You are on page 1of 15

Dyes and Pigments 192 (2021) 109405

Contents lists available at ScienceDirect

Dyes and Pigments


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

On the chemical reactivity of tricyanofuran(TCF)-based near-infrared


fluorescent redox probes – Effects of glutathione on the probe response and
product fluorescence
Przemysław Siarkiewicz a, Radosław Michalski b, Adam Sikora b, Renata Smulik-Izydorczyk b,
Marcin Szala a, Aleksandra Grzelakowska a, Julia Modrzejewska a, Asha Bailey c, Jacek E. Nycz e,
Balaraman Kalyanaraman c, d, Jan Grzegorz Malecki e, Jacek Zielonka c, d, **,
Radosław Podsiadły a, *
a
Institute of Polymer and Dye Technology, Faculty of Chemistry, Lodz University of Technology, Stefanowskiego 12/16, 90-924, Lodz, Poland
b
Institute of Applied Radiation Chemistry, Faculty of Chemistry, Lodz University of Technology, Zeromskiego 116, 90-924, Lodz, Poland
c
Department of Biophysics, Poland
d
Cancer Center Redox & Bioenergetics Shared Resource, Medical College of Wisconsin, 8701 Watertown Plank Road, Milwaukee, WI, 53226, United States
e
Institute of Chemistry, University of Silesia in Katowice, Ul. Szkolna 9, Katowice, 40-007, Poland

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

Keywords: Recent research towards the development of the redox probes for in vivo applications focuses on near-infrared
Redox probes fluorescent sensors and tricyanofuran-based fluorophores gained popularity thanks to their favorable spectral
Tricyanofuran-based probes properties. The tricyanofuran-based boronate probe (TCF-BA) has been proposed for specific fluorescent
Fluorescence
detection of selected biological oxidants in vitro and in vivo. Here, we report the detailed chemical reactivity of
Peroxynitrite
Thiol adducts
TCF-BA toward hydrogen peroxide, hypochlorite, and peroxynitrite in the presence and absence of glutathione, a
major small molecule biothiol present intracellularly at millimolar concentrations. We demonstrate that, at the
physiologically relevant concentration of glutathione, the TCF-BA probe forms an adduct, resulting in decreased
reactivity of the probe toward the oxidants tested. Only peroxynitrite efficiently oxidizes TCF-BA in the presence
of GSH. Furthermore, the fluorescent phenolic oxidation product, TCF-OH, also reacts with glutathione, which
results in a decreased fluorescence intensity. This observation suggests that the results reported with TCF-based
probes may be affected by the changes in intracellular glutathione, in addition to the desired analyte. We also
report a modified probe (TCF-BA-2) derived from 1-naphthalene boronic acid, which has similar reactivity to­
ward peroxynitrite. Although the TCF-BA-2 probe also reacts with glutathione, the absorption spectrum of its
oxidation product, TCF–OH–2, is not influenced by glutathione and, therefore, can be applied for real-time
monitoring of peroxynitrite formation in biological systems.

1. Introduction autofluorescence and cell photodamage [1,2], long wavelength red or


near infrared (NIR) fluorogenic chemical probes are desired for in vivo
Fluorescence microscopy and imaging are indispensable in studies of applications. In the last decade, numerous probes containing the tri­
cell function and intracellular analytes in vitro and in vivo. To provide cyanofuran (TCF) ring (see Scheme 2) have been synthesized and
deeper tissue light penetration and decrease background characterized [3–15] with the aim of detecting a specific analyte in vivo.

* Corresponding author.
** Corresponding author. Institute of Polymer and Dye Technology, Faculty of Chemistry, Lodz University of Technology, Stefanowskiego 12/16, 90-924, Lodz,
Poland.
E-mail addresses: przemyslaw.siarkiewicz@dokt.p.lodz.pl (P. Siarkiewicz), radoslaw.michalski@p.lodz.pl (R. Michalski), adam.sikora@p.lodz.pl (A. Sikora),
renata.smulik-izydorczyk@p.lodz.pl (R. Smulik-Izydorczyk), marcin.szala@p.lodz.pl (M. Szala), aleksandra.grzelakowska@p.lodz.pl (A. Grzelakowska), julia.
modrzejewska@dokt.p.lodz.pl (J. Modrzejewska), baileyaa1211@gmail.com (A. Bailey), jacek.nycz@us.edu.pl (J.E. Nycz), balarama@mcw.edu
(B. Kalyanaraman), gmalecki@us.edu.pl (J.G. Malecki), jzielonk@mcw.edu (J. Zielonka), radoslaw.podsiadly@p.lodz.pl (R. Podsiadły).

https://doi.org/10.1016/j.dyepig.2021.109405
Received 28 January 2021; Received in revised form 7 April 2021; Accepted 18 April 2021
Available online 1 May 2021
0143-7208/© 2021 Elsevier Ltd. All rights reserved.
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

The p-amino [6–8] or several p-hydroxyl [8–15] derivatives have been (k ~ 1 M− 1s− 1 [47,49]), AA-OOH (k ~ 10 M− 1s− 1 [52]), and HOCl (k ~
applied as fluorescent sensors for a range of analytes in cells, including 104 M− 1s− 1 [47,49]).
carbon monoxide (CO) [16], nitroxyl (HNO) [12], selected oxidants In addition to fast reaction kinetics, another advantage of boronic
[3–5,17,18], hydrazine [19], albumin [6,20], biothiols [11,13,15,21, probes (of the general structure depicted as Ar–B(OH)2 in Scheme 1) as
22], F− [23], K+ [7], and several metal ions: Cu2+ [10,24], Hg2+ [25] tools for ONOO− detection is their specific reaction mechanism with this
and palladium (Pd0) [26]. Also, TCF-based probes have been designed to oxidant (Scheme 1). The major, phenolic product Ar-OH (~85–99%
monitor cellular pH [9] and enzymatic activity of NADPH-quinone yield), common for all listed oxidants, is formed via the heterolytic
oxidoreductase (NQO1) [14], alkaline phosphatase [27], and carbox­ cleavage of the O–O bond in the anionic adduct. However, homolytic
ylesterase 2 [28]. cleavage of the peroxide bond results in the formation of an Ar–B
The p-phenyl-boronic acid pinacol ester linked via a double bound to (OH)2O•− radical anion, which further fragments to a phenyl-type
the TCF ring (TCF-BE, Scheme 2) has been proposed for detection of radical (Ar•) and nitrogen dioxide (•NO2). Stable, non-radical products
selected reactive oxygen species (ROS), including H2O2 [4,17], HOCl [5] Ar-NO2 and Ar–H are formed in the reaction of Ar• with •NO2 or
or ONOO− [3], via oxidation to the TCF-OH red fluorescent product. hydrogen donor, respectively [57]. Although the yield of the minor
However, none of mentioned works have reported competing for product(s) is relatively low (~1–15%), their formation may serve as a
oxidation of TCF-BE between ROS. Interestingly, ROS are generated unique footprint, providing a diagnostic marker for ONOO− formation.
upon one- or two-electron reduction of molecular oxygen to superoxide This is a significant advantage over the use of the nitrotyrosine marker,
radical anion (O•− 2 ) and hydrogen peroxide (H2O2), respectively. Sub­ as the latter can be also formed via ONOO− -independent pathways.
sequent reactions of those species may result in the formation of, e.g., Changing the phenyl ring in Ar–B(OH)2 for the appropriate fluorophore
peroxynitrite (ONOO− ) [29], hydroxyl radical (•OH), or hypochlorous (e.g., coumarin) or luciferin backbone results in a fluorogenic probe
acid (HOCl). At physiological conditions, intracellular ROS levels are (coumarin boronic acid, CBA [55,58]) or a bioluminogenic probe
regulated by antioxidant enzymes and non-enzymatic antioxidants, such (luciferin boronic acid, LBA [59]), since in both cases an identical
as superoxide dismutases (SOD), peroxiredoxins (Prx), glutathione oxidative conversion occurs, and ONOO− -specific products were
peroxidases, catalase (CAT), glutathione (GSH), or ascorbate. The detected.
increased production of ROS or the defective defense mechanisms leads Here, we present the results of the detailed investigation of the
to oxidative stress, which is widely believed to be cytotoxic. However, products formed during oxidation of TCF-BE by different biologically
O•−
2 and H2O2 generated at low levels may be involved in relevant oxidants and identify the products of nitration and chlorination
redox-signaling mechanisms, e.g., promoting cancer cell proliferation of the oxidation product, TCF-OH, to help interpret the in vivo fluores­
[30–34]. It was shown that the occurrence, growth, and metastasis of cence results obtained with this probe and identify the oxidants
tumors, and even the apoptosis, necrosis, and autophagy of tumor cells, involved. Furthermore, we show that the presence of GSH significantly
are all closely related to ROS production [35]. ROS also have an impact affects the probe performance in the detection of the oxidants studied.
on the tumor microenvironment. At moderate concentrations, ROS The chemical mechanisms beyond the observed effects of GSH are
activate cancer cell survival signaling and immune evasion, whereas at described. Our investigations brought about a novel probe derived from
high concentrations, ROS can cause cancer cell apoptosis [36]. 1-naphthalene boronic acid TCF-BA-2 (Scheme 2). The colorimetric
Recently, we reported that tumor growth is accompanied by response of this probe is not influenced by GSH and, therefore, can be
increased ROS formation and revealed differences in oxidant formation applied in cell culture or other systems compatible with spectrophoto­
in the inner and outer sections of tumor tissue, respectively, demon­ metric detection.
strating tumor redox heterogeneity [37]. Several factors—such as the
type and maturity of the tumor, the biological and metabolic environ­ 2. Material and methods
ment, and the location, identity, and level of ROS—may determine
whether ROS act as tumor-promoting or tumor-inhibiting agents [38]. It 2.1. Materials
was also shown that there is no linear relationship between antioxidant
supplementation and cancers prevention, with the clinical studies 3-Hydroxy-3-methyl-2-butanone, 2-hydroxy-4′ -(2-hydroxyethoxy)-
yielding mixed results [39–41]. In this context, to leverage ROS in 2-methylpropiophenone (EtxPhMP), malononitrile, lithium, benzalde­
effective anti-tumor strategies, it is important to provide noninvasive hyde (4-FPh-H), 4-hydroxybenzaldehyde (4-FPh-OH), 4-nitrobenzalde­
tools to image and quantify ROS levels in tumor tissue and its hyde (4-FPh-NO2), 4-hydroxy-3-nitrobenzaldehyde (4-FPh-OH-3-
microenvironment. NO2), 4-hydroxy-3-chlorobenzaldehyde (4-FPh-OH-3-Cl), 4-formylphe­
The basic chemical properties of ROS and their reactivity toward nylboronic acid pinacol ester (4-FPh-BE), 4-formylphenylboronic acid
cellular components result in a short lifetime of those species and pre­ (4-FPh-BA), 4-hydroxy-1-naphthaldehyde (4-FNaph-OH), 1-naphthal­
vent from them from significantly accumulating. Low steady-state con­ dehyde (4-FNaph-H), and Pd(dppf)Cl2 were purchased from Sigma­
centrations and unfavorable spectroscopic properties make direct –Aldrich (Poznan, Poland). 4-Formyl-naphthalene-1-boronic acid (4-
detection of ROS in cells practically impossible. The use of chemical or FNaph-BA) was purchased from Combi-Blocks (San Diego, CA). Sol­
genetically encoded probes that produce easily detectable oxidation vents used for syntheses were reagent grade. (Z)-1-[N-[3-aminopropyl]-
products has been proposed and widely applied to overcome those N-[4-(3-aminopropylammonio)butyl]-amino]diazen-1-ium-1,2-diolate
limitations. Furthermore, to identify ROS in biological systems, the use (Spermine-NONOate) and (Z)-1-[N-(3-aminopropyl)-N-(n-propyl)
of probes that are selective and/or form species-specific products has amino]diazen-1-ium-1,2-diolate (PAPA-NONOate) were purchased
been recommended [42]. from Cayman Chemical (Ann Arbor, MI). Hypoxanthine (HX), xanthine
Fluorogenic and luminogenic boronates have been recently devel­ oxidase (XO), GSH, CAT, bovine SOD, and penta(carboxymethyl)-
oped as tools for detection of selected ROS and reactive nitrogen species diethylenetriamine (DTPA) were purchased from Sigma-Aldrich (Mil­
(RNS) in cell-free systems and cultured cells in vitro and in animals in vivo waukee, WI).
[2,4,37,42–47]. The boronate-based probes are oxidized by H2O2 [1,45,
47–49], peroxymonocarbonate (HCO−4 ) [50,51], selected amino acid 2.2. Synthesis
hydroperoxides (AA-OOH) and protein hydroperoxides (Pr-OOH) [52],
HOCl [2,47,49,53–56], and ONOO− [3,4,49,53–55], to the same major All synthesized compounds containing the TCF ring were purified by
phenolic product. These oxidants differ, however, in their reaction ki­ column chromatography using silica gel (0.063–200 mm, 70–230
netics with boronate-based probes [47,49,53]. ONOO− is the fastest mesh). 1H nuclear magnetic resonance (NMR) and 13C NMR spectra
reported oxidant (k ~ 106 M− 1s− 1 [47,49,55]) in comparison with H2O2 were recorded on Bruker AVANCE 400 spectrometer at 400 MHz (1H) or

