You are on page 1of 13

Free Radical Biology and Medicine 179 (2022) 34–46

Contents lists available at ScienceDirect

Free Radical Biology and Medicine


journal homepage: www.elsevier.com/locate/freeradbiomed

Water-soluble cationic boronate probe based on coumarin imidazolium


scaffold: Synthesis, characterization, and application to cellular
peroxynitrite detection
Aleksandra Grzelakowska a, *, Julia Modrzejewska a, Jolanta Kolińska a, Marcin Szala a,
Monika Zielonka b, Karolina Dębowska c, Małgorzata Zakłos-Szyda d, Adam Sikora c,
Jacek Zielonka b, **, Radosław Podsiadły a, ***
a
Institute of Polymer and Dye Technology, Faculty of Chemistry, Lodz University of Technology, Stefanowskiego 16, 90-537, Lodz, Poland
b
Department of Biophysics, Medical College of Wisconsin, 8701 Watertown Plank Road, Milwaukee, WI, 53226, United States
c
Institute of Applied Radiation Chemistry, Faculty of Chemistry, Lodz University of Technology, Żeromskiego 116, 90-924, Lodz, Poland
d
Institute of Molecular and Industrial Biotechnology, Faculty of Biotechnology and Food Sciences, Lodz University of Technology, Stefanowskiego 2/22, 90-537, Lodz,
Poland

A R T I C L E I N F O A B S T R A C T

Keywords: Peroxynitrite (ONOO− ) has been implicated in numerous pathologies associated with an inflammatory compo­
Water-soluble fluorescent probe nent, but its selective and sensitive detection in biological settings remains a challenge. Here, the development of
Coumarin-based probe a new water-soluble and cationic boronate probe based on a coumarin-imidazolium scaffold (CI-Bz-BA) for the
Mitochondria-targeted two-photon fluorescent
fluorescent detection of ONOO− in cells is reported. The chemical reactivity of the CI-Bz-BA probe toward
probe
Redox probe
selected oxidants known to react with the boronate moiety was characterized, and the suitability of the probe for
Peroxynitrite the direct detection of ONOO− in cell-free and cellular system is reported. Oxidation of the probe results in the
formation of the primary hydroxybenzyl product (CI-Bz-OH), followed by the spontaneous elimination of the
quinone methide moiety to produce the secondary phenol (CI–OH), which is accompanied by a red shift in the
fluorescence emission band from 405 nm to 481 nm. CI-Bz-BA reacts with ONOO− stoichiometrically with a rate
constant of ~1 × 106 M-1s-1 to form, in addition to the major phenolic product CI–OH, the minor nitrated
product CI-Bz-NO2, which is not formed by other oxidants tested or via myeloperoxidase-catalyzed oxidation/
nitration. Both CI–OH and CI-Bz-NO2 products were also formed in the presence of cogenerated fluxes of nitric
oxide and superoxide radical anion produced during decomposition of a SIN-1 donor. Using RAW 264.7 cells, we
demonstrate the ability of the probe to report endogenously produced ONOO− via fluorescence measurements,
including plate reader real time monitoring and two-photon fluorescence imaging. Liquid chromatography/mass
spectrometry analyses of cell extracts and media confirmed the formation of both CI–OH and CI-Bz-NO2 in
macrophages activated to produce ONOO− . We propose the use of a combination of real-time monitoring of
probe oxidation using fluorimetry and fluorescence microscopy with liquid chromatography/mass spectrometry-
based product identification for rigorous detection and quantitative analyses of ONOO− in biological systems.

1. Introduction reaction between nitric oxide (•NO) and superoxide radical anion
(O2•‒) [1–3], has been considered as a key oxidizing and nitrating
Among various reactive oxygen and nitrogen species in biological species formed in various pathophysiological states, especially involving
systems, peroxynitrite (ONOO− ), the product of a diffusion-controlled inflammatory components [4–7]. Recent literature points to the

* Corresponding author.
** Corresponding author.
*** Corresponding author.
E-mail addresses: aleksandra.grzelakowska@p.lodz.pl (A. Grzelakowska), julia.modrzejewska@dokt.p.lodz.pl (J. Modrzejewska), jolanta.kolinska@p.lodz.pl
(J. Kolińska), marcin.szala@p.lodz.pl (M. Szala), mzielonka@mcw.edu (M. Zielonka), karolina.debowska@p.lodz.pl (K. Dębowska), malgorzata.zaklos-szyda@p.
lodz.pl (M. Zakłos-Szyda), adam.sikora@p.lodz.pl (A. Sikora), jzielonk@mcw.edu (J. Zielonka), radoslaw.podsiadly@p.lodz.pl (R. Podsiadły).

https://doi.org/10.1016/j.freeradbiomed.2021.12.260
Received 23 October 2021; Received in revised form 7 December 2021; Accepted 14 December 2021
Available online 17 December 2021
0891-5849/© 2021 The Author(s). Published by Elsevier Inc. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

involvement of ONOO− in pulmonary hypertension [8], liver and kidney


injury [9–12], and the mechanism of immunosuppressive effects of
myeloid-derived suppressor cells (MDSCs) in cancer [13]. ONOO− can
react directly with thiols, carbon dioxide, heme proteins, iron–sulfur
centers, and zinc thiolates [14,15]. In addition, the products of the
decomposition of ONOO− under physiological conditions include highly
oxidizing and/or nitrating hydroxyl radical (•OH), carbonate radical
anion (CO3•–), and nitrogen dioxide (•NO2), which may directly modify
and damage cell components [10]. The accurate detection of ONOO− in
different cellular and extracellular compartments, and the estimation of
its formation rates, is required to understand how nitric oxide-derived
oxidants affect biological processes, including mitochondrial dysfunc­
tion and cell death. The detection of cellular ONOO− is challenging due
to its short lifetime in biological systems [16] and low steady-state
concentration [17,18]. Initially, indirect methods, such as analysis of
nitrotyrosine residues of modified proteins [19,20] or low-throughput Fig. 1. Boronate probe based on 1-ethyl-3-(4-methylene-7-hydroxy-2H-chro­
electrochemical approaches [21], were used for its detection. Nitrative men-2-one) imidazolium chloride (CI-Bz-BE).
modifications of proteins and lipids have been identified as biomarkers
for reactive nitrogen species in numerous diseases [22,23]. With the the case of ONOO− , the oxidation takes place nearly a million times
constant development and improvement of LC-MS-based methods, the faster than with H2O2 (k(ONOO⁻) ~106 M-1s-1; k(H₂O₂) ~100 M-1s-1) [34].
detection of nitrated products on endogenous molecules have gained Reaction of the boronate probes with ONOO− involves nucleophilic
prominence and helped identify targeted for oxidative/nitrative modi­ addition of the oxidant to the boron atom, followed by nitrite elimina­
fication of cellular components. Tyrosyl moieties in proteins may be tion and deboronation to form a corresponding phenolic product. With a
nitrated, however, via different biochemical pathways including not few exceptions reported, ONOO− -induced oxidation of aryl boronates
only peroxynitrite, but also via heme peroxidase (e.g. myeloperoxidase, (boronic acids, ArB(OH)2, or boronic esters, ArBPin) typically proceeds
eosinophil peroxidase)-catalyzed nitration in the presence of nitrite and via two pathways: the major, nonradical pathway leads to the formation
hydrogen peroxide [24,25]. of corresponding phenol (ArOH), and the minor, radical pathway leads
A more recent approach to determine ONOO− in biological systems is to formation of phenyl radicals and radical-derived minor products
based on employing low-molecular-weight probes that, by a fast reac­ (ArH, ArNO2) [3,30,35–39]. Depending on the compound structure, the
tion with ONOO− , form stable and characteristic products [26,27]. It is contribution of the radical pathway in the oxidation of boronate may
preferable to use fluorescent probes that react with cellular oxidants to range from <1% to 20%. The minor pathway is highly specific for the
form the products, which can be detected using fluorescence-based reaction of boronates with ONOO− . Replacement of the boronate moiety
techniques and allow for sensitive detection of reactive species in bio­ in boronic compounds by the nitro group strongly supports an inter­
logical systems. mediary role of ONOO− and may serve as a unique marker for ONOO−
A fluorescent probe must fulfill many requirements, such as lack of formation in cell-free and cellular systems [34]. In case of bor­
toxicity, analyte specificity, and solubility in aqueous solutions, to be onobenzylated probes, the formation of the fluorescent product is a
used in biological systems [28,29]. Also, the amount of the detected two-step process [30,31,40]. The reaction of oxidant with the boronate
product should be proportional to the amount of the oxidant produced. group in probe leads to the corresponding primary phenol, which is not
Furthermore, the probe should not affect the system being investigated stable at physiological pH and upon spontaneous elimination of quinone
(i.e., it cannot induce reactive oxygen species formation or change the methide (QM) leads to the formation of a final, stable phenolic product
redox status of the cell). Preferably, the probe should be selective toward (Scheme S1).
a single oxidant. In case of a lack of selectivity, different oxidants should Lipophilic cationic agents have been shown to accumulate in cell
form different, oxidant-specific products, easily distinguishable by the mitochondria in response to the mitochondrial membrane potential [41,
detection technique used. It is, therefore, important to establish what 42]. This property was utilized in developing positively charged probes
species may react with the probe under various conditions, and the for mitochondrial membrane potential, but also to target various mo­
relative rates of those reactions. The product(s) of the reaction between lecular probes and bioactive agents to mitochondria [43–45]. The most
the probe and the oxidant should be sufficiently stable to accumulate common strategy to target the chemical agents to mitochondria is their
over the time of the experiment to allow quantitative analyses [26,29]. conjugation to a triphenylphosphonium cation (TPP+). Other organic
Most reported redox probes constructed via masking of hydroxyl or cations, where the positive charge is delocalized over or shielded by
amino groups in fluorophore require the use of an organic solvent to aromatic rings, have also been shown to accumulate in mitochondria
prepare the stock solutions and to improve their solubility in water. [46,47]. Recently, the coumarin imidazolium derivative containing the
However, organic cosolvent may result in cytotoxicity or interfere with 2,4-dinitrobenzenesulfonyl group as a thiol-sensing moiety was pro­
the assays, for example, by scavenging the analyte of interest. It was posed as a mitochondria-targeted two-photon fluorescent probe for the
shown that DMSO may compete with boronate probes for hypochlorous detection of mitochondrial thiols [48]. Upon reaction with thiols, the
acid (HOCl) [30,31]. Therefore, the development of water-soluble yet probe was converted into the hydroxycoumarin imidazolium (CI–OH)
cell-permeable probes would be highly beneficial. product. The fluorescence colocalization measurements confirmed a
Over the last decade, numerous molecular probes have been pro­ high degree of mitochondrial localization of the CI–OH product, with a
posed to measure ONOO− , with different selectivity and sensitivity [20, Pearson’s correlation coefficient of 0.9, similar to a TPP+-linked
32,33]. Among them, boronate-based molecular probes emerge as su­ hydroxycoumarin standard [48]. This provides an opportunity to target
perior to other chemical sensors reported to date for this oxidant coumarin-based sensors to mitochondria by conjugation of the coumarin
detection due to their rapid and direct reaction with ONOO− [26,32,34]. fluorophore with the alkylimidazolium cation via a methylene bridge.
Boronate probes react stoichiometrically with ONOO− to yield the cor­ Due to favorable photophysical and photochemical properties,
responding hydroxylated compound as the major reaction product. including high fluorescence yield and good photostability, the coumarin
Boronates can also be converted to the same major phenolic product by skeleton often is chosen as the basis for the construction of various types
HOCl, hydrogen peroxide (H2O2), and some other biologically relevant of chemosensors, by combing the 2H-chromen-2-one moiety with an
oxidants, but the reaction with ONOO− is kinetically most favorable. In analyte-sensing moiety [49]. One of the best-studied commercially

