are presented in Figure 2.
UV absorption spectra of the aqueous solutions of the studied fluoroquinolones (10 µM) were recorded on a Perkin Elmer Lambda 650 (Waltham, MA, USA) UV-Vis spectrophotometer (a slit—2 nm) at the temperature of 20 °C in the presence of all the selected TEMPO and PROXYL nitroxides used at concentrations 0–0.25 mM.
Fluorescence emission spectra as well as all fluorescence intensity measurements were performed with the use of a Cary Eclipse Varian (Agilent, Santa Clara, CA, USA) spectrofluorometer—equipped with a temperature controller and a 1.0 cm multicell holder; excitation and emission slits −2.5 or 5 nm, depending on fluoroquinolone’s quantum yield—at the temperature of 20 °C (in one case, i.e., danofloxacin/4-hydroxy-TEMPO system, additionally at 30 and 40 °C) in the presence of all the selected TEMPO and PROXYL nitroxides used at concentrations 0–2.5 mM. Excitation wavelength in case of steady-state experiments for all fluoroquinolones was set at 340 nm, based on absorption spectra, which are presented in Figure S4 in Supplementary Materials. The fluorescence intensity values were always measured at the maximum of the emission (470 nm for marbofloxacin; 423 nm for ciprofloxacin; 428 nm for norfloxacin; 440 nm for danofloxacin; 437 nm for enrofloxacin; 468 nm for levofloxacin; 475 nm for ofloxacin; fluorescence emission spectra of all studied antibiotics are presented in Figure S5 in Supplementary Materials). In the performed titration experiments, 2 mL of each fluoroquinolone at 10 µM was titrated with five 10 μL aliquots of each nitroxide solution at 0.1 M, and after gentle stirring, the corresponding fluorescence intensity was measured. The Stern–Volmer equation was used to determine the type of fluorescence quenching involving the studied fluoroquinolones and nitroxides:
F 0 F = 1 + K S V [ Q ] = 1 +   k q τ 0 [ Q ] .
In this equation, F0 and F are the fluorescence intensities in the absence and presence of quencher, respectively; kq is the bimolecular quenching rate constant; τ0 is the lifetime of the fluorophore in the absence of quencher; [Q] is the concentration of quencher. The Stern–Volmer quenching constant is given by KSV = kqτ0. If the quenching is known to be dynamic, the Stern–Volmer constant is represented by KD. In that case, a linear plot of F0/F versus [Q] yields an intercept of one on the y-axis and a slope equal to KD. Furthermore, for the system, in the case of which the most significant fluorescence intensity changes were observed (danofloxacin + 4-hydroxy-TEMPO)—to confirm the observed fluorescence quenching mechanism—the decrease in fluorescence intensity of the antibiotic under the action of the nitroxide was recorded at three different temperatures, namely 20, 30 and 40 °C with an accuracy of ±0.01 °C. Since the fluorescence measurements may be adversely affected by the inner-filter effects [44,45]—which can be caused by the absorption by the nitroxides of light at the excitation and emission wavelengths of fluoroquinolones—all experimentally determined values of fluorescence intensities were corrected using the following equation:
F corr = F obs · 10 A ex + A em 2 ,
where Fcorr and Fobs are the corrected and observed fluorescence intensities, respectively; whereas, Aex and Aem are the sum of the absorbance of the appropriate antibiotic and the nitroxide radical at the excitation and emission wavelength, respectively.
Fluorescence lifetimes of all studied fluoroquinolone antibiotics were measured using the single photon counting technique (at the maximum of emission chosen by monochromator; c = 10 µM) with the use of a FluoTime 300 (PicoQuant, Berlin, Germany) high performance fluorescence lifetime spectrometer (PicoQuant) at 20 °C. The excitation source was subnanosecond pulsed diode: PLS-340 nm.

4. Conclusions

The paper presents the results of our studies on the interactions between chosen fluoroquinolone antibiotics and various TEMPO and PROXYL nitroxides. It has been proven that fluorescence of all investigated fluoroquinolones is quenched by the stable radicals. The results of steady-state fluorescence and time-resolved measurements made it possible to determine bimolecular quenching rate constants for all the systems studied. The mechanism of fluorescence quenching was found to be totally dynamic/collisional and diffusion-limited due to the increase in non-radiative processes, such as the internal conversion or intersystem crossing. Furthermore, as the chosen nitroxides affected fluorescence of fluoroquinolone antibiotics in different ways, an influence of the structure and the type of substituents in the molecules of both fluoroquinolones and stable radicals on the quenching efficiency has been determined and discussed. The performance of fluorescence quenching experiments in pure water and methanol enabled us to state that the more polar solvent, the lower efficiency of quenching, which may arise from the solvation and/or viscosity effect.