2
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

Scheme 1. The mechanism of ONOO− -induced oxidation of arylboronate compounds Ar–B(OH)2 [49,57].

2.3. General preparation of dyes and chemicals used in analyses

The stock solutions of oxidants (HOCl, and H2O2) were prepared


freshly before each experiment and theirs concentrations were deter­
mined by spectrophotometry, using appropriate molar extinction coef­
ficient (ε292 = 350 M− 1cm− 1 in 0.1 M NaOH for HOCl and ε240 = 43.6
M− 1cm− 1 in water for H2O2) [45]. Peroxynitrite was synthesized in re­
action of 0.6 M nitrite with 0.7 M hydrogen peroxide at pH 13 [45].
Excess H2O2 was removed by passage through a column of MnO2 and the
solution was frozen at − 20 ◦ C. The liquid over the frozen solid was
collected, aliquoted into 1.5 mL tubes and stored at − 80 ◦ C. The con­
centration of peroxynitrite was determined spectrally at 302 nm (ε302 =
1.7 × 103 M− 1cm− 1), after dilution in 0.1 M NaOH to ~10 mM con­
centration, immediately prior to each experiment. During high perfor­
mance liquid chromatography (HPLC) analyses it was noticed that
Scheme 2. Synthesized compounds containing the TCF ring.
TCF-BE (pinacolate ester) undergoes fast hydrolysis to the boronic
acid form (TCF-BA) upon dilution in the aqueous phosphate buffer.
Therefore, although we added TCF-BE to the investigated mixtures,
100 MHz (13C), respectively. Compounds were dissolved in DMSO‑d6,
TCF-BA is the species that is tested.
and tetramethylsilane (TMS) was added as an internal reference for 1H
NMR and 13C NMR spectra. Chemical shifts (δ) are reported in ppm, and
coupling constant J values in hertz (Hz). The crystallographic data of 2.4. Spectroscopic studies
compounds TCF-NO2 and TCF-H are shown in the Supplementary In­
formation (Figures S1 and S2, Tables S1 and S2). 2-Dicyanomethylene-3- Absorption and fluorescence spectra were recorded using a Jasco UV-
cyano-4,5,5-trimethyl-2,5-dihydrofuran (TCF) was synthesized accord­ VIS-NIR V–670 spectrophotometer (Jasco, Japan) and a FLS-920 spec­
ing to the procedure described elsewhere [60] and used for subsequent trofluorometer (Edinburgh Instruments, UK), respectively. The fluores­
syntheses without further purification. Scheme 3 shows in condensed cence quantum yield of the dye (ΦDYE) was calculated from the equation,
form the synthetic pathways. For detailed synthetic procedures, melting GDYE ⋅η2DYE
points and also NMR and HRMS spectra of synthesized compounds the ΦDYE = ΦST (1)
GST ⋅η2ST
reader is referred to the Supplementary Information of this article.
where the subscripts ST and DYE denote standard and tested fluorescent
dye, respectively; Φ is the fluorescence quantum yield; G is the gradient
from the plot of integrated fluorescence intensity versus absorbance; and

3
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

Scheme 3. Synthetic pathways and reagents used to obtain the (A) TCF-BE, TCF-BA probe and TCF-OH, TCF-H, TCF-NO2, TCF-OH-3-NO2, TCF-OH-3-Cl dyes, (B)
TCF-BA-2, and TCF-H-2, TCF-OH-2 dyes, (C) TCF-Etx dye. The description of synthetic procedures is detailed in Supplementary Information.

η is the refractive index of the solvent. The reference systems used were 50 mM) with DTPA (100 μM), HX/XO (1 mM/70 μU/mL), PAPA-
rhodamine 101 in ethanol (ΦST = 100% [61]) or 4-methylumbelliferone NONOate (50 μM), and CAT (100 U/mL). This system produced 1.08
in water (ΦST = 35.6% [62]). μM/min flux of ONOO− .
The dissociation constants of the adducts were determined spectro­ The reactivity of the TCF-BA probe and TCF-OH dye toward 1O2 was
scopically by measuring the absorption spectra of probes or dyes in investigated according to a protocol described previously [21]. Briefly, a
aqueous solutions (pH 7.4) at different concentrations of GSH. solution comprising the TCF-BA probe (10 μM) or TCF-OH dye (10 μM),
sensitizer - Rose Bengal (10 μM), and CAT (100 U/mL) in a phosphate
2.5. Reaction with singlet oxygen and enzymatically generated ROS/RNS buffer (pH 7.4, 50 mM) was placed in a quartz cell and illuminated with
visible light using a 200-W xenon lamp at room temperature with
All solutions were prepared using deionized water (Millipore Milli-Q continuous bubbling with oxygen. The illuminating light was passed
system). Due to the poor solubility of probes in water, all experiments through a water filter and a 400 nm cutoff filter.
were performed with the addition of acetonitrile (MeCN) at up to 5% (v/
v). •NO was generated from thermal decomposition of PAPA-NONOate. 2.6. Pulse radiolysis study

NO fluxes were determined from the rate of decomposition of the
donor, measured following the decrease of its characteristic absorbance Pulse radiolysis experiments were carried out with high-energy (8
at 250 nm (ε = 8.1 × 103 M− 1cm− 1). At a temperature of 25 ◦ C, 50 μM of MeV) 7 ns electron pulses generated from an ELU-6 linear electron
PAPA-NONOate produced 1.08 μM/min •NO flux in solutions containing accelerator [46]. The aqueous solution of potassium thiocyanate (0.01
a phosphate buffer (pH 7.4, 50 mM) and DTPA (100 μM). The flux of O•− 2 M) saturated with nitrous oxide was used to determine the dose absor­
− 7
was produced by the XO-catalyzed oxidation of hypoxanthine in the bed per pulse, assuming G[(SCN)•− 2 ] = 6.2 × 10 mol/J, and
presence of CAT and determined by monitoring the cytochrome c(Fe3+) 3 − 1 − 1
ε[(SCN)2 ] = 7.6 × 10 M cm (G represents the radiation chemical
•−

reduction reflected in an increase in absorbance at 550 nm (using a yield, and ε is the molar absorption coefficient at 475 nm). The dose
difference in the values of the extinction coefficients between reduced delivered per pulse was within the range of 20 Gy.
and oxidized cytochrome of 2.1 × 104 M− 1 cm− 1). 70 μU/mL of XO in ONOO− was generated by irradiation of aerated solutions containing
the presence of 1 mM HX and 100 U/mL CAT produced 0.3 μM/min flux 10 mM nitrite, 1 M methanol, TCF-BA or TCF-BA-2 probe (10–40 μM),
of O•−
2 . H2O2 flux was generated from the same flux of O2 in the absence
•−
and a phosphate buffer (pH 7.4, 50 mM). Irradiation produces nearly
of CAT and presence of SOD (100 μg/mL). ONOO− flux was produced equal amounts of •NO and O•− 2 radicals from the primary products of
from co-generated fluxes of O•−
2 and NO in a phosphate buffer (pH 7.4,

water radiolysis (e−aq, •OH, H•) via reactions 1–7:

4
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

e−aq + NO−2 → NO•2−


2 (reaction 1, k1 = 4 × 109 M− 1s− 1) oxidation product, TCF-OH, reacts with GSH, resulting in a decrease in
fluorescence intensity [3]. Addition of thiols to the vinyl bond of the
NO•2− + H2PO−4 → •NO + HPO2− − 8 TCF-based probe was also explored as a basis for cellular GSH mea­
2 4 + OH (reaction 2, k2 = 1.5 × 10
M− 1s− 1) surements [13]. No studies have been, however, reported on the per­
formance of TCF-based probes for other analytes in the presence of GSH.
OH + CH3OH → •CH2OH + H2O (reaction 3, k3 = 9.7 × 108 M− 1s− 1) To clarify the oxidation chemistry of TCF-BA probe in GSH-free solution

and at physiologically relevant GSH concentrations, we firstly synthe­


H• + CH3OH → •CH2OH + H2 (reaction 4, k4 = 2.6 × 106 M− 1s− 1)
sized TCF-BA probe, its major oxidation product (the TCF-OH dye), and