35
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

available pro-coumarin sensors used for the detection of ONOO− is spectrophotometry, using the extinction coefficient values of 350 M-
1
coumarin-7-boronic acid (CBA) [3,50–53]. cm1 (at 292 nm, in 0.1 M NaOH) and 39.4 M-1cm-1 (at 240 nm, in
Here, we designed and synthesized a new water-soluble florescent water), respectively. Solutions of oxidants were added directly to the
boronate probe based on a coumarin scaffold, bearing a mitochondria- buffered reaction mixtures. All measurements were performed in
targeting imidazolium cationic moiety (CI-Bz-BA, Fig. 1). The chemi­ aqueous solutions of phosphate buffer (100 mM, pH 7.4) containing
cal reactivity of this probe toward selected biological oxidants is re­ diethylenetriaminepentaacetic acid (dtpa, 10 μM).
ported. The minor, yet specific products of the reaction between the In reactions involving HOCl, dtpa was omitted to avoid scavenging of
probe and ONOO− have been identified, and the authentic standards HOCl by dtpa. To investigate the effect of carbon dioxide (CO2) on the
were synthesized (Scheme S2-S3). The results are complemented by a yield of the products of the reaction of CI-Bz-BA with ONOO− , we car­
detailed spectroscopic characterization of the probe and its oxidation/ ried out the reaction in the presence of sodium bicarbonate (NaHCO3).
nitration products, followed by application of the probe to detect ONOO− , HOCl, or H2O2 was rapidly mixed with CI-Bz-BA using a vortex
ONOO− formed in situ in cell-free and cellular systems. mixer (10 s) and incubated for 5 min, 15 min, and 24 h, respectively,
before HPLC and spectroscopic analyses.
2. Materials and methods
2.5. HPLC analyses of the products of the CI-Bz-BA reaction with
2.1. Materials oxidants in cell-free system

Myeloperoxidase (MPO) and 3-morpholinosydnonimine hydrochlo­ HPLC analyses of CI-Bz-BA and its oxidation products (CI–OH, CI-
ride (SIN-1) were from Calbiochem (San Diego, CA). Catalase (CAT, Bz-H, CI-Bz-NO2) were performed using a Shimadzu instrument (Kyoto,
from bovine liver) and superoxide dismutase (SOD, from bovine eryth­ Japan) equipped with absorption and fluorescence detectors. The sam­
rocytes) were obtained from Sigma Aldrich (St. Louis, MO). ONOO− was ples (10 μl) were separated using a reverse-phase column (Phenomenex,
prepared as reported previously [54] and stored at − 80 ◦ C. H2O2 and Kinetex Biphenyl, 50 mm × 4.6 mm, 2.6 μm) equilibrated 0.1% (v/v)
HOCl were from Merck (Kenilworth, NJ). Other reagents and starting trifluoroacetic acid aqueous solution containing 5% (v/v) MeCN.
materials for chemical syntheses were purchased from commercial Gradient elution was performed at a flow rate of 1.5 mL/min with
vendors and used without further purification. ultraviolet–visible absorption detection (λ = 254 nm and 330 nm). The
compounds were eluted by raising MeCN concentration (v/v) from 5%
2.2. Synthesis to 55% over 5.5 min followed by an increase to 100% from 5.5 to 6 min.
Under these conditions, CI–OH shows a retention time of 2.50 min, CI-
Detailed synthetic procedures for all compounds shown are pre­ Bz-BA of 3.65 min, CI-Bz-H of 4.75 min, and CI-Bz-NO2 of 4.80 min.
sented in the Supplementary Information.
2.6. Oxidation of CI-Bz-BA induced by ONOO− produced from SIN-1
2.3. Spectroscopic study
SIN-1 slowly decomposed in aqueous solution at pH 7.4 producing
The absorption spectra were measured using a Jasco V–670 both •NO and O2•–, which react in a diffusion-limited reaction, yielding
UV–vis–NIR spectrophotometer (Jasco, Japan). The steady-state emis­ ONOO− [56–58]. SIN-1 stock solution (25 mM) was prepared in 50 mM
sion spectra, fluorescence excitation/emission matrix spectra, and HCl.
fluorescence lifetimes were recorded by a FLS-920 spectrofluorometer SIN-1 (50 μM) was added to a solution of CI-Bz-BA (10 μM) in a
(Edinburgh Instruments, Livingston, UK). Fluorescence spectra were phosphate buffer (pH 7.4, 100 mM) containing dtpa (10 μM), and the
taken with a slit width of 0.5 nm. A picosecond-pulsed light-emitting changes of fluorescence intensity during the generation of •NO and O2•–
diode (EPLED 340, Edinburgh Instruments) was used as an excitation obtained from the decomposition of SIN-1 in the presence or absence of
light source in the time-resolved measurements. The diode had a peak CAT (100 U/mL) and/or SOD (0.1 mg/mL) were monitored for 1 h in the
wavelength at 341.2 nm and a pulse width of 835.7 ps. The excitation kinetic mode.
and emission bandpasses were 2.0 nm. The instrument profile was ob­ The HPLC analysis was used to monitor the decay of SIN-1 and CI-Bz-
tained by replacing the sample with Ludox as a scatter. The data were BA, and to identify the products formed. To initiate the reaction, a SIN-1
analyzed by a least-squares reconvolution procedure using the software solution (100 μM) was combined with a phosphate buffer solution (100
package provided by the Edinburgh Instruments. Fluorescence lifetimes mM, pH 7.4) containing dtpa (10 μM) and CI-Bz-BA (200 μM). The
were calculated from intensity decay analyses [39]. Fluorescence analysis was performed using a Shimadzu instrument equipped with an
quantum yields were determined by the relative method using the absorption detector, as described above. Changes in the peak area were
procedure described previously [40]. Quinine sulfate was used as a used to determine the progress of the reaction over the time.
reference standard [41,42].
2.7. Oxidation of CI-Bz-BA induced by H2O2 in the presence of MPO and
2.4. Preparation of solutions NaNO2

For all experiments, stock solutions of CI-Bz-BE, CI–OH, CI-Bz-H, MPO stock solution (5 μM) was prepared in a phosphate buffer (100
and CI-Bz-NO2 were prepared by dissolving the solid compounds in mM, pH 7.4), aliquoted, and stored at − 80 ◦ C. The stock solutions of
DMSO (10 mM) except for the measurements involving HOCl, for which H2O2 (100 mM) and nitrite (NaNO2, 1 M) were prepared in distilled
the CI-Bz-BE probe was dissolved in water (1 mM). Although all pre­ water. CI-Bz-BA (100 μM) was incubated with MPO (0–15 nM), H2O2
pared compounds are soluble in water, in accordance with the recom­ (500 μM), and NaNO2 (5 mM) in a phosphate buffer (100 mM, pH 7.4)
mendations [55], the concentrated stock solutions were prepared in containing dtpa (10 μM) at room temperature for 60 min. Reactions
DMSO to avoid hydrolysis of the probe in aqueous media when stored were terminated by adding CAT (100 U/mL), and the products were
for extended periods of time. immediately analyzed by HPLC, as described above.
The working solutions of ONOO− , HOCl, and H2O2 were freshly
prepared before each experiment and kept on ice. The concentration 2.8. Kinetic studies
ONOO− was determined by spectrophotometry, using the extinction
coefficient value 1.7 × 103 M-1cm-1 (at 302 nm, in 0.1 M NaOH). The The determination of the rate constants of CI-Bz-BA with the
concentrations of HOCl and H2O2 were also determined by different oxidants was carried out under pseudo-first-order conditions,