Supplementary Materials

The following are available online at https://www.mdpi.com/1422-0067/22/2/885/s1. Figure S1: UV absorption spectra of danofloxacin (10 µM) in the presence of increasing concentrations of 4-hydroxy-TEMPO (0–0.25 mM in the measured sample) (A) and pure 4-hydroxy-TEMPO at analogous concentrations in aqueous solution (B), Figure S2: Fluorescence emission spectra of danofloxacin (10 µM) in the presence of increasing concentrations of 4-hydroxy-TEMPO (0–2.5 mM in the measured sample); λex = 340 nm, Figure S3: Stern–Volmer plots from the steady-state fluorescence quenching of danofloxacin (10 µM) by 4-hydroxy-TEMPO in aqueous solutions at different temperatures (20, 30 and 40 °C); λex = 340 nm. The results are shown as mean ± standard deviation (S.D.) of three independent experiments. In all experiments, standard deviations were less than 4%, and thus, for a better clarity, error bars were omitted, Figure S4: UV absorption spectra of the studied fluoroquinolone antibiotics (10 µM), Figure S5: Fluorescence emission spectra of the studied fluoroquinolone antibiotics (10 µM); λex = 340 nm.

Author Contributions

Conceptualization, K.Ż.; methodology, K.Ż.; investigation, K.Ż. and I.B.; writing—original draft preparation, K.Ż., I.B.; supervision, K.Ż., W.W. and L.C.; funding acquisition, K.Ż. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Polish National Science Centre under Grant No. 2016/23/D/ST4/01576 (K.Ż.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on reasonable request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Forrester, A.R.; Thomson, R.H. Stable Nitroxide Radicals. Nat. Cell Biol. 1964, 203, 74–75. [Google Scholar] [CrossRef]
  2. Żamojć, K.; Jacewicz, D.; Zdrowowicz, M.; Chmurzyński, L. Kinetics of the reaction between 1,3-diphenylisobenzofuran and nitrogen dioxide studied by steady-state fluorescence. Res. Chem. Intermed. 2012, 39, 3023–3031. [Google Scholar] [CrossRef]
  3. Samuni, A.; Winkelsberg, D.; Pinson, A.; Hahn, S.M.; Mitchell, J.B.; Russo, A. Nitroxide stable radicals protect beating cardiomyocytes against oxidative damage. J. Clin. Investig. 1991, 87, 1526–1530. [Google Scholar] [CrossRef] [PubMed]
  4. Żamojć, K.; Zdrowowicz, M.; Rudnicki-Velasquez, P.B.; Krzymiński, K.; Zaborowski, B.; Niedziałkowski, P.; Jacewicz, D.; Chmurzyński, L. The development of 1,3-diphenylisobenzofuran as a highly selective probe for the detection and quantitative determination of hydrogen peroxide. Free Radic. Res. 2017, 51, 38–46. [Google Scholar] [CrossRef] [PubMed]
  5. Śledziński, Z.; Woźniak, M.; Antosiewicz, J.; Lezoche, E.; Familiari, M.; Bertoli, E.; Greci, L.; Brunelli, A.; Mazera, N.; Wajda, Z. Protective effect of 4-hydroxy-TEMPO, a low molecular weight superoxide dismutase mimic, on free rad-ical toxicity in experimental pancreatitis. Int. J. Pancreatol. 1995, 18, 153–160. [Google Scholar] [PubMed]
  6. Chateauneuf, J.; Lusztyk, J.; Ingold, K.U. Absolute rate constants for the reactions of some carbon-centered radicals with 2,2,6,6-tetramethyl-1-piperidinoxyl. J. Org. Chem. 1988, 53, 1629–1632. [Google Scholar] [CrossRef]
  7. Samuni, A.; Goldstein, S.; Russo, A.; Mitchell, J.B.; Krishna, M.C.; Neta, P. Kinetics and Mechanism of Hydroxyl Radical and OH-Adduct Radical Reactions with Nitroxides and with Their Hydroxylamines. J. Am. Chem. Soc. 2002, 124, 8719–8724. [Google Scholar] [CrossRef]
  8. Goldstein, S.; Samuni, A. Kinetics and Mechanism of Peroxyl Radical Reactions with Nitroxides. J. Phys. Chem. A 2007, 111, 1066–1072. [Google Scholar] [CrossRef]