CH2OH + O2 → •O2CH2OH (reaction 5, k5 = 4.9 × 109 M− 1s− 1) the anticipated minor products of its reactions with ONOO− (TCF-NO2,
TCF-H, TCF-OH-3-NO2) and with HOCl (TCF-OH-3-Cl). Scheme 3 pre­
O2CH2OH + HPO2−
• •−
4 → O2 + CH2O + H2PO4 (reaction 6, k6 = 2 × 10
6
sents all of the synthetic pathways in a condensed form. TCF-derived
− 1 − 1
M s ) compounds were synthesized by modifying the published method [60]
O•− • − 9 − 1 − 1 using TCF and the appropriate benzaldehyde as the starting materials.
2 + NO → ONOO (reaction 7, k7 = 4-7 × 10 M s )
The results of our investigation (see section 3.3) encouraged us to design
The product build-up kinetics were monitored at 585 and 630 nm for and synthesize a novel probe, TCF-BA-2, and its oxidation product, the
TCF-BA and TCF-BA-2, respectively. The recorded kinetic traces were TCF-OH-2 dye. As previously discussed, these compounds were syn­
fitted to a function corresponding to the first order product formation. thesized via a simple condensation of TCF with 4-formylnaphthalene-1-­
Results from the fitting of at least three kinetic traces were averaged to boronic acid or 4-formyl-1-naphthol, respectively. To clarify the
determine the observed first-order rate constant. The second order rate reactivity of TCF-derived compounds with GSH, we also synthesized
constants were obtained from the slopes of the plots of the observed first- TCF-Etx dye, an analogue lacking the styryl unit (Scheme 3).
order rate constants versus the concentration of the studied probe.
3.2. Spectroscopic characterization of TCF-derived compounds

2.7. Stopped-flow measurements The spectroscopic characterization of probes and dyes in aqueous
solution at physiological pH is mandatory for the correct interpretation
2.7.1. Kinetics of TCF-BA reaction with HOCl of the signals obtained when using the fluorogenic probes in cell culture
The rate constant of the reaction of TCF-BA with HOCl was deter­ models. Therefore, we determined the absorption and fluorescence
mined using an Applied Photophysics SX 20 stopped flow spectropho­ emission spectra of the synthesized compounds in aqueous solutions
tometer equipped with an absorption detector (Surrey, UK). HOCl containing a phosphate buffer (pH 7.4, 50 mM) and acetonitrile (20%).
(0.4–2.0 mM) was rapidly mixed with the solution containing TCF-BA The results of spectroscopic characterization are summarized in Table 1.
(40 μM), phosphate buffer (pH 7.4, 100 mM), and MeCN (10%). The From these data, one can conclude that the TCF-BA probe can serve not
thermostated cell (25 ◦ C) with a 10-mm optical pathway was used for only as a fluorogenic probe but also as a naked-eye colorimetric sensor,
kinetic measurements. The kinetic traces were analyzed in the same way since a replacement of the boronate group by hydroxyl group causes a
as described above for pulse radiolysis experiments. large bathochromic (almost 190 nm) and hyperchromic effect (ε587 =
1.85 × 105 M− 1cm− 1). The nitrated and chlorinated dyes, TCF-OH-3-
2.7.2. Kinetics of the reaction between the TCF-BA-SG adduct and NO2 and TCF-OH-3-Cl, exhibit absorption and emission bands located in
ONOO− the same region as the TCF-OH dye. This implicates the necessity of the
The reaction mixture contained 0–25 μM TCF-BA probe, 1 mM GSH, high-performance liquid chromatography (HPLC) analyses to identify
5% MeCN, and a phosphate buffer (pH 7.4, 50 mM) with 100 μM DTPA. the specific products formed in the reaction between the TCF-BA probe
After 1 h of incubation, the CBA probe (10 μM) and ONOO− (5 μM) were and ONOO− or HOCl (see the next section).
added. Spectra were collected 15 min after adding ONOO− with the aid
of a Varian Cary Eclipse spectrofluorometer equipped with a plate
reader accessory (Agilent Technologies, Santa Clara, CA). The solutions
were excited at 332 nm, and the emitted light intensity was measured at Table 1
470 nm (PMT voltage = 670 V, emission/excitation slit = 5 nm). Each Absorption and fluorescence emission properties of studied compounds in
point represents the average value of four samples. The error bars aqueous solution containing a phosphate buffer (pH 7.4, 50 mM) and MeCN
represent standard deviations. The rate constant was determined using (20%).
the competition kinetic approach described by the following equation: Absorption Emission
( )
1 1 1 kTCF− BA− SG [TCF − BA − SG] λmax/nm (ε/103 λmax/nm (ε/103 λex λem Φ (%)
= + × (2) M− 1cm− 1) M− 1cm− 1) (nm) (nm)
Fi F0 F0 kCBA [CBA] i
TCF-BA 407 (12.7) – –
The rate constant of the CBA reaction with ONOO− is equal to k = TCF-OH 452, 448 (48.5)c, 587 607 0.052a
1.1 × 106 M− 1s− 1 [55]. 587 (185.0)c 588 (181.0)d
TCF-H 400 (16.8) – –
TCF-NO2 389 (36.3) – –
3. Results TCF-OH- 539 (64.3) c
414 (60.0) , 539 607 0.069a
3-NO2 537 (203.0)d
3.1. Synthesis of TCF-derived compounds TCF-OH- 585 (54.0) 438 (44.0)c, 585 614 0.094a
3-Cl 584 (234.0)d
TCF-BA- 310 (21.0),
Several papers described the syntheses of fluorogenic boronate – –
2 440 (7.6)
probes derived from p-phenyl-boronic acid ester linked via double TCF-OH- 629 (141.0) c
512 (52.3) , 629 650 0.025a
bound with the appropriate strong-electron acceptor unit, e.g., the TCF 2 632 (101.0)d
ring. The TCF-based probe (TCF-BE) has been proposed for selective TCF-Etx 386 (31.5) 386 470 0.052b
detection of H2O2 [4,17], HOCl [5], or ONOO− [3]. Here, we performed a
Rodamine 101 in ethanol (ΦST = 1.0 [61]) used as the standard.
detailed stoichiometric, kinetic and product analysis studies to define b
4-methylumbelliferone in water (ΦST = 0.356 [62]) used as the standard.
the selectivity of the probe and its potential applicability as a redox c
in DMSO/water (1/5 v/v) at pH = 2 from Ref. [9].
probe in biological systems. It has been also recently reported that the d
in DMSO/water (1/5 v/v) at pH = 10 from Ref. [9].

5
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

3.3. Oxidation of the TCF-BA probe by biologically relevant oxidants in TCF-Ph• phenyl-type radical and •NO2) and TCF-H (TCF-Ph• reduction
GSH-free solution product); nitrated the TCF-OH dye, TCF-OH-3-NO2; and performed
HPLC analyses of the reaction mixture (Fig. 2A). HPLC analyses showed
After determining the spectroscopic properties of the TCF-BA probe that the TCF-NO2 and TCF-H compounds in the reaction between
and TCF-OH dye, we examined the reactivity of these two TCF-derived TCF-BA and ONOO− are formed in a yield of below 1%. Moreover, our
compounds toward several ROS/RNS in GSH-free systems. Analyses of experiments with 2-methyl-2-nitroso-propane as the spin trap did not
fluorescence and absorption spectra of TCF-BA and TCF-OH in the show any electron paramagnetic resonance (EPR) signal of a phenyl
presence of O•– 2 , H2O2, NO, and ONOO fluxes (Figs. S3–S5) showed
• −
radical adduct, while under the same conditions, we observed an intense
that TCF-OH dye is consumed in the presence of singlet oxygen or EPR spectrum of a phenyl radical spin adduct when acetylphenylboronic
ONOO− , while the TCF-BA probe responds primarily to ONOO− and, to acid was used instead of TCF-BA (Fig. S6), in agreement with previous
a smaller extent, to H2O2 or •NO-generating system. Therefore, we report [57]. Instead of TCF-NO2 and TCF-H, we identified
decided to identify and quantify the products formed in the reaction TCF-OH-3-NO2 as a minor product formed in the reaction mixture,
between ONOO− and TCF-BA or TCF-OH. A bolus addition of ONOO− to which we attributed to the nitration of the phenolic product formed. We
the solution of the TCF-BA probe led to the appearance of a red color and have confirmed that TCF-OH undergoes nitration to TCF-OH-3-NO2 in
fluorescence (Fig. 1) due to the formation of TCF-OH [3]. the presence of ONOO− (Fig. 2).
Next, we investigated the stoichiometry and products of the reaction Next, we investigated the reaction between the TCF-BA probe and
between the probe and ONOO− (Fig. 2). HPLC analyses confirmed that HOCl. The stoichiometric analysis of this reaction is shown in Fig. 3B.
ONOO− oxidized the TCF-BA probe to the TCF-OH dye. Analysis of the Similar to ONOO− , HOCl oxidized the TCF-BA probe, forming TCF-OH
data in Fig. 2 revealed that complete consumption of the probe required at 1:1 stoichiometry in a 100% yield. As in the case of ONOO− , the
a slight excess of ONOO− . It is well known that ONOO− oxidizes boronic amount of the TCF-OH product decreased when an excess of the oxidant
compounds, giving corresponding phenol as the major, non-radical end was added. The disappearance of TCF-OH in a reaction mixture can be
product (~85–99% yield) [45,47,49,53,57]. The minor, radical attributed to the chlorination reaction leading to the formation of
pathway leads to a phenyl-type radical, nitrogen dioxide (•NO2), and chlorinated derivative(s). This is supported by the detection of a new
stable products formed from them (~1–15% yield) (Scheme 1) [45,47, HPLC peak assigned to TCF-OH-3-Cl (Fig. 3A) of the same retention time
49,53,57,63]. Oxidation of TCF-BA by ONOO− resulted in the formation as the synthesized authentic standard and of the product formed when
TCF-OH in an almost quantitative (99%) yield. Additionally, it is evident TCF-OH was mixed with HOCl. Chlorinated products formed from the
that TCF-OH was consumed when an excess of ONOO− was present in boronate-derived phenols [45,47] or from hydroethidine [64,65] can be
the reaction mixture. In order to clarify the reactions that take place used as specific markers for HOCl.
between ONOO− and the TCF-BA probe or TCF-OH product, we syn­ Because the second-order rate constant of the reaction of arylboronic
thesized the expected stable end-products formed in the radical minor acids with H2O2 is of the order of 1 M− 1s− 1 [47,49], the stoichiometric
pathway, TCF-NO2 (the anticipated product of the recombination of analysis of the reaction between the boronic probe and H2O2 requires

Fig. 1. (A) Absorption and (B) emission


spectra of the TCF-BA probe (100 μM)
before and after mixing it with ONOO−
in a phosphate buffer (pH 7.4, 50 mM)
containing MeCN (50%) and DTPA (10
μM). Absorption and emission spectra
were recorded after 5 × and 50 × dilu­
tion of reaction mixture, respectively.
The final concentration of the TCF-BA
probe was 20 μM (absorption) and 2 μM
(emission, λexc 550 nm). (C) Absorption
spectra of the TCF-BA probe (25 μM)
and TCF-OH dye (30 μM). (D) Emission
spectrum of the TCF-OH dye (2 μM) in a
phosphate buffer (pH 7.4, 50 mM) con­
taining of MeCN (50%) (λexc 550 nm).