36
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

using an excess of either the probe (in the case of ONOO− and HOCl) or formic acid. The compounds were eluted by increasing the MeCN con­
the oxidant (in the case of H2O2) as described in the Supplementary centration in the mobile phase from 5% to 100% over 6 min. The flow
Information. rate was set at 0.5 mL/min, and the flow was diverted to waste during
The rate constant for reaction of CI-Bz-BA with ONOO− was also the first 1.5 min and after 6.5 min, counting from the time of injection.
determined via competition kinetics assays utilizing boronate probes CI–OH (retention time 2.1 min), CI-Bz-BE (retention time 5.3 min), CI-
CBA and peroxy crimson-1 (PC1) as the competitors, using the reported Bz-BA (retention time 3.4 min), CI-Bz-H (retention time 4.5 min), and
rate constants of 1.1 × 106 M-1s-1 [3] and 9.9 × 105 M-1s-1 [40], CI-Bz-NO2 (retention time 4.55 min) were detected as positive ions
respectively. The details of the competition kinetics analyses are using the multiple reaction monitoring mode, with the primary/frag­
described in the Supplementary Information. ment ion pairs 487.2 > 270.05, 405.1 > 270.1, 271.0 > 147.05, 406.0 >
270.0, and 361.00 > 270.0.
2.9. Cell culture experiments
2.11. Two-photon fluorescence imaging
RAW 264.7 cells (ATCC, USA) were maintained in Dulbecco’s
Modified Eagle’s Medium (Gibco, USA) containing 10% (v/v) fetal For two-photon fluorescence microscopy, cells were seeded over­
bovine serum, L-glutamine (2 mM), penicillin (100 U/ml), and strepto­ night into 35 mm glass bottom dishes (MatTek Corporation, Ashland,
mycin (0.1 mg/ml), and grown at 37 ◦ C in a 5% carbon dioxide MA), followed by activation to produce ONOO− , as described above.
atmosphere. Cells were imaged using a Carl Zeiss LSM510 META detection system
To determine a cytotoxic potential of CI-Bz-BA and CI–OH, cells (Jena, Germany). Images were collected with a 40 × objective and
were incubated with the compounds for 4 h or 24 h before measure­ image acquisition software. Two-photon excitation properties of the
ments of cell viability. Two separate assays, the MTT and PrestoBlue, CI–OH fluorophore were characterized previously [48]. However,
were used to determine cell metabolic activity. before imaging ONOO− production with the CI-Bz-BE probe, the
A stock solution of the CI-Bz-BE probe (10 mM in DMSO) was added two-photon illumination parameters were optimized using cells loaded
directly to the cell media to obtain a 20 μM concentration, resulting in with the fluorescent product, CI–OH, and the excitation wavelength of
0.2% (v/v) final concentration of DMSO. 735 nm; emission light in the range 465–550 nm was selected.
Oxidation of CI-Bz-BA was monitored in a 96-well fluorescence plate
reader. The cells were seeded in black 96-well plates at a concentration 3. Results and discussion
of 3 × 104 cells per well (in 150 μL total volume) and incubated over­
night at 37 ◦ C in a 5% carbon dioxide atmosphere. To induce iNOS 3.1. Synthesis of coumarins
expression and •NO production, cells were incubated for 8 h with
interferon γ (IFN-γ, 50 U/ml) and lipopolysaccharide (LPS, 0.5 μg/ml) in The CI-Bz-BE probe and the anticipated major and minor products of
Dulbecco’s Modified Eagle’s Medium. Subsequently, the medium was its reaction with ONOO− were obtained via two simple steps presented
replaced by the DPBS buffer supplemented with sodium pyruvate (0.33 in Scheme S2. Detailed synthetic protocols are included in the Supple­
mM) and glucose (5.56 mM, DPBS-GP, Gibco, USA), and O2•– production mentary Information. The compounds were obtained with good yields,
was stimulated by the addition of phorbol 12-myristate 13-acetate ranging from 73% to 91%. NMR spectroscopy and mass spectrometry
(PMA, 1 μM). Concomitant stimulation of •NO and O2•– production re­ studies confirmed the chemical structures of the compounds (See
sults in ONOO− generation by RAW 264.7 cells, as described previously Experimental section in Supplementary Information and Figs. S1–S8.)
[35,53]. At the time of the addition of PMA, the CI-Bz-BA probe (20 μM)
also was added. Where indicated, L-N-nitroarginine methyl ester hy­ 3.2. Spectroscopic characterization of the probe and its oxidation/
drochloride (L-NAME, 1 mM, Cayman Chemical, USA) or CAT (1000 nitration products
U/mL) were added before stimulation with PMA. Immediately after the
addition of DPBS-GP containing the probe with or without PMA, The spectroscopic properties of coumarins are presented in Table S1.
L-NAME, and CAT, the plate was placed in a plate reader prewarmed to UV–vis absorption and fluorescence spectra at pH of 7.4 were charac­
37 ◦ C. Total fluorescence intensities were acquired using a FLUOstar terized for all products (Figs. S12-S13). In aqueous solutions, CI-Bz-BE
Omega plate reader (BMG LABTECH, Germany) equipped with the (pinacolate ester) undergoes hydrolysis to the boronic acid form (CI-Bz-
appropriate excitation and emission filters (λex = 355 nm, λem = 520 BA). Thus, the spectroscopic properties in an aqueous solution pertain to
nm). The instrument was kept at 37 ◦ C during the measurements, and CI-Bz-BA.
fluorescence intensity was read from the bottom of each well. Changes in Both the probe and the anticipated products carrying the coumarin
the fluorescence intensity were monitored for 2 h in the kinetic mode. imidazolium scaffold show absorption bands in the ultraviolet region
located at 323–330 nm. The presence of benzyl groups does not cause
2.10. LC-MS analyses of extraction of CI-Bz-BA oxidation products any significant change in the absorption spectra as compared with
CI–OH. The fluorescence spectra of compounds CI–OH, CI-Bz-BA, and
To identify the oxidation products of CI-Bz-BA formed in cells by CI-Bz-H show one fluorescence maximum, while compound CI-Bz-NO2
endogenously generated ONOO− , RAW 264.7 macrophages were seeded possess two maxima. For all compounds, the fluorescence emission
in 10-cm culture dishes (1.5 × 106 cells/dish). Cells were stimulated to maximum is located in the visible spectral range. The incorporation of
produce ONOO− using same treatments as in the plate reader experi­ benzyl groups causes a blue shift of the emission wavelength (λem) and a
ments. After the addition of CI-Bz-BA (20 μM) in DPBS-GP, dishes with decrease in fluorescence lifetime when compared with CI–OH. The
cells were incubated for 1 h at 37 ◦ C in a carbon dioxide-free incubator. strong electron-withdrawing substituent –NO2 (compound CI-Bz-NO2)
The oxidation products of CI-Bz-BA by endogenously generated ONOO− largely decreases the fluorescence intensity, compared with the unsub­
were extracted according to a slightly modified version of a published stituted system CI-Bz-H and the CI-Bz-BA probe. As shown in Fig. S12,
procedures [35,55,59]. both the probe and its main oxidation product exhibit fluorescence. A
Coumarin-based products were analyzed by liquid chromatogra­ comparison between the fluorescence emission spectra (Fig. 12B) or the
phy–mass spectrometry (LC-MS) performed using a Shimadzu LC-MS fluorescence excitation-emission matrix spectra (Fig. S14) of CI-Bz-BA
8030 triple quadrupole mass detector coupled to a Shimadzu Nexera 2 and CI–OH indicates clear distinction of the emission bands, suggesting
ultra-high-performance liquid chromatography system. The samples that the fluorescence of the probe should not interfere with the fluo­
were injected on a Raptor Biphenyl column (Restek, 100 mm × 2.1 mm, rescence of its oxidation product. The introduction of the imidazolium
2.7 μm) equilibrated with 10% of MeCN in water containing 0.1% of moiety to the coumarin skeleton results in a red shift of the emission

37
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

comparison with 7-hydroxycoumarin (COH) (Table S1, Fig. S15).