  9. Blinco, J.P.; Keddie, D.J.; Wade, T.; Barker, P.J.; George, G.A.; Bottle, S. Profluorescent nitroxides: Sensors and stabilizers of radical-mediated oxidative damage. Polym. Degrad. Stab. 2008, 93, 1613–1618. [Google Scholar] [CrossRef] [Green Version]
  10. Keana, J.F.W. Newer aspects of the synthesis and chemistry of nitroxide spin labels. Chem. Rev. 1978, 78, 37–64. [Google Scholar] [CrossRef]
  11. Żamojć, K.; Zdrowowicz, M.; Hać, A.; Witwicki, M.; Rudnicki-Velasquez, P.B.; Wyrzykowski, D.; Wiczk, W.; Chmurzyński, L. Dihydroxy-Substituted Coumarins as Fluorescent Probes for Nanomolar-Level Detection of the 4-Amino-TEMPO Spin Label. Int. J. Mol. Sci. 2019, 20, 3802. [Google Scholar] [CrossRef] [Green Version]
  12. Krishna, M.C.; Russo, A.; Mitchell, J.B.; Goldstein, S.; Dafni, H.; Samuni, A. Do nitroxide antioxidants act as scav-engers of O2•- or as SOD mimics? J. Biol. Chem. 1996, 271, 26026–26031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Weil, J.T.; Van Der Veen, J.; Olcott, H.S. Stable Nitroxides as Lipid Antioxidants. Nat. Cell Biol. 1968, 219, 168–169. [Google Scholar] [CrossRef] [PubMed]
  14. Conte, M.; Ma, Y.; Loyns, C.; Price, P.; Rippon, D.; Chechik, V. Mechanistic insight into TEMPO-inhibited polymerisation: Simultaneous determination of oxygen and inhibitor concentrations by EPR. Org. Biomol. Chem. 2009, 7, 2685–2687. [Google Scholar] [CrossRef] [PubMed]
  15. Żamojć, K.; Zdrowowicz, M.; Wiczk, W.; Jacewicz, D.; Chmurzynski, L. Dihydroxycoumarins as highly selective fluorescent probes for the fast detection of 4-hydroxy-TEMPO in aqueous solution. Rsc Adv. 2015, 5, 63807–63812. [Google Scholar] [CrossRef]
  16. Gelvan, D.; Saltman, P.; Powell, S.R. Cardiac reperfusion damage prevented by a nitroxide free radical. Proc. Natl. Acad. Sci. USA 1991, 88, 4680–4684. [Google Scholar] [CrossRef] [Green Version]
  17. Cella, J.A.; Kelley, J.A.; Kenehan, E.F. Nitroxide-catalyzed oxidation of alcohols using m-chloroperbenzoic acid. New method. J. Org. Chem. 1975, 40, 1860–1862. [Google Scholar] [CrossRef]
  18. Khudyakov, I.V.; Turro, N.J. A Tr Esr study of the quenching of photoexcited dioxouranium (VI) salts by stable nitroxyl free radicals. Res. Chem. Intermed. 1993, 19, 15–30. [Google Scholar] [CrossRef] [Green Version]
  19. Lion, Y.; Delmelle, M.; Van De Vorst, A. New method of detecting singlet oxygen production. Nat. Cell Biol. 1976, 263, 442–443. [Google Scholar] [CrossRef]
  20. Ishii, K.; Hirose, Y.; Fujitsuka, H.; Ito, O.; Kobayashi, N. Time-Resolved EPR, Fluorescence, and Transient Absorption Studies on Phthalocyaninatosilicon Covalently Linked to One or Two TEMPO Radicals. J. Am. Chem. Soc. 2001, 123, 702–708. [Google Scholar] [CrossRef]
  21. A Divo, A.; Sartorelli, A.C.; Patton, C.L.; Bia, F.J. Activity of fluoroquinolone antibiotics against Plasmodium falciparum in vitro. Antimicrob. Agents Chemother. 1988, 32, 1182–1186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Robinson, A.A.; Belden, J.B.; Lydy, M.J. Toxicity of fluoroquinolone antibiotics to aquatic organisms. Environ. Toxicol. Chem. 2005, 24, 423–430. [Google Scholar] [CrossRef] [PubMed]
  23. Hartmann, A.; Alder, A.C.; Koller, T.; Widmer, R.M. Identification of fluoroquinolone antibiotics as the main source of umuc genotoxicity in native hospital wastewater. Environ. Toxicol. Chem. 1998, 17, 377. [Google Scholar] [CrossRef]
