Significance
Hydrogen peroxide plays an important role in the fine balance between physiological and pathological processes. To detect this diffusible small molecule, we expanded the scope of organic triggers in developing borinic acids as an alternative and more sensitive trigger than the most conventional boronate-based sensors. We discovered that borinic acid is 10,000-fold more reactive than its boronic counterpart toward H2O2-mediated oxidation. An accurate determination of oxidation kinetic constants and computational experiments corroborate this higher reactivity. This improvement also proved effective for in-cell detection of exogenously as well as endogenously produced H2O2. We believe borinic acids represent a new and efficient tool allowing for the development of new devices for a better understanding of H2O2-mediated signaling processes.
Keywords: borinic acid trigger, kinetic study, hydrogen peroxide detection, fluorescent probes, activity-based sensing
Abstract
Hydrogen peroxide (H2O2) is responsible for numerous damages when overproduced, and its detection is crucial for a better understanding of H2O2-mediated signaling in physiological and pathological processes. For this purpose, various “off–on” small fluorescent probes relying on a boronate trigger have been prepared, and this design has also been involved in the development of H2O2-activated prodrugs or theranostic tools. However, this design suffers from slow kinetics, preventing activation by H2O2 with a short response time. Therefore, faster H2O2-reactive groups are awaited. To address this issue, we have successfully developed and characterized a prototypic borinic-based fluorescent probe containing a coumarin scaffold. We determined its in vitro kinetic constants toward H2O2-promoted oxidation. We measured 1.9 × 104 m−1⋅s−1 as a second-order rate constant, which is 10,000-fold faster than its well-established boronic counterpart (1.8 m−1⋅s−1). This improved reactivity was also effective in a cellular context, rendering borinic acids an advantageous trigger for H2O2-mediated release of effectors such as fluorescent moieties.
Reactive oxygen species (ROS) are involved in various physiological processes. In particular, hydrogen peroxide (H2O2) plays a critical role in the regulation of numerous biological activities as a signaling molecule (1, 2). However, aberrant production or accumulation of H2O2 leads to oxidative stress conditions, which can cause lesions associated with aging, cancer (3), and several neurodegenerative diseases such as Alzheimer’s or Parkinson’s (4, 5). Differentiation of physiological or abnormal conditions is closely connected with slight changes in H2O2 levels. However, the generation and degradation of H2O2 are variable within different cellular compartments, and this small molecule is highly diffusive, rendering the capture of small H2O2 fluctuations and the study of its spatial and temporal dynamics difficult. Therefore, the development of selective and sensitive H2O2-reactive tools for applications in a biological context represents a challenge for a better understanding of H2O2-mediated signaling in physiological and pathological processes or the use of H2O2 activation for the release of biological effectors (6).
Numerous strategies have been developed to implement H2O2-reactive molecular triggers, as exemplified by “off–on” small fluorescent probes. Such probes have attracted particular attention due to their easy implementation, high expected signal-to-noise ratio, and compatibility with standard equipment present in cellular biology research environments (7–9). Activation in such a context is triggered or modified by H2O2-mediated transformation of a suitable chemical species. Several approaches have been explored including probes based on arylsulfonyl ester perhydrolysis (10), oxidation of arylboronates (11), Baeyer–Villiger oxidation of diketones (12), Tamao oxidation of silanes (13), a tandem Payne–Dakin reaction (14) or a seleno-Mislow–Evans rearrangement (15). Among them, designs based on the boronate esters oxidation pioneered by Chang are the most explored, due to their remarkable stability, low toxicity profile, ease of preparation, and specificity toward H2O2, as illustrated in recent reviews (16–18). Upon reaction with H2O2, these compounds undergo an oxidative conversion into aryl borate esters that further hydrolyze into the corresponding phenols along with borate esters or boric acid (Scheme 1A). This conversion turns on probe fluorescence or activates drug release either directly or via the degradation of a self-immolative spacer. This chemospecific and biologically compatible reaction allowed, for instance, developing highly selective fluorescent probes for H2O2 imaging in cells (19–23). However, H2O2-triggered conversion of boronic acids to phenols is still not completely satisfactory in a biological context (24) since most of these probes have second-order reaction rate constants of 0.1 to 1.0 m−1⋅s−1 (14). In cells, H2O2 is present in the 1 to 100 nm concentration range in physiological conditions and could reach up to 100 μm under oxidative stress conditions (25). Therefore, most of the boronate-based systems need an incubation time longer than 30 min for activation at an H2O2 concentration of 100 μm. At such a time scale, H2O2 typically diffuses over a distance of 2 mm (evaluated as (DH2O2τ)0.5 with DH2O2 = 1.7 × 10−9 m2⋅s−1 from ref. 26 and τ = 30 min). Hence, to improve spatial resolution for H2O2 imaging, alternative H2O2 triggers with rapid reaction rates allowing real-time activation by H2O2 are still required.
Scheme 1.
(A) Current boronic acid (R = H) or boronate (R,R = tetramethylethylene) as H2O2-responsive group releasing a hydroxyaryl as effector and a boric acid or a borate ester respectively. (B) This study: a borinic acid as H2O2-responsive group releasing a hydroxyaryl as effector and a boronic acid.
