Next Article in Journal
Ring-Shaped Piezoelectric 5-DOF Robot for Angular-Planar Motion
Next Article in Special Issue
HT-29 Colon Cancer Cell Electromanipulation and Assessment Based on Their Electrical Properties
Previous Article in Journal
Preparation and Characterization of PLG Microparticles by the Multiple Emulsion Method for the Sustained Release of Proteins
Previous Article in Special Issue
High-Throughput Dispensing of Viscous Solutions for Biomedical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Quantification of H2O2 Released by Airway Cells Growing in Different Culture Media

1
Department of Engineering, University of Palermo, 90128 Palermo, Italy
2
Institute of Traslational Pharmacology (IFT), National Research Council of Italy (CNR), 90146 Palermo, Italy
3
Ri.MED Foundation, 90146 Palermo, Italy
4
Dipietro Group, 96010 Melilli, Italy
*
Authors to whom correspondence should be addressed.
Micromachines 2022, 13(10), 1762; https://doi.org/10.3390/mi13101762
Submission received: 26 September 2022 / Revised: 11 October 2022 / Accepted: 12 October 2022 / Published: 18 October 2022
(This article belongs to the Special Issue Biosensors for Biomedical and Environmental Applications)

Abstract

:
Quantification of oxidative stress is a challenging task that can help in monitoring chronic inflammatory respiratory airway diseases. Different studies can be found in the literature regarding the development of electrochemical sensors for H2O2 in cell culture medium to quantify oxidative stress. However, there are very limited data regarding the impact of the cell culture medium on the electrochemical quantification of H2O2. In this work, we studied the effect of different media (RPMI, MEM, DMEM, Ham’s F12 and BEGM/DMEM) on the electrochemical quantification of H2O2. The used electrode is based on reduced graphene oxide (rGO) and gold nanoparticles (AuNPs) and was obtained by co-electrodeposition. To reduce the electrode fouling by the medium, the effect of dilution was investigated using diluted (50% v/v in PBS) and undiluted media. With the same aim, two electrochemical techniques were employed, chronoamperometry (CH) and linear scan voltammetry (LSV). The influence of different interfering species and the effect of the operating temperature of 37 °C were also studied in order to simulate the operation of the sensor in the culture plate. The LSV technique made the sensor adaptable to undiluted media because the test time is short, compared with the CH technique, reducing the electrode fouling. The long-term stability of the sensors was also evaluated by testing different storage conditions. By storing the electrode at 4 °C, the sensor performance was not reduced for up to 21 days. The sensors were validated measuring H2O2 released by two different human bronchial epithelial cell lines (A549, 16HBE) and human primary bronchial epithelial cells (PBEC) grown in RPMI, MEM and BEGM/DMEM media. To confirm the results obtained with the sensor, the release of reactive oxygen species was also evaluated with a standard flow cytometry technique. The results obtained with the two techniques were very similar. Thus, the LSV technique permits using the proposed sensor for an effective oxidative stress quantification in different culture media and without dilution.

1. Introduction

The imbalance between the production of reactive oxygen species (ROS) from both endogenous and exogenous sources and antioxidant defense systems (including superoxide dismutase, catalase and glutathione peroxidase) leads to oxidative stress [1,2,3,4]. ROS production can be increased by various physiological or pathological conditions. If the antioxidant enzymes fail to rebalance this ROS production, an accumulation occurs, causing cell damage. This contributes to inflammation, aging, cancer and several chronic diseases, including chronic obstructive pulmonary disease (COPD) [5,6,7]. Cigarette smoke is a strong inducer of oxidative stress and represents the main risk factor for COPD. Accordingly, therapies aimed at reducing oxidative burden or increasing antioxidant defences are useful in COPD management exacerbations and in preserving lung functions [8].
ROS include superoxide anion, hydroxyl radical and H2O2, among others [9,10,11]. Their half-life time is very short due to their high reactivity, and they are difficult to quantify [12,13]. Among ROS, H2O2 has the longest half-life time as well as the ability to cross biological membranes and induce damage in the extracellular space [14,15,16]. Thus, the quantification of H2O2 in cellular supernatants appears to be a convenient way to monitor the oxidative status of the cell [17,18].
Nowadays, H2O2 quantification is carried out by different laboratory-based techniques such as fluorometric and colorimetric assays, nuclear magnetic resonance spectroscopy and liquid chromatography [19,20,21,22,23]. Among the drawbacks of these techniques are their high cost, long analysis time, requirements of highly skilled personnel, and more importantly, their inability to facilitate in situ analysis to provide real-time results. Indeed, a sample must be collected from the cell culture to quantify H2O2 released by cells. This makes it very challenging to continuously monitor oxidative stress in real-time during cell growth [24]. Electrochemical sensors are perfect candidates to minimize all these drawbacks because they can be applied for in situ and real-time analysis, while offering good performances in terms of sensitivity, selectivity and limit of detection [25,26,27,28,29,30,31,32,33,34,35]. To improve the performance of an electrochemical cell, the use of nanostructured electrodes ensures a high active surface area and promotes a high current density [36,37,38,39,40,41,42,43].
In our previous work, we developed and studied the features of a nanostructured electrode made of reduced graphene oxide (rGO) and gold nanoparticles (Au-NPs) for the chronoamperometric detection of H2O2 released by the human macrophages cell line, THP1, grown in a Roswell Park Memorial Institute (RPMI) 1640 medium [44]. This sensor could be used to measure H2O2 release from other types of cell cultures as well. Different media can be used to grow different cells or the same cells, and the same culture medium can be used to grow different cells [45,46,47]. Usually, a medium contains a very wide range of chemicals, stabilizers and nutrients [48,49,50,51] and can differ from the other in nutrients and growth factors [48,52,53,54]. The most commonly used media include RPMI 1640 [55], Eagle Minimum Essential Medium (MEM) [49], Dulbecco’s Modified Eagle Medium (DMEM) [56], Ham’s F-10 & F12 [57], Medium 199 [58], and Iscove’s Modified Dulbecco’s Medium (IMDM) [59]. Considering their complex composition, an influence on the electrochemical detection of H2O2 by the different growth media can be expected. For this reason, we have carried out a systematic investigation of the effect of the cell culture media on the electrochemical response of the sensor for the detection of H2O2 released by epithelial cells from central and distal airways. In recent years, a number of different nanostructures have been studied for the electrochemical detection of hydrogen peroxide, such as Pt NPs [60], Cu6(SC7H4NO)6 nanoclusters [61], MnO2 nanosheets [62] and Co3O4 nanowires [63]. Our sensor consists of an indium tin oxide/poly-ethylene terephthalate (ITO-PET) substrate modified by electrodeposited AuNPs and rGO. These active materials were selected because they were demonstrated to have improved detection sensitivity for H2O2 compared to sensors consisting of either AuNPs or rGO only [44,64,65,66,67].
To date, very few papers have reported on the influence of cell culture media on the performance of electrochemical sensors [68,69]. In this work, the effect of different culture media was studied in detail. Previously, AuNPs-rGO-based sensors were optimized and tested [44] in RPMI medium by chronoamperometry (CH). Using CH, it was necessary to dilute the sample with phosphate-buffered saline (PBS; pH 7.4) by 50% in volume to avoid fouling of the sensor surface, which could yield a low detection sensitivity. Starting from these results, the electrochemical sensor described in [44] was used here with the aim of performing a systematic investigation on the influence of the following parameters on the quantification of H2O2:
-
Cell culture medium (MEM, DMEM, RPMI, Ham’s F12 and bronchial epithelial cell growth media (BEGM)/DMEM);
-
Electrochemical quantification technique (CH and LSV);
-
Medium dilution (undiluted vs. diluted in PBS (50% v/v));
-
Operating temperature (25 °C and 37 °C);
-
Interferents (uric acid, sodium chloride, lactic acid, glucose and HEPES);
-
Storage condition (immersed in PBS or in deionized water and stored at 4 and 20 °C, sealed under vacuum and stored at 4 and 20 °C or stored at 4 and 20 °C in air).
Considering that the media have a very complex matrix, a fouling of the surface of the sensor is expected during its operation. To overcome this problem, a dilution of medium can be a solution. However, for a sensor that must operate in situ (in the culture plate during cell growth), a different approach is necessary, such as the use of a fast electrochemical technique that permits performing the analysis before the appearance of the fouling phenomena. For this reason, in this work, an LSV technique was also used, and the results were compared to those of conventional CH [70,71,72,73].
In addition, the sensor was validated quantifying the H2O2 released from different cell lines. In particular, human primary bronchial epithelial cells (PBEC), human bronchial epithelial cell line (16HBE) and adenocarcinoma alveolar basal epithelial cell line (A549) were cultured and stimulated with both pro-oxidant (cigarette smoke extract, (CSE) [74]) and antioxidant (resveratrol [75]) stimuli. The release of H2O2 release was also quantified with the electrochemical sensor, and results were compared with the data obtained by flow-cytometry using the same cells stained with Carboxy-H2DCFDA and MitoSOX Red probe, which detect intracellular ROS and mitochondrial superoxide, respectively. Both techniques revealed a significant increase in ROS in cells exposed to CSE. Resveratrol, an antioxidant molecule, reverted this effect. The AuNPs-rGO-based sensor offers high sensitivity and selectivity, a short response time (<60 s) and can be applied to real-time, in situ monitoring of H2O2 release.