  24. Redgrave, L.S.; Sutton, S.B.; Webber, M.A.; Piddock, L.J.V. Fluoroquinolone resistance: Mechanisms, impact on bacteria, and role in evolutionary success. Trends Microbiol. 2014, 22, 438–445. [Google Scholar] [CrossRef]
  25. Paton, J.; Reeves, D. Fluoroquinolone Antibiotics. Drugs 1988, 36, 193–228. [Google Scholar] [CrossRef]
  26. Kümmerer, K.; Al-Ahmad, A.; Mersch-Sundermann, V. Biodegradability of some antibiotics, elimination of the genotoxicity and affection of wastewater bacteria in a simple test. Chemosphere 2000, 40, 701–710. [Google Scholar] [CrossRef]
  27. Mamalis, N. The increasing problem of antibiotic resistance. J. Cataract. Refract. Surg. 2007, 33, 1831–1832. [Google Scholar] [CrossRef]
  28. Albini, A.; Monti, S. Photophysics and photochemistry of fluoroquinolones. Chem. Soc. Rev. 2003, 32, 238–250. [Google Scholar] [CrossRef]
  29. Smith, A.; Pennefather, P.M.; Kaye, S.B.; Hart, C.A. Fluoroquinolones. Drugs 2001, 61, 747–761. [Google Scholar] [CrossRef]
  30. Ball, P.; Tillotson, G. Tolerability of Fluoroquinolone Antibiotics Past, Present and Future. Drug Saf. 1995, 13, 343–358. [Google Scholar] [CrossRef]
  31. Seedher, N.; Agarwal, P. Complexation of fluoroquinolone antibiotics with human serum albumin: A fluorescence quenching study. J. Lumin. 2010, 130, 1841–1848. [Google Scholar] [CrossRef]
  32. Żamojć, K.; Wiczk, W.; Zaborowski, B.; Makowski, M.; Pranczk, J.; Jacewicz, D.; Chmurzyński, L. Fluorescence quenching of fluoroquinolone antibiotics by 4-hydroxy-TEMPO in aqueous solution. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 133, 887–891. [Google Scholar] [CrossRef] [PubMed]
  33. Park, H.-R.; Kim, T.H.; Bark, K.-M. Physicochemical properties of quinolone antibiotics in various environments. Eur. J. Med. Chem. 2002, 37, 443–460. [Google Scholar] [CrossRef]
  34. Viola, G.; Facciolo, L.; Canton, M.; Vedaldi, D.; Dall’Acqua, F.; Aloisi, G.G.; Amelia, M.; Barbafina, A.; Elisei, F.; Latterini, L. Photophysical and Phototoxic Properties of the Antibacterial Fluoroquinolones Levofloxacin and Moxifloxacin. Chem. Biodivers. 2004, 1, 782–801. [Google Scholar] [CrossRef] [PubMed]
  35. Lakowicz, J.R. (Ed.) Principles of Fluorescence Spectroscopy, 3rd ed.; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  36. Wang, J.; Liu, Z.; Liu, J.; Liu, S.; Shen, W. Study on the interaction between fluoroquinolones and erythrosine by absorption, fluorescence and resonance Rayleigh scattering spectra and their application. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2008, 69, 956–963. [Google Scholar] [CrossRef]
  37. Wang, J.; Kong, L.; Shen, W.; Hu, X.; Shen, Y.; Liu, S. Synergistic fluorescence quenching of quinolone antibiotics by palladium(ii) and sodium dodecyl benzene sulfonate and the analytical application. Anal. Methods 2014, 6, 4343–4352. [Google Scholar] [CrossRef]
  38. Neves, P.; Sousa, I.; Winterhalter, M.; Gameiro, P. Fluorescence quenching as a tool to investigate quinolone anti-biotic interactions with bacterial protein OmpF. J. Membr. Biol. 2009, 227, 133–140. [Google Scholar] [CrossRef]
  39. Fei, Y.; Lu, G.; Fan, G.; Wu, Y. Spectroscopic studies on the binding of a new quinolone antibacterial agent: Sin-afloxacin to DNA. Anal. Sci. 2009, 25, 1333–1338. [Google Scholar] [CrossRef] [Green Version]
  40. Zhao, X.; Hu, Z.; Yang, X.; Cai, X.; Wang, Z.; Xie, X. Noncovalent interactions between fluoroquinolone antibiotics with dissolved organic matter: A 1H NMR binding site study and multi-spectroscopic methods. Environ. Pollut. 2019, 248, 815–822. [Google Scholar] [CrossRef]