3.3. Chemical reactivity of CI-Bz-BA probe toward selected biological


oxidants

The reaction of CI-Bz-BA with ONOO− leads to the batochromic shift


of the fluorescence of the solution, with the spectrum of the product
resembling that of authentic CI–OH (Fig. 2). A single isosbestic point
was observed at 450 nm for this oxidative conversion. There is a distinct
linear increase in fluorescence intensity at 481 nm and depletion of the
intensity at 405 nm with successive addition of ONOO− at doses up to
10 μM, followed by complete disappearance of the band due to the probe
and saturation of the intensity due to the product. This suggests a 1:1
reaction stoichiometry between the probe and the oxidant.
Next, the HPLC analyses were performed to for the identification and
quantitative analyses of the reaction products formed during the reac­
tion between the CI-Bz-BA probe and ONOO− . Both product identifi­
cation and quantitative analyzes were done using independently
synthesized standards. HPLC analyses confirmed that CI–OH is the
Fig. 2. The fluorescence spectra changes of CI-Bz-BA (10 μM) after incubation
major product of the reaction of CI-Bz-BA with ONOO− (Fig. 3A), while
in the presence of ONOO− (0–30 μM) in a phosphate buffer (100 mM, pH 7.4 in
additional, minor products (e.g., CI-Bz-NO2) are also formed, even when
the presence of 10 μM dtpa). Inset: Fluorescence intensity measured at 405 nm
and 481 nm for a solution of CI-Bz-BA (10 μM) in a phosphate buffer (100 mM,
the probe was present in excess of the oxidant (Fig. 3C). Titration of the
pH 7.4 in the presence of 10 μM dtpa) after the addition of ONOO− (0–30 μM). probe with ONOO− results in the product profiles shown in Fig. 3B and
Data are means ± standard deviation of three independent experiments. The D, indicating the formation of both phenolic (major) and radical-
emission spectra (excitation at 330 nm; ex/em slit: 0.5 nm) were collected after mediated products (minor) in a dose-dependent manner. The maximal
5 min incubation of CI-Bz-BA with the oxidant. yield of the CI–OH product was close to 92%, and a slight excess of the
oxidant was required for complete consumption of the CI-Bz-BA probe.
band with a slight decrease in the fluorescence quantum yield in These data agree with the fluorescence measurements conducted under
the same experimental conditions (Fig. 2). The HPLC analyses revealed

Fig. 3. Oxidation of the CI-Bz-BA probe by ONOO− . (A, C) The HPLC chromatograms of the incubation mixtures containing CI-Bz-BA (100 μM) and ONOO− (0, 80,
and 200 μM) in a phosphate buffer (100 mM, pH 7.4 in the presence of 10 μM dtpa). (B, D) HPLC-based titration of CI-Bz-BA (100 μM) with ONOO− (0–300 μM) in a
phosphate buffer (100 mM, pH 7.4 in the presence of 10 μM dtpa). Data are means ± standard deviation of three independent experiments. The HPLC traces were
collected after 5 min incubation of CI-Bz-BA with oxidant using the absorption detector set at 330 nm.

38
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

Scheme 1. The major and minor products of the reaction of CI-Bz-BA and ONOO− .

Fig. 4. Kinetics of the reaction of CI-Bz-BA with ONOO− . (A) Kinetic traces of CI-Bz-BA fluorescence decay observed at 405 nm. (B) The dependence of pseudo-first-
order rate constants of the reaction of Cl-Bz-BA probe with ONOO− on the probe concentration. Solutions consisted of the CI-Bz-BA probe (25–50 μM), phosphate
buffer (0.1 M, pH 7.4 in the presence of 200 μM dtpa), and ONOO− (5 μM). Data are means ± standard deviation of three independent experiments.

the build-up of two partially overlapping peaks with retention times of produce CI–OH as the stable end product (Scheme 1). QM-NO2 is ex­
4.75 and 4.80 min, corresponding to CI-Bz-H and CI-Bz-NO2, respec­ pected to react with water to form 4-hydroxy-3-nitrobenzyl alcohol
tively, the two products anticipated for the radical pathway of the re­ (OH-Bz-3NO2–OH). However, OH-Bz-3NO2–OH may be also formed as
action. We expected that the •NO2 radical formed in the minor pathway a product of the nitration of 4-hydroxybenzyl alcohol (OH-Bz-OH),
and/or during decomposition of excess ONOO− would lead to nitration which is formed upon elimination of QM by CI-Bz-OH. As shown in
of both the primary (CI-Bz-OH) and possibly the secondary (CI–OH) Fig. 3C and D, with an increase in ONOO− concentration, especially at or
phenolic products, depending on the time required for CI-Bz-OH to above the concentration of the probe, a new product eluting at 2.81 min
eliminate QM and form CI–OH. The nitrated CI–OH standard was not appeared in the HPLC traces. The retention time of this product was
prepared; however, after the reaction of CI-Bz-BA with ONOO− , we identical to that of the independently synthesized standard (Scheme S3),
detected in the reaction mixture the same product as obtained when OH-Bz-3NO2–OH. The same product was also formed when OH-Bz-OH
CI–OH was mixed with ONOO− (Fig. S16). This confirms that the was mixed with ONOO− (Fig. S17).
elimination of the QM moiety is fast enough to produce CI–OH, when The rate constant of the reaction between CI-Bz-BA and ONOO− was
the •NO2 radical is still present in the solution. determined using the stopped-flow technique due to its expected high
We hypothesized that nitration of the primary phenolic product CI- value and the instability of ONOO− at pH 7.4. Because the fluorescent
Bz-OH by •NO2 would produce CI-Bz-OH-NO2, which after the elimi­ product CI–OH is formed via a multi-step process involving the self-
nation of QM-NO2 through a self-immolative reaction mechanism would immolation of QM, the pseudo-first-order rate constants were

39
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

M-1s-1 and 5.2 × 105 M-1s-1, obtained from the competition assays using
CBA and PC1 competitors, respectively, are ca. two-fold lower than the
value from the stopped-flow experiment, which may be due to the
above-mentioned limitations of the stopped-flow experiments and/or
other factors. Nevertheless, the determined values of the rate constant
are close to the values reported for other boronate probes and indicate
the ability of CI-Bz-BA to rapidly intercept and detect ONOO− .
The oxidative conversion of the CI-Bz-BA probe by ONOO− was also
investigated in the system co-generating •NO and O2•‒. To this end, the
CI-Bz-BA probe was incubated with SIN-1 in a phosphate buffer (pH 7.4)
at room temperature. Fig. 5 depicts the changes of fluorescence intensity
of CI-Bz-BA at 481 nm over time in the presence and absence of SIN-1
and/or CAT and SOD. The addition of SIN-1 to the solution of CI-Bz-
BA at pH 7.4 resulted in a strong and time-dependent increase in fluo­
rescent intensity, which was partially inhibitable by SOD, but not by
CAT. This suggests that ONOO− formed in situ from •NO and O2•‒ is
responsible for oxidation of CI-Bz-BA to the corresponding fluorescent
product.
To identify and quantify the products formed during the incubation
of CI-Bz-BA and SIN-1, the reaction mixtures were analyzed by HPLC. As
Fig. 5. Changes of fluorescence intensity of an aqueous solution of CI-Bz-BA
(10 μM) due to co-generation of •NO and O2•– from the decomposition of SIN-1
shown in Fig. S18, both the major product, CI–OH, and the minor
(50 μM) in the presence or absence of CAT (100 U/mL) and/or SOD (0.1 mg/ products, CI-Bz-H and CI-Bz-NO2, were detected, and both CI–OH and
mL) in a phosphate buffer (100 mM, pH 7.4) containing dtpa (10 μM). Inset: CI-Bz-NO2 increased almost linearly with the time of incubation within
The rates of increase in the fluorescence signal intensity from incubations the 4-h period. These results further confirm that ONOO− , whether
contained CI-Bz-BA and SIN-1 in the presence or absence of CAT and/or SOD. added as a bolus or generated in situ, can be detected using the CI-Bz-BA
The solutions containing CI-Bz-BA were excited at 330 nm, and the emitted probe. Furthermore, in addition to the major and fluorescent product,
light intensity was measured at 481 nm (excitation/emission slit = 0.5 nm). CI–OH, the reaction results in the formation of ONOO− -specific nitrated
Data are means ± standard deviation of three independent experiments. product, CI-Bz-NO2.
As ONOO− − dependent reactions are often altered in the presence of
obtained by fitting the decay of the fluorescence signal recorded at 405 CO2 [61–63], we investigated the effects of varying the levels of HCO3−
nm, following the consumption of the probe (Fig. 4). From the slopes of on the extent of oxidation of the CI-Bz-BA probe. With increasing HCO3−
the plots of the observed first-order rate constants versus the concen­ concentration, the yield of the CI–OH and CI-Bz-NO2 products
tration of the reactant, the second-order rate constant was determined to decreased (Fig. S19), suggesting that the probe competes with CO2 for
be equal (1.3 ± 0.1) × 106 M-1s-1. This value is close to the values re­ ONOO− . This is in contract with many other probes reported for
ported for other boronate-based probes [30,31,34–36,40,60]. The lim­ ONOO− , which detect ONOO− -derived radicals rather than ONOO− it­
itation of the stopped-flow-based monitoring the product consumption self [64]. Moreover, the results obtained indicate that, even in the
was the relatively small difference in the fluorescence intensity before presence of physiological concentrations of HCO3− , the CI-Bz-BA probe
and after the reaction. This is because the probe was used in excess of can still intercept ONOO− - to produce the corresponding phenol,
ONOO− and possibly due to smaller differences in fluorescence intensity CI–OH, as well as the ONOO− -specific nitrated product, CI-Bz-NO2.
between the probe and the primary phenol, as compared with CI–OH. Boronate probes have been originally proposed and utilized for the
The rate constant of the reaction of CI-Bz-BA with ONOO− was also detection of H2O2 and later shown to be also able to report HOCl [3,30,
determined using the competition kinetic approach (Fig. S22). The 31,34–37,60,65–67]. Therefore, the reactivity of the CI-Bz-BA probe
coumarin boronate (CBA) and resorufin boronate (PC1) derivatives toward these oxidants was investigated using fluorescence spectroscopy
were used as competitors in kinetic experiments. The values, 5.4 × 105 and HPLC analyses.