To address this issue, we envisioned the use of borinic acids, structures in which one of the boron–oxygen bonds of the boronic acid is replaced by a boron–carbon bond. Due to these electronic modifications, borinic acids exhibit more electrophilic properties (27–30) compared to their boronic acid counterparts and could be more prone to rapid oxidation. These structures have been mainly exploited as catalysts in various reactions such as epoxide ring opening (31), hydrosilylation (32), transamidation (33), aldol reaction (34, 35), C–H activation (36, 37), selective monoalkylation, acylation and sulfonation of diols (38, 39), or regioselective glycosylation reactions (40–42). Surprisingly, the reactivity of these borinic species remains underexplored (43–45), probably due to their limited synthetic access (46–49). They were usually obtained through the addition of strong organometallic reagents (RLi/RMgBr) onto boron-based electrophiles such as trialkylborates, boron halides, diborane, or boronate esters. To date, a detailed study of the reactivity of borinic acids toward oxidation including reaction with H2O2 has not been reported and their use as triggers for the direct release of a probe or an effector has not been considered.
Herein, we report the design, synthesis, and evaluation of a borinic-triggered prototypic probe prone to direct and rapid activation by the H2O2 molecule (Scheme 1B). We establish a detailed kinetic analysis of the H2O2-promoted oxidation of this borinic acid as well as a comparative study with its corresponding boronic analog. Furthermore, we demonstrate the shorter response time of the borinic trigger compared to the boronic trigger against H2O2-mediated oxidation in a cellular environment.
Results and Discussion
Borinic Acids React Faster with H2O2 than Boronic Acids.
First, in order to validate the proof of concept, we compared the reactivity of simple model compounds toward H2O2, i.e., diphenylborinic acid, synthesized from addition of phenyl lithium onto pinacol phenylboronate (47, 50), and the commercially available phenylboronic acid. These two compounds were submitted to oxidation with one equivalent of H2O2 in deuterated phosphate-buffered saline (PBS, pH 7.4). The reaction progress was monitored using 1H NMR spectroscopy. These early experiments were extremely encouraging since we observed a complete oxidation of the diphenylborinic acid into phenol and phenylboronic acid within 2 min, whereas 2 h were required for full conversion of the phenylboronic acid into phenol under the same conditions (SI Appendix, Fig. S1). The experimental results revealing the superior reactivity of borinic acid were clearly corroborated by density functional theory (DFT) calculations performed at the M062X/6-311G++(d,p) level (see SI Appendix for details), showing a difference of activation energy for the initial nucleophilic addition of the perhydroxyl anion onto the boron center of both diphenylborinic and phenylboronic acids of 6.5 kcal⋅mol−1 in favor of the former.
Design and Synthesis of a Model Fluorogenic Borinic Probe.
Capitalizing on these preliminary results, we then designed a borinic acid-based fluorogenic probe in which the boron atom is substituted by a phenyl moiety and a 4-methylcoumarin scaffold as a fluorescent chromophore. The synthetic route involved a four-step sequence in which the key reaction relies on the addition of phenyl lithium onto a coumarin-based pinacol boronic ester (Scheme 2A). Thus, treatment of 4-methylumbelliferone (4-MU) with trifluoromethanesulfonic anhydride afforded the corresponding triflate 1 in 97% yield, which was subjected to a Pd-catalyzed cross-coupling reaction with bis(pinacolato)diboron under microwave irradiation according to Chang’s conditions (20, 51). The resulting pinacol boronic ester 2 was then reacted with a solution of phenyl lithium to afford the borinic acid, which was immediately converted into the corresponding N,N-dimethylaminoethyl ester 3. This compound was isolated by precipitation in 24% yield over two steps. Acidic hydrolysis then provided the expected borinic acid 4 in 67% yield. The corresponding 4-methylcoumarin boronic acid 5 was also synthesized for comparison. It was easily obtained from the pinacol ester using recently reported pinacol exchange reaction with methylboronic acid (Scheme 2B) (52).
Scheme 2.
(A) Synthesis of 4-methylcoumarin–based borinic acid 4. Conditions: a) Tf2O (1.1 equiv), Pyr (2.2 equiv), CH2Cl2, RT, 2 h, 97%; b) bis(pinacolato)diboron (1.2 equiv), KOAc (1.5 equiv), Pd(dppf)Cl2·CH2Cl2 (10 mol %), toluene, microwave irradiation, 120 °C, 3 h, 65%; c) PhLi in n-Bu2O (1.9 m, 1.1 equiv), −78 °C, 4 h then HCl in Et2O (2 m, 7 equiv), −78 °C to room temperature (RT); d) N,N-dimethylethanolamine (2 equiv), Et2O, RT, 12 h, 24%; e) HCl aq 1 m, EtOAc, RT, 30 min, 67%. (B) Synthesis of 4-methylcoumarin boronic acid 5: f) MeB(OH)2, TFA, CH2Cl2, RT, 4 h, 87%.
This Probe Confirms Higher Reactivity of Borinic Acids.
H2O2-mediated oxidation of 4 and 5 was first monitored by 1H NMR spectroscopy in deuterated PBS buffer (pH 7.4). As previously observed with the phenyl-based model compounds, the borinic derivative was found to be more reactive than its boronic counterpart. Thus, the addition of H2O2 triggers full oxidation of the borinic acid 4 in less than 2 min. In contrast, the boronic acid 5 is gradually converted into 4-MU to reach complete conversion within 2.5 h (SI Appendix, Fig. S2). Due to the unsymmetrical nature of the borinic acid probe, the oxidative rearrangement of 4 leads to the formation of two different alcohol/boronic acid couples, i.e., either 4-MU/phenylboronic acid as the expected cleavage products or phenol/4-methylcoumarin boronic acid 5 as the undesired cleavage products. The ratio of products was determined by 1H NMR integration and was evaluated to be 20/80, against the direct release of 4-MU (Fig. 1).
Fig. 1.
Regioselectivity of the oxidation of probe 4 determined by 1H NMR spectroscopy (600 MHz) in deuterated PBS buffer (pH 7.4, 5% DMSO). 1H NMR spectra of the borinic acid 4 (A) and 4 with H2O2-urea complex (1 equiv) after 2 min (B). (C) The two different couples of alcohol/boronic acid resulting from the oxidation reaction. 4-MU = 4-methylumbelliferone.
Study of the Oxidation Kinetics of the Model Borinic Probe with H2O2.
Having demonstrated the higher reactivity of the borinic trigger, we then undertook the accurate determination of the kinetic constants of the oxidation reactions (Fig. 2A). To calculate these parameters, we took advantage of the profluorescent properties of 4 and used real-time measurement of the fluorescence emission performed under pseudo-first-order conditions (excess of oxidant). Initial H2O2-promoted oxidation experiments were conducted with boronic acid 5 at different temperatures and H2O2 concentrations. As expected, reaction with H2O2 leads to an increase in fluorescence emission at 450 nm, indicating the release of the 4-MU chromophore (Fig. 2B).
Fig. 2.
Kinetic study of oxidation of borinic acid 4 and boronic acid 5. (A) Simplified kinetic scheme of oxidation of the borinic acid 4. (B and C) Time-dependent fluorescence intensity evolution of 5 μM 5 (B) or 4 (C) upon addition of 100 μm H2O2-urea solution in PBS buffer 1× (pH 7.4, 0.04% DMSO) at 310 K (pseudo-first-order conditions). The fluorescence emission was normalized and was recorded at λem = 450 nm (λex = 405 nm). a.u., arbitrary units.
For kinetic analysis of the experimental data (see SI Appendix for the determination), a model including the different steps involved in the 4-MU release was taken into account allowing determination of the second-order rate constants k3 associated with the boronic acid oxidation as 1.8 m−1⋅s−1 and 6.3 m−1⋅s−1 at 293 and 310 K, respectively (SI Appendix, Fig. S7). These results are consistent with those previously reported in the literature (53–56). In comparison, oxidation of borinic derivative 4 shows a different behavior. Upon addition of H2O2, fluorescence rapidly increases, which suggests an oxidation of the borinic acid probe 4 on the minute time scale affording directly 4-MU through the desired, although minor, oxidation pathway. After a few minutes, fluorescence increases more slowly on a time scale, which is in line with the formation of 4-MU from the boronic acid 5 obtained as a major product from oxidizing probe 4 (Fig. 2C). Upon considering that several reactions are rapid on the time scale of the overall oxidation process, we could account for the oxidation of the borinic acid 4 with the reduced mechanism shown in Fig. 2A. After integration of the kinetic results obtained above for the boronic acid 5, we extracted 1.9 × 104 and 4.2 × 104 m−1⋅s−1 for the rate constants k1 at 293 and 310 K, respectively (SI Appendix, Fig. S10). These results point to an oxidation of the borinic trigger, which is about 10,000 times faster than for its boronic counterpart. This represents a major improvement compared to boronate scaffolds and recently reported allylic selenide trigger oxidation (k1 9.2 m−1⋅s−1) (15).
Stability and Selectivity of Borinic Trigger.
In order to explore the potential applicability of borinic acids as new H2O2 sensors under physiological conditions, the stability of diphenylborinic acid as a model was evaluated in various biorelevant media. Hence, this model proved to be stable upon shelf storage, as well as in PBS buffer (pH 7.4) and in Dulbecco’s modified Eagle’s medium (DMEM). Addition of glucose, glutathione, or bovine serum albumin (BSA) to the solution in PBS buffer had almost no impact on the stability of borinic acid on an hour time scale (less than 10% of degradation after 8 h; SI Appendix, Fig. S11). In addition, the presence of a reducing agent like glutathione had no influence on the outcome of the H2O2-promoted oxidation reaction since only the borinic acid was oxidized in the presence of 1 equivalent of H2O2 (SI Appendix, Fig. S14). This result highlights the superior reactivity of the borinic trigger toward H2O2 compared to thiol derivatives. Then, the selectivity of the borinic probe 4 toward several ROS was examined in order to verify that this higher reactivity does not result in a decrease in selectivity toward H2O2. For numerous relevant ROS, including superoxide, tert-butyl hydroperoxide, as well as hydroxyl and tert-butyl radical, negligible fluorescent changes were detected (Fig. 3A), indicating that borinic acid 4 exhibits a high selectivity toward H2O2, which is similar to the corresponding boronic acid 5 (SI Appendix, Fig. S15) as previously reported (57).
Fig. 3.