  41. Żamojć, K.; Wiczk, W.; Chmurzyński, L. The influence of the type of substituents and the solvent on the interac-tions between different coumarins and selected TEMPO analogues–Fluorescence quenching studies. Chem. Phys. 2018, 513, 188–194. [Google Scholar] [CrossRef]
  42. Van Doorslaer, X.; Dewulf, J.; Van Langenhove, H.; Demeestere, K. Fluoroquinolone antibiotics: An emerging class of environmental micropollutants. Sci. Total Environ. 2014, 500, 250–269. [Google Scholar] [CrossRef] [PubMed]
  43. Żamojć, K.; Wiczk, W.; Zaborowski, B.; Jacewicz, D.; Chmurzyński, L. Fluorescence quenching of 7-amino-4-methylcoumarin by different TEMPO derivatives. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 136, 1875–1880. [Google Scholar] [CrossRef] [PubMed]
  44. Kubista, M.; Sjöback, R.; Eriksson, S.; Albinsson, B. Experimental correction for the inner-filter effect in fluorescence spectra. Analyst 1994, 119, 417–419. [Google Scholar] [CrossRef]
  45. Puchalski, M.M.; Morra, M.J.; Von Wandruszka, R. Assessment of inner filter effect corrections in fluorimetry. Anal. Bioanal. Chem. 1991, 340, 341–344. [Google Scholar] [CrossRef]
Figure 1. Stern–Volmer plots for the steady-state fluorescence quenching of the studied fluoroquinolones by different 2,2,6,6-tetramethylpiperidinyl-N-oxyl (TEMPO) and 2,2,5,5-tetramethylpyrrolidynyl-N-oxyl (PROXYL) derivatives in aqueous solutions at 20 °C. F0 and F are the fluorescence intensities of the fluoroquinolones in the absence and presence of the nitroxides, respectively. The results are shown as a mean of three independent experiments. In all experiments, standard deviations were less than 4%, and thus, for better clarity, error bars were omitted.
Figure 1. Stern–Volmer plots for the steady-state fluorescence quenching of the studied fluoroquinolones by different 2,2,6,6-tetramethylpiperidinyl-N-oxyl (TEMPO) and 2,2,5,5-tetramethylpyrrolidynyl-N-oxyl (PROXYL) derivatives in aqueous solutions at 20 °C. F0 and F are the fluorescence intensities of the fluoroquinolones in the absence and presence of the nitroxides, respectively. The results are shown as a mean of three independent experiments. In all experiments, standard deviations were less than 4%, and thus, for better clarity, error bars were omitted.
Ijms 22 00885 g001
Figure 2. Molecular structures of the selected nitroxide radicals and fluoroquinolone antibiotics (with their commercial names).
Figure 2. Molecular structures of the selected nitroxide radicals and fluoroquinolone antibiotics (with their commercial names).
Ijms 22 00885 g002
Table 1. Stern–Volmer dynamic quenching constants (KD), linear correlation coefficients (R2) and bimolecular quenching rate constants (kq) recovered for the steady-state fluorescence quenching of the studied fluoroquinolones by various TEMPO and PROXYL nitroxides in aqueous solutions at 20 °C. The results are shown as a mean ±standard deviation (S.D.) of three independent experiments.
Table 1. Stern–Volmer dynamic quenching constants (KD), linear correlation coefficients (R2) and bimolecular quenching rate constants (kq) recovered for the steady-state fluorescence quenching of the studied fluoroquinolones by various TEMPO and PROXYL nitroxides in aqueous solutions at 20 °C. The results are shown as a mean ±standard deviation (S.D.) of three independent experiments.
Nitroxide RadicalFluoroquinolone
(Fluorescence Lifetime ± S.D.)
KD ± S.D.