Fig. 6. The fluorescence spectra changes of CI-Bz-BA (10 μM) after incubation in the presence of (A) H2O2 and (B) HOCl (0–30 μM) in a phosphate buffer (100 mM,
pH 7.4). The emission spectra (excitation at 330 nm; ex/em slit: 0.5 nm) were collected after 24 h and 15 min incubation of CI-Bz-BA with H2O2 and HOCl,
respectively. Insets: Data are means ± standard deviation of three independent experiments.

40
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

Fig. 7. Oxidation of CI-Bz-BA by H2O2. (A–D) HPLC analyses of the products formed from the reaction of the CI-Bz-BA probe and H2O2. Incubation mixtures
consisting of CI-Bz-BA (100 μM) and H2O2 [(A,B) 10 mM and (C) 0, 80, and 200 μM] in a phosphate buffer (100 mM, pH 7.4) containing dtpa (10 μM). The probe and
its oxidation products were detected using an absorption detector at (A, C) 330 nm and (B) 254 nm. (D) HPLC-based titration of the CI-Bz-BA probe (100 μM) with
H2O2 (0–300 μM) in a phosphate buffer (100 mM, pH 7.4) containing dtpa (10 μM). Data are means ± standard deviation of three independent experiments. The
HPLC traces were collected after 24-h incubation of CI-Bz-BA with oxidant using an absorption detector set at 330 nm.

Fig. 8. Oxidation of CI-Bz-BA by HOCl. (A,B) HPLC analyses of the products formed from the reaction of the CI-Bz-BA probe and HOCl. Incubation mixtures
consisting of CI-Bz-BA (100 μM) and HOCl (0, 80, and 200 μM) in a phosphate buffer (100 mM, pH 7.4) without dtpa. The probe and its oxidation products were
detected using an absorption detector at 330 nm. (D) HPLC-based titration of the CI-Bz-BA probe (100 μM) with HOCl (0–300 μM) in a phosphate buffer (100 mM, pH
7.4) without dtpa. Data are means ± standard deviation of three independent experiments. The HPLC traces were collected after 15-min incubation of CI-Bz-BA with
oxidant using an absorption detector set at 330 nm.

41
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

boronate-based redox probes [3,31,34–36,40,60].


Time-dependent HPLC-based analyses of the reaction mixtures of CI-
Bz-BA and H2O2 indicate that before CI–OH is formed, elution of a peak
at 3.85 min is detected. Based on our previous report [30], we assigned
this peak to the primary phenolic product (CI-Bz-OH) formed during the
oxidation of CI-Bz-BA. Decomposition of CI-Bz-OH into CI–OH is
accompanied by elimination of QM, which reacts with water to form
OH-Bz-OH. This product was detected at 1.70 min, coeluting with the
authentic standard of OH-Bz-OH (Fig. 7). Because no other peaks were
detected, it can be concluded that upon oxidation by H2O2, CI-Bz-BA
undergoes conversion to CI–OH as the sole product of the reaction. To
determine the reaction stoichiometry, CI-Bz-BA was incubated with
different concentrations of H2O2 and the HPLC analyses were performed
after 24-h incubation. The results indicate that the reaction between
CI-Bz-BA and H2O2 is stoichiometric, as the yield of the probe con­
sumption and product formation is close to 100% with respect to the
amount of H2O2. Moreover, we did not observe degradation of the
CI-Bz-BA-derived product, CI–OH, over the course of experimentation,
even after 24-h incubation with an excess of the oxidant.
HPLC analyses of the reaction mixtures of the CI-Bz-BA probe with
HOCl indicate that the maximum yield of CI–OH reached ca. 70% under
the experimental conditions used (Fig. 8). The amount of CI–OH
decreased when an excess of the oxidant was added. The disappearance
of CI–OH in a reaction mixture can be attributed to the chlorination
reaction leading to the formation of chlorinated derivative(s). This is
supported by the detection of a new HPLC peak of the same retention
time as the product formed when CI–OH was mixed with HOCl
(Fig. S20). 3-Chloro-4-hydroxybenzyl alcohol (OH-Bz-3Cl–OH), the
product of hypothetical chlorination of primary phenolic product, CI-
Bz-OH, which is formed from of the oxidation of CI-Bz-BA and/or
chlorination of OH-Bz-OH, which is formed by the decomposition of CI-
Bz-OH, was not detected under the experimental conditions used
(Fig. S21).

3.4. Oxidation of CI-Bz-BA by the MPO/H2O2 system

One of the limitations of the common assays to detect ONOO− is the


inability to distinguish between ONOO− and MPO/H2O2/NO2− -medi­
ated oxidation/nitration. To investigate whether CI-Bz-BA can differ­
entiate between ONOO− and MPO/H2O2/NO2− , the reaction mixtures
Fig. 9. Comparison of the products formed from the oxidation of CI-Bz-BA by were analyzed by HPLC (Fig. 9). Although the major product, CI–OH,
ONOO− and MPO. (A) HPLC traces of the products detected during the reaction was formed in all the systems tested, due to the presence of ONOO− or
between the CI-Bz-BA probe (100 μM) and ONOO− (200 μM), H2O2 (500 μM), H2O2, this experiment clearly identified CI-Bz-H and CI-Bz-NO2 as
H2O2 (500 μM) + MPO (10 nM), H2O2 (500 μM) + NaNO2 (5 mM), and H2O2 specific products for ONOO− . The formation of ONOO− − specific minor
(500 μM) + NaNO2 (5 mM) + MPO (10 nM) in a phosphate buffer (100 mM, pH products occurred in the presence of ONOO− but not in the presence of
7.4) with dtpa (10 μM). The products were determined 1 h after adding H2O2 at MPO/H2O2/NO2− , which is consistent with the data reported previously
330 nm. (B) Results of titration of CI-Bz-BA (100 μM) with MPO. CI-Bz-BA was
for the ortho-mito-phenylboronic acid probe [68]. On the other hand,
incubated with MPO (0–15 nM) in the presence of H2O2 (500 μM) and NaNO2
MPO in the presence of nitrite leads to nitration of the primary phenolic
(5 mM) in a phosphate buffer (100 mM, pH 7.4) with dtpa (10 μM) for 1 h. Data
are means ± standard deviation of three independent experiments.
product CI-Bz-OH, formed from of oxidation of CI-Bz-BA by H2O2
and/or to nitration of OH-Bz-OH, which is formed by decomposition of
CI-Bz-OH, as evidenced by the formation of OH–3NO2-Bz-OH. Incu­
Similar to the oxidation by ONOO− , the reactions with H2O2 and
bation of OH-Bz-OH with the MPO/H2O2/NO2− system leads to the
HOCl lead to a shift in the fluorescence spectrum, with the product
formation of the same nitration product (Fig. S17).
exhibiting similar spectrum to that of CI–OH (Fig. 6). Analyses of the
fluorescence spectra of the reaction mixtures of CI-Bz-BA with HOCl
3.5. Application of the probe to detect cellular ONOO−
point to a lower yield of the fluorescent product(s) compared with those
for ONOO− and H2O2, and the consumption of the fluorescent product in
Before application for ONOO− detection, the effect of CI-Bz-BA
the presence of excess HOCl.
probe and the CI–OH oxidation product on cell viability was tested with
The reaction kinetics for CI-Bz-BA oxidation by HOCl and H2O2 was
the use of the MTT and PrestoBlue assays. RAW 264.7 cells were treated
investigated under pseudo-first-order conditions, using an excess of
with various concentrations (0–200 μM) of the studied compounds for 4
either the probe in the case of HOCl or of the oxidant in the case of H2O2,
or 24 h before the viability assessment (Fig. S25). CI-Bz-BA and CI–OH
and monitoring the decay of the fluorescence signal due to the probe
did not affect cell viability in the concentration range of 0–100 μM
(Figs. S23 and 24). The rate constants for the reactions between CI-Bz-
whether measured after 4 or 24 h of incubation.
BA and H2O2 or HOCl were determined to be 1.9 ± 0.1 and (1.4 ± 0.1) ×
Next, the probe was tested in RAW 264.7 cells activated to produce
104 M-1s-1, respectively, significantly lower than observed for ONOO− . •
NO and/or O2•‒. Stimulation of RAW 264.7 cells with LPS/IFN-γ and
These values are, however, in the same range as those reported for other
PMA leads to the formation of ONOO− [53]. As shown in Fig. 10A,

42
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

Fig. 10. Real-time monitoring of CI-Bz-BA oxidation by activated RAW 264.7 macrophages. (A) The increase in fluorescence signal intensity from incubations
containing RAW 264.7 macrophages activated by different stimulators as shown. (B) Rate of increase in the fluorescence signal intensity from RAW 264.7 mac­
rophages activated by different stimulators in the absence or in the presence of L-NAME and CAT. Incubations contained CI-Bz-BA (20 μM) and RAW 264.7 mac­
rophages in DBPS-GP buffer in the presence of different stimulators (λex = 355 nm, λem = 520 nm). Data are means ± standard deviation of three independent
experiments.