(A) Selectivity of probe 4 toward H2O2 vs. other ROS. Bars represent fluorescence responses of 5 μM probe 4 to 100 μM various ROS at 5 (white), 15 (light gray), 30 (dark gray), and 60 min (black) in PBS buffer 1× (pH 7.4) at 25 °C. Emission intensity was recorded at λem = 450 nm (λex = 360 nm). (B) Fluorescence lifetime images. Fluorescence images of COS7gp91-p22 cells with 10 μm probe 4 (Top) or 5 (Bottom) prior to (control) and upon exposure to 100 μm exogenous H2O2 for 30 s, 45 s, and 165 s. Images were acquired using a DAPI filter (λem = 447 nm, λex = 387 nm). (Scale bar, 30 μm.) (C) Time-dependent fluorescence intensity of 10 μm borinic acid 4 (red curve) or 10 μm boronic acid 5 (blue curve) in transfected COS7gp91/p22 cells. Negative controls were carried out in nontransfected COS7gp91/p22 cells with the borinic acid 4 (light red curve) and the boronic acid 5 (light blue curve). Data were recorded in PBS buffer 1× (pH 7.4, 0.02% DMSO) at 310 K, at λem = 450 nm (λex = 360 nm). a.u., arbitrary units.
Probe Activation in Cells Using Exogenously Added H2O2.
We next investigated the ability of the borinic trigger to detect H2O2 in a cellular environment. First, we compared the real-time responses of probes 4 and 5 upon exposure to exogenous H2O2 in COS7gp91-p22 model cells. Images clearly displayed higher fluorescence intensity with the borinic acid 4 (Fig. 3 B, Top) compared to the boronic acid 5 (Fig. 3 B, Bottom) within 30 s after addition of H2O2. These data suggest that the borinic acid trigger is capable of detecting intracellular H2O2 with a very fast response time.
Detection of Endogenously Produced H2O2.
We then evaluated the capacity of this borinic trigger to release the fluorophore through activation by endogenously produced H2O2. For this purpose, we used COS7gp91-p22 cells transfected with a chimeric protein, rendering these cells capable of continuously producing the superoxide radical anion through an activated NADPH oxidase (see SI Appendix) (58). Since superoxide spontaneously dismutates into H2O2 in aqueous buffer (59), these model cells provide an interesting system to compare the reactivity of both fluorogenic probes. We were pleased to observe that the borinic acid 4 rapidly generated a significant increase of the fluorescence signal within the first 15 min of the experiment in contrast to the boronic acid 5 (Fig. 3C). The ratio of the two slopes indicates that 4-MU is released 10-fold faster with the borinic trigger than with the boronic one, which is in a reasonable agreement with the theoretical value determined by the kinetic analysis (see SI Appendix, section 3.2.3). Moreover, the fluorescence increase was arrested in the presence of either diphenyleneiodonium chloride (DPI), a NADPH oxidase inhibitor, or catalase, an enzyme which catalyzes H2O2 dismutation (SI Appendix, Fig. S18). These results clearly support the borinic acid trigger as a valuable tool for a real-time and specific monitoring of endogenously produced H2O2 in a physiological environment.
Conclusion
We have described an example of a borinic-based molecular device enabling direct and fast fluorophore release upon H2O2-mediated oxidation. Through detailed kinetic studies, we demonstrated that this coumarin-containing borinic sensor responded to H2O2 in an unprecedented short time. This improved reactivity is also effective when H2O2 is endogenously produced by cells. Optimizations are now underway to control the regioselectivity of this oxidation and get improved sensors in terms of spectral properties and brightness for real-time imaging in a biological context. This work has revealed the superior reactivity of the borinic function toward H2O2-promoted oxidation compared to the corresponding boronic counterpart, rendering this trigger a promising tool for H2O2-mediated activation. We believe that this trigger could find applications not only for H2O2 sensors but also in various fields such as drug release and theranostic applications.
Materials and Methods
Synthesis.
The syntheses of borinic acids are described in SI Appendix. All chemical reagents were purchased from chemical suppliers and used without further purifications. The 2-aminoethanol was distilled before use. Dry dichloromethane was obtained using a Glass Technology GT S100 device. Tetrahydrofuran (THF) and diethyl ether (Et2O) were distilled from Na/benzophenone prior to use. Reactions were monitored by NMR spectroscopy or analytical thin-layer chromatography (TLC). TLC was performed over Merck 60 F254 with detection by ultraviolet light and/or by charring with sulfuric acid, KMnO4, or phosphomolybdic acid solutions. Silica gel 60 (40 to 63 μm) was used for flash column chromatography. NMR spectra were recorded on Bruker Advance I 300, 360 and Advance II 600-MHz spectrometers, using the residual nondeuterated solvent as an internal standard. For 11B experiments, chemical shifts are given in parts per million (ppm) relative to BF3•Et2O (0 ppm) as an external standard. High-resolution mass spectra were recorded on a MicroTOFq mass spectrometer equipped with an electrospray interface (electrospray ionization positive or negative mode).
H2O2-Mediated Oxidation of Borinic Acids by 1H NMR.
H2O2-triggered oxidations of boronic and borinic derivatives were followed by 1H NMR spectroscopy performed on a Bruker Advance II 600-MHz spectrometer at 20 °C. Boronic or borinic acids reference samples were prepared in a deuterated PBS buffer 1× (pH 7.4) containing 5% deuterated dimethyl sulfoxide (DMSO-d6) (vol/vol) at 2 mm concentration (V= 500 µL). The H2O2-mediated oxidation was performed using 1 equivalent of oxidant. For this purpose, each sample was prepared by mixing 4 mm solutions of boronic or borinic acids in deuterated PBS buffer 1× (pH 7.4) containing 10% DMSO-d6 (vol/vol) (250 µL) and a 4 mm deuterated aqueous H2O2.urea solution (250 µL). The first spectrum was recorded on Bruker Advance II 600-MHz spectrometer after 2 min upon addition of H2O2 and the progress of the reaction was then monitored as a function of time.