[mol−1·dm3]
R2kq ± S.D.
[mol−1·dm3·s−1]
TEMPOEnrofloxacin
(1.95 ± 0.02 ns)
48.32 ± 1.010.998(2.48 ± 0.08) × 1010
4-methoxy-TEMPO40.29 ± 0.241.000(2.07 ± 0.03) × 1010
4-hydroxy-TEMPO66.92 ± 1.620.998(3.43 ± 0.12) × 1010
4-oxo-TEMPO24.71 ± 0.540.998(1.27 ± 0.04) × 1010
4-acetamido-TEMPO56.48 ± 0.341.000(2.90 ± 0.05) × 1010
4-carboxy-TEMPO27.33 ± 0.241.000(1.40 ± 0.03) × 1010
3-carboxy-PROXYL29.52 ± 0.141.000(1.51 ± 0.02) × 1010
3-carbamoyl-PROXYL47.75 ± 1.930.994(2.45 ± 0.12) × 1010
3-carbamoyl-PROXYL (MeOH)49.26 ± 0.920.999(2.53 ± 0.07) × 1010
TEMPO methacrylate21.03 ± 1.180.988(1.08 ± 0.07) × 1010
TEMPOCiprofloxacin
(1.54 ± 0.01 ns)
18.93 ± 1.900.961(1.23 ± 0.13) × 1010
4-methoxy-TEMPO25.50 ± 1.020.994(1.66 ± 0.08) × 1010
4-hydroxy-TEMPO26.60 ± 1.700.984(1.73 ± 0.12) × 1010
4-oxo-TEMPO31.73 ± 1.450.992(2.06 ± 0.11) × 1010
4-acetamido-TEMPO13.07 ± 0.510.994(8.49 ± 0.39) × 109
4-carboxy-TEMPO20.02 ± 0.880.992(1.30 ± 0.07) × 1010
3-carboxy-PROXYL23.10 ± 0.620.997(1.50 ± 0.05) × 1010
3-carbamoyl-PROXYL4.11 ± 0.130.996(2.67 ± 0.10) × 109
3-carbamoyl-PROXYL (MeOH)4.15 ± 0.180.993(2.69 ± 0.13) × 109
TEMPO methacrylate64.30 ± 3.350.989(4.18 ± 0.24) × 1010
TEMPODanofloxacin
(5.97 ± 0.05 ns)
52.94 ± 1.670.996(8.87 ± 0.35) × 109
4-methoxy-TEMPO57.51 ± 1.810.996(9.63 ± 0.38) × 109
4-hydroxy-TEMPO127.3 ± 0.31.000(2.13 ± 0.02) × 1010
4-oxo-TEMPO47.77 ± 2.350.990(8.00 ± 0.46) × 109
4-acetamido-TEMPO106.6 ± 1.50.999(1.79 ± 0.04) × 1010
4-carboxy-TEMPO48.96 ± 0.660.999(8.20 ± 0.18) × 109
3-carboxy-PROXYL51.46 ± 0.471.000(8.62 ± 0.15) × 109
3-carbamoyl-PROXYL68.56 ± 1.890.997(1.15 ± 0.04) × 1010
3-carbamoyl-PROXYL (MeOH)84.70 ± 2.500.996(1.42 ± 0.05) × 1010
TEMPO methacrylate22.00 ± 1.860.972(3.69 ± 0.34) × 109
TEMPOMarbofloxacin
(3.54 ± 0.02 ns)
24.58 ± 0.650.997(6.94 ± 0.22) × 109
4-methoxy-TEMPO47.07 ± 2.840.986(1.33 ± 0.09) × 1010
4-hydroxy-TEMPO31.86 ± 0.820.997(9.00 ± 0.28) × 109
4-oxo-TEMPO61.13 ± 1.580.997(1.73 ± 0.05) × 1010
4-acetamido-TEMPO15.86 ± 0.830.989(4.48 ± 0.26) × 109
4-carboxy-TEMPO21.92 ± 0.680.996(6.19 ± 0.23) × 109
3-carboxy-PROXYL24.05 ± 2.090.971(6.79 ± 0.63) × 109
3-carbamoyl-PROXYL9.13 ± 1.070.948(2.58 ± 0.32) × 109
3-carbamoyl-PROXYL (MeOH)19.85 ± 1.140.987(5.61 ± 0.35) × 109
TEMPO methacrylate74.23 ± 2.310.996(2.10 ± 0.08) × 1010
TEMPONorfloxacin(1.62 ± 0.02 ns)29.56 ± 1.220.993(1.82 ± 0.10) × 1010
4-methoxy-TEMPO31.24 ± 1.120.995(1.93 ± 0.09) × 1010
4-hydroxy-TEMPO10.81 ± 0.210.998(6.67 ± 0.21) × 109