2. Materials and Methods

2.1. Reagents

Flexible indium tin oxide/polyethylene terephthalate substrate (60 Ω cm−2) and graphene oxide (4 mgmL−1) were purchased from Sigma Aldrich (St. Louis, MO, USA) and Graphenea, respectively. The following reagents were purchased from AlphaAesar: KAuCl4, 2-propanol, sodium acetate, glacial acetic acid, PBS tablet, 30% (v/v) H2O2, 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES), lactic acid, uric acid, sodium nitrate, sodium chloride, glucose.
MEM, RPMI-1640, DMEM, Fetal Bovine Serum (FBS), nonessential amino acids, L-glutamine, gentamicin, streptomycin and penicillin were obtained from Euroclone, while BEGM and Ham’s F12 Medium from Lonza Bioscience (see Supplementary Material for their composition). The probe 6-carboxy-2′,7′-dichlorodihydrofluorescein diacetate (H2DCFDA, C-2938) and MitoSOX™ Red mitochondrial superoxide indicator were bought from Life Technologies.

2.2. Sensor Fabrication and Electrochemical Detection of H2O2

The working electrode, made of AuNPs and rGO, was synthesized as previously described [44]. Briefly, the ITO-PET substrate was ultrasonically cleaned with pure iso-propanol and deionized water for 10 min. This electrode was then inserted into a home-made 3D printed cell (Figure S1), where the exposed active area of the electrode was 0.785 cm2. In this cell, a Pt wire was used as counter electrode while a saturated calomel electrode (SCE) as reference.
A Princeton Applied Research potentiostat/galvanostat (PARSTAT, mod. 2273) was used for both fabrication and characterization. AuNPs and rGO were potentiostatically co-deposited on the working electrode at −800 mV vs. SCE for 200 s in a solution containing 0.5 mgmL−1 GO and 0.5 mM KAuCl4 in acetate buffer (pH 5.4).
The effect of 5 different culture media (MEM, DMEM, Ham’s F12, RPMI and BEGM/DMEM (B/D)) on H2O2 quantification was evaluated. The AuNPs-rGO-based sensor was calibrated using both CH (at a constant potential of −800 mV vs. SCE) and LSV (in the potential range from +200 mV to −1200 mV, with a scan rate of 25 mV s−1). For CH tests, the medium was diluted with 50% of PBS (pH 7.4), while for LSV measurement, the sensor was calibrated in both diluted and undiluted media. Each calibration was carried out at least 3 times with 3 different electrodes.
The effect of temperature on the sensor performance was also evaluated. Particularly, H2O2 was quantified at room temperature and 37 °C. This temperature was chosen because cells are conventionally cultured at 37 °C.
The sensor selectivity was studied using only LSV as an electrochemical technique because, in our previous work, it was already found that the electrode was selective also using CH. LSV was carried out in the presence of 0.5 mM of H2O2 and 5 mM of different interfering species (uric acid, sodium chloride, lactic acid, glucose and HEPES) that can be found in media.
The sensor stability was evaluated by measuring the output after 21 days of storing under different conditions (immersed in PBS or in deionized water and stored at 4 and 20 °C, sealed under vacuum and stored at 4 and 20 °C, stored at 4 and 20 °C in air).

2.3. Cell Culture Test

Cell line of immortalized human normal bronchial epithelial (16HBE) and cell line of lung adenocarcinoma (A549) (Interlab Cell Line Collection) were used.
16HBE were cultured in MEM with 10% fetal bovine serum (FBS) (heat-deactivated 56 °C, 30 min), 1% non-essential amino acids, 2 mM L-glutamine and 0.5% gentamicin [76]. A549 were cultured in DMEM with 10% FBS (heat deactivated at 56 °C, 30 min), streptomycin and penicillin, 1% nonessential amino acids and 2 mM L-glutamine. All cited components were obtained from Euroclone.
16HBE and A549 were seeded in 6-well plates (500,000 cells/well) and the adherent cell culture monolayers were maintained in a humidified ambient, at 37 °C and with 5% CO2. Once reached the confluence (approximately 2 million cells/well), cells were stimulated with CSE at different concentrations (20% for 16HBE and 2.5% for A549) for 24 h. The 16HBE cells were also treated with resveratrol (40 μM), an antioxidant molecule, for 24 h in the presence or absence of CSE.
Human primary bronchial epithelial cells (PBECs) (American Type Culture Collection (ATCC), Manassas V.A.; PCS-300-010) were also used. PBECs were differentiated using the air–liquid interface (ALI) culture. Cells were seeded on 0.4 μm pore sized 12-well transwell plates (40,000 cells/well) (Corning Costar, Cambridge, MA, USA) coated with human fibronectin (Santa Cruz Biotecnology, Dallas, TX, USA), bovine albumin fraction V (Sigma Aldrich, St. Louis, MO, USA) and PureCol® (Advanced BioMatrix, Carlsbad, CA, USA) using B/D mixture (1:1, v/v BEGM/DMEM). The B/D medium contains 25 µM Hepes, bronchial epithelial cell growth supplement, 100 UmL−1 penicillin, 100 μgmL−1 streptomycin and 15 ngmL−1 retinoic acid receptor agonist EC23 (Tocris, Bristol, UK). Cells were cultured as submerged until confluence (approximately 500,000 cells/well), then the apical medium was removed and differentiated at ALI for 15 days. The differentiation of the cells into a more complex tissue containing different cell types is confirmed by the presence of cilia beating, by the mucus secretion and by the measurement of the trans-epithelial electrical resistance (TEER > 500 Ω cm2). After 15 days, PBECs were stimulated with CSE 20% for 24 h.
CSE was prepared by burning two cigarettes (3R4F-Kentucky—The Tobacco Research Institute, University of Kentucky) without any filter in 20 mL of PBS using a Watson–Marlow 323 E/D peristaltic pump (Rotterdam, The Netherlands). The solution was filter-sterilized (0.22 μm pore filter) and considered to be 100% of CSE. This was further diluted in the cell culture medium to reach the specific concentration for each experiment.
Flow cytometry was used to measure intracellular ROS and mitochondrial superoxide. After the stimulation, the cells were harvested and stained with 1 μM of 6-carboxy-2′, 7′-dichlorodihydrofluorescein diacetate (Carboxy-H2DCFDA) probe to measure intracellular ROS (30 min, ambient temperature) and with 3 μM of MitoSOX™ Red probe, which is specific for mitochondrial superoxide (15 min, at 37 °C) [77,78]. The flow cytometer CytoFLEX (Beckman Coulter, Brea, CA, USA) was used in these assays. The results were expressed as the mean of fluorescence intensity (MFI).

3. Results and Discussion

The sensor tested in this work consists of AuNPs and rGO that were co-electrodeposited on ITO/PET substrate. In this sensor, AuNPs and rGO act as active materials for the quantification of H2O2, while ITO acts only as conducting material. Due to its poor electrocatalytic properties, ITO does not contribute to the sensitivity of the sensor but simply makes the PET substrate conductive, thus making possible the electrodeposition of Au and rGO and the subsequent electrochemical characterization tests. Thus, the sensing material consists of only Au-NPs (about 33 nm) and r-GO flakes (about 10 × 18 μm). According to the results obtained in [44], AuNPs-rGO-based sensors tested in RPMI as medium diluted with 50% of PBS displayed for the detection of H2O2 a limit of detection (LOD) of 6.55 µM and an average sensitivity of 0.064 μAμM−1cm−2. In the previous study [44], we also demonstrated that using CH, the dilution of medium (50%) is necessary to have good sensitivity because, in medium alone, a decrease in sensitivity of about an 80% was observed due to fouling phenomena. In this work, the same experiments were carried out using other culture media. In particular, in each test, at the start of CH PBS alone was present in the cell. After the stabilization of the current, an equal volume of medium was added to have a dilution of 50%. The addition of the medium causes a spike in the current due to the instantaneous change in the solution composition at the electrode interface. After the transient, the current stabilizes, and the value is used as the blank current. In subsequent tests, the culture medium injected into the cell contains different amounts of H2O2. This causes a variation in the current, which depends on the concentration of H2O2.
Figure 1 (see also Figure S2) shows the effect of increasing H2O2 concentration on the CH experiments carried out using diluted MEM (1:1 v/v in PBS) as the solution. As expected, the higher the H2O2 concentration, the higher the output current density. A similar behavior was obtained in the other tested media. Considering that sensor calibration depends on the current value, careful and objective evaluations must be made. Thus, to build the calibration line, the output current density was selected based on the slope of the i-t curve. In particular, the current density corresponding to a slope equal to or lower than 50 nAs−1 was selected. Operating in this way, the value of the current used for the calibration of the sensor is independent of the operator’s choice, and above all, it is independent of the precise time in which the measurement is made. With this procedure, all the i-t curves were processed, including that obtained for the blank. The current measured for the blank was subtracted from the values obtained by varying the concentration of H2O2, and the obtained data were used to construct the calibration lines, as shown in Figure 2.
Figure 2 shows the calibration plots based on the CH signal obtained in different culture media at −800 mV vs. SCE. As can be observed, the electrode sensitivity (estimated by the slope of each linear calibration expression) was affected by the medium composition. Chronoamperometric detection of H2O2 maintains a similar sensitivity with MEM, DMEM and Ham’s F12, while the sensitivity dropped down to 0.0138 µAµM−1cm−2 using B/D medium. For each medium, the LOD was calculated by measuring the standard deviation of the blank using the following equation [79]:
LOD = 3.3 × SD/S
where SD is the standard deviation of the blank, and S is the electrode sensitivity. The standard deviation of the blank using MEM, DMEM and RPMI was very similar (ranging from 0.128 to 0.157 µAcm−2), while it was much higher using Ham’s F12 and B/D (0.406 and 0.51 µAcm−2).
In Table 1, the analytical parameters estimated using CH carried out in diluted culture media are tabulated.
From the results reported in Table 1, it can be concluded that the best medium for chronoamperometry quantification of H2O2 is the RPMI. This result is attributable to the different media compositions.
The same experiments were carried out using LSV as the electrochemical technique. The potential was scanned from +200 to −1200 mV vs. SCE at a scan rate of 25 mV/s. The linear scan voltammograms obtained in a blank solution and in MEM de-aerated using a continuous N2 flux are shown in Figure 3a,b, respectively. Figure 3b shows a featureless voltammogram, whereas Figure 3a depicts two peaks at −0.5 and −0.8 V, which are likely to be related to dissolved oxygen.
The first peak at −0.5 V is attributed to the reduction in dissolved oxygen to H2O2, following Equation (2) [80,81]
O 2 + 2 H 2 O + 2 e H 2 O 2 + 2 OH
The second peak at −0.8 V is related to the reduction in H2O2 to water, following Equation (3) [82,83]:
H 2 O 2 + 2 H + + 2 e 2 H 2 O
Thus, even in the absence of H2O2, the presence of dissolved oxygen at about −400 mV generates a small amount of H2O2 that is then revealed at more cathodic potentials [84]. When different concentrations of H2O2 are spiked into the solution, the corresponding voltammograms are displayed in Figure 3c. In these voltammograms, the first peak at −0.4 V does not increase in intensity, while the peak at −0.8 V increases with H2O2 concentration [85]. To illustrate this point more clearly, only linear scan voltammograms for low H2O2 concentration are shown in Figure 3d. These results are a further confirmation that the peak at about −400 mV is attributable to the electro-generation of H2O2 from dissolved oxygen reduction. According to Nernst’s equation, at high H2O2 concentration, the peak initially located at −0.8 V shifts to −0.9 V. A similar behavior was observed using both diluted and undiluted culture media. Figure 4 shows the obtained calibration lines using the media alone. The sensor performances are reported in Table 2. Interestingly, the sensitivity is much higher using LSV compared to CH (Table 1). For all studied media, the sensitivity increases by a factor of 3–4.
Using MEM, RPMI and DMEM, the effect of dilution is almost negligible (Table 2), while for Ham’s F12 and B/D, there is still an increase in sensitivity after dilution. Furthermore, in those media, better results were obtained with LSV compared to CH. This effect could be attributed to the fastness of the LSV compared to CH. In fact, the whole LSV experiments lasted less than 60 s, while CH experiments lasted for at least 600 s, due to the requirement for signal stabilization. During stabilization, H2O2 reacts at the electrode surface along with all the other chemical species present in the culture media. This may strongly affect the signal output due to a bio-fouling effect, especially with more complex mediums such as Ham’s F12 or B/D. Thus, the AuNPs-rGO-based sensor can work with every kind of medium with both LSV or CH measurements but, depending on the medium and the expected H2O2 concentration, it is necessary to use LSV or CH as the electrochemical technique to have a high sensitivity. In particular, the sensor can detect low concentrations of H2O2 in RPMI, MEM and DMEM medium with both LSV and CH, while for B/D, CH can be used when high concentrations of H2O2 are expected. LSV must be used to detect low concentrations of H2O2 (Table 1 and Table 2). From these results, it can be concluded that the proposed sensor is adaptable for use in different media.
Considering that the final goal of our research is to use the sensor directly on the plate during the culture of cells, the influence of operating temperature was also studied. Particularly, as the cell lines are cultured in incubators at a controlled temperature of 37 °C, the electrode was also tested at this temperature in order to simulate the operation in the plate [86]. Figure 5a,b shows the linear scan voltammograms obtained at 37 °C and the corresponding calibration line, respectively. The output current density increases with temperature, while electrode sensitivity is almost constant. In fact, a value of 0.127 μAμM−1cm−2 was obtained compared to 0.125 μAμM−1cm−2 obtained at room temperature (Table 2). The higher cathodic current density in Figure 5b, compared to that of Figure 4b, is expected at higher temperature due to the more favorable kinetics [87,88]. Thus, the sensor can also operate in the plate during the cell growth.
To fully characterize the electrode features for H2O2 quantification, a selectivity test was carried out towards different chemicals that could be found in the different culture media or that could be generated by the different cell lines. Particularly, 5 mM of the interfering species (sodium chloride, sodium nitrate, glucose, lactic acid, HEPES and uric acid) were added to the solution containing 0.5 mM of H2O2. This is a conservative condition due to the high concentration of interfering species compared to the concentration of H2O2 ([H2O2]/[interfering] = 10) [89]. Figure 6 shows the results both as LSV curves (Figure 6a) and as a ratio between the peak current measured in the absence and in the presence of the interfering species (Figure 6b). In all conditions, the interference is negligible (lower than 5%) in terms of both current intensity and potential. This high selectivity of the sensor is expected because the potential at −0.8 V is way off the redox peaks of interferences. Only in the case of uric acid was a very small shift in potential value (−30 mV) observed. This interference was probably due to the different pH of the solution. In fact, uric acid must be solubilized in a 0.1 M KOH due to its low solubility, and thus, its injection in the electrochemical cell can modify the solution pH [90]. Similar behavior was found in the other media.
Finally, to study the stability of the AuNPs-rGO-based sensors over time, the effect of different storage methods was evaluated by measuring the current density for 5 mM H2O2 before and after 21 days of storage. In particular, electrodes were stored in the following conditions:
-
Immersed in PBS or in deionized water and stored at 4 and 20 °C;
-
Sealed under vacuum and stored at 4 and 20 °C;
-
Stored at 4 and 20 °C in air.
The results are summarized in Table 3. All the storage conditions with the electrode immersed in liquid solutions (PBS or deionized water) were found to be almost destructive for the electrode, with a signal reduction ≥20%. A similar result was found storing the electrode in air at room temperature, with a signal reduction of 36%.
The best storing conditions were found to be at 4 °C in air and vacuum at both 4 and 20 °C. In these conditions, the change in current density was lower than 10%. Considering that the sensor has a reproducibility of about 5% (see Figure 4), these storage methods are suitable for storing it without appreciable deterioration in performance. Thus, it can be concluded that the sensors stored in these conditions are stable for at least 21 days after their production. This result agrees with other studies [91,92], where the same storage method was used for similar sensors (based on AuNPs and rGO but on a different substrate).
The AuNPs-rGO-based sensor was used to quantify the H2O2 released by A549, 16HBE and PBEC cells grown in RPMI, MEM and B/D, respectively. Considering the higher sensitivity obtained with LSV as the electrochemical technique, these samples were analyzed using this technique. A549 and PBEC were tested in two different conditions: untreated cells (NT) and treated with CSE as a pro-oxidant stimulus. 16HBE cell line was tested after the treatment with CSE and with an antioxidant molecule resveratrol (RES) and a combination of pro- and antioxidant stimula (RES + CSE). Figure 7a shows the linear scan voltammograms obtained in 16HBE cells. Particularly, Figure 7a shows the whole LSV curves obtained with 16HBE cell line, while background-subtracted voltammograms are shown in the inset. As expected, in all the studied cell lines, CSE treatment led to an increase in H2O2 production, while the treatment with RES, tested only in 16HBE cells, thanks to its antioxidant action, reverted the CSE-induced increase in H2O2 release. The results obtained in all experiments are summarized in Table 4.
To confirm the results obtained with the sensor, we used the same cells in which we tested the release of H2O2 in the culture medium to evaluate cellular oxidative stress assessing the production of intracellular ROS and mitochondrial superoxide after stimulation with cigarette smoke (pro-oxidant) and RES (antioxidant), the latter only in 16HBE. For this purpose, we used Carboxy-H2DCFDA and Mitosox Red probe, commonly used in research laboratories, which bound the intracellular ROS and mitochondrial superoxide, respectively, and evaluated their expression by flow cytometry. As Figure 7 shows, treating 16HBE (b), A549 (c, e) and PBEC (d, f) with CSE increases the production of intracellular ROS and mitochondrial superoxide, while RES reduces cigarette smoke-induced oxidative stress in 16HBE (b). These results are in line with those obtained with AuNPs-rGO-based sensors, confirming the possibility of using the latter as a replacement for more expensive and time-consuming laboratory techniques for monitoring cellular oxidative stress.

4. Conclusions

In this work, we have shown the effect of different parameters on the quantification of H2O2 using a AuNPs-rGO-based sensor. In particular, the detection was performed using different cell culture media (diluted and undiluted), different electrochemical techniques (CH and LSV) and temperature operations (25 and 37 °C). Furthermore, the effect of different interferents and storage conditions was studied. Using CH, the AuNPs-rGO-based sensor showed a sensitivity ranging from 0.033 to 0.064 μAμM−1cm−2 using MEM, DMEM or RPMI as medium, while the sensitivity decreased using Ham’s F12 or B/D. The LOD was about 10 µM with RPMI, MEM and DMEM, while it increased to up to 50–80 µM in Ham’s F12 and B/D. LSV was also used to quantify H2O2, and results showed a higher sensitivity (about four times) in all the studied medium, with a consistent decrease in LOD. The selectivity test showed an excellent anti-interference property of the AuNPs-rGO-based sensor. The effect of electrode storage for 21 days was also studied in different conditions. The results showed that the best storage method consists of in storage at 4 °C in air and in the dark. In these conditions, the current density changed by only 2% after 21 days, confirming the long-term stability of AuNPs-rGO-based sensor. To use the electrode directly in the cell culture plate, the effect of operating temperature was studied, and a negligible effect was found, showing an increase in current density with a constant sensitivity.
The sensor was used to quantify H2O2 in different cell cultures. In particular, the H2O2 released from A549, 16HBE and PBEC cells exposed to pro-oxidant and antioxidant treatments was measured. The sensor was able to quantify the variation of H2O2 released from different cell types after different treatments. These results were validated by flow cytometry, a technique providing a quantitative measure of intracellular oxidative stress. Thus, the AuNPs-rGO-based sensor can be effectively used to quantify oxidative stress in cell culture medium in a fast, easy, cheap, reproducible, and sensitive way.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/mi13101762/s1, Tables with media composition. Figure S1: Photo of the 3D printed cell. Figure S2: Effect of increasing H2O2 concentration on CH experiment carried out at −800 mV vs. SCE using diluted MEM medium.

Author Contributions

Conceptualization, B.P., S.D.V., M.F. and M.R.G.; methodology, B.P., S.D.V., M.F. and M.R.G.; validation, B.P., S.D.V., M.F. and M.R.G.; formal analysis, C. Z., L.B. and M.G.B.; investigation, L.B., S.D.V., M.F. and M.G.B.; data curation, L.B., S.D.V., C.Z., M.F. and M.G.B.; writing—original draft preparation, B.P. and S.D.V.; writing—review and editing, B.P., C.C., S.D.V., C.Z., E.P. and R.I.; supervision, C.C., M.R., G.A., E.P. and R.I; project administration, M.R., G.A., E.P. and R.I.; funding acquisition, C.C., B.P., E.P. and R.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Fondazione Ri.Med, Italian National Research Council and University of Palermo and has been financed by the Project “SeNSO” (n. 082651290364, linea di intervento 1.1.5 del P.O. FESR Sicilia 2014/2020).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AuNPs-rGOsensor
16HBEhuman normal bronchial epithelial cell line
A549lung adenocarcinoma cell line
ALIair-liquid interface culture
AuNPsgold nanoparticles
BEGMbronchial epithelial cell growth media (BEGM)
B/DBEGM/DMEM (1:1) medium
Carboxy-H2DCFDA6-carboxy-2′,7′-dichlorodihydrofluorescein diacetate
CHchronoamperometry
COPDchronic obstructive pulmonary disease
CSEcigarette smoke extract
DMEMDulbecco’s modified eagle medium (DMEM)
FBSfetal bovine serum (FBS)
HEPES4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
ITO-PETindium tin oxide/polyethylene terephthalate substrate
LODlimit of detection
LSVlinear scan voltammetry
MEMEagle’s minimum essential medium (MEM),
MFImean of fluorescence intensity
PBECshuman primary bronchial epithelial cells
PBSPhosphate-buffered saline
RESresveratrol
rGOreduces graphene oxide
ROSreactive oxygen species
SCEsaturated calomel electrode
RPMI-1640Roswell Park Memorial Institute-1640 medium,
TEERtrans-epithelial electrical resistance

References

  1. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative Stress: Harms and Benefits for Human Health. Oxid. Med. Cell Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M.T.D.; Mazur, M.; Telser, J. Free radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem. Cell Biol. 2007, 39, 44–84. [Google Scholar] [CrossRef] [PubMed]
  3. Lin, L.-S.; Wang, J.-F.; Song, J.; Liu, Y.; Zhu, G.; Dai, Y.; Shen, Z.; Tian, R.; Song, J.; Wang, Z.; et al. Cooperation of endogenous and exogenous reactive oxygen species induced by zinc peroxide nanoparticles to enhance oxidative stress-based cancer therapy. Theranostics 2019, 9, 7200–7209. [Google Scholar] [CrossRef] [PubMed]
  4. Kohen, R.; Nyska, A. Invited Review: Oxidation of Biological Systems: Oxidative Stress Phenomena, Antioxidants, Redox Reactions, and Methods for Their Quantification. Toxicol. Pathol. 2002, 30, 620–650. [Google Scholar] [CrossRef] [Green Version]
  5. Cipollina, C.; Bruno, A.; Fasola, S.; Cristaldi, M.; Patella, B.; Inguanta, R.; Vilasi, A.; Aiello, G.; La Grutta, S.; Torino, C.; et al. Cellular and Molecular Signatures of Oxidative Stress in Bronchial Epithelial Cell Models Injured by Cigarette Smoke Extract. Int. J. Mol. Sci. 2022, 23, 1770. [Google Scholar] [CrossRef]
  6. Furukawa, S.; Fujita, T.; Shimabukuro, M.; Iwaki, M.; Yamada, Y.; Nakajima, Y.; Nakayama, O.; Makishima, M.; Matsuda, M.; Shimomura, I. Increased oxidative stress in obesity and its impact on metabolic syndrome. J. Clin. Investig. 2004, 114, 1752–1761. [Google Scholar] [CrossRef]
  7. Ganguli, G.; Mukherjee, U.; Sonawane, A. Peroxisomes and Oxidative Stress: Their Implications in the Modulation of Cellular Immunity During Mycobacterial Infection. Front. Microbiol. 2019, 10, 1121. [Google Scholar] [CrossRef] [Green Version]
  8. Ferraro, M.; Di Vincenzo, S.; Sangiorgi, C.; Barone, S.L.; Gangemi, S.; Lanata, L.; Pace, E. Carbocysteine Modifies Circulating miR-21, IL-8, sRAGE, and fAGEs Levels in Mild Acute Exacerbated COPD Patients: A Pilot Study. Pharmaceuticals 2022, 15, 218. [Google Scholar] [CrossRef]
  9. Apel, K.; Hirt, H. Reactive oxygen species: Metabolism, oxidative stress, and signal transduction. Annu. Rev. Plant Biol. 2004, 55, 373–399. [Google Scholar] [CrossRef] [Green Version]
  10. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 2002, 7, 405–410. [Google Scholar] [CrossRef]
  11. Garg, M.; Gupta, A.; Sharma, A.L.; Singh, S. Advancements in 2D Materials Based Biosensors for Oxidative Stress Biomarkers. ACS Appl. Bio Mater. 2021, 4, 5944–5960. [Google Scholar] [CrossRef] [PubMed]
  12. Dickinson, B.C.; Chang, C.J. Chemistry and biology of reactive oxygen species in signaling or stress responses. Nat. Chem. Biol. 2011, 7, 504–511. [Google Scholar] [CrossRef] [Green Version]
  13. Griendling, K.; Touyz, R.M.; Zweier, J.L.; Dikalov, S.; Chilian, W.; Chen, Y.-R.; Harrison, D.G.; Bhatnagar, A. Measurement of Reactive Oxygen Species, Reactive Nitrogen Species, and Redox-Dependent Signaling in the Cardiovascular System: A Scientific Statement From the American Heart Association. Circ. Res. 2016, 119, e39–e75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Chance, B.; Sies, H.; Boveris, A. Hydroperoxide metabolism in mammalian organs. Physiol. Rev. 1979, 59, 527–605. [Google Scholar] [CrossRef] [PubMed]
  15. Bienert, G.P.; Schjoerring, J.K.; Jahn, T.P. Membrane transport of hydrogen peroxide. Biochim. Biophys. Acta 2006, 1758, 994–1003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Waghray, M.; Cui, Z.; Horowitz, J.; Subramanian, I.M.; Martinez, F.J.; Toews, G.B.; Thannickal, V.J. Hydrogen peroxide is a diffusible paracrine signal for the induction of epithelial cell death by activated myofibroblasts. FASEB J. 2005, 19, 1–16. [Google Scholar] [CrossRef]
  17. Balamurugan, T.; Mani, V.; Hsieh, C.-C.; Huang, S.-T.; Peng, T.-K.; Lin, H.-Y. Real-time tracking and quantification of endogenous hydrogen peroxide production in living cells using graphenated carbon nanotubes supported Prussian blue cubes. Sens. Actuators B Chem. 2018, 257, 220–227. [Google Scholar] [CrossRef]
  18. Marquitan, M.; Clausmeyer, J.; Actis, P.; Córdoba, A.L.; Korchev, Y.; Mark, M.D.; Herlitze, S.; Schuhmann, W. Intracellular Hydrogen Peroxide Detection with Functionalised Nanoelectrodes. ChemElectroChem 2016, 3, 2125–2129. [Google Scholar] [CrossRef] [Green Version]
  19. Katerji, M.; Filippova, M.; Duerksen-Hughes, P. Approaches and Methods to Measure Oxidative Stress in Clinical Samples: Research Applications in the Cancer Field. Oxidative Med. Cell Longev. 2019, 2019, 1–29. [Google Scholar] [CrossRef] [Green Version]
  20. Xiao, Y.; Meierhofer, D. Are Hydroethidine-Based Probes Reliable for Reactive Oxygen Species Detection? Antioxid. Redox Signal. 2019, 31, 359–367. [Google Scholar] [CrossRef]
  21. Wojtala, A.; Bonora, M.; Malinska, D.; Pinton, P.; Duszynski, J.; Wieckowski, M.R. Methods to Monitor ROS Production by Fluorescence Microscopy and Fluorometry. In Methods in Enzymology; Elsevier: Amsterdam, The Netherlands, 2014; pp. 243–262. [Google Scholar]
  22. Zhang, Y.; Dai, M.; Yuan, Z. Methods for the detection of reactive oxygen species. Anal. Methods 2018, 10, 4625–4638. [Google Scholar] [CrossRef]
  23. Stephenson, N.A.; Bell, A.T. Quantitative analysis of hydrogen peroxide by 1H NMR spectroscopy. Anal. Bioanal. Chem. 2005, 381, 1289–1293. [Google Scholar] [CrossRef] [PubMed]
  24. Halliwell, B.; Whiteman, M. Measuring reactive species and oxidative damage in vivo and in cell culture: How should you do it and what do the results mean? Br. J. Pharmacol. 2004, 142, 231–255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Mazzara, F.; Patella, B.; Aiello, G.; Sunseri, C.; Inguanta, R. Ascorbic Acid determination using linear sweep voltammetry on flexible electrode modified with gold nanoparticles and reduced graphene oxide. In Proceedings of the 2020 IEEE 20th Mediterranean Electrotechnical Conference (MELECON), Palermo, Italy, 16–18 June 2020; pp. 406–410. [Google Scholar]
  26. Yu, Y.; Pan, M.; Peng, J.; Hu, D.; Hao, Y.; Qian, Z. A review on recent advances in hydrogen peroxide electrochemical sensors for applications in cell detection. Chin. Chem. Lett. 2022, 33, 4133–4145. [Google Scholar] [CrossRef]
  27. Patella, B.; Sortino, A.; Mazzara, F.; Aiello, G.; Drago, G.; Torino, C.; Vilasi, A.; O’Riordan, A.; Inguanta, R. Electrochemical detection of dopamine with negligible interference from ascorbic and uric acid by means of reduced graphene oxide and metals-NPs based electrodes. Anal. Chim. Acta 2021, 1187, 339124. [Google Scholar] [CrossRef]
  28. Rojas, D.; Hernández-Rodríguez, J.F.; Della Pelle, F.; Escarpa, A.; Compagnone, D. New trends in enzyme-free electrochemical sensing of ROS/RNS. Application to live cell analysis. Mikrochim. Acta 2022, 189, 1–22. [Google Scholar] [CrossRef]
  29. Patella, B.; Piazza, S.; Sunseri, C.; Inguanta, R. Nio thin film for mercury detection in water by square wave anodic stripping voltammetry. Chem. Eng. Trans. 2017, 60, 1–6. [Google Scholar] [CrossRef]
  30. O’Sullivan, B.; O’Sullivan, S.; Narayan, T.; Shao, H.; Patella, B.; Seymour, I.; Inguanta, R.; O’Riordan, A. A direct comparison of 2D versus 3D diffusion analysis at nanowire electrodes: A finite element analysis and experimental study. Electrochim. Acta 2022, 408, 139890. [Google Scholar] [CrossRef]
  31. Mazzara, F.; Patella, B.; D’Agostino, C.; Bruno, M.; Carbone, S.; Lopresti, F.; Aiello, G.; Torino, C.; Vilasi, A.; O’Riordan, A.; et al. PANI-Based Wearable Electrochemical Sensor for pH Sweat Monitoring. Chemosensors 2021, 9, 169. [Google Scholar] [CrossRef]
  32. Murphy, A.; Seymour, I.; Rohan, J.; OrRiordan, A.; OrConnell, I. Portable Data Acquisition System for Nano and Ultra-Micro Scale Electrochemical Sensors. IEEE Sens. J. 2020, 21, 3210–3215. [Google Scholar] [CrossRef]
  33. Patella, B.; Moukri, N.; Regalbuto, G.; Cipollina, C.; Pace, E.; Di Vincenzo, S.; Aiello, G.; O’Riordan, A.; Inguanta, R. Electrochemical Synthesis of Zinc Oxide Nanostructures on Flexible Substrate and Application as an Electrochemical Immunoglobulin-G Immunosensor. Materials 2022, 15, 713. [Google Scholar] [CrossRef] [PubMed]
  34. Daly, R.; Narayan, T.; Shao, H.; O’Riordan, A.; Lovera, P. Platinum-Based Interdigitated Micro-Electrode Arrays for Reagent-Free Detection of Copper. Sensors 2021, 21, 3544. [Google Scholar] [CrossRef] [PubMed]
  35. Patella, B.; Sortino, A.; Aiello, G.; Sunseri, C.; Inguanta, R. Reduced graphene oxide decorated with metals nanoparticles electrode as electrochemical sensor for dopamine. In Proceedings of the 2019 IEEE International Conference on Flexible and Printable Sensors and Systems (FLEPS 2019), Glasgow, UK, 8–10 July 2019; pp. 1–3. [Google Scholar]
  36. Thangadurai, T.D.; Manjubaashini, N. Progressions in chemical and biological analytes sensing technology based on nanostructured materials: A comprehensive review. Mater. Sci. Eng. B 2021, 271, 115307. [Google Scholar] [CrossRef]
  37. Buccheri, B.; Ganci, F.; Patella, B.; Aiello, G.; Mandin, P.; Inguanta, R. Ni-Fe alloy nanostructured electrodes for water splitting in alkaline electrolyser. Electrochim. Acta 2021, 388, 138588. [Google Scholar] [CrossRef]
  38. Arenz, M.; Mayrhofer, K.J.J.; Stamenkovic, V.; Blizanac, B.B.; Tomoyuki, T.; Ross, P.N.; Markovic, N.M. The Effect of the Particle Size on the Kinetics of CO Electrooxidation on High Surface Area Pt Catalysts. J. Am. Chem. Soc. 2005, 127, 6819–6829. [Google Scholar] [CrossRef] [PubMed]
  39. Dhara, K.; Mahapatra, D.R. Recent advances in electrochemical nonenzymatic hydrogen peroxide sensors based on nanomaterials: A review. J. Mater. Sci. 2019, 54, 12319–12357. [Google Scholar] [CrossRef]
  40. Patella, B.; Russo, R.; O’Riordan, A.; Aiello, G.; Sunseri, C.; Inguanta, R. Copper nanowire array as highly selective electrochemical sensor of nitrate ions in water. Talanta 2021, 221, 121643. [Google Scholar] [CrossRef]
  41. Zeng, K.; Zhang, D. Recent progress in alkaline water electrolysis for hydrogen production and applications. Prog. Energy Combust. Sci. 2010, 36, 307–326. [Google Scholar] [CrossRef]
  42. Inguanta, R.; Ferrara, G.; Piazza, S.C. Sunseri Nanostructure fabrication by template deposition into anodic alumina membranes. Chem. Eng. Trans. 2009, 17, 957–962. [Google Scholar] [CrossRef]
  43. Trujillo, R.; Barraza, D.; Zamora, M.; Cattani-Scholz, A.; Madrid, R. Nanostructures in Hydrogen Peroxide Sensing. Sensors 2021, 21, 2204. [Google Scholar] [CrossRef]
  44. Patella, B.; Buscetta, M.; Di Vincenzo, S.; Ferraro, M.; Aiello, G.; Sunseri, C.; Pace, E.; Inguanta, R.; Cipollina, C. Electrochemical sensor based on rGO/Au nanoparticles for monitoring H2O2 released by human macrophages. Sens. Actuators B Chem. 2021, 327, 128901. [Google Scholar] [CrossRef]
  45. Arora, M. Cell Culture Media: A Review. Mater. Methods 2013, 3, 175. [Google Scholar] [CrossRef]
  46. Vis, M.A.M.; Ito, K.; Hofmann, S. Impact of Culture Medium on Cellular Interactions in in vitro Co-culture Systems. Front. Bioeng. Biotechnol. 2020, 8, 911. [Google Scholar] [CrossRef] [PubMed]
  47. Ackermann, T.; Tardito, S. Cell Culture Medium Formulation and Its Implications in Cancer Metabolism. Trends Cancer 2019, 5, 329–332. [Google Scholar] [CrossRef]
  48. Yao, T.; Asayama, Y. Animal-cell culture media: History, characteristics, and current issues. Reprod. Med. Biol. 2017, 16, 99–117. [Google Scholar] [CrossRef] [Green Version]
  49. Eagle, H. The Specific Amino Acid Requirements of a Human Carcinoma Cell (Strain hela) in Tissue Culture. J. Exp. Med. 1955, 102, 37–48. [Google Scholar] [CrossRef] [Green Version]
  50. Eagle, H. The Specific Amino Acid Requirements of A Mammalian Cell (Strain L.) on Tissue Culture. J. Biol. Chem. 1955, 214, 839–852. [Google Scholar] [CrossRef]
  51. Kleinman, H.; Luckenbill-Edds, L.; Cannon, F.; Sephel, G. Use of extracellular matrix components for cell culture. Anal. Biochem. 1987, 166, 1–13. [Google Scholar] [CrossRef]
  52. McKee, T.J.; Komarova, S.V. Is it time to reinvent basic cell culture medium? Am. J. Physiol. Physiol. 2017, 312, C624–C626. [Google Scholar] [CrossRef] [Green Version]
  53. Williams, G.; Weisburger, E.K.; Weisburger, J. Isolation and long-term cell culture of epithelial-like cells from rat liver. Exp. Cell Res. 1971, 69, 106–112. [Google Scholar] [CrossRef]
  54. Schubert, A.-K.; Smink, J.J.; Pumberger, M.; Putzier, M.; Sittinger, M.; Ringe, J. Standardisation of basal medium for reproducible culture of human annulus fibrosus and nucleus pulposus cells. J. Orthop. Surg. Res. 2018, 13, 209. [Google Scholar] [CrossRef] [PubMed]
  55. Moore, G.E.; Ito, E.; Ulrich, K.; Sandberg, A.A. Culture of human leukemia cells. Cancer 1966, 19, 713–723. [Google Scholar] [CrossRef]
  56. Dulbecco, R.; Freeman, G. Plaque production by the polyoma virus. Virology 1959, 8, 396–397. [Google Scholar] [CrossRef]
  57. Ham, R.G.; Sattler, G.L. Clonal growth of differentiated rabbit cartilage cells. J. Cell Physiol. 1968, 72, 109–114. [Google Scholar] [CrossRef] [PubMed]
  58. Morgan, J.F.; Morton, H.J.; Parker, R.C. Nutrition of Animal Cells in Tissue Culture. I. Initial Studies on a Synthetic Medium. Exp. Biol. Med. 1950, 73, 1–8. [Google Scholar] [CrossRef]
  59. Iscove, N.N.; Melchers, F. Complete replacement of serum by albumin, transferrin, and soybean lipid in cultures of lipopolysaccharide-reactive B lymphocytes. J. Exp. Med. 1978, 147, 923–933. [Google Scholar] [CrossRef] [Green Version]
  60. Zhang, C.; Zhang, R.; Gao, X.; Cheng, C.; Hou, L.; Li, X.; Chen, W. Small Naked Pt Nanoparticles Confined in Mesoporous Shell of Hollow Carbon Spheres for High-Performance Nonenzymatic Sensing of H2O2 and Glucose. ACS Omega 2018, 3, 96–105. [Google Scholar] [CrossRef] [Green Version]