DFT Calculations.
An energy diagram for the two first steps of the oxidation addition of H2O2 to phenylboronic acid or diphenylborinic acid was established using DFT calculations, which were performed with Gaussian 09. The geometries of all the structures were optimized at the M062X/6-311G++(d,p) level, and the vibrational-frequency calculations confirmed that these conformations were local minima or maxima, as expected. Intrinsic reaction coordinate calculations were carried out from each transition state to verify whether the reactant and the product were connected to the same transition state. The solvation effect was taken into account by including a few water molecules that make hydrogen bonds with the heteroatoms present in each compound. All energies are listed in SI Appendix. Our calculations corroborated the higher reactivity of borinic acid compared to boronic acid toward oxidation reaction since the nucleophilic addition is much more favorable.
Kinetic Studies.
Oxidation reactions were monitored by fluorescence using a Photon Technology International QuantaMaster QM-1 spectrofluorimeter equipped with a Peltier cell holder (TLC50; Quantum Northwest). All measurements of protonation constants were performed in Britton–Robinson buffer 1× and kinetic analyses were performed in PBS buffer 1× (pH 7.4). All solutions were prepared using water purified through a Direct-Q 5 (Millipore). pH measurements were performed on a Standard pH meter PHM210 Radiometer Analytical (calibrated with aqueous buffers at pH 4 and 7 or 10) with a Crison 5208 electrode, which is accurate over the 0 to 14 pH range. Ultraviolet-visible absorption spectra were recorded in 1-cm × 1-cm quartz cuvettes (Hellma) on a diode array ultraviolet-visible spectrophotometer (Evolution array; Thermo Scientific) at 293 K. The molar absorption coefficients were extracted while checking the validity of the Beer–Lambert law. The absorbance evolutions as a function of pH were analyzed with the SPECFIT/32 Global Analysis System (version 3.0 for 32-bit Windows systems) to extract the pKa of the investigated compounds.
Selectivity of Probes toward Various ROS.
Fluorescence emission intensity of 5 μM probe in the presence of 100 μm of various ROS in PBS buffer 1× (pH 7.4) was recorded at λem = 450 nm (λex = 360 nm) at 25 °C. Experiments were performed in triplicate and the results were averaged. H2O2 and superoxide (O2•−) were delivered from solid H2O2⋅urea and solid KO2, respectively. tert-butyl hydroperoxide (TBHP) was delivered from 70% aqueous solution. Hydroxyl radical (•OH) and tert-butoxy radical (•OtBu) were generated by the Fenton reaction of 2.5 mm Fe2+ with 100 µm H2O2 or 100 µm TBHP, respectively.
Cell Culture.
COS7gp91-p22 cells were kindly provided by M. Dinauer, Washington University in St. Louis, St. Louis, MO. These cells stably express two membrane subunits of the NADPH oxidase (gp91 and p22), affording an inactive enzyme. Upon transfection with a chimeric protein named trimera and composed of the domains necessary for the activation of the oxidase, the cells continuously produce superoxide radical anion (see below for details). COS7gp91-p22 cells were grown in cell DMEM culture medium containing 10% of fetal bovine serum (Gibco) and supplemented with the selecting antibiotics G418 (1.8 mg/mL; Sigma-Aldrich) and Puromycin (1 µg/mL; Sigma-Aldrich) at 37 °C. For cell splitting, cells are detached from the culture surface. We used the commercial Cell Dissociation Buffer Enzyme Free (Gibco) to avoid any damage on the extracellular parts of the membranous gp91 and p22. In order to induce a continuous NADPH oxidase activity, COS7gp91-p22 cells were transfected with the complementary DNA (cDNA) coding for a chimeric protein, also called trimera, composed of all the domains necessary for the activation of the membranous catalytic core of the NADPH oxidase. The trimera contains the N terminus of p47phox (amino acids 1 to 286), the N terminus of p67phox (amino acids 1 to 212), and a full-length Rac1 (amino acids 1 to 192) with the Q61L mutation that mimic the guanosine triphosphate (GTP) bound form. The day before transfection, cells were seeded in 48-well plates and transfected at 85 to 90% of confluency using X-tremeGENE HP DNA transfection reagent following strictly the manufacturer’s protocol (Roche) with the plasmid coding for a fluorescent protein labeled trimera, Citrine-trimera. Twenty-four hours after transfection, the medium was removed and cells were washed twice with Dulbecco’s PBS before the acquisition in 250 µL of PBS with glucose (PBSG).
Cell Imaging: Nontransfected COS7gp91-p22 Cell with Exogenously Added H2O2.
Nontransfected COS7gp91-p22 cells were plated on an eight-well chamber slide (Ibidi) and loaded with 10 µm of probes for 15 min (295 μL PBSG buffer, pH 7.4, and 0.02% DMSO) prior to imaging the control. Cells were then treated with 5 μL of a 6 mm H2O2 solution (final concentration 100 µm). Images were immediately recorded on a Leica inverted SP8 microscope equipped with a ×63 oil immersion objective (numerical aperture 1.4) using a DAPI filter (λex = 387 nm, λem = 447 nm). Fiji software was used to analyze all images and to quantify fluorescence.
Detection of Endogenously Produced H2O2.
Probes were added at a final concentration of 10 µM to transfected COS7gp91-p22 cells. Fluorescence was recorded at 450 nm upon excitation at 360 nm with a Synergy H1 microplate reader (BioTek) at 37 °C. Catalase was added at a final concentration of 2.5 μg/mL and DPI at 50 µm.
Supplementary Material
Acknowledgments
We thank Prof. Arnaud Gautier for preliminary cellular studies, the Ministère de l’Enseignement Supérieur, de la Recherche et de l’Innovation, for a PhD grant to R.O., and IdEx Paris Saclay for a PhD grant to B.G.-F. This work was supported by IdEx Paris Saclay (Initiative de Recherche Stratégique BioProbe). We also thank the French Agency for Research for funding preliminary studies (Grant ANR-12-BS07-0022 ROSAS and Grant ANR-21-CE07-0047). The CHARM3AT LabEx is also acknowledged for financial support to M.P.
Footnotes
The authors declare no competing interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2107503118/-/DCSupplemental.
Data Availability
All study data are included in the article and/or SI Appendix.
References
- 1.Veal E. A., Day A. M., Morgan B. A., Hydrogen peroxide sensing and signaling. Mol. Cell 26, 1–14 (2007). [DOI] [PubMed] [Google Scholar]
- 2.Di Marzo N., Chisci E., Giovannoni R., The role of hydrogen peroxide in redox-dependent signaling: Homeostatic and pathological responses in mammalian cells. Cells 7, 156–172 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Finkel T., Serrano M., Blasco M. A., The common biology of cancer and ageing. Nature 448, 767–774 (2007). [DOI] [PubMed] [Google Scholar]
- 4.Andersen J. K., Oxidative stress in neurodegeneration: Cause or consequence? Nat. Med. 10 (suppl.), S18–S25 (2004). [DOI] [PubMed] [Google Scholar]
- 5.Park L., et al. , Nox2-derived radicals contribute to neurovascular and behavioral dysfunction in mice overexpressing the amyloid precursor protein. Proc. Natl. Acad. Sci. U.S.A. 105, 1347–1352 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Peiró Cadahía J., Previtali V., Troelsen N. S., Clausen M. H., Prodrug strategies for targeted therapy triggered by reactive oxygen species. MedChemComm 10, 1531–1549 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Zheng D.-J., Yang Y.-S., Zhu H.-L., Recent progress in the development of small-molecule fluorescent probes for the detection of hydrogen peroxide. TrAC. Trends Analyt. Chem. 118, 625–651 (2019). [Google Scholar]
- 8.Guo H., Aleyasin H., Dickinson B. C., Haskew-Layton R. E., Ratan R. R., Recent advances in hydrogen peroxide imaging for biological applications. Cell Biosci. 4, 64 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Chen X., Tian X., Shin I., Yoon J., Fluorescent and luminescent probes for detection of reactive oxygen and nitrogen species. Chem. Soc. Rev. 40, 4783–4804 (2011). [DOI] [PubMed] [Google Scholar]
- 10.Maeda H., et al. , Fluorescent probes for hydrogen peroxide based on a non-oxidative mechanism. Angew. Chem. Int. Ed. Engl. 43, 2389–2391 (2004). [DOI] [PubMed] [Google Scholar]
- 11.Lo L.-C., Chu C.-Y., Development of highly selective and sensitive probes for hydrogen peroxide. Chem. Commun. (Camb.) 21, 2728–2729 (2003). [DOI] [PubMed] [Google Scholar]
- 12.Abo M., et al. , Development of a highly sensitive fluorescence probe for hydrogen peroxide. J. Am. Chem. Soc. 133, 10629–10637 (2011). [DOI] [PubMed] [Google Scholar]
- 13.Zhou X., et al. , Chemoselective alteration of fluorophore scaffolds as a strategy for the development of ratiometric chemodosimeters. Angew. Chem. Int. Ed. Engl. 56, 4197–4200 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Ye S., Hu J. J., Yang D., Tandem Payne/Dakin reaction: A new strategy for hydrogen peroxide detection and molecular imaging. Angew. Chem. Int. Ed. Engl. 57, 10173–10177 (2018). [DOI] [PubMed] [Google Scholar]
- 15.Pham D., et al. , Fluorogenic probe using a Mislow-Evans rearrangement for real-time imaging of hydrogen peroxide. Angew. Chem. Int. Ed. Engl. 59, 17435–17441 (2020). [DOI] [PubMed] [Google Scholar]
- 16.Lippert A. R., Van de Bittner G. C., Chang C. J., Boronate oxidation as a bioorthogonal reaction approach for studying the chemistry of hydrogen peroxide in living systems. Acc. Chem. Res. 44, 793–804 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Chen X., et al. , Recent progress in the development of fluorescent, luminescent and colorimetric probes for detection of reactive oxygen and nitrogen species. Chem. Soc. Rev. 45, 2976–3016 (2016). [DOI] [PubMed] [Google Scholar]
- 18.Liu Y., Jiao C., Lu W., Zhang P., Wang Y., Research progress in the development of organic small molecule fluorescent probes for detecting H2O2. RSC Adv. 9, 18027–18041 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Chang M. C. Y., Pralle A., Isacoff E. Y., Chang C. J., A selective, cell-permeable optical probe for hydrogen peroxide in living cells. J. Am. Chem. Soc. 126, 15392–15393 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Dickinson B. C., Huynh C., Chang C. J., A palette of fluorescent probes with varying emission colors for imaging hydrogen peroxide signaling in living cells. J. Am. Chem. Soc. 132, 5906–5915 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Lin V. S., Dickinson B. C., Chang C. J., Boronate-based fluorescent probes: Imaging hydrogen peroxide in living systems. Methods Enzymol. 526, 19–43 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Wu W., et al. , Bioluminescent probe for hydrogen peroxide imaging in vitro and in vivo. Anal. Chem. 86, 9800–9806 (2014). [DOI] [PubMed] [Google Scholar]
- 23.Wang P., Wang K., Chen D., Mao Y., Gu Y., A novel colorimetric and near-infrared fluorescent probe for hydrogen peroxide imaging in vitro and in vivo. RSC Adv. 5, 85957–85963 (2015). [Google Scholar]
- 24.Kalyanaraman B., Hardy M., Zielonka J., A critical review of methodologies to detect reactive oxygen and nitrogen species stimulated by NADPH oxidase enzymes: Implications in pesticide toxicity. Curr. Pharmcol. Rep. 2, 193–201 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Sies H., Hydrogen peroxide as a central redox signaling molecule in physiological oxidative stress: Oxidative eustress. Redox Biol. 11, 613–619 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Kern D. M. H., The polarography and standard potential of the oxygen-hydrogen peroxide couple. J. Am. Chem. Soc. 76, 4208–4214 (1954). [Google Scholar]
- 27.Ishihara K., Kurihara H., Yamamoto H., Bis(pentafluorophenyl)borinic acid as a highly effective oppenauer oxidation catalyst for allylic and benzylic alcohols. J. Org. Chem. 62, 5664–5665 (1997). [Google Scholar]
- 28.Wang G., Garrett G. E., Taylor M. S., Borinic acid-catalyzed, regioselective ring opening of 3,4-epoxy alcohols. Org. Lett. 20, 5375–5379 (2018). [DOI] [PubMed] [Google Scholar]
- 29.Chen J. L.-Y., Scott H. K., Hesse M. J., Willis C. L., Aggarwal V. K., Highly diastereo- and enantioselective allylboration of aldehydes using α-substituted allyl/crotyl pinacol boronic esters via in situ generated borinic esters. J. Am. Chem. Soc. 135, 5316–5319 (2013). [DOI] [PubMed] [Google Scholar]
- 30.Chen J. L.-Y., Aggarwal V. K., Highly diastereoselective and enantiospecific allylation of ketones and imines using borinic esters: Contiguous quaternary stereogenic centers. Angew. Chem. Int. Ed. Engl. 53, 10992–10996 (2014). [DOI] [PubMed] [Google Scholar]
- 31.Wang G., Taylor M. S., Borinic acid-catalyzed regioselective ring-opening of 3,4- and 2,3-epoxy alcohols with halides. Adv. Synth. Catal. 360, 398–403 (2020). [Google Scholar]
- 32.Chardon A., Rouden J., Blanchet J., Borinic acid mediated hydrosilylations: Reductions of carbonyl derivatives. Eur. J. Org. Chem. 2019, 995–998 (2020). [Google Scholar]
- 33.Dine T. M., Evans D., Rouden J., Blanchet J., Formamide synthesis through borinic acid catalysed transamidation under mild conditions. Chemistry 22, 5894–5898 (2016). [DOI] [PubMed] [Google Scholar]
- 34.Mori Y., Kobayashi J., Manabe K., Kobayashi S., Use of boron enolates in water. The first boron enolate-mediated diastereoselective aldol reactions using catalytic boron sources. Tetrahedron 58, 8263–8268 (2002). [Google Scholar]
- 35.Lee D., Newman S. G., Taylor M. S., Boron-catalyzed direct aldol reactions of pyruvic acids. Org. Lett. 11, 5486–5489 (2009). [DOI] [PubMed] [Google Scholar]
- 36.Dimakos V., Su H. Y., Garrett G. E., Taylor M. S., Site-selective and stereoselective C-H alkylations of carbohydrates via combined diarylborinic acid and photoredox catalysis. J. Am. Chem. Soc. 141, 5149–5153 (2019). [DOI] [PubMed] [Google Scholar]
- 37.Hubrich J., Himmler T., Rodefeld L., Ackermann L., Ruthenium(II)-catalyzed C-H arylation of anilides with boronic acids, borinic acids and potassium trifluoroborates. Adv. Synth. Catal. 357, 474–480 (2015). [Google Scholar]
- 38.Chan L., Taylor M. S., Regioselective alkylation of carbohydrate derivatives catalyzed by a diarylborinic acid derivative. Org. Lett. 13, 3090–3093 (2011). [DOI] [PubMed] [Google Scholar]