4-oxo-TEMPO35.47 ± 0.970.997(2.19 ± 0.09) × 1010
4-acetamido-TEMPO13.57 ± 0.870.984(8.38 ± 0.64) × 109
4-carboxy-TEMPO33.06 ± 0.390.999(2.04 ± 0.05) × 1010
3-carboxy-PROXYL34.46 ± 1.330.994(2.13 ± 0.11) × 1010
3-carbamoyl-PROXYL16.54 ± 0.740.992(1.02 ± 0.06) × 1010
3-carbamoyl-PROXYL (MeOH)29.20 ± 0.540.999(1.80 ± 0.06) × 1010
TEMPO methacrylate39.74 ± 1.520.994(2.45 ± 0.12) × 1010
TEMPOLevofloxacin(7.10 ± 0.06 ns)41.65 ± 1.340.996(5.87 ± 0.24) × 109
4-methoxy-TEMPO45.11 ± 1.650.995(6.35 ± 0.29) × 109
4-hydroxy-TEMPO22.12 ± 0.670.996(3.12 ± 0.12) × 109
4-oxo-TEMPO57.08 ± 1.780.996(8.04 ± 0.32) × 109
4-acetamido-TEMPO12.24 ± 1.030.978(1.72 ± 0.16) × 109
4-carboxy-TEMPO28.40 ± 0.500.999(4.00 ± 0.10) × 109
3-carboxy-PROXYL31.78 ± 0.181.000(4.48 ± 0.06) × 109
3-carbamoyl-PROXYL28.62 ± 0.640.998(4.03 ± 0.12) × 109
3-carbamoyl-PROXYL (MeOH)30.64 ± 1.300.993(4.32 ± 0.22) × 109
TEMPO methacrylate49.06 ± 1.150.998(6.91 ± 0.22) × 109
TEMPOOfloxacin(7.01 ± 0.05 ns)43.74 ± 0.810.999(6.24 ± 0.16) × 109
4-methoxy-TEMPO46.56 ± 1.450.996(6.64 ± 0.25) × 109
4-hydroxy-TEMPO21.33 ± 0.510.998(3.04 ± 0.09) × 109
4-oxo-TEMPO55.53 ± 2.270.993(7.92 ± 0.38) × 109
4-acetamido-TEMPO27.10 ± 0.810.996(3.87 ± 0.14) × 109
4-carboxy-TEMPO28.11 ± 0.680.998(4.00 ± 0.13) × 109
3-carboxy-PROXYL32.45 ± 0.361.000(4.63 ± 0.08) × 109
3-carbamoyl-PROXYL25.35 ± 1.270.990(3.62 ± 0.21) × 109
3-carbamoyl-PROXYL (MeOH)27.90 ± 1.070.994(3.98 ± 0.18) × 109
TEMPO methacrylate61.78 ± 2.260.995(8.81 ± 0.39) × 109
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Żamojć, K.; Bylińska, I.; Wiczk, W.; Chmurzyński, L. Fluorescence Quenching Studies on the Interactions between Chosen Fluoroquinolones and Selected Stable TEMPO and PROXYL Nitroxides. Int. J. Mol. Sci. 2021, 22, 885. https://doi.org/10.3390/ijms22020885

AMA Style

Żamojć K, Bylińska I, Wiczk W, Chmurzyński L. Fluorescence Quenching Studies on the Interactions between Chosen Fluoroquinolones and Selected Stable TEMPO and PROXYL Nitroxides. International Journal of Molecular Sciences. 2021; 22(2):885. https://doi.org/10.3390/ijms22020885

Chicago/Turabian Style

Żamojć, Krzysztof, Irena Bylińska, Wiesław Wiczk, and Lech Chmurzyński. 2021. "Fluorescence Quenching Studies on the Interactions between Chosen Fluoroquinolones and Selected Stable TEMPO and PROXYL Nitroxides" International Journal of Molecular Sciences 22, no. 2: 885. https://doi.org/10.3390/ijms22020885

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop