Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Mar 1.
Published in final edited form as: J Electroanal Chem (Lausanne). 2022 Jan 31;908:116101. doi: 10.1016/j.jelechem.2022.116101

Ferrocene-Grafted Carbon Nanotubes for Sensitive Non-Enzymatic Electrochemical Detection of Hydrogen Peroxide

Bo Wu 1, Sanjida Yeasmin 1, Ye Liu 1, Li-Jing Cheng 1,*
PMCID: PMC8896809  NIHMSID: NIHMS1778479  PMID: 35250418

Abstract

Sensitive detection of hydrogen peroxide (H2O2) residue in aseptic packaging at point of use is critical to food safety. We present a sensitive non-enzymatic, amperometric H2O2 sensor based on ferrocene-functionalized multi-walled carbon nanotubes (MWCNT-FeC) and facile screen-printed carbon electrodes (SPCEs). The sensor utilizes the covalently grafted ferrocene as an effective redox mediator and the MWCNT networks to provide a large active surface area for efficient electrocatalytic reactions. The electrocatalytic MWCNT-FeC modified electrodes feature a high-efficiency electron transfer and a high electrocatalytic activity towards H2O2 reduction at a low potential of −0.15 V vs. Ag/AgCl. The decreased operating potential improves the selectivity by inherently eliminating the cross-reactivity with other electroactive interferents, such as dopamine, glucose, and ascorbic acid. The sensor exhibits a wide linear detection range from 1 μM to 1 mM with a detection limit of 0.49 μM (S/N=3). The covalently functionalized electrodes offered highly reproducible and reliable detection, providing a robust property for continuous, real-time H2O2 monitoring. Furthermore, the proposed sensor was successfully employed to determine H2O2 levels in spiked packaged milk and apple juice with satisfactory recoveries (94.33–97.62%). The MWCNT-FeC modified SPCEs offered a facile, cost-effective method for highly sensitive and selective point-of-use detection of H2O2.

Keywords: Hydrogen peroxide, non-enzymatic, ferrocene carbon nanotube, amperometric biosensor, low potential

1. Introduction

Over the past few decades, the determination of hydrogen peroxide (H2O2) residue has been critical in the food industry and environmental analysis [13]. Hydrogen peroxide is commonly used as a preservative in milk and applied for sterilization in aseptic food packaging due to its inherent bactericidal and sporicidal properties [4]. However, repeated exposure to H2O2 residue in food or beverage may increase the risk of many health problems, including skin irritation and gastrointestinal tract, cyanosis, and cardiac arrest [5]. According to the U.S. Food and Drug Administration (FDA) regulation (21 CFR 178.1005), the H2O2 residue level in the aseptically-packaged products must not exceed 0.5 ppm (~15 μM) [1, 6]. Thus, real-time monitoring of H2O2 levels in food production or detecting H2O2 residue at point of use is critical to food safety in the food and beverage industry. Conventional H2O2 sensing relies on spectroscopy [7, 8] and enzyme-based fluorescence or chemiluminescence assays [9, 10]. Although these techniques provide sensitive and selective detections, they generally rely on expensive lab-based equipment, require sample pre-treatments, and suffer from low stability due to enzymes’ intrinsic nature, making them unsuitable for on-site routine analysis of H2O2 [11]. On the contrary, electrochemical sensors offer rapid, simple, cost-effective approaches for sensitive and selective detection [1215]. Furthermore, they also allow direct real-time analysis without sample pre-treatment procedures [11, 16, 17]. Among them, enzyme-based electrochemical biosensors provide high sensitivity and low detection limits. However, the complicated enzyme immobilization and stabilization processes, the environment-dependent enzyme activity, and the high cost hinder their general applications [18]. Therefore, it is highly desirable to develop an enzyme-free H2O2 sensor that can overcome these issues by incorporating electrocatalytic nanomaterials.

Various approaches have been reported to construct a non-enzymatic sensor based on oxidation or reduction of H2O2 [1921]. The oxidation of H2O2 often suffers from high over-potential, which simultaneously produces oxidative responses from any non-target species in the biofluids, including ascorbate, uric acid, dopamine, etc., leading to poor sensing specificity [4]. The sensors based on Prussian Blue for catalytic reduction of H2O2 have been demonstrated to exhibit a low detection limit and good selectivity. However, Prussian Blue-contained electrodes were found to exhibit low operational stability and risk of cyanide leakage, limiting their practical application [19, 22]. On the other hand, ferrocene is an alternative catalyst that supports a low over-potentials for the redox reaction of H2O2. Ferrocene becomes one of the most exploited catalysts for non-enzymatic sensors because it supports fast electron-transfer rates, good stability of two redox states, and, more importantly, excellent electrocatalytic activity towards both the reduction and oxidation of H2O2 [23]. Efforts have been made for immobilizing ferrocene onto the sensor surface. Physical adsorption of ferrocene/ferrocenium (FeC/FeC+) redox couples on electrodes were applied to facilitate H2O2 detection [21] but yielded limited stability because they are likely to detach from the electrode surface [24]. Polymers with covalently attached FeC redox mediators have been reported to overcome the stability issue with proper electrochemical properties [1, 25]. However, the polymer backbone tends to swell during sensing, increasing the distance between FeC mediators and hindering the electron transfer [2628]. Moreover, these sensors require tedious and complicated fabrication processes [29]. Therefore, developing a facile and robust sensor that covalently immobilizes redox reagents is crucial to enhance sensing performance and reliability. Conductive nanomaterials covalently bound with redox mediators, such as ferrocene-graphene nanosheets [21], can be promising in accelerating electron transfer processes. Compared to graphene, multi-walled carbon nanotubes have been proven superior in supporting a larger surface area and strong mechanical property [30]. Electrode material aside, most H2O2 sensors were demonstrated on stand-alone glassy carbon electrodes [3133], which can hardly be miniaturized for point-of-use applications. In comparison, screen-printed carbon electrodes (SPCEs) can be scaled and integrated to form a low-cost, disposable coplanar three-electrode system suitable for mass production [34].