6
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

Fig. 2. Reaction between the TCF-BA probe and ONOO− in an aqueous solution containing a phosphate buffer (pH 7.4, 50 mM), DTPA (10 μM), and MeCN (5%). (A)
HPLC chromatograms of the standards (10 μM) and reaction mixtures of the TCF-BA probe (20 μM) and TCF-OH dye (20 μM) with ONOO− . The traces were collected
using an absorption detector set at 440 nm. (B) HPLC-based titration of the TCF-BA probe (20 μM) with ONOO− .

Fig. 3. Reaction between the TCF-BA probe and HOCl in an aqueous solution containing a phosphate buffer (pH 7.4, 50 mM), DTPA (10 μM), and MeCN (5%). (A)
HPLC chromatograms of the standards and reaction mixtures of the TCF-BA probe (20 μM) and TCF-OH dye (20 μM) with HOCl. The traces were collected using an
absorption detector set at 440 nm. (B) HPLC-based titration of the TCF-BA probe (20 μM) with HOCl.

the reaction mixture to have a long (≥24 h) incubation time, especially determined. Using time-resolved techniques, pulse radiolysis, and
for low concentrations of H2O2, to ensure completion of the reaction. stopped-flow, we determined the second-order rate constants for the
Tests carried out under analogous conditions as for HOCl and ONOO− (i. reaction of the TCF-BA probe with ONOO− (Fig. 5A) and HOCl (Fig. 5B),
e., for a 20 μM concentration of the probe in a phosphate buffer with 5% respectively. Because the reaction of boronates with H2O2 is relatively
of MeCN) showed the instability of TCF-BA during a 24-h incubation slow [45,47,49], time resolution of regular spectrophotometers is suf­
period. Since it was difficult to precisely determine the concentration of ficient to monitor the reaction kinetics of TCF-BA with H2O2 (Fig. 5C).
the residual boronic probe, we repeated experiments with a higher Table 1 summarizes the results of the kinetic experiments. ONOO− ox­
concentration of the probe in an aqueous solution containing a phos­ idizes the TCF-BA probe with a higher rate constant of (1.8 ± 0.3) × 106
phate buffer and MeCN (50%). Under those conditions, we observed that M− 1s− 1 than HOCl of (2.4 ± 0.1) × 104 M− 1s− 1 or H2O2 of (1.9 ± 0.4)
30% of the TCF-BA undergoes spontaneous decomposition, but an M− 1s− 1. These rate constants are typical for aromatic boronic acids [47,
excess of H2O2 still was required for complete consumption of the probe 49,54]. The spectroscopic (Fig. S7) and kinetic (Fig. S8) experiments
(Fig. 4). H2O2 converted TCF-BA to TCF-OH as the sole detectable also showed that TCF-OH reacts with H2O2 with a rate constant of (5.21
product (Fig. 4A). ± 0.04) × 10− 2 M− 1s− 1, which is 40 times lower than that for the re­
Finally, the rate constants of the TCF-BA probe reactions with action between H2O2 and TCF-BA. However, it can be observed that,
selected biological oxidants (ONOO− , HOCl, and H2O2) were upon the reaction of TCF-BA with an excess of H2O2 (100-fold excess),

7
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

Fig. 4. Reaction between the TCF-BA probe and H2O2 in an aqueous solution containing a phosphate buffer (pH 7.4, 50 mM), DTPA (10 μM), and MeCN (50%). (A)
HPLC chromatograms of the standards and reaction mixtures of the TCF-BA probe with H2O2. (B) HPLC-based titration of the TCF-BA probe (100 μM) with H2O2. The
traces were collected using an absorption detector set at 440 nm.

Fig. 5. Kinetics of the reaction between TCF-BA and studied oxidants. The pseudo-first order rate constants were obtained by fitting the build-up of the signal
recorded at 585 nm. (A) The dependence of pseudo-first order rate constants of the reaction of TCF-BA with ONOO− on the TCF-BA concentration. The pseudo-first
order rate constants were obtained by fitting the build-up of the signal at 585 nm obtained by pulse radiolysis of aqueous solution containing TCF-BA (10–30 μM), 10
mM NaNO2, 1 M CH3OH, and a phosphate buffer (pH 7.4, 50 mM). The optical length of the cuvette was 1 cm, and the solutions received a radiation dose of 20 Gy.
(B) The dependence of the pseudo-first order rate constants of the reaction of TCF-BA probe with HOCl on the initial concentration of HOCl. (C) The dependence of
pseudo-first order rate constants of the TCF-BA reaction with H2O2 on the initial concentration of H2O2. Solutions consisted of TCF-BA (20 μM), a phosphate buffer
(pH 7.4, 50 mM), DTPA (100 μM), MeCN (10%), and H2O2 (200–400 μM).

TCF-OH was formed and subsequently underwent consumption measurements (Fig. 6A and B) revealed that, as the concentration of GSH
(Fig. S7). in the solution increases, the intensity of the absorption bands of the
TCF-BA probe and TCF-OH dye decreases. Moreover, a new absorption
band with a maximum at 339 nm appears. In order to confirm that GSH
3.4. Oxidation of the probe in the presence of GSH forms an adduct to the double bound, we synthesized a TCF-Etx dye (see
Scheme 2). Because TCF-Etx, an analogue without a styryl unit, behaved
The use of a fluorogenic probe (e.g., TCF-BA) for the determination of in a different way (Fig. 6F), we conclude that the studied TCF-BA probe
biologically important oxidants (ONOO− , HOCl, H2O2) also requires the and TCF-OH dye might react with GSH via a Michael addition to the
recognition of the reactivity of the probe and the resulting fluorescent double bond, yielding two diastereoisomers of GSH adduct TCF-SG (R)
product (e.g., TCF-OH) with tripeptide GSH (L-γ-glutamyl-L-cysteinyl- and TCF-SG (S) (Fig. 6E). Such reactivity was utilized recently in
glycine), which serves as a cellular low molecular antioxidant and a GSH-sensing probe developed by Peng Hou and co-workers [13].
substrate for antioxidant enzymes (glutathione peroxidase and gluta­ Based on UV–vis absorption spectra of TCF-BA and TCF-OH recorded
redoxin). Based on our experience developing probes for thiols [66–69], at different GSH concentrations (Fig. 6A and B), the dissociation con­
recent report on the TCF-OH fluorescence decrease in the presence of stants Kdis for TCF-BA-SG and TCF-OH-SG were estimated at 5.6 × 10− 5
GSH, and taking into account high intracellular concentration (1–10 and 5.6 × 10− 4, respectively (Fig. S9). The comparison of these disso­
mM) of GSH, we anticipated that GSH may form a Michael adduct with ciation constants allows the conclusion that the oxidation of the boronic
tricyanofuran-based probes and dyes (e.g., TCF-BA and TCF-OH). The acid group in the TCF-BA-SG adduct at submillimolar concentrations of
electron-withdrawing group (e.g., TCF) conjugated to the carbon–carbon GSH may result in partial release the TCF-OH fluorescent dye.
double (C–– C) bond is a suitable site to attach nucleophiles, for example,
The formation of the TCF-BA-SG and TCF-OH-SG adducts may not
GSH. Such reactions are usually catalazed by presence of base in solu­ only affect the probe fluorescence response but also its reactivity toward
tions [13,70–72]. Therefore, we determined whether GSH interacts with oxidants. Therefore, we decided to study the effect of GSH on TCF-BA
the TCF-BA probe and TCF-OH product. The spectrophotometric

8
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

Fig. 6. Electronic absorption spectra of TCF-derived probes (A) TCF-BA (25 μM) and (C) TCF-BA-2 (20 μM), and dyes (B) TCF-OH (8 μM), (D) TCF-OH-2 (6 μM), and
(F) TCF-Etx (4 μM) before and after mixing them with GSH in a phosphate buffer (pH 7.4, 50 mM) containing of MeCN (5%) and DTPA (10 μM). (E) Stereoisomers
formed in the reaction of TCF-BA with GSH.

oxidation by ONOO− . For comparison, we also measured colorimetric of the reaction mixture of TCF-BOR with ONOO− in the presence of GSH
response of TCF-BA-SG after addition of HOCl or H2O2 under the same (Fig. 7A) show that GSH did not significantly affect the consumption of
conditions. Only bolus addition of ONOO− led to the appearance of a the probe by ONOO− [46]. However, it is evident from Fig. 7B that GSH
slight red color due to the formation of TCF-OH (Fig. S12B). A lower inhibits the consumption the TCF-OH dye by ONOO− . Using the
intensity of color is caused by the subsequent reaction of TCF-OH with competition method, we determined the second-order rate constants for
GSH. The lack of response of the probe to HOCl and H2O2 could be the reaction between ONOO− and the TCF-BA-SG adduct. The adduct is
explained by direct scavenging of those oxidants by GSH. HPLC analyses oxidized by ONOO− with a lower rate constant (6.2 ± 0.6) × 105 M− 1s− 1

Fig. 7. (A) Loss of the TCF-BA probe


concentration in the reaction of 20 μM
TCF-BA with 10 μM ONOO− . (B) Loss of
the TCF-OH dye concentration in the
reaction of 20 μM TCF-OH with 10 μM
ONOO− . Reactions were performed in
an aqueous solution containing 50 mM
PB (pH 7.4), 100 μM DTPA, 10% MeCN,
and without or with the addition of 0.4
mM GSH. The concentrations of TCF-BA
and TCF-OH were determined based on
the HPLC measurements. The traces
were collected using an absorption de­
tector set at 440 nm. Data are the
means ± standard deviation of three
independent experiments.

9
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

and tested the possibility of the formation of a specific nitro-naphthalene


product, anticipated for the potential minor, radical pathway of the
reaction between TCF-BA-2 and ONOO− . The TCF-BA-2 probe reacts
with ONOO− with a rate constant of (6.0 ± 0.2) × 105 M− 1s− 1, and HPLC
analyses (Fig. 11) showed formation of the TCF-OH-2 dye as a sole re­
action product. Moreover, similar to the results observed for TCF-BA,
the EPR spin trapping experiment with 2-methyl-2-nitroso-propane
(Fig. S6, trace c) did not produce detectable levels of the spin adduct.
This indicates that the reactions of both the TCF-BA and TCF-BA-2
probes with ONOO− proceed predominantly via the non-radical
pathway, resulting in the formation of the corresponding phenolic
products.
We also determined whether TCF-BA-2 and TCF-OH-2 can form
corresponding adducts with GSH. Spectrophotometric analyses (Fig. 6C)
showed that the TCF-BA-2 probe forms an adduct with GSH (Kdis = 7.1
× 10− 5), whereas the absorption band of TCF-OH-2 dye did not change
with an increased concentration of GSH (Fig. 6D). Like in the case of
TCF-BA, in the presence of GSH only ONOO− is able to induce the
colorimetric response of the TCF-BA-2 probe (Fig. S13B).
Fig. 8. The dependence of the reciprocal of fluorescence intensity monitored at
Finally, we compared oxidation of the TCF-BA and TCF-BA-2 probes
470 nm on the [TCF-BA-SG]/[CBA] ratio.
in an enzymatic HX/XO/spermine-NONOate system, which produced a
flux of ONOO− in situ (Figs. S14 and S15), and tested the effect of the
(Fig. 8, Table 2) than the TCF-BA probe, but the rate constant still is high presence of GSH on the signal detected. Again, it can be clearly seen in
enough to outcompete other pathways of ONOO− decay, including Fig. 12 that the TCF-BA probe response is attenuated in the presence of
self-decay and scavenging by GSH. Collectively, these results demon­ GSH whereas the colorimetric monitoring of the oxidation of the TCF-
strate that, at physiologically relevant GSH concentrations, the TCF-BA BA-2 probe by ONOO− is not affected by GSH. The lower fluorescent
probe may undergo oxidation to TCF-OH only in a reaction with intensity observed in an enzymatic system with GSH is caused by the
ONOO− . formation of the TCF-OH-SG adduct.
Because the signal intensity from the TCF-BA probe depends on the
amount of GSH (Fig. 6), we decided to modify the structure of the probe
4. Discussion
in an attempt to minimize the effect of GSH. The reaction between GSH
and hydroxyl bearing styryl compound is more probable for phenol state
Identification and quantification of ROS/RNS is extremely impor­
than phenolate. Because of that we searched for analogical dye which
tant, since these species play a crucial role in many pathophysiological
has much lower pKa and a good solubility in a aqueous solution con­
processes. For example, depending on their concentrations, ROS may
taining a small amount of organic solvent (e.g. ~ 10% MeCN). The latter
show antitumor activity or may initiate cancer proliferation, angiogen­
requirement resulted in the resignation from the synthesis of the probe
esis, immune evasion, and tumor growth and metastasis. Understanding
based on the TCF–OH–3-NO2 dye. TCF-OH-2 with naphthalene moiety
the mechanisms of ROS/RNS formation and localization may allow the
fulfilled all the requirements, i.e. pK = 5.3 [9] and sufficient solubility in
development of novel, effective redox-active drugs that inhibit tumor
our system. We synthesized the TCF-BA-2 probe and TCF-OH-2 dye
growth and prevent metastasis. A commonly used method for detecting
(Scheme 2), and, like before, we spectroscopically characterized those
ROS/RNS in vitro and in vivo is the bioluminescence (BLI) or fluorescence
compounds and measured the reactivity of the TCF-BA-2 probe toward
imaging (FLI). BLI and FLI are based on suitable probes that, upon re­
ONOO− . The spectroscopic studies (Table 1, Fig. 9C and D) revealed that
action with an oxidant, release a luciferase bioluminescent substrate or a
replacing the benzene ring with a naphthalene moiety led to the
highly fluorescent compound, respectively. The appearance of the
red-shift of the absorption and fluorescence emission bands of
bioluminescence or fluorescence signal and its intensity reflect the
TCF-OH-2 in comparison with TCF-OH. After adding ONOO− to the
presence and amount of the oxidant, respectively. However, interpre­
solution of TCF-BA-2, we observed the solution become an intense blue
tation of the BLI or FLI signal is not trivial, and several factors should be
color due to the formation of TCF-OH-2 dye (Fig. 9A and Fig. S13A). The
considered. First, a significant challenge is to develop a selective probe
lower value of the Φfl of TCF-OH-2, in comparison with TCF-OH, but
that reacts with the oxidant with a 1:1 stoichiometry. Second, the rate
high value of the extinction coefficient suggest that TCF-BA-2 may
constant of the reaction of the oxidant of interest with the probe should
perform better as a colorimetric probe than as a fluorogenic sensor.
be high enough to enable a sufficient accumulation of the detectable
As for the TCF-BA probe, we determined the second-order rate
oxidation product, in competition with the oxidant’s reaction with other
constants for the reaction of ONOO− with TCF-BA-2 (Table 1, Fig. 10)
cellular targets (e.g., GSH, thiols, CO2, Prx, heme-containing proteins
[73,74]). Third, for in vivo applications, emission at far-red and/or at
NIR wavelengths, which are minimally absorbed by tissue, is preferred.
Table 2 For these reasons, the most promising luminogenic and fluorogenic
Determined values of the second-order rate constants for the reaction of TCF-BA
probes for the detection of several non-radical ROS/RNS, produced
with different oxidants and for the reaction of the TCF-BA-SG adduct and TCF-
under inflammatory conditions (i.e., H2O2, HOCl, ONOO− ), are based on
BA-2 with ONOO− at pH 7.4
the oxidative conversion of arylboronic acids or esters [75]. Masking of
ONOO− HOCl H2O2
hydroxyl or amino group(s) of a fluorophore or firefly luciferin de­
− 1 − 1
k (M s ) at 25 C ◦
rivatives by boronate or boronobenzyl moieties turns off the fluores­
TCF-BA (1.8 ± 0.3) × 106a (2.4 ± 0.1) × 104b (1.9 ± 0.4)c cence [46,48,53–55,58] or luminescence [59], respectively. The
TCF-BA-SG (6.2 ± 0.6) × 105b reaction of boronic/boronate group with H2O2, HOCl, and ONOO−
TCF-BA-2 (6.0 ± 0.2) × 105a proceeds with a 1:1 stoichiometry and uncages the hydroxyl group and
a
Pulse radiolysis. thus turns on the fluorescence/luminescence [46,48,53–55,58,59].
b
Stopped flow. Under physiological pH, boronate probes are oxidized by HOCl and
c
Regular spectrophotometer. ONOO− at a significantly higher rate than is the case of H2O2 [47,49,55].

10
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

Fig. 9. (A) Absorption and (B) emission


spectra of probe TCF-BA-2 (100 μM)
before and after mixing it with ONOO−
in a phosphate buffer (pH 7.4, 50 mM)
containing of MeCN (50%) and DTPA
(10 μM). Absorption and emission
spectra were recorded after 10 × and
50 × dilution of reaction mixture,
respectively. The final concentration of
the TCF-BA-2 probe was 18 μM (ab­
sorption) and 2 μM (emission, λexc 625
nm). (C) Absorption spectra of the TCF-
BA-2 probe (10 μM) and TCF-OH-2 dye
(6 μM). (D) Emission spectrum of the
TCF-OH-2 dye (2 μM) in a phosphate
buffer (pH 7.4, 50 mM) containing of
MeCN (20%) (λexc 625 nm).

Fig. 10. Kinetics of the reaction be­


tween TCF-BA-2 and ONOO− . (A) Ki­
netic traces of TCF-OH-2 build-up were
observed at 630 nm. (B) The depen­
dence of pseudo-first-order rate con­
stants of the reaction of TCF-BA-2 with
ONOO− on the TCF-BA-2 concentration.
The pseudo-first-order rate constants
were obtained by fitting the build-up of
the signal at 585 nm obtained by pulse
radiolysis of aqueous solution contain­
ing TCF-BA-2 (20–40 μM), 10 mM
NaNO2, 1 M CH3OH, and a phosphate
buffer (pH 7.4, 50 mM). The optical
length of the cuvette was 1 cm, and so­
lutions received a radiation dose of 20
Gy.

Several boronate probes were synthesized and characterized as far-red (TCF-NO2, TCF-H) were not detected. Spin trapping experiments also
or NIR fluorogenic probes for H2O2, HOCl, or ONOO− [75]. confirm that ONOO− -induced TCF-BA oxidation proceeds without
TCF-BA was developed as a selective probe for H2O2 [4,17], HOCl radical intermediates. This makes the TCF-BA probe similar to fluores­
[5], or ONOO− [3]. The data presented here show that this probe can be cein boronate (Fl-B) [47], which produced fluorescein in a 99% yield in
oxidized by all three oxidants with a 1:1 stoichiometry in GSH-free so­ the reaction with ONOO− [76–78].
lutions. HPLC analyses confirmed formation of TCF-OH dye as a sole In the case of excess of ONOO− and HOCl, we detect nitrated (TCF-
oxidation product. The anticipated phenyl radical-derived products OH-3-NO2) and chlorinated (TCF-OH-3-Cl) products, respectively.

11
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

Fig. 11. The reaction between the TCF-BA-2 probe and ONOO− in an aqueous solution containing a phosphate buffer (pH 7.4, 50 mM), DTPA (10 μM), and MeCN
(5%). (A) HPLC chromatograms of the standards (10 μM) and reaction mixtures of the TCF-BA-2 probe (20 μM) and TCF-OH-2 dye (20 μM) with ONOO− . The traces
were collected using an absorption detector set at 440 nm. (B) HPLC-based titration of the TCF-BA-2 probe (20 μM) with ONOO− .

Fig. 12. Oxidation of TCF-BA and TCF-


BA-2 probes by ONOO− generated from
co-generated O•–
2 from HX/XO, and NO

from spermine-NONOate. (A) The real-


time recorded fluorescent signal of
TCF-OH dye formed during TCF-BA
probe oxidation by ONOO− . (λexc 535
nm) (B) The real-time recorded absorp­
tion signal of TCF-OH-2 dye formed
during TCF-BA-2 probe oxidation by
ONOO− . Reactions were conducted in a
phosphate buffer (pH 7.4, 50 mM) con­
taining DTPA (100 μM), HX/XO and
spermine-NONOate (black) and with the
addition of 400 μM of GSH (red).

These compounds have similar spectroscopic characteristics as TCF-OH, accordingly adjusted.


which means that HPLC or liquid chromatography–mass spectrometry We have also identified a TCF-based boronate probe containing the
(LC-MS) should be used for their detection and identification. Among naphthalene moiety, TCF-BA-2, whose oxidative response is indepen­
the oxidants tested, the fastest colorimetric and fluorescence response of dent of GSH levels. TCF-BA-2 is oxidized to the appropriate 1-naphthol-
TCF-BA was observed for ONOO− . Like the boronates previously tested derived dye (TCF-OH-2) with a lower pKa (pKa = 5.3 [9]) than TCF-OH
[47,49,53,55], TCF-BA reacts with ONOO− with a rate constant several dye (pKa = 7.6 [9]). An alternative approach based on chlorination of
orders of magnitude greater than for H2O2 or HOCl (Table 1). The high the phenyl ring, resulting in a lower pKa value of the OH group in the
rate constant of the reaction of the TCF-BA probe with ONOO− should TCF-OH fluorophore, has been previously proposed to prevent its re­
enable the detection of ONOO− in the presence of its physiological action with GSH [11]. A negative charge of ionized hydroxyl group O−
scavengers (e.g., CO2 kONOO ~1 × 104 M− 1s− 1 [79], GSH kONOO = 1.36 can delocalize within chromophore. Thus, diminished partial positive
× 103 M− 1s− 1 [80], and albumin kONOO = 8.31 × 103 M− 1s− 1 [81]), with charges at the carbon atoms of the central double bond prevent forma­
the caveat that signal intensity may depend on cellular nucleophile tion of adducts with GSH. We propose to use the TCF-BA-2 probe as a
levels (GSH and other biothiols). We demonstrated that thiols form highly sensitive colorimetric probe for ONOO− in the presence or
Michael adduct with both TCF-BA and TCF-OH. Therefore, the quanti­ absence of GSH.
fication of ONOO− with the aid of the TCF-BA probe in cell culture In summary, we have characterized chemical reactivity of a TCF-
should be done using HPLC or LC-MS to take into account effects of GSH based boronate redox probe, TCF-BA, toward ONOO− , HOCl, and
and other nucleophiles on the signal intensity from TCF-OH. These re­ H2O2, and identified reactions (Scheme 4) affecting the fluorescent
sults also suggest that other chemical probes built on a fluorophore response of the probe. The red fluorescent product TCF-OH is formed in
similar to TCF-OH may also be reactive towards GSH, and thus, such a the reaction of TCF-BA with all three oxidants. In the case of ONOO−
possibility should be experimentally tested and data interpretation and HOCl, TCF-OH may undergo subsequent reactions to form nitrated

12
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

Scheme 4. Oxidative conversion of the studied TCF-BA and TCF-BA-2 probes and subsequent reaction.

TCF-OH-3-NO2 and chlorinated product TCF-OH-3-Cl, respectively. We acquisition, Writing – review & editing. Jan Grzegorz Malecki: Meth­
demonstrated that both the TCF-BA probe and TCF-OH dye react with odology, Investigation. Jacek Zielonka: Conceptualization, Investiga­
GSH to form TCF-BA-SG and TCF-OH-SG adducts, which affects the tion, Methodology, Formal analysis, Resources, Writing – review &
chemical reactivity and colorimetric and fluorescence response of the editing, Supervision, Funding acquisition. Radosław Podsiadły:
probe. Although the TCF-BA-2 probe also reacts with GSH, the absorp­ Conceptualization, Investigation, Methodology, Resources, Writing –
tion spectrum of its oxidation product, TCF-OH-2, is not influenced by review & editing, Funding acquisition, Supervision. All authors
GSH and, therefore, can be applied for real-time monitoring of ONOO− reviewed the results and approved the final version of the manuscript.
formation in the presence of GSH.
Declaration of competing interest
Author contributions
The authors declare that they have no known competing financial
Przemysław Siarkiewicz: Investigation, Methodology, Formal anal­ interests or personal relationships that could have appeared to influence
ysis. Radosław Michalski: Methodology, Formal analysis, Funding the work reported in this paper.
acquisition. Adam Sikora: Methodology, Investigation, Formal analysis,
Funding acquisition. Renata Smulik-Izydorczyk: Methodology, Investi­
Acknowledgements
gation, Formal analysis. Marcin Szala: Methodology, Investigation.
Aleksandra Grzelakowska: Methodology, Investigation. Julia Mod­
This work was supported by the Polish National Science Centre
rzejewska: Investigation. Asha Bailey: Investigation. Jacek E. Nycz:
within the SONATA BIS 6 program (grant no. 2016/22/E/ST4/00549 to
Methodology, Investigation. Balaraman Kalyanaraman: Funding
R.P.), the Polish National Science Centre within the SONATA BIS