probe can be used to selectively monitor ONOO− production by mac­


rophages activated to co-generate •NO and O2•‒.
Since the two-photon excitation properties of CI–OH fluorophore
have been characterized previously [48], two-photon fluorescence mi­
croscopy was applied to image probe oxidation in activated macro­
phages. As shown in Fig. 11, upon excitation at 735 nm,
LPS/IFN-γ/PMA-stimulated cells have shown significant two-photon
fluorescence response (465–550 nm) in comparison with the control,
non-stimulated cells. Based on the published report confirming a high
degree of mitochondrial localization of the CI–OH product [48], we
propose that the punctate fluorescence observed in Fig. 12 may be due to
the presence of the CI–OH product in cell mitochondria. Further studies
are needed, however, to determine whether the probe detected mito­
chondrial ONOO− or the product formed in different site(s) translocated
to mitochondria, in the investigated system.
Next, the products of CI-Bz-BA oxidation were analyzed by LC-MS in
cell lysates and cell culture media. As shown in Fig. 12, incubation of
macrophages activated to produce •NO and O2•‒ in the presence of CI-
Bz-BA resulted in the formation of CI–OH and ONOO− -specific product
CI-Bz-NO2. These products were detected both intracellularly and in the
cell media, and were inhibitable by preincubation of the cells with L-
NAME.
Fig. 11. Representative two-photon microscopy images of RAW 264.7 macro­
These results confirm that CI-Bz-BA is a suitable probe for detecting
phages incubated with CI-Bz-BA (20 μM) (control, left panel); stimulated with
LPS (1 μg/mL), IFN-γ (50 U/mL), and PMA (1 μM); and then incubated with CI- ONOO− generated in cellular systems and that CI-Bz-NO2 may be used
Bz-BA (20 μM) (right panel). The incubation time was 90 min. The images were as an ONOO− -specific marker to confirm the identity of the oxidant.
obtained using 735 nm as the excitation wavelength and 465–550 nm emission
windows, at 40 × magnification. (DIC, differential interference contrast). 4. Conclusions

stimulation of RAW 264.7 macrophages with LPS/IFN-γ and PMA in the We report the development of a novel fluorescent probe, CI-Bz-BA,
presence of the CI-Bz-BA probe led to a strong and time-dependent in­ which shows a kinetic preference for ONOO− in cell-free and cell-based
crease in fluorescence intensity. The addition of LPS and IFN-γ or PMA systems. The probe was synthesized with a simple two-step methodology
alone to incubations containing macrophages and CI-Bz-BA caused no under microwave conditions, and provided a good overall yield and high
increase in fluorescence as compared with the control cells. This sug­ purity. The unequivocal advantage of using the CI-Bz-BA probe is its
gests that ONOO− formed from co-generated •NO and O2•‒ was solubility in water, and for this reason it is not necessary to use an
responsible for probe oxidation. To confirm the identity of the species organic cosolvent that can interfere with the oxidants. The chemical
responsible for CI-Bz-BA oxidation, the effects of L-NAME, as a nitric reactivity of the CI-Bz-BA probe was studied in detail, including its re­
oxide synthase inhibitor, and CAT, an enzymatic H2O2 scavenger, on the action kinetics and reactivity toward various oxidants (ONOO− , H2O2,
rate of probe oxidation were tested. While L-NAME caused a significant HOCl). The three oxidants evaluated are able to oxidize CI-Bz-BA to
decrease of fluorescence derived from the probe oxidation, CAT had no corresponding phenolic product CI–OH, resulting in a red shift of its
effect on the rate of probe oxidation, indicating that ONOO− rather than fluorescence emission band. However, the rate constant of the reaction
H2O2 was the oxidant detected. The results indicate that the CI-Bz-BA with ONOO− is more than 5 and 2 orders of magnitude higher than that
of H2O2 and HOCl, respectively. The profile of the products formed from

43
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

Fig. 12. LC-MS analyses of products formed from the oxidation of CI-Bz-BA activated to produce ONOO− RAW 264.7 macrophages. RAW 264.7 macrophages were
activated using LPS (0.5 μg/mL), IFN-γ (50 U/mL), and PMA (1 μM), and incubated for 1 h with CI-Bz-BA (20 μM) in DPBS supplemented with glucose (5.56 mM) and
sodium pyruvate (0.33 mM), as described in the Experimental section. LC-MS traces of CI-Bz-BA, CI–OH, and CI-Bz-NO2 in the (A) cell lysates and (B) media.

the reaction between ONOO− and the CI-Bz-BA probe is highly specific ONOO− formation in a variety of biologically relevant systems. Based on
for this oxidant. The major pathway leads to the formation of CI–OH, the presence of the positive charge and published report [48], we
whereas under the conditions used the minor pathway yields two anticipate that the applications of the probe may be also extended for the
products: CI-Bz-H and CI-Bz-NO2. Reaction of the probe and ONOO− , detection of mitochondrial oxidants, including ONOO− .
added as bolus or generated in situ from superoxide and nitric oxide
fluxes, led to the formation of a characteristic nitrated product CI-Bz- Author contributions
NO2. CI-Bz-BA competes with CO2 for ONOO− and probe oxidation
products were observed under biologically-relevant levels of CO2. These A.G. wrote the draft of the manuscript. All authors contributed to and
data underscore the usefulness of the probe for ONOO− detection in gave approval to the final version of the manuscript.
biological systems. Since the nitrobenzene-type product is not generated
during incubation of boronate probe with MPO/H2O2/nitrite systems,
Declaration of competing interest
its detection demonstrates the presence of ONOO− . On the other hand,
nitration of the primary phenolic product, CI-Bz-OH, which is formed
The authors declare no conflict of interest.
from of oxidation of CI-Bz-BA and/or nitration of OH-Bz-OH, which is
formed by the decomposition of CI-Bz-OH and results in the production
Acknowledgments
of OH-Bz-3NO2–OH, cannot distinguish between ONOO− and •NO2
formed from the MPO/H2O2/NO2− system.
This work was supported by a grant from the Polish National Science
Incubation of CI-Bz-BA with RAW 264.7 macrophages activated to
Centre (NCN) within the SONATA BIS 6 program (Grant No. 2016/22/
produce ONOO− yielded the corresponding fluorescent phenolic com­
E/ST4/00549).
pound, CI–OH, as well as the ONOO− -specific nitrated product, CI-Bz-
The LC-MS analyses were performed at the Medical College of Wis­
NO2. These studies demonstrate that CI-Bz-BA fulfills the requirements
consin Cancer Center Redox and Bioenergetics Shared Resource. The
to be classified as an efficient ONOO− probe applicable to the study of
authors thank Dr. Suresh Kumar (Medical College of Wisconsin) for his

44
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

help with the two-photon fluorescence imaging experiments. lipid peroxidation. Formation of novel nitrogen-containing oxidized lipid
derivatives, J. Biol. Chem. 269 (1994) 26066–26075.
A.G. was supported by a grant from the Polish National Science
[24] M.-L. Brennan, W. Wu, X. Fu, Z. Shen, W. Song, H. Frost, C. Vadseth, L. Narine,
Centre (NCN) within the MINIATURA 4 program (Grant No. 2020/04/ E. Lenkiewicz, M.T. Borchers, A.J. Lusis, J.J. Lee, N.A. Lee, H.M. Abu-Soud,
X/ST4/01002). H. Ischiropoulos, S.L. Hazen, A Tale of Two Controversies: defining both the role of
A.G. thanks Klaudia Salamon for initial contributions. peroxidases in nitrotyrosine formation in vivo using eosinophil peroxidase and
myeloperoxidase-deficient mice, and the nature of peroxidase-generated reactive
nitrogen species, J. Biol. Chem. 277 (2002) 17415–17427.
Appendix A. Supplementary data [25] S. Pfeiffer, A. Lass, K. Schmidt, B. Mayer1, Protein tyrosine nitration in mouse
peritoneal macrophages activated in vitro and in vivo: evidence against an
essential role of peroxynitrite, Faseb. J. 15 (2001) 2355–2364.
Supplementary data related to this article can be found at htt [26] J. Zielonka, B. Kalyanaraman, Small-molecule luminescent probes for the detection
ps://doi.org/10.1016/j.freeradbiomed.2021.12.260. of cellular oxidizing and nitrating species, Free Radic. Biol. Med. 128 (2018) 3–22.
[27] Y. Mikhed, K. Bruns, S. Schildknecht, M. Jörg, M. Dib, M. Oelze, K.J. Lackner,
T. Münzel, V. Ullrich, A. Daiber, Formation of 2-nitrophenol from salicylaldehyde
References as a suitable test for low peroxynitrite fluxes, Redox Biol. 7 (2016) 39–47.
[28] E. Oliveira, E. Bértolo, C. Núñez, V. Pilla, H.M. Santos, J. Fernández-Lodeiro,
[1] R. Kissner, T. Nauser, P. Bugnon, P.G. Lye, W.H. Koppenol, formation and A. Fernández-Lodeiro, J. Djafari, J.L. Capelo, C. Lodeiro, Green and red fluorescent
properties of peroxynitrite as studied by laser flash photolysis, high-pressure dyes for translational applications in imaging and sensing analytes: a dual-color
stopped-flow technique, and pulse radiolysis, Chem. Res. Toxicol. 10 (1997) flag, ChemistryOpen 7 (2018) 9–52.
1285–1292. [29] L.M. Hyman, K.J. Franz, Probing oxidative stress: small molecule fluorescent
[2] S. Goldstein, G. Czapski, The reaction of NO⋅ with O2⋅− and HO2⋅− : a pulse sensors of metal ions, reactive oxygen species, and thiols, Coord. Chem. Rev. 256
radiolysis study, Free Radic. Biol. Med. 19 (1995) 505–510. (2012) 2333–2356.
[3] J. Zielonka, A. Sikora, J. Joseph, B. Kalyanaraman, Peroxynitrite is the major [30] J. Zielonka, R. Podsiadły, M. Zielonka, M. Hardy, B. Kalyanaraman, On the use of
species formed from different flux ratios of co-generated nitric oxide and peroxy-caged luciferin (PCL-1) probe for bioluminescent detection of inflammatory
superoxide: direct reaction with boronate-based fluorescent probe, J. Biol. Chem. oxidants in vitro and in vivo – identification of reaction intermediates and oxidant-
285 (2010) 14210–14216. specific minor products, Free Radic. Biol. Med. 99 (2016) 32–42.
[4] R. Radi, Peroxynitrite, a stealthy biological oxidant, J. Biol. Chem. 288 (2013) [31] N. Rios, L. Piacenza, M. Trujillo, A. Martínez, V. Demicheli, C. Prolo, M.N. Álvarez,
26464–26472. G.V. López, R. Radi, Sensitive detection and estimation of cell-derived
[5] P. Pacher, J.S. Beckman, L. Liaudet, Nitric oxide and peroxynitrite in health and peroxynitrite fluxes using fluorescein-boronate, Free Radic. Biol. Med. 101 (2016)
disease, Physiol. Rev. 87 (2007) 315–424. 284–295.
[6] M.N. Möller, N. Rios, M. Trujillo, R. Radi, A. Denicola, B. Alvarez, Detection and [32] C. Prolo, N. Rios, L. Piacenza, M.N. Álvarez, R. Radi, Fluorescence and
quantification of nitric oxide–derived oxidants in biological systems, J. Biol. Chem. chemiluminescence approaches for peroxynitrite detection, Free Radic. Biol. Med.
294 (2019) 14776–14802. 128 (2018) 59–68.
[7] R. Radi, Oxygen radicals, nitric oxide, and peroxynitrite: redox pathways in [33] C.C. Winterbourn, The challenges of using fluorescent probes to detect and
molecular medicine, Proc. Natl. Acad. Sci. U.S.A. 115 (2018) 5839–5848. quantify specific reactive oxygen species in living cells, Biochim. Biophys. Acta
[8] Z. Daneva, C. Marziano, M. Ottolini, Y.-L. Chen, T.M. Baker, M. Kuppusamy, 1840 (2014) 730–738.
A. Zhang, H.Q. Ta, C.E. Reagan, A.D. Mihalek, R.B. Kasetti, Y. Shen, B.E. Isakson, R. [34] A. Sikora, J. Zielonka, K. Dębowska, R. Michalski, R. Smulik-Izydorczyk, J. Pięta,
D. Minshall, G.S. Zode, E.A. Goncharova, V.E. Laubach, S.K. Sonkusare, Caveolar R. Podsiadły, A. Artelska, K. Pierzchała, B. Kalyanaraman, Boronate-based probes
peroxynitrite formation impairs endothelial TRPV4 channels and elevates for biological oxidants: a novel class of molecular tools for redox biology, Front.
pulmonary arterial pressure in pulmonary hypertension, Proc. Natl. Acad. Sci. U.S. Chem. 8 (2020), 580899.
A. 118 (2021), e2023130118. [35] A. Grzelakowska, M. Zielonka, K. Dębowska, J. Modrzejewska, M. Szala, A. Sikora,
[9] M. Kizilgun, Y. Poyrazoglu, Y. Oztas, H. Yaman, E. Cakir, T. Cayci, O.E. Akgul, Y. J. Zielonka, R. Podsiadły, Two-photon fluorescent probe for cellular peroxynitrite:
G. Kurt, H. Yaren, Z.I. Kunak, E. Macit, E. Ozkan, M.Y. Taslipinar, T. Turker, fluorescence detection, imaging, and identification of peroxynitrite-specific
A. Ozcan, Beneficial effects of N-acetylcysteine and ebselen on Renal Ischemia/ products, Free Radic. Biol. Med. 169 (2021) 24–35.
Reperfusion injury, Ren. Fail. 33 (2011) 512–517. [36] A. Sikora, J. Zielonka, M. Lopez, J. Joseph, B. Kalyanaraman, Direct oxidation of
[10] A. Guven, B. Uysal, O. Akgul, H. Cermik, G. Gundogdu, I. Surer, H. Ozturk, boronates by peroxynitrite: mechanism and implications in fluorescence imaging of
A. Korkmaz, Scavenging of peroxynitrite Reduces Renal Ischemia/Reperfusion peroxynitrite, Free Radic. Biol. Med. 47 (2009) 1401–1407.
injury, Ren. Fail. 30 (2008) 747–754. [37] M. Szala, A. Grzelakowska, J. Modrzejewska, P. Siarkiewicz, D. Słowiński,
[11] H. Jaeschke, A. Ramachandran, The role of oxidant stress in acetaminophen- M. Świerczyńska, J. Zielonka, R. Podsiadły, Characterization of the reactivity of
induced liver injury, Curr. Opin. Toxicol. 20–21 (2020) 9–14. luciferin boronate - a probe for inflammatory oxidants with improved stability,
[12] K. Du, A. Ramachandran, H. Jaeschke, Oxidative stress during acetaminophen Dyes Pigments 183 (2020), 108693.
hepatotoxicity: sources, pathophysiological role and therapeutic potential, Redox [38] A. Sikora, J. Zielonka, J. Adamus, D. Debski, A. Dybala-Defratyka, B. Michalowski,
Biol. 10 (2016) 148–156. J. Joseph, R.C. Hartley, M.P. Murphy, B. Kalyanaraman, Reaction between
[13] S. Feng, X. Cheng, L. Zhang, X. Lu, S. Chaudhary, R. Teng, C. Frederickson, M. peroxynitrite and triphenylphosphonium-substituted arylboronic acid Isomers:
M. Champion, R. Zhao, L. Cheng, Y. Gong, H. Deng, X. Lu, Myeloid-derived identification of diagnostic marker products and biological implications, Chem.
suppressor cells inhibit T cell activation through nitrating LCK in mouse cancers, Res. Toxicol. 26 (2013) 856–867.
Proc. Natl. Acad. Sci. U.S.A. 115 (2018) 10094–10099. [39] A. Sikora, J. Zielonka, M. Lopez, A. Dybala-Defratyka, J. Joseph, A. Marcinek,
[14] M. Trujillo, G. Ferrer-Sueta, R. Radi, Chapter Ten - Kinetic Studies on Peroxynitrite B. Kalyanaraman, Reaction between peroxynitrite and boronates: EPR spin-
Reduction by Peroxiredoxins, in: E. Cadenas, L. Packer (Eds.), Methods in trapping, HPLC analyses, and quantum mechanical study of the free radical
Enzymology, Academic Press, 2008, pp. 173–196. pathway, Chem. Res. Toxicol. 24 (2011) 687–697.
[15] R. Radi, Oxygen radicals, nitric oxide, and peroxynitrite: redox pathways in [40] K. Dębowska, D. Dębski, B. Michałowski, A. Dybala-Defratyka, T. Wójcik,
molecular medicine, Proc. Natl. Acad. Sci. U.S.A. 115 (2018) 5839–5848. R. Michalski, M. Jakubowska, A. Selmi, R. Smulik, Ł. Piotrowski, J. Adamus,
[16] Y. Zuo, Y. Jiao, C. Ma, C. Duan, A novel fluorescent probe for hydrogen peroxide A. Marcinek, S. Chlopicki, A. Sikora, Characterization of fluorescein-based
and its application in bio-imaging, Molecules 26 (2021) 3352. monoboronate probe and its application to the detection of peroxynitrite in
[17] N. Ríos, C. Prolo, M.N. Álvarez, L. Piacenza, R. Radi, Chapter 21 - Peroxynitrite endothelial cells treated with doxorubicin, Chem. Res. Toxicol. 29 (2016) 735–746.
Formation and Detection in Living Cells, in: L.J. Ignarro, B.A. Freeman (Eds.), [41] J. Zielonka, J. Joseph, A. Sikora, M. Hardy, O. Ouari, J. Vasquez-Vivar, G. Cheng,
Nitric Oxide, third ed., Academic Press, 2017, pp. 271–288. M. Lopez, B. Kalyanaraman, Mitochondria-targeted triphenylphosphonium-based
[18] D. Rane, E.J. Carlson, Y. Yin, B.R. Peterson, Chapter One - Fluorescent Detection of compounds: syntheses, mechanisms of action, and therapeutic and diagnostic
Peroxynitrite during Antibody-dependent Cellular Phagocytosis, in: D. applications, Chem. Rev. 117 (2017) 10043–10120.
M. Chenoweth (Ed.), Methods in Enzymology, Academic Press, 2020, pp. 1–35. [42] M.P. Murphy, Targeting lipophilic cations to mitochondria, Biochim. Biophys. Acta
[19] G. Ferrer-Sueta, N. Campolo, M. Trujillo, S. Bartesaghi, S. Carballal, N. Romero, 1777 (2008) 1028–1031.
B. Alvarez, R. Radi, Biochemistry of peroxynitrite and protein tyrosine nitration, [43] R.A.J. Smith, R.C. Hartley, M.P. Murphy, Mitochondria-targeted small molecule
Chem. Rev. 118 (2018) 1338–1408. therapeutics and probes, Antioxidants Redox Signal. 15 (2011) 3021–3038.
[20] P. Wardman, Fluorescent and luminescent probes for measurement of oxidative [44] G. Cheng, M. Hardy, J. Zielonka, K. Weh, M. Zielonka, K.A. Boyle, M. Abu Eid,
and nitrosative species in cells and tissues: progress, pitfalls, and prospects, Free D. McAllister, B. Bennett, L.A. Kresty, M.B. Dwinell, B. Kalyanaraman,
Radic. Biol. Med. 43 (2007) 995–1022. Mitochondria-targeted magnolol inhibits OXPHOS, proliferation, and tumor
[21] A. Vasilescu, M. Gheorghiu, S. Peteu, Nanomaterial-based electrochemical sensors growth via modulation of energetics and autophagy in melanoma cells, Cancer
and optical probes for detection and imaging of peroxynitrite: a review, Treat. Res. Commun. 25 (2020) 100210.
Microchim. Acta 184 (2017) 649–675. [45] A.R. Chowdhury, J. Zielonka, B. Kalyanaraman, R.C. Hartley, M.P. Murphy, N.
[22] F.J. Schopfer, P.R.S. Baker, B.A. Freeman, NO-dependent protein nitration: a cell G. Avadhani, Mitochondria-targeted paraquat and metformin mediate ROS
signaling event or an oxidative inflammatory response? Trends Biochem. Sci. 28 production to induce multiple pathways of retrograde signaling: a dose-dependent
(2003) 646–654. phenomenon, Redox Biol. 36 (2020), 101606.
[23] H. Rubbo, R. Radi, M. Trujillo, R. Telleri, B. Kalyanaraman, S. Barnes, M. Kirk, B. [46] G. Battogtokh, Y.S. Choi, D.S. Kang, S.J. Park, M.S. Shim, K.M. Huh, Y.-Y. Cho, J.
A. Freeman, Nitric oxide regulation of superoxide and peroxynitrite-dependent Y. Lee, H.S. Lee, H.C. Kang, Mitochondria-targeting drug conjugates for cytotoxic,