In this work, we demonstrate a facile, sensitive, non-enzymatic H2O2 sensor based on ferrocene-grafted multi-walled carbon nanotube (MWCNT-FeC) nanocomposites. The nanocomposite offers several advantages, including strong catalytic properties towards H2O2 reduction, outstanding electron transfer property, and large electroactive surface area. The unique features enable the MWCNT-FeC modified SPCE to detect H2O2 with high sensitivity, wide dynamic range, and high stability, compared with the H2O2 sensors reported to date. More importantly, the sensor operates at a low reduction potential of −0.15 V for H2O2 sensing that inherently avoids the cross-reactivity with common electroactive interferences, such as glucose, ascorbic acid, and dopamine, leading to a highly selective H2O2 detection. The practical application was validated by determining the H2O2 levels in spiked aseptic milk and juice. The results suggest that the proposed MWCNT-FeC modified SPCEs are promising for practical H2O2 detection.

2. Materials and methods

2.1. Reagents and Apparatus

Ferrocene carboxaldehyde (FeC-CHO), ferrocyanide/ferricyanide ([Fe(CN)6]3/4−), glucose, and L-ascorbic acid were purchased from Sigma-Aldrich. Hydrogen peroxide (H2O2, 30% solution) was purchased from Macron Fine Chemicals. Chitosan (CS, 85% deacetylated), sodium cyanoborohydride (NaCNBH3) was purchased from Alfa Aesar. 3-Hydroxytyramine hydrochloride (dopamine, DA) was purchased from Acros Organics. Phosphate-buffered saline (PBS, 10X, pH 7.4) was purchased from Fisher Scientific. Plasma-created amine-functionalized multi-walled carbon nanotubes (MWCNT-NH2) were procured from Cheap Tubes, Inc. (USA). Carbon ink (CI-2042) and silver/silver chloride ink (CI-4002) were purchased from Engineered Materials Systems, Inc. All other chemicals were of analytical grade and used without further purification. All the solutions were prepared using double distilled water.

2.2. Synthesis of MWCNTs-FeC nanocomposite

Ferrocene grafted multi-walled carbon nanotubes (MWCNTs-FeC) were synthesized using a procedure modified from a reported method for conjugation of FeC on polymer [35]. The process started with dispersing 30 mg MWCNT-NH2 in 2.5 mL 0.1 M acetic acid solution. FeC-CHO (35 mg) was dissolved in 5 mL methanol solution and added to the MWCNT-NH2 dispersion. After being fully sonicated for 10 min, the reaction mixture was stirred for 4 hours at 35 °C, followed by adding NaCNBH3 (50 mg) to the mixture and continuing the reaction for another 24 hours. Subsequently, the products were extracted by centrifugation and rinsed thoroughly with double distilled water and methanol three times. Finally, the products were dried under vacuum and stored as a powder for further use.

2.3. Preparation of screen-printed carbon electrodes

The flexible SPCEs were fabricated on polyethylene terephthalate (PET) substrates using a homemade stainless-steel screen-printing mask. The three-step manufacturing process started with transferring carbon ink through a mask onto the PET substrate to provide electrical contacts. The electrode connections, 3 mm diameter working electrode, and auxiliary electrode were subsequently baked at 110 °C for 15 min. The reference electrode was printed with silver/silver chloride ink, then cured at 110 °C for 10 min [36]. Finally, polyimide (Kapton) tape was used to delimit the sensor area, leaving contact pads and connections on the other side of the sensor.

2.4. Preparation of MWCNTs-FeC/CS composite-modified electrode

The 1 mg synthesized MWCNT-FeC was dispersed in a 1 mL, 0.5 mg mL−1 chitosan solution prepared in 0.2 M acetic acid, followed by a 20 min probe sonication. The resultant 4 μL MWCNT-FeC/CS homogeneous solution was drop-casted onto the working electrode of SPCE, followed by 30 min baking at 50 ℃. CS improved the uniformity of the MWCNT-FeC in the drop-casting process. The modified electrodes were stored at 4 ℃ for future use. The fabrication process of the MWCNT-FeC/CS composite-modified SPCE is presented in Fig. S1. Control samples (MWCNT-NH2/SPCE) prepared by drop-casting untreated MWCNT-NH2/CS composite on SPCE were created to investigate the effect of FeC modification on the electrocatalytic property of the modified electrodes.

2.5. Material and device characterizations

Scanning electron microscopy (SEM) images and energy-dispersive X-ray spectroscopy (EDX) were acquired on a QUANTA 600F scanning electron microscope (FEI, USA). Electrochemical measurements, including cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), and chronoamperometry (CA), were performed with a three-electrode system using potentiostat (Gamry Reference 600+) electrochemical system driven by Gamry Framework™ Applications Version 6.3. CV measurements were conducted with a scan rate of 100 mV s−1 between −0.3 V and 0.5 V in 1X PBS solution at pH 7.0. The CV results indicate that the sensor performed most efficiently at an optimum voltage of −0.15 V vs. Ag/AgCl, selected as the applied potential for amperometric measurements. The sensor’s amperometric response to H2O2 was recorded at a working potential of −0.15 V (vs. Ag/AgCl) with various H2O2 concentrations in a 1X PBS solution (scheme shown in Fig. 1). The sensing data was taken at 30 seconds when the current reached a steady state.

Figure 1.

Figure 1.

Printable H2O2 sensor enabled by ferrocene grafted multi-walled carbon nanotubes (MWCNT-FeC) and its H2O2 sensing mechanism.

2.6. H2O2 detection in commercial products

Milk and apple juice purchased from a local grocery store were firstly spiked with H2O2 to reach concentrations above, equal, and below 15 μM. After thoroughly mixing for 5 min, the H2O2 spiked samples were diluted with 10X PBS in a 9:1 ratio. After 5 min shaking, the mixture was centrifuged at 7000 rpm for 15 min, and the top supernatant was collected for H2O2 detection under the optimized conditions.