  61. Gao, X.; He, S.; Zhang, C.; Du, C.; Chen, X.; Xing, W.; Chen, S.; Clayborne, A.; Chen, W. Single Crystal Sub-Nanometer Sized Cu6(SR)6 Clusters: Structure, Photophysical Properties, and Electrochemical Sensing. Adv. Sci. 2016, 3, 1600126. [Google Scholar] [CrossRef]
  62. He, S.; Zhang, B.; Liu, M.; Chen, W. Non-enzymatic hydrogen peroxide electrochemical sensor based on a three-dimensional MnO2 nanosheets/carbon foam composite. RSC Adv. 2014, 4, 49315–49323. [Google Scholar] [CrossRef]
  63. Liu, M.; He, S.; Chen, W. Co3O4 nanowires supported on 3D N-doped carbon foam as an electrochemical sensing platform for efficient H2O2 detection. Nanoscale 2014, 6, 11769–11776. [Google Scholar] [CrossRef]
  64. Klekotka, E.; Kasztelan, M.; Palys, B. Factors Influencing the Electrocatalytic Properties of Graphene Oxide—Gold Nanoparticles Hybrid System. ChemElectroChem 2021, 8, 3080–3088. [Google Scholar] [CrossRef]
  65. Cheng, C.; Zhang, C.; Gao, X.; Zhuang, Z.; Du, C.; Chen, W. 3D Network and 2D Paper of Reduced Graphene Oxide/Cu2O Composite for Electrochemical Sensing of Hydrogen Peroxide. Anal. Chem. 2017, 90, 1983–1991. [Google Scholar] [CrossRef] [Green Version]
  66. Zhang, R.; Chen, W. Recent advances in graphene-based nanomaterials for fabricating electrochemical hydrogen peroxide sensors. Biosens. Bioelectron. 2017, 89, 249–268. [Google Scholar] [CrossRef]
  67. Ju, J.; Chen, W. In Situ Growth of Surfactant-Free Gold Nanoparticles on Nitrogen-Doped Graphene Quantum Dots for Electrochemical Detection of Hydrogen Peroxide in Biological Environments. Anal. Chem. 2014, 87, 1903–1910. [Google Scholar] [CrossRef]
  68. Chmayssem, A.; Petit, L.; Verplanck, N.; Mourier, V.; Vignoud, S.; Vrana, N.E.; Mailley, P. Characterization of the Impact of Classical Cell-culture Media on the Response of Electrochemical Sensors. Electroanalysis 2022, 34, 1201–1211. [Google Scholar] [CrossRef]
  69. Oliveira, M.; Conceição, P.; Kant, K.; Ainla, A.; Diéguez, L. Electrochemical Sensing in 3D Cell Culture Models: New Tools for Developing Better Cancer Diagnostics and Treatments. Cancers 2021, 13, 1381. [Google Scholar] [CrossRef]
  70. Gulaboski, R.; Mirčeski, V.; Kappl, R.; Hoth, M.; Bozem, M. Review—Quantification of Hydrogen Peroxide by Electrochemical Methods and Electron Spin Resonance Spectroscopy. J. Electrochem. Soc. 2019, 166, G82–G101. [Google Scholar] [CrossRef]
  71. Scholz, F. Voltammetric techniques of analysis: The essentials. ChemTexts 2015, 1, 17. [Google Scholar] [CrossRef] [Green Version]
  72. Arrigan, D.W.M. Electrochemical Strategies in Detection Science; Royal Society of Chemistry: Cambridge, UK, 2015. [Google Scholar]
  73. Bard, A.J.; Faulkner, L.R. Electrochemical Methods: Fundamentals and Applications, 2nd ed.; Wiley: New York, NY, USA, 2001. [Google Scholar]
  74. Taylor, M.; Carr, T.; Oke, O.; Jaunky, T.; Breheny, D.; Lowe, F.; Gaça, M. E-cigarette aerosols induce lower oxidative stress in vitro when compared to tobacco smoke. Toxicol. Mech. Methods 2016, 26, 465–476. [Google Scholar] [CrossRef]
  75. Conte, E.; Fagone, M.; Fruciano, E.; Gili, M.; Iemmolo, C. Vancheri, Anti-inflammatory and antifibrotic effects of resveratrol in the lung. Histol. Histopathol. 2015, 30, 523–529. [Google Scholar] [CrossRef]
  76. Ferraro, M.; Gjomarkaj, M.; Siena, L.; DI Vincenzo, S.; Pace, E. Formoterol and fluticasone propionate combination improves histone deacetylation and anti-inflammatory activities in bronchial epithelial cells exposed to cigarette smoke. Biochim. Biophys. Acta (BBA)—Mol. Basis Dis. 2017, 1863, 1718–1727. [Google Scholar] [CrossRef]
  77. Cipollina, C.; DI Vincenzo, S.; Siena, L.; Di Sano, C.; Gjomarkaj, M.; Pace, E. 17-oxo-DHA displays additive anti-inflammatory effects with fluticasone propionate and inhibits the NLRP3 inflammasome. Sci. Rep. 2016, 6, 37625. [Google Scholar] [CrossRef]
  78. Cipollina, C.; DI Vincenzo, S.; Gerbino, S.; Siena, L.; Gjomarkaj, M.; Pace, E. Dual anti-oxidant and anti-inflammatory actions of the electrophilic cyclooxygenase-2-derived 17-oxo-DHA in lipopolysaccharide- and cigarette smoke-induced inflammation. Biochim. et Biophys. Acta (BBA)—Gen. Subj. 2014, 1840, 2299–2309. [Google Scholar] [CrossRef]
  79. Patella, B.; Aiello, G.; Drago, G.; Torino, C.; Vilasi, A.; O’Riordan, A.; Inguanta, R. Electrochemical detection of chloride ions using Ag-based electrodes obtained from compact disc. Anal. Chim. Acta 2021, 1190, 339215. [Google Scholar] [CrossRef]
  80. Shruthi, T.K.; Swain, G.M. Detection of H2O2 from the Reduction of Dissolved Oxygen on TCP-Coated AA2024-T3: Impact on the Transient Formation of Cr(VI). J. Electrochem. Soc. 2019, 166, C3284–C3289. [Google Scholar] [CrossRef]
  81. Jiang, K.; Back, S.; Akey, A.J.; Xia, C.; Hu, Y.; Liang, W.; Schaak, D.; Stavitski, E.; Nørskov, J.K.; Siahrostami, S.; et al. Highly selective oxygen reduction to hydrogen peroxide on transition metal single atom coordination. Nat. Commun. 2019, 10, 3997. [Google Scholar] [CrossRef] [Green Version]
  82. Tian, J.; Liu, Q.; Ge, C.; Xing, Z.; Asiri, A.M.; Al-Youbi, A.O.; Sun, X. Ultrathin graphitic carbon nitride nanosheets: A low-cost, green, and highly efficient electrocatalyst toward the reduction of hydrogen peroxide and its glucose biosensing application. Nanoscale 2013, 5, 8921–8924. [Google Scholar] [CrossRef]
  83. Tran, H.V.; Huynh, C.D.; Tran, H.V.; Piro, B. Cyclic voltammetry, square wave voltammetry, electrochemical impedance spectroscopy and colorimetric method for hydrogen peroxide detection based on chitosan/silver nanocomposite. Arab. J. Chem. 2018, 11, 453–459. [Google Scholar] [CrossRef]
  84. Cai, X.; Tanner, E.E.L.; Lin, C.; Ngamchuea, K.; Foord, J.S.; Compton, R.G. The mechanism of electrochemical reduction of hydrogen peroxide on silver nanoparticles. Phys. Chem. Chem. Phys. 2017, 20, 1608–1614. [Google Scholar] [CrossRef]
  85. Katsounaros, I.; Schneider, W.B.; Meier, J.C.; Benedikt, U.; Biedermann, P.U.; Auer, A.A.; Mayrhofer, K.J.J. Hydrogen peroxide electrochemistry on platinum: Towards understanding the oxygen reduction reaction mechanism. Phys. Chem. Chem. Phys. 2012, 14, 7384–7391. [Google Scholar] [CrossRef]
  86. Takagi, M.; Ueda, K. Comparison of the optimal culture conditions for cell growth and tissue plasminogen activator production by human embryo lung cells on microcarriers. Appl. Microbiol. Biotechnol. 1994, 41, 565–570. [Google Scholar] [CrossRef]
  87. Hwang, J.-Y.; Lai, J.-Y. The effect of temperature on limiting current density and mass transfer in electrodialysis. J. Chem. Technol. Biotechnol. 2007, 37, 123–132. [Google Scholar] [CrossRef]
  88. Oje, A.I.; Ogwu, A. Effect of temperature on the electrochemical performance of silver oxide thin films supercapacitor. J. Electroanal. Chem. 2021, 882, 115015. [Google Scholar] [CrossRef]