13
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

program (grant no. 2015/18/E/ST4/00235 to A.S.), the Polish National and tissues. Sensor Actuator B Chem 2019;297:126731. https://doi.org/10.1016/j.
snb.2019.126731.
Science Centre within the SONATA program (grant no. 2018/31/D/
[19] Hao Y, Zhang Y, Ruan K, Meng F, Li T, Guan J, Du L, Qu P, Xu M. A highly selective
ST4/03494 to R.M.), Institutional Research Grant IRG #16-183-31 from long-wavelength fluorescent probe for hydrazine and its application in living cell
the American Cancer Society and the MCW Cancer Center to J.Z., and imaging. Spectrochim Acta 2017;184:355–60. https://doi.org/10.1016/j.
NIH NCI (grant no. R01 CA208648) to B.K. saa.2017.04.041.
[20] Choudhury R, Patel SR, Ghosh A. Selective detection of human serum albumin by
This work was supported by the National Center for Research and near infrared emissive fluorophores: insights into structure-property relationship.
Development (Warsaw, Poland) within the grant InterChemMed J Photochem Photobiol Chem 2019;376:100–7. https://doi.org/10.1016/j.
(POWR.03.02.00–00-I029/16). jphotochem.2019.02.038.
[21] Meng Y-L, Xin Z-H, Jia Y-J, Kang Y-F, Ge L-P, Zhang C-H, Dai M-Y. A near-infrared
fluorescent probe for direct and selective detection of cysteine over homocysteine
Appendix A. Supplementary data and glutathione. Spectrochim Acta 2018;202:301–4. https://doi.org/10.1016/j.
saa.2018.05.060.
[22] Zhang B, Zhang H, Zhong M, Wang S, Xu Q, Cho D-H, Qiu H. A novel off-on
Supplementary data to this article can be found online at https://doi. fluorescent probe for specific detection and imaging of cysteine in live cells and in
org/10.1016/j.dyepig.2021.109405. vivo. Chin Chem Lett 2020;31:133–5. https://doi.org/10.1016/j.
cclet.2019.05.061.
[23] Zhu B, Kan H, Liu J, Liu H, Wei Q, Du B. A highly selective ratiometric visual and
References red-emitting fluorescent dual-channel probe for imaging fluoride anions in living
cells. Biosens Bioelectron 2014;52:298–303. https://doi.org/10.1016/j.
[1] Yuan L, Lin WY, Zheng KB, He LW, Huang WM. Far-red to near infrared analyte- bios.2013.09.010.
responsive fluorescent probes based on organic fluorophore platforms for [24] Xu J, Wang Z, Liu C, Xu Z, Zhu B, Wang N, Wang K, Wang J. A colorimetric and
fluorescence imaging. Chem Soc Rev 2013;42:622–61. https://doi.org/10.1039/ fluorescent probe for the detection of Cu2+ in a complete aqueous solution. Anal
C2CS35313J. Sci 2018;34:453–7. https://doi.org/10.2116/analsci.17P517.
[2] Weissleder R. A clearer vision for in vivo imaging. Nat Biotechnol 2001;19:316–7. [25] Zhu B, Wang W, Liu L, Jiang H, Du B, Wei Q. A highly selective colorimetric and
https://doi.org/10.1038/86684. long-wavelength fluorescent probe for Hg2+. Sensor Actuator B Chem 2011;191:
[3] Sedgwick AC, Han HH, Gardiner JE, Bull SD, He XP, James TD. Long-wavelength 605–11. https://doi.org/10.1016/j.snb.2013.10.028.
fluorescent boronate probes for the detection and intracellular imaging of [26] Teng X, Tian M, Zhang J, Tang L, Xin J. A TCF-based colorimetric and fluorescent
peroxynitrite. Chem Commun 2017;53:12822–5. https://doi.org/10.1039/ probe for palladium detection in an aqueous solution. Tetrahedron Lett 2018;59:
c7cc07845e. 2804–8. https://doi.org/10.1016/j.tetlet.2018.06.016.
[4] Choudhury R, Ricketts AT, Molina DG, Paudel P. A boronic acid based [27] Gwynne L, Sedgwick AC, Gardiner JE, Williams GT, Kim G, Lowe JP, Maillard J-Y,
intramolecular charge transfer probe for colorimetric detection of hydrogen Jenkins ATA, Bull SD, Sessler JL, Yoon J, James TD. Long wavelength TCF-based
peroxide. Tetrahedron Lett 2019;60:151258. https://doi.org/10.1016/j. fluorescent probe for the detection of alkaline phosphatase in live cells. Front Chem
tetlet.2019.151258. 2019;7:255. https://doi.org/10.3389/fchem.2019.00255.
[5] Shu W, Yan L, Wang Z, Liu J, Zhang S, Liu C, Zhu B. A novel visual and far-red [28] Feng L, Liu Z-M, Xu L, Lv X, Ning J, Hou J, Ge G-B, Cui J-N, Yang L. A highly
fluorescent dual-channel probe for the rapid and sensitive detection of selective long-wavelength fluorescent probe for the detection of human
hypochlorite in aqueous solution and living cells. Sens Actuators, B 2015;221: carboxylesterase 2 and its biomedical applications. Chem Commun 2014;50:
1130–6. https://doi.org/10.1016/j.snb.2015.07.066. 14519–22. https://doi.org/10.1039/C4CC06642A.
[6] Wang YR, Feng L, Xu L, Li Y, Wang DD, Hou J, Zhou K, Jin Q, Ge GB, Cui JN, [29] Radi R. Oxygen radicals, nitric oxide, and peroxynitrite: redox pathways in
Yang L. A rapid-response fluorescent probe for the sensitive and selective detection molecular medicine. Proc Natl Acad Sci USA 2018;115:5839–48. https://doi.org/
of human albumin in plasma and cell culture supernatants. Chem Commun 2016; 10.1073/pnas.1804932115.
52:6064–7. https://doi.org/10.1039/C6CC00119J. [30] Gorrini C, Harris IS, Mak TW. Modulation of oxidative stress as an anticancer
[7] Zhou X, Su F, Tian Y, Youngbull C, Johnson RH, Meldrum DR. A new highly strategy. Nat Rev Drug Discov 2013;12:931–47. https://doi.org/10.1038/nrd4002.
selective fluorescent K+ sensor. J Am Chem Soc 2011;133:18530–3. https://doi. [31] Chio IIC, Tuveson DA. ROS in Cancer: the burning question. Trends Mol Med 2017;
org/10.1021/ja207345s. 23:411–29. https://doi.org/10.1016/j.molmed.2017.03.004.
[8] Lord SJ, Consley NR, Lee HD, Samuel R, Liu N, Twieg RJ, Moerner WE. [32] Idelchik M, Begley U, Begley TJ, Melendez JA. Mitochondrial ROS control of
A photoactivatable push-pull fluorophore for single-molecule imaging in live cells. cancer. Semin Canc Biol 2017;47:57–66. https://doi.org/10.1016/j.
J Am Chem Soc 2008;130:9204–5. https://doi.org/10.1021/ja802883k. semcancer.2017.04.005.
[9] Ipuy M, Billon C, Micouin G, Samarut J, Andraud C, Bretonnière Y. Fluorescent [33] Weinberg F, Hamanaka R, Wheaton WW, Weinberg S, Joseph J, Lopez M,
push-pull pH-responsive probes for ratiometric detection of intracellular pH. Org Kalyanaraman B, Mutlu GM, Budinger GR, Chandel NS. Mitochondrial metabolism
Biomol Chem 2014;12:3641–8. https://doi.org/10.1039/C4OB00147H. and ROS generation are essential for Kras-mediated tumorigenicity. Proc Natl Acad
[10] Nguyen KH, Hao Y, Zeng K, Wei X, Yuan S, Li F, Fan S, Xu M, Liu Y. A reaction- Sci USA 2010;107:8788–93. https://doi.org/10.1073/pnas.1003428107.
based long-wavelength fluorescent probe for Cu2+ detection and imaging in living [34] Diebold L, Chandel NS. Mitochondrial ROS regulation of proliferating cells. Free
cells. J Photochem Photobiol Chem 2018;358:201–6. https://doi.org/10.1016/j. Radic Biol Med 2016;100:86–93. https://doi.org/10.1016/j.
jphotochem.2018.03.023. freeradbiomed.2016.04.198.
[11] Sedgwick AC, Gardiner JE, Kim G, Yevglevskis M, Lloyd MD, Jenkins ATA, Bull SD, [35] Jia P, Dai C, Cao P, Sun D, Ouyang R, Miao Y. The role of reactive oxygen species in
Yoon J, James TD. Long-wavelength TCF-based fluorescence probes for the tumor treatment. RSC Adv 2020;10:7740–50. https://doi.org/10.1039/
detection and intracellular imaging of biological thiols. Chem Commun 2018;54: c9ra10539e.
4786–9. https://doi.org/10.1039/C8CC01661E. [36] Aggarwal V, Tuli HS, Varol A, Thakral F, Yerer MB, Sak K, Varol M, Jain A,
[12] Liu C, Wang Y, Tang C, Liu F, Ma Z, Zhao Q, Wang Z, Zhu B, Zhang X. A reductant- Khan MA, Sethi G. Role of reactive oxygen species in cancer progression: molecular
resistant ratiometric, colorimetric and far-red fluorescent probe for rapid and mechanisms and recent advancements. Biomolecules 2019;9:735–61. https://doi.
ultrasensitive detection of nitroxyl. J Mater Chem B 2017;5:3557–64. https://doi. org/10.3390/biom9110735.
org/10.1039/C6TB03359H. [37] Cheng G, Pan J, Podsiadly R, Zielonka J, Garces AM, Machado LGDD, Bennett B,
[13] Hou P, Sun J, Wang H, Liu L, Zou L, Chen S. TCF-imidazo[1,5-α]pyridine: a McAllister D, Dwinell MB, You M, Kalyanaraman B. Increased formation of reactive
potential robust ratiometric fluorescent probe for glutathione detection with high oxygen species during tumor growth: ex vivo low-temperature EPR and in vivo
selectivity. Sens Actuators, B 2019;304:127244. https://doi.org/10.1016/j. bioluminescence analyses. Free Radic Biol Med 2020;147:167–74. https://doi.org/
snb.2019.127244. 10.1016/j.freeradbiomed.2019.12.020.
[14] Zhu Y, Han J, Zhang Q, Zhao Z, Wang J, Xu X, Hao H, Zhang J. A highly selective [38] Wang J, Yi J. Cancer cell killing via ROS: to increase or decrease, that is the
fluorescent probe for human NAD(P)H:quinone oxidoreductase 1 (hNQO1) question. Canc Biol Ther 2008;7:1875–84. https://doi.org/10.4161/cbt.7.12.7067.
detection and imaging in living tumor cells. RSC Adv 2019;9:26729–33. https:// [39] Omenn GS, Goodman GE, Thornquist MD, Balmes J, Cullen MR, Glass A, Keogh JP,
doi.org/10.1039/C9RA05650E. Meyskens FL, Valanis B, Williams JH, Barnhart S, Hammar S. Effects of a
[15] Yang M, Fan J, Sun W, Du J, Long S, Peng X. Simultaneous visualization of combination of beta carotene and vitamin A on lung cancer and cardiovascular
cysteine/homocysteine and glutathione in living cells and Daphnia magna via dual- disease. N Engl J Med 1996;334:1150–5. https://doi.org/10.1056/
signaling fluorescent chemosensor. Dyes Pigments 2019;168:189–96. https://doi. NEJMcibr1405701.
org/10.1016/j.dyepig.2019.04.056. [40] Piskounova E, Agathocleous M, Murphy MM, Hu Z, Huddlestun SE, Zhao Z,
[16] Wang Z, Zhao Z, Liu C, Geng Z, Duan Q, Jia P, Li Z, Zhu H, Zhu B, Sheng W. A long- Leitch AM, Johnson TM, DeBerardinis RJ, Morrison SJ. Oxidative stress inhibits
wavelength ultrasensitive colorimetric fluorescent probe for carbon monoxide distant metastasis by human melanoma cells. Nature 2015;527:186–91. https://
detection in living cells. Photochem Photobiol Sci 2019;18:1851–7. https://doi. doi.org/10.1038/nature15726.
org/10.1039/C9PP00222G. [41] Vance TM, Su J, Fontham ET, Koo SI, Chun OK. Dietary antioxidants and prostate
[17] Xu F, Tang W, Kang S, Song J, Duan X. A highly sensitive and photo-stable cancer: a review. Nutr Canc 2013;65:793–801. https://doi.org/10.1080/
fluorescent probe for endogenous intracellular H2O2 imaging in live cancer cells. 01635581.2013.806672.
Dyes Pigments 2018;153:61–6. https://doi.org/10.1016/j.dyepig.2018.02.015. [42] Hardy M, Zielonka J, Karoui H, Sikora A, Michalski R, Podsiadły R, Lopez M,
[18] Liu C, Jia P, Wu L, Li Z, Zhu H, Wang Z, Deng S, Shu W, Zhang X, Yu Y, Zhu B. Vasquez-Vivar J, Kalyanaraman B, Ouari O. Antioxidants Redox Signal 2018;28:
Rational design of a highly efficient two-photon fluorescent probe for tracking 1416–32. https://doi.org/10.1089/ars.2017.7398.
intracellular basal hypochlorous acid and its applications in identifying tumor cells