- 39.Lee D., Williamson C. L., Chan L., Taylor M. S., Regioselective, borinic acid-catalyzed monoacylation, sulfonylation and alkylation of diols and carbohydrates: Expansion of substrate scope and mechanistic studies. J. Am. Chem. Soc. 134, 8260–8267 (2012). [DOI] [PubMed] [Google Scholar]
- 40.Beale T. M., Taylor M. S., Synthesis of cardiac glycoside analogs by catalyst-controlled, regioselective glycosylation of digitoxin. Org. Lett. 15, 1358–1361 (2013). [DOI] [PubMed] [Google Scholar]
- 41.Tanaka M., Takahashi D., Toshima K., 1,2-cis-α-Stereoselective glycosylation utilizing a glycosyl-acceptor-derived borinic ester and its application to the total synthesis of natural glycosphingolipids. Org. Lett. 18, 5030–5033 (2016). [DOI] [PubMed] [Google Scholar]
- 42.D’Angelo K. A., Taylor M. S., Borinic acid-catalyzed stereo- and site-selective synthesis of β-glycosylceramides. Chem. Commun. (Camb.) 53, 5978–5980 (2017). [DOI] [PubMed] [Google Scholar]
- 43.Winkle D. D., Schaab K. M., Suzuki reaction of a diarylborinic acid: One-pot preparation and cross-coupling of bis(3,5-dimethylphenyl)borinic acid. Org. Process Res. Dev. 5, 450–451 (2001). [Google Scholar]
- 44.Dimitrijević E., Cusimano M., Taylor M. S., Synthesis of benzannulated heterocycles by twofold Suzuki-Miyaura couplings of cyclic diarylborinic acids. Org. Biomol. Chem. 12, 1391–1394 (2014). [DOI] [PubMed] [Google Scholar]
- 45.Wang C., Zhang J., Tang J., Zou G., A sequential Suzuki coupling approach to unsymmetrical aryl s-triazines from cyanuric chloride. Adv. Synth. Catal. 359, 2514–2519 (2017). [Google Scholar]
- 46.Brown H. C., Srebnik M., Cole T. E., Improved procedures for the preparation of boronic and borinic esters. Organometallics 5, 2300–2303 (1986). [Google Scholar]
- 47.Hofer A., et al. , Design, synthesis and pharmacological characterization of analogs of 2-aminoethyl diphenylborinate (2-APB), a known store-operated calcium channel blocker, for inhibition of TRPV6-mediated calcium transport. Bioorg. Med. Chem. 21, 3202–3213 (2013). [DOI] [PubMed] [Google Scholar]
- 48.Marciasini L., Cacciuttolo B., Vaultier M., Pucheault M., Synthesis of borinic acids and borinate adducts using diisopropylaminoborane. Org. Lett. 17, 3532–3535 (2015). [DOI] [PubMed] [Google Scholar]
- 49.Guan C., Huang L., Ren C., Zou G., Development of a telescoped process for preparation of N,O-chelated diarylborinates. Org. Process Res. Dev. 22, 824–828 (2018). [Google Scholar]
- 50.El Dine T. M., Rouden J., Blanchet J., Borinic acid catalysed peptide synthesis. Chem. Commun. (Camb.) 51, 16084–16087 (2015). [DOI] [PubMed] [Google Scholar]
- 51.Ishiyama T., Itoh Y., Kitano T., Miyaura N., Synthesis of arylboronates via the palladium(0)-catalyzed cross-coupling reaction of tetra(alkoxo)diborons with aryl triflates. Tetrahedron Lett. 38, 3447–3450 (1997). [Google Scholar]
- 52.Hinkes S. P. A., Klein C. D. P., Virtues of volatility: A facile transesterification approach to boronic acids. Org. Lett. 21, 3048–3052 (2019). [DOI] [PubMed] [Google Scholar]
- 53.Sikora A., Zielonka J., Lopez M., Joseph J., Kalyanaraman B., Direct oxidation of boronates by peroxynitrite: Mechanism and implications in fluorescence imaging of peroxynitrite. Free Radic. Biol. Med. 47, 1401–1407 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.LaButti J. N., Gates K. S., Biologically relevant chemical properties of peroxymonophosphate (=O3POOH). Bioorg. Med. Chem. Lett. 19, 218–221 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Wu W., et al. , Bioluminescent probe for hydrogen peroxide imaging in vitro and in vivo. Anal. Chem. 86, 9800–9806 (2014). [DOI] [PubMed] [Google Scholar]
- 56.Rios N., et al. , Sensitive detection and estimation of cell-derived peroxynitrite fluxes using fluorescein-boronate. Free Radic. Biol. Med. 101, 284–295 (2016). [DOI] [PubMed] [Google Scholar]
- 57.Du L., Li M., Zheng S., Wang B., Rational design of a fluorescent hydrogen peroxide probe based on the umbelliferone fluorophore. Tetrahedron Lett. 49, 3045–3048 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Masoud R., et al. , Conversion of NOX2 into a constitutive enzyme in vitro and in living cells, after its binding with a chimera of the regulatory subunits. Free Radic. Biol. Med. 113, 470–477 (2017). [DOI] [PubMed] [Google Scholar]
- 59.Winterbourn C. C., Hampton M. B., Livesey J. H., Kettle A. J., Modeling the reactions of superoxide and myeloperoxidase in the neutrophil phagosome: Implications for microbial killing. J. Biol. Chem. 281, 39860–39869 (2006). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All study data are included in the article and/or SI Appendix.