  89. Yang, L.; Wang, B.; Qi, H.; Gao, Q.; Li, C.-Z.; Zhang, C. Highly Sensitive Electrochemical Sensor for the Determination of 8-Hydroxy-2′-deoxyguanosine Incorporating SWCNTs-Nafion Composite Film. J. Sensors 2015, 2015, 504869. [Google Scholar] [CrossRef] [Green Version]
  90. Wilcox, W.R.; Khalaf, A.; Weinberger, A.; Kippen, I.; Klinenberg, J.R. Solubility of uric acid and monosodium urate. Med. Biol. Eng. Comput. 1972, 10, 522–531. [Google Scholar] [CrossRef]
  91. Lu, X.-Y.; Li, J.; Kong, F.-Y.; Wei, M.-J.; Zhang, P.; Li, Y.; Fang, H.-L.; Wang, W. Improved Performance for the Electrochemical Sensing of Acyclovir by Using the rGO–TiO2–Au Nanocomposite-Modified Electrode. Front. Chem. 2022, 10, 410. [Google Scholar] [CrossRef]
  92. Hong, J.; Wang, Y.; Zhu, L.; Jiang, L. An Electrochemical Sensor Based on Gold-Nanocluster-Modified Graphene Screen-Printed Electrodes for the Detection of β-Lactoglobulin in Milk. Sensors 2020, 20, 3956. [Google Scholar] [CrossRef]
Figure 1. Effect of increasing H2O2 concentration on CH experiment carried out at −800 mV vs. SCE using diluted MEM medium.
Figure 1. Effect of increasing H2O2 concentration on CH experiment carried out at −800 mV vs. SCE using diluted MEM medium.
Micromachines 13 01762 g001
Figure 2. Calibration line obtained with CH at −800 mV vs. SCE in (a) diluted MEM, (b) diluted DMEM, (c) diluted Ham’s F12 and (d) diluted B/D medium (n = 3).
Figure 2. Calibration line obtained with CH at −800 mV vs. SCE in (a) diluted MEM, (b) diluted DMEM, (c) diluted Ham’s F12 and (d) diluted B/D medium (n = 3).
Micromachines 13 01762 g002
Figure 3. Linear scan voltammetry experiments carried out in: (a) MEM blank solution, (b) de-aerated MEM blank solution, (c,d) aerated MEM blank solution with different H2O2 concentrations. In all experiments, undiluted MEM was used (n = 3).
Figure 3. Linear scan voltammetry experiments carried out in: (a) MEM blank solution, (b) de-aerated MEM blank solution, (c,d) aerated MEM blank solution with different H2O2 concentrations. In all experiments, undiluted MEM was used (n = 3).
Micromachines 13 01762 g003
Figure 4. Calibration line obtained with LSV in: (a) RPMI alone, (b) MEM alone, (c) DMEM alone, (d) Ham’s F12 alone, (e) B/D alone (n = 3).
Figure 4. Calibration line obtained with LSV in: (a) RPMI alone, (b) MEM alone, (c) DMEM alone, (d) Ham’s F12 alone, (e) B/D alone (n = 3).
Micromachines 13 01762 g004aMicromachines 13 01762 g004b
Figure 5. (a) LSV experiments at increasing H2O2 concentration and (b) corresponding calibration line. The tests were performed in pure MEM and at 37 °C (n = 3).
Figure 5. (a) LSV experiments at increasing H2O2 concentration and (b) corresponding calibration line. The tests were performed in pure MEM and at 37 °C (n = 3).
Micromachines 13 01762 g005
Figure 6. (a) LSV experiments carried out in MEM alone in the presence of 0.5 mM H2O2 and 5 mM of interfering species and (b) corresponding interference of each chemical on sensor output (n = 3).
Figure 6. (a) LSV experiments carried out in MEM alone in the presence of 0.5 mM H2O2 and 5 mM of interfering species and (b) corresponding interference of each chemical on sensor output (n = 3).
Micromachines 13 01762 g006
Figure 7. (a) LSV experiments with 16HBE cell line and ROS detection by flow cytometry using Carboxy-H2DCFDA probe in (b) 16HBE, (c) A549, and (d) PBEC and using Mitosox Red probe in (e) A549, and (f) PBEC. In the inset of (a) the LSV curves after baseline subtraction. * p value < 0.05 Unpaired t-test (n = 3).
Figure 7. (a) LSV experiments with 16HBE cell line and ROS detection by flow cytometry using Carboxy-H2DCFDA probe in (b) 16HBE, (c) A549, and (d) PBEC and using Mitosox Red probe in (e) A549, and (f) PBEC. In the inset of (a) the LSV curves after baseline subtraction. * p value < 0.05 Unpaired t-test (n = 3).
Micromachines 13 01762 g007
Table 1. Features of the AuNPs-rGO-based sensors using CH at −800 mV vs. SCE as the electrochemical technique in different culture media diluted with PBS.
Table 1. Features of the AuNPs-rGO-based sensors using CH at −800 mV vs. SCE as the electrochemical technique in different culture media diluted with PBS.
MediumSensitivity
µAµM−1cm−2
Limit of Detection
µM
Linear Range
µM
Diluted RPMI0.0646.5525–5000
Diluted MEM0.029913.0510–5000
Diluted DMEM0.03315.3710–5000
Diluted Ham’s F120.026451.61100–5000
Diluted B/D0.013880.14250–5000
Table 2. Performance of the AuNPs-rGO-based sensors using LSV as the electrochemical technique in different culture media diluted and not diluted with PBS.
Table 2. Performance of the AuNPs-rGO-based sensors using LSV as the electrochemical technique in different culture media diluted and not diluted with PBS.
MediumSensitivity
µAµM−1cm−2
Limit of Detection
µM
Linear Range
µM
Pure RPMI0.1343.1220–6000
Diluted RPMI0.1343.1220–6000
Pure MEM0.1254.5720–6000
Diluted MEM0.1254.5720–6000
Pure DMEM0.1085.720–6000
Diluted DMEM0.1055.820–6000
Pure Ham’s F120.1116.7750–6000
Diluted Ham’s F120.0823.0650–6000
Pure B/D0.0626.71100–6000
Diluted B/D0.1115.250–6000
Table 3. Stability of the AuNPs-rGO-based sensors after 21 days of storage.
Table 3. Stability of the AuNPs-rGO-based sensors after 21 days of storage.
SampleCurrent Density
µAcm−2
Difference
%
Fresh electrode240 ± 120%
Immersed in deionized water at 4 °C189 ± 8.5−21.2%
Immersed in deionized water at 20 °C212 ± 11.7−12%
Immersed in PBS at 4 °C157 ± 7.5−35%
Immersed in PBS at 20 °C171 ± 8.9−29%
Stored in air at 4 °C245 ± 9.8+2%
Stored in air at 20 °C153 ± 9.2−36%
Vacuum at 4 °C259 ± 14.2+7.9%
Vacuum at 20 °C222 ± 9.3−7.5%
Table 4. Real samples analysis tested using AuNPs-rGO-based sensors with LSV.
Table 4. Real samples analysis tested using AuNPs-rGO-based sensors with LSV.
SampleNT
(µM)
CSE
(µM)
RES
(µM)
RES + CSE
(µM)
16 HBE-MEM10.63 ± 1.2841.5 ± 11.89.7 ± 2.115.58 ± 3.58
A549-DMEM12.92 ± 1.0415.8 ± 1.15--
PBEC-B/D42.6 ± 7.1129 ± 21.5--
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Patella, B.; Vincenzo, S.D.; Zanca, C.; Bollaci, L.; Ferraro, M.; Giuffrè, M.R.; Cipollina, C.; Bruno, M.G.; Aiello, G.; Russo, M.; et al. Electrochemical Quantification of H2O2 Released by Airway Cells Growing in Different Culture Media. Micromachines 2022, 13, 1762. https://doi.org/10.3390/mi13101762

AMA Style

Patella B, Vincenzo SD, Zanca C, Bollaci L, Ferraro M, Giuffrè MR, Cipollina C, Bruno MG, Aiello G, Russo M, et al. Electrochemical Quantification of H2O2 Released by Airway Cells Growing in Different Culture Media. Micromachines. 2022; 13(10):1762. https://doi.org/10.3390/mi13101762

Chicago/Turabian Style

Patella, Bernardo, Serena Di Vincenzo, Claudio Zanca, Luciano Bollaci, Maria Ferraro, Maria Rita Giuffrè, Chiara Cipollina, Maria Giuseppina Bruno, Giuseppe Aiello, Michele Russo, and et al. 2022. "Electrochemical Quantification of H2O2 Released by Airway Cells Growing in Different Culture Media" Micromachines 13, no. 10: 1762. https://doi.org/10.3390/mi13101762

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

Article Metrics

Back to TopTop