14
P. Siarkiewicz et al. Dyes and Pigments 192 (2021) 109405

[43] Andina D, Leroux J-C, Luciani P. Ratiometric fluorescent probes for the detection of [62] Abdel-Mottaleb MSA, El-Sayed BA, Abo-Aly MM, El-Kady MY. Fluorescence-
reactive oxygen species. Chem Eur J 2017;23:13549–73. https://doi.org/10.1002/ properties and excited state interactions of 7-hydroxy-4-methylcoumarin laser dye.
chem.201702458. J Photochem Photobiol Chem 1989;46:379–90. https://doi.org/10.1016/1010-
[44] Xing P, Feng Y, Niu Y, Li Q, Zhang Z, Dong L, Wang C. A water-soluble, two-photon 6030(89)87054-6.
probe for imaging endogenous hypochlorous acid in live tissue. Chem Eur J 2018; [63] Sikora A, Zielonka J, Adamus J, Debski D, Dybala-Defratyka A, Michalowski B,
24:5748–53. https://doi.org/10.1002/chem.201800249. Joseph J, Hartley RC, Murphy MP, Kalyanaraman B. Reaction between
[45] Zielonka J, Podsiadły R, Zielonka M, Hardy M, Kalyanaraman B. On the use of peroxynitrite and triphenylphosphonium-substituted arylboronic acid isomers:
peroxy-caged luciferin (PCL-1) probe for bioluminescent detection of inflammatory identification of diagnostic marker products and biological implications. Chem Res
oxidants in vitro and in vivo - identification of reaction intermediates and oxidant- Toxicol 2013;26:856–67. https://doi.org/10.1021/tx300499c.
specific minor products. Free Radic Biol Med 2016;99:32–42. https://doi.org/ [64] Maghzal GJ, Cergol KM, Shengule SR, Suarna C, Newington D, Kettle AJ, Payne RJ,
10.1016/j.freeradbiomed.2016.07.023. Stocker R. Assessment of myeloperoxidase activity by the conversion of
[46] Dębowska K, Dębski D, Michałowski B, Dybala-Defratyka A, Wójcik T, Michalski R, hydroethidine to 2-chloroethidium. J Biol Chem 2014;289:5580–95. https://doi.
Jakubowska M, Selmi A, Smulik R, Piotrowski Ł, Adamus J, Marcinek A, org/10.1074/jbc.M113.539486.
Chlopicki S, Sikora A. Characterization of fluorescein-based monoboronate probe [65] Talib J, Maghzal GJ, Cheng D, Stocker R. Detailed protocol to assess in vivo and ex
and its application to the detection of peroxynitrite in endothelial cells treated with vivo myeloperoxidase activity in mouse models of vascular inflammation and
doxorubicin. Chem Res Toxicol 2016;29:735–46. https://doi.org/10.1021/acs. disease using hydroethidine. Free Radic Biol Med 2016;97:124–35. https://doi.
chemrestox.5b00431. org/10.1016/j.freeradbiomed.2016.05.004.
[47] Rios N, Piacenza L, Trujillo M, Martínez A, Demicheli V, Prolo C, Álvarez MN, [66] Kolińska J, Grzelakowska A. Characterization of a novel styrylbenzimidazolium
López GV, Radi R. Sensitive detection and estimation of cell-derived peroxynitrite based dye and its application in the detection of biothiols. Luminescence 2020:1–9.
fluxes using fluorescein-boronate. Free Radic Biol Med 2016;101:284–95. https:// https://doi.org/10.1002/bio.3956.
doi.org/10.1016/j.freeradbiomed.2016.08.033. [67] Grzelakowska A, Kolińska J, Mąkiewicz M. The synthesis and spectroscopic
[48] Miller EW, Albers AE, Pralle A, Isacoff EY, Chang CJ. Boronate-based fluorescent characterisation of 3-formyl-2-quinolones in the presence of biothiols. Color
probes for imaging cellular hydrogen peroxide. J Am Chem Soc 2005;127: Technol 2018;134:440–9. https://doi.org/10.1111/cote.12355.
16652–9. https://doi.org/10.1021/ja054474f. [68] Kolińska J, Grzelakowska A, Sokołowska J. Novel 7-maleimido-2(1H)-quinolones
[49] Sikora A, Zielonka J, Lopez M, Joseph J, Kalyanaraman B. Direct oxidation of as potential fluorescent sensors for the detection of sulphydryl groups. Color
boronates by peroxynitrite: mechanism and implications in fluorescence imaging of Technol 2018;134:148–55. https://doi.org/10.1111/cote.12326.
peroxynitrite. Free Radic Biol Med 2009;47:1401–7. https://doi.org/10.1016/j. [69] Kowalska A, Kolińska J, Podsiadły R, Sokołowska J. Dyes derived from 3-formyl-2
freeradbiomed.2009.08.006. (1H)-quinolone - synthesis, spectroscopic characterisation, and their behaviour in
[50] Truzzi DR, Augusto O. Influence of CO2 on hydroperoxide metabolism. In: the presence of sulfhydryl and non-sulfhydryl amino acids. Color Technol 2015;
Vissers MC, Hampton M, Kettle AJ, editors. Hydrogen peroxide metabolism in 131:157–64. https://doi.org/10.1111/cote.12140.
health and disease. Boca Raton, FL: CRC Press; 2017. p. 81–99. [70] Schwöbel JAH, Wondrousch D, Koleva YK, Madden JC, Cronin MTD,
[51] Rios N, Radi R, Kalyanaraman B, Zielonka J. Tracking isotopically labeled oxidants Schüürmann G. Prediction of Michael-type acceptor reactivity toward glutathione.
using boronate-based redox probes. J Biol Chem 2020;295:6665–76. https://doi. Chem Res Toxicol 2010;23:1576–785. https://doi.org/10.1021/tx100172x.
org/10.1074/jbc.RA120.013402. [71] Bernardes A, Pérez CN, Mayer M, da Silva CC, Martins FT, Perjési P. Study of
[52] Michalski R, Zielonka J, Gapys E, Marcinek A, Joseph J, Kalyanaraman B. Real- reactions of two Mannich bases derived of 4’-hydroxychalcones with glutathione
time measurements of aminoacid and protein hydroperoxides using coumarin by RP-TLC, RP-HPLC and RP-HPLC-ESI-MS analysis. J Braz Chem Soc 2017;28:
boronic acid. J Biol Chem 2014;289:22536–53. https://doi.org/10.1074/jbc. 1048–62. https://doi.org/10.21577/0103-5053.20160260.
M114.553727. [72] Baker LMS, Baker PRS, Golin-Bisello F, Schopfer FJ, Fink M, Woodcock SR,
[53] Zielonka J, Sikora A, Hardy M, Joseph J, Dranka BP, Kalyanaraman B. Boronate Branchaud BP, Radi R, Freeman BA. Nitro-fatty acid reaction with glutathione and
probes as diagnostic tools for real time monitoring of peroxynitrite and cysteine. Kinetic analysis of thiol alkylation by a Michael addition reaction. J Biol
hydroperoxides. Chem Res Toxicol 2012;25:1793–9. https://doi.org/10.1021/ Chem 2007;282:31085–93. https://doi.org/10.1074/jbc.M704085200.
tx300164j. [73] Ferrer-Sueta G, Radi R. Chemical biology of peroxynitrite: kinetics, diffusion, and
[54] Zielonka J, Zielonka M, Sikora A, Adamus J, Joseph J, Hardy M, Ouari O, radicals. ACS Chem Biol 2009;4:161–77. https://doi.org/10.1021/cb800279q.
Dranka BP, Kalyanaraman B. Global profiling of reactive oxygen and nitrogen [74] Storkey C, Davies MJ, Pattison DI. Reevaluation of the rate constants for the
species in biological systems: high-throughput real-time analyses. J Biol Chem reaction of hypochlorous acid (HOCl) with cysteine, methionine, and peptide
2012;287:2984–95. https://doi.org/10.1074/jbc.M111.309062. derivatives using a new competition kinetic approach. Free Radic Biol Med 2014;
[55] Zielonka J, Sikora A, Joseph J, Kalyanaraman B. Peroxynitrite is the major species 73:60–6. https://doi.org/10.1016/j.freeradbiomed.2014.04.024.
formed from different flux ratios of co-generated nitric oxide and superoxide: direct [75] Sikora A, Zielonka J, Dębowska K, Michalski R, Smulik-Izydorczyk R, Pięta J,
reaction with boronate-based fluorescent probe. J Biol Chem 2010;285:14210–6. Podsiadły R, Artelska A, Pierzchała K, Kalyanaraman B. Boronate-based probes for
https://doi.org/10.1074/jbc.M110.110080. inflammatory oxidants - a novel class of molecular tools for redox biology. Front
[56] Shu W, Yan L, Wang Z, Liu J, Zhang S, Liu C, Zhu B. A novel visual and far-red Chem 2020;8:843–75. https://doi.org/10.3389/fchem.2020.580899.
fluorescent dual-channel probe for the rapid and sensitive detection of [76] Kalyanaraman B, Darley-Usmar V, Davies KJA, Dennery PA, Forman HJ,
hypochlorite in aqueous solution and living cells. Sensor Actuator B Chem 2015; Grisham MB, Mann GE, Moor K, Roberts II LJ, Ischiropoulos H. Measuring reactive
221:1130–6. https://doi.org/10.1016/j.snb.2015.07.066. oxygen and nitrogen species with fluorescent probes: challenges and limitations.
[57] Sikora A, Zielonka J, Lopez M, Dybala-Defratyka A, Joseph J, Marcinek A, Free Radic Biol Med 2012;52:1–6. https://doi.org/10.1016/j.
Kalyanaraman B. Reaction between peroxynitrite and boronates: EPR spin- freeradbiomed.2011.09.030.
trapping, HPLC analyses, and quantum mechanical study of the free radical [77] Radi R, Cosgrove TP, Beckman JS, Freeman BA. Peroxynitrite-induced luminol
pathway. Chem Res Toxicol 2011;24:687–97. https://doi.org/10.1021/tx100439a. chemiluminescence. Biochem J 1993;290:51–7. https://doi.org/10.1042/
[58] Smulik R, Dębski D, Zielonka J, Michałowski B, Adamus J, Marcinek A, bj2900051.
Kalyanaraman B, Sikora A. Nitroxyl (HNO) reacts with molecular oxygen and [78] Wardman P. Chapter Fourteen. Methods to measure the reactivity of peroxynitrite-
forms peroxynitrite at physiological pH. Biological implications. J Biol Chem 2014; derived oxidants toward reduced fluoresceins and rhodamines. Methods Enzymol
289:35570–81. https://doi.org/10.1074/jbc.M114.597740. 2008;441:261–82. https://doi.org/10.1016/S0076-6879(08)01214-7.
[59] Szala M, Grzelakowska A, Modrzejewska J, Siarkiewicz P, Słowiński D, [79] Denicola A, Freeman BA, Trujillo M, Radi R. Peroxynitrite reaction with carbon
Świerczyńska M, Zielonka J, Podsiadły R. Characterization of the reactivity of dioxide/bicarbonate: kinetics and influence on peroxynitrite-mediated oxidations.
luciferin boronate - a probe for inflammatory oxidants with improved stability. Arch Biochem Biophys 1996;333:49–58. https://doi.org/10.1006/abbi.1996.0363.
Dyes Pigments 2020;183:108693. https://doi.org/10.1016/j.dyepig.2020.108693. [80] Trujillo M, Radi R. Peroxynitrite reaction with the reduced and the oxidized forms
[60] Gopalan P, Katz HE, McGee DJ, Erben C, Zielinski T, Bousquet D, Muller D, of lipoic acid: new insights into the reaction of peroxynitrite with thiols. Arch
Grazul J, Olsson Y. Star-shaped azo-based dipolar chromophores: design, synthesis, Biochem Biophys 2002;397:91–8. https://doi.org/10.1006/abbi.2001.2619.
matrix compatibility, and electro-optic activity. J Am Chem Soc 2004;126:1741–7. [81] Alvarez B, Ferrer-Sueta G, Freeman BA, Radi R. Kinetics of peroxynitrite reaction
https://doi.org/10.1021/ja039768k. with amino acids and human serum albumin. J Biol Chem 1999;274:842–8.
[61] Karstens T, Kobs K. Rhodamine B and rhodamine 101 as reference substances for https://doi.org/10.1074/jbc.274.2.842.
fluorescence quantum yield measurements. J Phys Chem 1980;84:1871–2. https://
doi.org/10.1021/j100451a030.

15

You might also like