45
A. Grzelakowska et al. Free Radical Biology and Medicine 179 (2022) 34–46

anti-oxidizing and sensing purposes: current strategies and future perspectives, [59] R. Michalski, D. Thiebaut, B. Michałowski, M.M. Ayhan, M. Hardy, O. Ouari,
Acta Pharm. Sin. B. 8 (2018) 862–880. M. Rostkowski, R. Smulik-Izydorczyk, A. Artelska, A. Marcinek, J. Zielonka,
[47] H. Wang, B. Fang, B. Peng, L. Wang, Y. Xue, H. Bai, S. Lu, N.H. Voelcker, L. Li, B. Kalyanaraman, A. Sikora, Oxidation of ethidium-based probes by biological
L. Fu, W. Huang, Recent advances in chemical biology of mitochondria targeting, radicals: mechanism, kinetics and implications for the detection of superoxide, Sci.
Front. Chem. 9 (2021) 321. Rep. 10 (2020), 18626.
[48] Y. Li, K.-N. Wang, B. Liu, X.-R. Lu, M.-F. Li, L.-N. Ji, Z.-W. Mao, Mitochondria- [60] P. Siarkiewicz, R. Michalski, A. Sikora, R. Smulik-Izydorczyk, M. Szala,
targeted two-photon fluorescent probe for the detection of biothiols in living cells, A. Grzelakowska, J. Modrzejewska, A. Bailey, J.E. Nycz, B. Kalyanaraman, J.
Sens. Actuators B Chem. 255 (2018) 193–202. G. Malecki, J. Zielonka, R. Podsiadły, On the chemical reactivity of tricyanofuran
[49] D. Cao, Z. Liu, P. Verwilst, S. Koo, P. Jangjili, J.S. Kim, W. Lin, Coumarin-based (TCF)-based near-infrared fluorescent redox probes – effects of glutathione on the
small-molecule fluorescent chemosensors, Chem. Rev. 119 (2019) 10403–10519. probe response and product fluorescence, Dyes Pigments 192 (2021), 109405.
[50] R. Smulik, D. Dębski, J. Zielonka, B. Michałowski, J. Adamus, A. Marcinek, [61] A. Denicola, B.A. Freeman, M. Trujillo, R. Radi, Peroxynitrite reaction with carbon
B. Kalyanaraman, A. Sikora, Nitroxyl (HNO) reacts with molecular oxygen and dioxide/bicarbonate: kinetics and Influence on peroxynitrite-mediated oxidations,
forms peroxynitrite at physiological pH: biological implications, J. Biol. Chem. 289 Arch. Biochem. Biophys. 333 (1996) 49–58.
(2014) 35570–35581. [62] O. Augusto, S. Goldstein, J.K. Hurst, J. Lind, S.V. Lymar, G. Merenyi, R. Radi,
[51] D. Dębski, R. Smulik, J. Zielonka, B. Michałowski, M. Jakubowska, K. Dębowska, Carbon dioxide-catalyzed peroxynitrite reactivity – the resilience of the radical
J. Adamus, A. Marcinek, B. Kalyanaraman, A. Sikora, Mechanism of oxidative mechanism after two decades of research, Free Radic. Biol. Med. 135 (2019)
conversion of Amplex® Red to resorufin: pulse radiolysis and enzymatic studies, 210–215.
Free Radic. Biol. Med. 95 (2016) 323–332. [63] M.G. Bonini, O. Augusto, Carbon dioxide stimulates the production of thiyl,
[52] J. Zielonka, A. Sikora, M. Hardy, J. Joseph, B.P. Dranka, B. Kalyanaraman, sulfinyl, and disulfide radical anion from thiol oxidation by peroxynitrite, J. Biol.
Boronate probes as diagnostic tools for real time monitoring of peroxynitrite and Chem. 276 (2001) 9749–9754.
hydroperoxides, Chem. Res. Toxicol. 25 (2012) 1793–1799. [64] P. Wardman, Fluorescent and luminescent probes for measurement of oxidative
[53] J. Zielonka, M. Zielonka, A. Sikora, J. Adamus, J. Joseph, M. Hardy, O. Ouari, B. and nitrosative species in cells and tissues: progress, pitfalls, and prospects, Free
P. Dranka, B. Kalyanaraman, Global profiling of reactive oxygen and nitrogen Radic. Biol. Med. 43 (2007) 995–1022.
species in biological systems, J. Biol. Chem. 287 (2012) 2984–2995. [65] J. Modrzejewska, M. Szala, A. Grzelakowska, M. Zakłos-Szyda, J. Zielonka,
[54] D.S. Bohle, P.A. Glassbrenner, B. Hansert, Synthesis of Pure R. Podsiadły, Novel boronate probe based on 3-Benzothiazol-2-yl-7-hydroxy-
Tetramethylammonium Peroxynitrite, Methods in Enzymology, Academic Press, chromen-2-one for the detection of peroxynitrite and hypochlorite, Molecules 26
1996, pp. 302–311. (2021) 5940.
[55] J. Zielonka, A. Sikora, J. Adamus, B. Kalyanaraman, Detection and differentiation [66] H. Iwashita, E. Castillo, M.S. Messina, R.A. Swanson, C.J. Chang, A tandem
between peroxynitrite and hydroperoxides using mitochondria-targeted activity-based sensing and labeling strategy enables imaging of transcellular
arylboronic acid, Methods Mol. Biol. 1264 (2015) 171–181. hydrogen peroxide signaling, Proc. Natl. Acad. Sci. U.S.A. 118 (2021),
[56] R.J. Singh, N. Hogg, J. Joseph, E. Konorev, B. Kalyanaraman, The peroxynitrite e2018513118.
generator, SIN-1, becomes a nitric oxide donor in the presence of electron [67] A.R. Lippert, G.C. Van de Bittner, C.J. Chang, Boronate oxidation as a
acceptors, Arch. Biochem. Biophys. 361 (1999) 331–339. bioorthogonal reaction approach for studying the chemistry of hydrogen peroxide
[57] J.L. Trackey, T.F. Uliasz, S.J. Hewett, SIN-1-induced cytotoxicity in mixed cortical in living systems, Acc. Chem. Res. 44 (2011) 793–804.
cell culture: peroxynitrite-dependent and -independent induction of excitotoxic cell [68] J. Zielonka, M. Zielonka, L. VerPlank, G. Cheng, M. Hardy, O. Ouari, M.M. Ayhan,
death, J. Neurochem. 79 (2001) 445–455. R. Podsiadły, A. Sikora, J.D. Lambeth, B. Kalyanaraman, Mitigation of NADPH
[58] A.J. Gow, Q. Chen, M. Gole, M. Themistocleous, V.M.-Y. Lee, H. Ischiropoulos, Two oxidase 2 activity as a strategy to inhibit peroxynitrite formation, J. Biol. Chem.
distinct mechanisms of nitric oxide-mediated neuronal cell death show thiol 291 (2016) 7029–7044.
dependency, Am. J. Physiol. Cell Physiol. 278 (2000) C1099–C1107.

46

You might also like