3. Result and discussion

3.1. Morphology Characterization of MWCNTs-FeC modified SPCE

The SEM image of a bare SPCE surface in Fig. 2(a) shows the porous surface morphology of carbon black. The coating of MWCNT-FeC composite on SPCE in Fig. 2(b) resulted in excellent coverage of woven mesh-like carbon nanotube networks over the electrode surface, creating a three-dimensional structure that increased the electroactive area for electrochemical reaction. The EDX analysis of the bare SPCE in Fig. 2(c) displays the presence of 83.67 at. % carbon and 16.29 at. % oxygen. On the other hand, the MWCNT-FeC modified SPCE surface contained additional peaks in the EDX spectrum corresponding to irons as shown in Fig. 2(d), yielding 92.81 at. % carbon, 6.97 at. % oxygen and 0.23 at. % iron. The presence of iron confirms the existence of ferrocene on the modified electrode. Also, the increased carbon percentage on the MWCNT-FeC modified electrode resulted from the presence of high-density carbon nanotubes.

Figure 2.

Figure 2.

SEM images of (a) SPCE and (b) MWCNT-FeC modified SPCE. Scale bars indicate 1 μm. EDX spectra of (c) SPCE and (d) MWCNT-FeC modified SPCE.

3.2. Electrochemical characterization of MWCNTs-FeC/SPCE

CV measurements were carried out to further verify the successful conjugation of ferrocene on the carbon nanotubes. Fig. 3(a) shows the CVs of the bare SPCE, MWCNT-NH2 modified SPCE, and MWCNT-FeC modified SPCE characterized in 1X PBS at a sweep rate of 100 mV s−1 without redox reagents in the solution. Both bare SPCE and MWCNT-NH2 modified SPCE did not present any indication of redox activity. However, the MWCNT-NH2 modified SPCE exhibited an increased baseline current due to the enlarged electroactive surface area contributed by the dense MWCNT networks. A pair of stable and well-defined redox peaks were observed on the MWCNT-FeC modified SPCE with the anodic and cathodic peak potentials at +0.117 V and +0.079 V vs. Ag/AgCl, respectively. The redox peaks were the electrochemical signatures of the immobilized FeC/FeC+ redox couple on MWCNTs, implying the successful conjugation of ferrocene to the MWCNTs [21, 37]. The broadening of the redox peaks was probably due to the interactions of π-π stacking between the FeC groups [38]. The modified electrode yielded a +98 mV formal potential calculated from the average of the cathodic and anodic peak potentials and 42 mV potential separation between reduction and oxidation peaks. The formal potential and peak potential separation of this work were lower than the counterparts reported previously [3941] due to the improved electron transfer rate and a larger electroactive surface area [42]. Fig. 3(b) shows the CVs of the MWCNT-FeC modified SPCE under various potential scan rates. The oxidation and reduction peak currents increase linearly with the scan rate in the range from 10 to 600 mV s−1, as shown in Fig. 3(c). The peak current Ip as a function of scan rate v agrees with Laviron’s theory, Ip = n2F2vAΓ/4RT, with A, F, R, and T being the surface area, Faraday constant, the universal gas constant, and temperature in Kelvin, respectively [43]. The n is the number of electrons involved in the catalytic reaction, and Γ is the electroactive coverage of the catalyst on the electrode. The linear relationship between the peak current Ip and scan rate v implies that the surface-confined redox reagents were confined on the electrode surface due to the covalently immobilized ferrocene on the MWCNT modified electrode. Compared to the previously reported CNT-FeC nanocomposites modified electrodes [4446], the proposed sensor provided a higher electroactive coverage of ferrocene (Γ = 2.12 × 10−9 mol cm−2) and a more prominent redox peak. The results can be attributed to the high-density amine functional groups created on the MWCNT surface, offering abundant conjugation sites for ferrocene grafting [47, 48]. The short functional linkers reduced the distance between the FeC redox center and the conductive MWCNTs, thus promoting electron transfer efficiency [49]. The interfacial properties of the electrodes were analyzed using electrochemical impedance spectroscopy (EIS) in the presence of [Fe(CN)6]3/4− redox reagents in the solution. As shown in Fig. 3(d), the charge transfer resistance (Rct) of the bare SPCE decreased from 2.3 kΩ to 146 Ω after being modified with MWCNT-NH2 due to the introduction of a high-density conductive MWCNTs network. The Rct further reduced to 97 Ω after FeC grafting. The reduced Rct across the electrode-electrolyte interface could result from the covalent conjugation of FeC on MWCNTs that brought the high-density FeC/FeC+ redox centers very close to the MWCNTs, further improving the charge transfer channel to the electron [41].

Figure 3.

Figure 3.

(a) CVs of (1) SPCE, (2) MWCNT-NH2 modified SPCE, and (3) MWCNT-FeC modified SPCE measured in 1X PBS at 100 mV s−1 scan rate. (b) CVs of MWCNT-FeC modified SPCE measured in 1X PBS at various scan rates (from inner to outer curves: 10, 50, 100, 200, 300, 400, 500, 600 mV s−1). (c) Anodic and cathodic peak currents vs. scan rate. (d) EIS of (1) SPCE, (2) MWCNT-NH2 modified SPCE, and (3) MWCNT-FeC modified SPCE measured in 0.1 M KCl containing 1.0 mM [Fe(CN)6]3−/4− (1:1). Inset: Randles equivalent circuit model for the extraction of charge transfer resistance Rct.

3.3. Hydrogen peroxide detection

The H2O2 electrocatalytic activity of the developed sensor was evaluated using cyclic voltammetry and chronoamperometry. The CVs in Fig. 4 show that the MWCNT-FeC modified SPCE significantly increased both reduction and oxidation of H2O2 compared with the MWCNT-NH2 modified SPCE. The feature was not observed in most of the reported modified electrodes, which only promotes either oxidation [50, 51] or reduction [52, 53] of H2O2. After introducing 1 mM H2O2, the current of the MWCNT-FeC modified electrode increased by 8.3 μA at −0.15 V and 2.8 μA at 0.6 V, whereas the current of the non-functionalized MWCNT-NH2 modified electrode increased by only 1 μA at −0.15 V and 0.8 μA at 0.6 V. The elevated current in MWCNT-FeC electrodes suggests ferrocene’s strong electrocatalytic activity towards H2O2 on the functionalized MWCNT surfaces for both the oxidation of H2O2 to O2 and the reduction to H2O. Overall, the improved electrocatalytic activity can be attributed to the large electron transfer through the modified carbon nanotube, the shorter distance between the mediators and electrode, and FeC’s strong electrocatalytic activity towards H2O2 reduction.

Figure 4.

Figure 4.

CVs of (a) MWCNT-NH2 modified SPCE and (b) MWCNT-FeC modified SPCE in response to various H2O2 concentrations: (1) 0 mM, (2) 1 mM, and (3) 10 mM in 1X PBS at 100 mV s−1 scan rate.

Apart from detection sensitivity, the operating potential for the amperometric measurements must be properly chosen to eliminate interferences from the common electroactive species and achieve highly selective H2O2 detection. The CVs of the MWCNT-FeC modified SPCE in the absence and presence of 0.5 mM H2O2 and other interferents, including glucose, ascorbic acid, and dopamine, are shown in Fig. 5(ad), respectively. Fig. 5(e) and 5(f) summarize the CV currents acquired at −0.15 V and 0.6 V electrode potentials vs. Ag/AgCl. At 0.6 V electrode potential, ascorbic acid and dopamine were easily oxidized, yielding the current changes larger than that contributed by the oxidation of H2O2 at the same potential, which agrees with our previous observation [54]. In contrast, at −0.15 V, the electrode actively reduced H2O2 while showed negligible responses to glucose and ascorbic acid, leading to a substantial current change compared with those produced by the three interferents, as shown in Fig. 5(e). The MWCNT-FeC modified electrode efficiently reduced H2O2 to H2O at a potential as low as −0.15 vs. Ag/AgCl and allowed specific detection of H2O2 without the simultaneous oxidization or reduction of the interferents, significantly improving the detection selectivity. Applying a more negative electrode potential did not considerably increase the current response. Therefore, this electrode potential was then chosen to examine the amperometric response against H2O2 reduction. The MWCNT-FeC modified SPCE enables a highly selective H2O2 detection with a much lower electrode potential than the previous works [55, 56]. The low operating potential for H2O2 detection also prevented the interference of O2 reduction in solution. As presented in Fig. S2, detection of H2O2 in the analytes purged with air and nitrogen gas for 10 min yielded similar responses. Because FeC did not present strong electrocatalytic activity towards oxygen reduction and the detection was performed at a small reduction potential, we did not observe the impact of O2 reduction on H2O2 detection.

Figure 5.

Figure 5.

CVs of MWCNT-FeC modified SPCE in the absence and the presence of 0.5 mM (a) H2O2, (b) glucose, (c) ascorbic acid, and (d) dopamine in 1X PBS at 100 mV s−1 scan rate. CV current change in response to 0.5 mM analytes obtained at −0.15 V (e) and 0.6 V (f) vs. Ag/AgCl as indicated by the circles in the CV curves (a-d).

Fig. 6(a) shows the amperometric recordings of H2O2 detection at various concentrations. The H2O2 analyte solution was loaded for detection at time zero with the sensor biased at −0.15 V. The resulting calibration curve shown in Fig. 6(b) summarizes the current levels acquired in Fig. 6(a) at 30 s. The sensor exhibited a linear response to H2O2 (R2 = 0.997) in a broad concentration range from 1 μM to 1 mM. The amperometric current change ΔI (μA) in response to H2O2 concentration c (mM) follows the relation ΔI = 9.26 c + 0.38. A high current density of 136.4 μA mM−1 cm−2 and a low detection limit of 0.49 μM H2O2 were obtained in the linear range at a signal-to-noise ratio of 3, low enough to detect H2O2 in practical applications. We observed a noticeable increase in current when H2O2 concentration increases from zero to 1 μM in the inset of Fig. 6(a), resulting in a current offset in the calibration curve after a straight-line fitting as shown in Fig. 6(b). The offset could be related to the reactive reduction of H2O2 on the MWCNT-FeC modified electrode. We observed that the presence of H2O2 in buffer solutions can produce a detectable net current even at zero electrode bias. The zero-bias current slightly increased with H2O2 concentration. The result implies that part of the sensing current measured at −0.15 V was contributed by the intrinsic reduction of H2O2 at zero voltage bias. At low concentrations, the zero-bias reduction current accounted for a relatively large share of the total sensing current. Further study is required to clarify the effect.

Figure 6.

Figure 6.

(a) Amperometric response of MWCNT-FeC modified SPCE to various H2O2 concentrations in 1X PBS (pH 7.0). (b) Calibration curve for H2O2 detection. (c) Calibration curves for glucose, ascorbic acid (AA), and dopamine (DA). The sensor electrode was biased at −0.15 V vs. Ag/AgCl.

Compared to the ferrocene-based H2O2 sensors reported previously [24], the MWCNT-FeC modified electrode showed a superior H2O2 detection sensitivity, lower detection limit, and improved selectivity due to the operation at low reduction potential. Fig. 6(c) shows that glucose and ascorbic acid displayed imperceptible amperometric current changes. Dopamine yielded a current level of less than a quarter of the sensing response of H2O2 of the same concentrations. It is worth noting that the dopamine concentrations in the actual samples are much lower than the values demonstrated here for selectivity validation.

The MWCNT-FeC modified SPCE also exhibited highly reproducible and repeatable H2O2 detection. As shown in Fig. 7(a), repeated measurements on the same device show a negligible change in the sensing results. Fig. 7(b) shows that the device-to-device variation collected from five sensors was measured to be only 3.7%, indicating the reproducibility of the H2O2 sensors. The sensor interrogation-regeneration plots in Fig. 7(c) show repeatable H2O2 sensing results during four cycles of alternating detection and refreshing processes in 1X PBS buffer, implying that the sensor can be reused after a simple regeneration process by rinsing it with 1X PBS buffer. Overall, the sensors were found to be reasonably stable over time. Fig. 7(d) shows a ~ 8% drop in sensing signal for the devices stored for ten days at 4°C. The shelf life is expected to be further improved with proper humidity control. The reproducible, stable, and reliable detection results signify that the proposed sensor is a promising candidate for practical measurements.

Figure 7.

Figure 7.

(a) Repeatability tests of the sensor. (b) Device-to-device variation tests. (c) Interrogation-regeneration plot for the sensor during alternating H2O2 sensing (S) and PBS rinsing (R) processes. (d) Stability of the sensor over 10 days. The sensors were tested in 0.5 mM H2O2 in 1X PBS.

Real-time, continuous H2O2 sensing was further evaluated using a sensor integrated with a PDMS microfluidic channel, as shown in Fig. 8(a). The printed sensor with the working electrode biased at −0.15 V vs. Ag/AgCl continuously monitored the H2O2 levels in an analyte inflow. The analyte was delivered to the sensor area with a constant flow rate using a syringe pump. Interferents, including ascorbic acid, glucose, and dopamine, were added at different time points to evaluate the detection selectivity. Fig. 8(b) shows the amperometric recordings of successive additions of H2O2 and the interferents. The results suggest that even the concentrations of the interferents were four times higher than H2O2, the interferents’ current responses are much lower than those contributed by H2O2. It implies the MWCNT-FeC modified SPCE sensor has an excellent selectivity towards H2O2.

Figure 8.

Figure 8.

(a) Image of a real-time sensing setup; (b) Chronoamperometry of the sensor responding to H2O2 and interferents. The working electrode is biased at −0.15 V vs. Ag/AgCl.

The MWCNT-FeC modified SPCE sensor presented in this paper exhibited a superior sensing performance and promising applications for continuous, real-time sensing. Table 1 compares this work with several representative electrochemical H2O2 sensors reported to date. The comparison indicates that the proposed MWCNT-FeC modified SPCE sensor featured superior sensing performance with low detection limit, wide linear dynamic range, high selectivity, low operating potential, and reproducibility.

Table 1.

Comparison of representative electrochemical H2O2 sensors with this work

Electrode Working voltage (V) Dynamic range (μM) LoD (μM) Interferents Real samples CH2O2 in real samples (μM) Ref.
RGO/CNTs-Pt/GCE −0.2 0.3–4000 0.31 DA, Cys, Fru, GSH, GLC, AA Milk 32–1218 [57]
[Cu(bmtc)2(H2O)]/SPCE −0.6 10–524 0.57 Urea, AA, GLC, UA Milk 10–30 [58]
LSG/AgNPs −0.56 100–10000 7.9 AA, GLC, KCl, NaCl Milk 200–500 [59]
BiVO4/TiO2 5 5–200 5 UA, AA, B12, GLY, ASN N/A N/A [60]
GNPs-PB/BG/GCE −0.05 9.6–143 3.2 AA, UA, GLC, GLY, APAP N/A N/A [61]
MWCNTs-FeC/SPCEs −0.15 1–1000 0.49 DA, AA, GLC Milk, apple juice 10–50 This work

GCE: glassy carbon electrode; Fru: fructose; Cys: cysteine; GSH: glutathione; [Cu(bmtc)2(H2O)]: copper (II) 1-(3-bromobenzyl)-5-methyl-1H-1,2,3-triazole-4-carboxylate; UA: uric acid; LSG: laser scribed graphene electrode; AgNPs: silver nanoparticles; NaCl: sodium chloride; GLY: glycine; PB: prussian blue; BG: bucky gel; GNPs: gold nanoparticles; ASN: asparagine; APAP: acetaminophen.

3.4. Determination of H2O2 in the real samples

The feasibility of sensing H2O2 in real samples using MWCNT-FeC modified SPCE was evaluated through a recovery study using commercial milk and apple juice samples. The samples were spiked with three known H2O2 concentrations, including 15 μM, the maximum H2O2 level allowed by U.S. FDA for aseptic packaging. As shown in Table 2, the recovery rates and RSD values were in the range of 94.33–97.62% and 1.93–4.46%, respectively, suggesting that the MWCNT-FeC modified SPCE sensor can be readily used to analyze H2O2 residues in commercially packaged beverages.

Table 2.

Determination of H2O2 in commercially packaged milk and apple juice samples using MWCNT-FeC/SPCE.

Samples Spiked (μM) After dilution (μM) Found (μM) RSD (%) Recovery (%)
Milk #1 10 9.0 8.52 4.20 94.66
Milk #2 15 13.5 12.95 2.36 95.93
Milk #3 50 45.0 42.87 1.93 95.27
Juice #1 10 9.0 8.49 3.84 94.33
Juice #2 15 13.5 12.99 2.78 96.22
Juice #3 50 45 43.93 4.46 97.62

Relative standard deviation (RSD) of 3 measurements.

4. Conclusion

In summary, we have successfully demonstrated a sensitive printable H2O2 sensor based on ferrocene-grafted carbon nanotubes for enzyme-free amperometric H2O2 sensors prepared by a facile and cost-effective method. Owing to the high-density FeC redox centers on MWCNT, efficient electron transfer between FeC and MWCNT, and large electrical conductance of the MWCNT network, the sensor exhibits excellent electrocatalytic activity towards H2O2 reduction with a low detection limit of 0.49 μM and a wide linear range from 1 μM to 1 mM covering the interested H2O2 level in aseptically-packaged products. Moreover, the sensor can operate at a much lower electrode potential of −0.15 V vs. Ag/AgCl, further improving detection selectivity. Furthermore, the high repeatability of the sensor allowed for real-time, continuous monitoring of H2O2 concentration in analyte flows. Successful determination of H2O2 in spiked beverages also displayed satisfactory recoveries. The H2O2 sensor based on MWCNT-FeC modified SPCE offers an excellent balance between simplicity, cost, and performance, which appears as a good candidate for routine sensing of H2O2 at the point of use.

Supplementary Material

1

Acknowledgments

The authors acknowledge partial financial support from the National Science Foundation (1810067) and the National Institutes of Health (1R21DE027170-01).

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1].Mattoussi M, Matoussi F, Raouafi N, Non-enzymatic amperometric sensor for hydrogen peroxide detection based on a ferrocene-containing cross-linked redox-active polymer, Sens. Actuators B 274 (2018) 412–418. [Google Scholar]
  • [2].Li G, Wang Y, Xu H, A hydrogen peroxide sensor prepared by electropolymerization of pyrrole based on screen-printed carbon paste electrodes, Sensors 7(3) (2007) 239–250. [Google Scholar]
  • [3].Saleh Ahammad A, Hydrogen peroxide biosensors based on horseradish peroxidase and hemoglobin, J Biosens Bioelectron S 9 (2013) 2. [Google Scholar]
  • [4].Hsu C-L, Chang K-S, Kuo J-C, Determination of hydrogen peroxide residues in aseptically packaged beverages using an amperometric sensor based on a palladium electrode, Food Control 19(3) (2008) 223–230. [Google Scholar]
  • [5].Watt BE, Proudfoot AT, Vale JA, Hydrogen peroxide poisoning, Toxicol. Rev 23(1) (2004) 51–57. [DOI] [PubMed] [Google Scholar]
  • [6].Nasirizadeh N, Shekari Z, Nazari A, Tabatabaee M, Fabrication of a novel electrochemical sensor for determination of hydrogen peroxide in different fruit juice samples, J Food Drug Anal 24(1) (2016) 72–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Nogueira RFP, Oliveira MC, Paterlini WC, Simple and fast spectrophotometric determination of H2O2 in photo-Fenton reactions using metavanadate, Talanta 66(1) (2005) 86–91. [DOI] [PubMed] [Google Scholar]
  • [8].Abbas M, Luo W, Zhu L, Zou J, Tang H, Fluorometric determination of hydrogen peroxide in milk by using a Fenton reaction system, Food Chem 120(1) (2010) 327–331. [Google Scholar]
  • [9].Juven BJ, Pierson MD, Antibacterial effects of hydrogen peroxide and methods for its detection and quantitation, J. Food Prot 59(11) (1996) 1233–1241. [DOI] [PubMed] [Google Scholar]
  • [10].Allain CC, Poon LS, Chan CS, Richmond W, Fu PC, Enzymatic determination of total serum cholesterol, Clin. Chem 20(4) (1974) 470–475. [PubMed] [Google Scholar]
  • [11].Kanyong P, Rawlinson S, Davis J, A non-enzymatic sensor based on the redox of ferrocene carboxylic acid on ionic liquid film-modified screen-printed graphite electrode for the analysis of hydrogen peroxide residues in milk, J. Electroanal. Chem 766 (2016) 147–151. [Google Scholar]
  • [12].Saber-Tehrani M, Pourhabib A, Husain SW, Arvand M, A simple and efficient electrochemical sensor for nitrite determination in food samples based on Pt nanoparticles distributed poly (2-aminothiophenol) modified electrode, Food Anal. Methods 6(5) (2013) 1300–1307. [Google Scholar]
  • [13].Chandran S, Lonappan LA, Thomas D, Jos T, Kumar KG, Development of an electrochemical sensor for the determination of amaranth: a synthetic dye in soft drinks, Food Anal. Methods 7(4) (2014) 741–746. [Google Scholar]
  • [14].Rasheed Z, Vikraman AE, Thomas D, Jagan JS, Kumar KG, Carbon-nanotube-based sensor for the determination of butylated hydroxyanisole in food samples, Food Anal. Methods 8(1) (2015) 213–221. [Google Scholar]
  • [15].Raoof JB, Teymoori N, Khalilzadeh MA, ZnO Nanoparticle Ionic Liquids Carbon Paste Electrode as a Voltammetric Sensor for Determination of Sudan I in the Presence of Vitamin B 6 in Food Samples, Food Anal. Methods 8(4) (2015) 885–892. [Google Scholar]
  • [16].Chaubey A, Malhotra B, Mediated biosensors, Biosens. Bioelectron 17(6–7) (2002) 441–456. [DOI] [PubMed] [Google Scholar]
  • [17].Hart JP, Crew A, Crouch E, Honeychurch KC, Pemberton RM, Some recent designs and developments of screen‐printed carbon electrochemical sensors/biosensors for biomedical, environmental, and industrial analyses, Anal. Lett 37(5) (2004) 789–830. [Google Scholar]
  • [18].Shamkhalichenar H, Choi J-W, Non-enzymatic hydrogen peroxide electrochemical sensors based on reduced graphene oxide, J. Electrochem. Soc 167(3) (2020) 037531. [Google Scholar]
  • [19].Karyakin AA, Advances of Prussian blue and its analogues in (bio) sensors, Curr. Opin. Electrochem 5(1) (2017) 92–98. [Google Scholar]
  • [20].Dey RS, Raj CR, Development of an amperometric cholesterol biosensor based on graphene− Pt nanoparticle hybrid material, J. Phys. Chem. C 114(49) (2010) 21427–21433. [Google Scholar]
  • [21].Fan L, Zhang Q, Wang K, Li F, Niu L, Ferrocene functionalized graphene: preparation, characterization and efficient electron transfer toward sensors of H 2 O 2, J. Mater. Chem 22(13) (2012) 6165–6170. [Google Scholar]
  • [22].Mohammad A, Yang Y, Khan M, Faustino P, Long-term stability study of Prussian blue—A quality assessment of water content and cyanide release, Clin. Toxicol 53(2) (2015) 102–107. [DOI] [PubMed] [Google Scholar]
  • [23].Chen W, Cai S, Ren Q-Q, Wen W, Zhao Y-D, Recent advances in electrochemical sensing for hydrogen peroxide: a review, Analyst 137(1) (2012) 49–58. [DOI] [PubMed] [Google Scholar]
  • [24].Zhao G-C, Xu M-Q, Zhang Q, A novel hydrogen peroxide sensor based on the redox of ferrocene on room temperature ionic liquid film, Electrochem. commun 10(12) (2008) 1924–1926. [Google Scholar]
  • [25].Pietschnig R, Polymers with pendant ferrocenes, Chem. Soc. Rev 45(19) (2016) 5216–5231. [DOI] [PubMed] [Google Scholar]
  • [26].Ju H, Gong Y, Zhu H, Electrolyte effects on electrochemical properties of osmium complex polymer modified electrodes, Anal. Sci 17(1) (2001) 59–63. [DOI] [PubMed] [Google Scholar]
  • [27].Calvo EJ, Danilowicz C, Diaz L, Enzyme catalysis at hydrogel-modified electrodes with redox polymer mediator, J. Chem. Soc., Faraday Trans 89(2) (1993) 377–384. [Google Scholar]
  • [28].Koide S, Yokoyama K, Electrochemical characterization of an enzyme electrode based on a ferrocene-containing redox polymer, J. Electroanal. Chem 468(2) (1999) 193–201. [Google Scholar]
  • [29].Bean LS, Heng LY, Yamin BM, Ahmad M, Photocurable ferrocene-containing poly (2-hydroxyl ethyl methacrylate) films for mediated amperometric glucose biosensor, Thin Solid Films 477(1–2) (2005) 104–110. [Google Scholar]
  • [30].Zheng D, Vashist SK, Dykas MM, Saha S, Al-Rubeaan K, Lam E, Luong JH, Sheu F-S, Graphene versus multi-walled carbon nanotubes for electrochemical glucose biosensing, Materials 6(3) (2013) 1011–1027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [31].Li S-J, Du J-M, Zhang J-P, Zhang M-J, Chen J, A glassy carbon electrode modified with a film composed of cobalt oxide nanoparticles and graphene for electrochemical sensing of H 2 O 2, Microchim. Acta 181(5–6) (2014) 631–638. [Google Scholar]
  • [32].Wang A-J, Zhang P-P, Li Y-F, Feng J-J, Dong W-J, Liu X-Y, Hydrogen peroxide sensor based on glassy carbon electrode modified with β-manganese dioxide nanorods, Microchim. Acta 175(1–2) (2011) 31. [Google Scholar]
  • [33].Kumar SA, Lo P-H, Chen S-M, Electrochemical analysis of H2O2 and nitrite using copper nanoparticles/poly (o-phenylenediamine) film modified glassy carbon electrode, J. Electrochem. Soc 156(7) (2009) E118. [Google Scholar]
  • [34].Migliorelli D, Generelli S, Glaser N, Viviani M, Mühlebach L, Junuzovic R, Screen-printed electrodes as a platform for smart and low-cost point of care devices, Multidisciplinary Digital Publishing Institute Proceedings 1(8) (2017) 842. [Google Scholar]
  • [35].Qiu J-D, Wang R, Liang R-P, Xia X-HJB, Bioelectronics, Electrochemically deposited nanocomposite film of CS-Fc/Au NPs/GOx for glucose biosensor application, 24(9) (2009) 2920–2925. [DOI] [PubMed] [Google Scholar]
  • [36].Nagar B, Balsells M, de la Escosura-Muñiz A, Gomez-Romero P, Merkoçi A, Fully printed one-step biosensing device using graphene/AuNPs composite, Biosens. Bioelectron 129 (2019) 238–244. [DOI] [PubMed] [Google Scholar]
  • [37].Wu B, Yeasmin S, Liu Y, Cheng L-J, Sensitive and selective electrochemical sensor for serotonin detection based on ferrocene-gold nanoparticles decorated multiwall carbon nanotubes, Sens. Actuators B 354 (2022) 131216. [Google Scholar]
  • [38].Sun N, Guan L, Shi Z, Li N, Gu Z, Zhu Z, Li M, Shao Y, Ferrocene peapod modified electrodes: preparation, characterization, and mediation of H2O2, Anal. Chem 78(17) (2006) 6050–6057. [DOI] [PubMed] [Google Scholar]
  • [39].Yan Y-M, Baravik I, Yehezkeli O, Willner I, Integrated electrically contacted glucose oxidase/carbon nanotube electrodes for the bioelectrocatalyzed detection of glucose, J. Phys. Chem. C 112(46) (2008) 17883–17888. [Google Scholar]
  • [40].Zribi B, Haghiri-Gosnet A-M, Bendounan A, Ouerghi A, Korri-Youssoufi H, Charge transfer and band gap opening of a ferrocene/graphene heterostructure, Carbon 153 (2019) 557–564. [Google Scholar]
  • [41].Elancheziyan M, Manoj D, Saravanakumar D, Thenmozhi K, Senthilkumar S, Amperometric sensing of catechol using a glassy carbon electrode modified with ferrocene covalently immobilized on graphene oxide, Microchim. Acta 184(8) (2017) 2925–2932. [Google Scholar]
  • [42].Ben Aoun S, Nanostructured carbon electrode modified with N-doped graphene quantum dots–chitosan nanocomposite: a sensitive electrochemical dopamine sensor, R. Soc. Open Sci 4(11) (2017) 171199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [43].Laviron E, General expression of the linear potential sweep voltammogram in the case of diffusionless electrochemical systems, J. Electroanal. Chem. Interf. Electrochem 101(1) (1979) 19–28. [Google Scholar]
  • [44].Diakowski PM, Xiao Y, Petryk MW, Kraatz H-B, Impedance based detection of chemical warfare agent mimics using ferrocene-lysine modified carbon nanotubes, Anal. Chem 82(8) (2010) 3191–3197. [DOI] [PubMed] [Google Scholar]
  • [45].Qiu J-D, Zhou W-M, Guo J, Wang R, Liang R-P, Amperometric sensor based on ferrocene-modified multiwalled carbon nanotube nanocomposites as electron mediator for the determination of glucose, Anal. Biochem 385(2) (2009) 264–269. [DOI] [PubMed] [Google Scholar]
  • [46].Kozma J, Papp S, Gyurcsányi RE, Solid-contact ion-selective electrodes based on ferrocene-functionalized multi-walled carbon nanotubes, Electrochem. commun 123 (2021) 106903. [Google Scholar]
  • [47].Chen Z, Dai XJ, Magniez K, Lamb PR, de Celis Leal DR, Fox BL, Wang X, Improving the mechanical properties of epoxy using multiwalled carbon nanotubes functionalized by a novel plasma treatment, Composites Part A: Applied Science and Manufacturing 45 (2013) 145–152. [Google Scholar]
  • [48].Chen C, Liang B, Lu D, Ogino A, Wang X, Nagatsu M, Amino group introduction onto multiwall carbon nanotubes by NH3/Ar plasma treatment, Carbon 48(4) (2010) 939–948. [Google Scholar]
  • [49].Huang K, Duclairoir F, Pro T, Buckley J, Marchand G, Martinez E, Marchon JC, De Salvo B, Delapierre G, Vinet F, Ferrocene and porphyrin monolayers on Si (100) surfaces: preparation and effect of linker length on electron transfer, ChemPhysChem 10(6) (2009) 963–971. [DOI] [PubMed] [Google Scholar]
  • [50].Zhou K, Zhu Y, Yang X, Luo J, Li C, Luan S, A novel hydrogen peroxide biosensor based on Au–graphene–HRP–chitosan biocomposites, Electrochim. Acta 55(9) (2010) 3055–3060. [Google Scholar]
  • [51].Wang M-Q, Zhang Y, Bao S-J, Yu Y-N, Ye C, Ni (II)-based metal-organic framework anchored on carbon nanotubes for highly sensitive non-enzymatic hydrogen peroxide sensing, Electrochim. Acta 190 (2016) 365–370. [Google Scholar]
  • [52].Yuan J, Xu S, Zeng H-Y, Cao X, Pan AD, Xiao G-F, Ding P-X, Hydrogen peroxide biosensor based on chitosan/2D layered double hydroxide composite for the determination of H2O2, Bioelectrochemistry 123 (2018) 94–102. [DOI] [PubMed] [Google Scholar]
  • [53].Sarkar A, Ghosh AB, Saha N, Bhadu GR, Adhikary B, Newly designed amperometric biosensor for hydrogen peroxide and glucose based on vanadium sulfide nanoparticles, ACS Appl. Nano Mater 1(3) (2018) 1339–1347. [Google Scholar]
  • [54].Bo W, Yeasmin S, Liu Y, Cheng L-J, Ferrocene Functionalized Gold Nanoparticles on Carbon Nanotube Electrodes for Portable Dopamine Sensor, ECS Meeting Abstracts, IOP Publishing, 2021, p. 1345. [Google Scholar]
  • [55].Cheng H, Guan L, Shi Z, Zhu Z, Gu Z, Li M, Bi‐Directional Electrocatalytic Detection of Dopamine at an Electrode Modified with Ferrocene‐Filled Carbon Nanotube Peapods, Electroanalysis 25(9) (2013) 2041–2044. [Google Scholar]
  • [56].Mohammadi SZ, Beitollahi H, Nikpour N, Hosseinzadeh R, Electrochemical sensor for determination of ascorbic acid using a 2-Chlorobenzoyl Ferrocene/Carbon nanotube paste electrode, Anal. Bioanal. Chem. Res 3(2) (2016) 187–194. [Google Scholar]
  • [57].Zhang Y, Cao Q, Zhu F, Xu H, Zhang Y, Xu W, Liao X, An Amperometric Hydrogen Peroxide Sensor Based on Reduced Graphene Oxide/Carbon Nanotubes/Pt NPs Modified Glassy Carbon Electrode, Int. J. Electrochem. Sci 15 (2020) 8771–8785. [Google Scholar]
  • [58].Quezada V, Martinez T, Nelson R, Pérez-Fehrmann M, Zaragoza G, Vizcarra A, Kesternich V, Hernández-Saravia LP, A novel platform of using copper (II) complex with triazole-carboxilated modified as bidentated ligand SPCE for the detection of hydrogen peroxide in milk, J. Electroanal. Chem 879 (2020) 114763. [Google Scholar]
  • [59].Aparicio-Martínez E, Ibarra A, Estrada-Moreno IA, Osuna V, Dominguez RB, Flexible electrochemical sensor based on laser scribed graphene/Ag nanoparticles for non-enzymatic hydrogen peroxide detection, Sens. Actuators B 301 (2019) 127101. [Google Scholar]
  • [60].Derbali M, Othmani A, Kouass S, Touati F, Dhaouadi H, BiVO4/TiO2 nanocomposite: electrochemical sensor for hydrogen peroxide, Materials Research Bulletin 125 (2020) 110771. [Google Scholar]
  • [61].Keihan AH, Karimi RR, Sajjadi S, Wide dynamic range and ultrasensitive detection of hydrogen peroxide based on beneficial role of gold nanoparticles on the electrochemical properties of prussian blue, J. Electroanal. Chem 862 (2020) 114001. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1

RESOURCES