Next Article in Journal
An Implantable Magneto-Responsive Poly(aspartamide) Based Electrospun Scaffold for Hyperthermia Treatment
Next Article in Special Issue
Defect Density-Dependent pH Response of Graphene Derivatives: Towards the Development of pH-Sensitive Graphene Oxide Devices
Previous Article in Journal
Efficient Charge Transfer Channels in Reduced Graphene Oxide/Mesoporous TiO2 Nanotube Heterojunction Assemblies toward Optimized Photocatalytic Hydrogen Evolution
Previous Article in Special Issue
Copper Cobalt Sulfide Structures Derived from MOF Precursors with Enhanced Electrochemical Glucose Sensing Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Electrochemical Sensing of Hydrogen Peroxide (H2O2) Released from Cancer Cells

1
Natural and Medical Sciences Research Center, University of Nizwa, P.O. Box 33, Birkat Al Mauz, Nizwa 616, Oman
2
Division of Pharmacy Practice and Policy, School of Pharmacy, University of Nottingham, Nottingham NG7 2RD, UK
3
Department of Pharmaceutical Chemistry, College of Pharmacy, King Khalid University, Abha 62529, Saudi Arabia
4
Institute of Medicinal and Pharmaceutical Chemistry, Technische Universitaet Braunschweig, 38106 Braunschweig, Germany
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(9), 1475; https://doi.org/10.3390/nano12091475
Submission received: 16 February 2022 / Revised: 22 March 2022 / Accepted: 23 March 2022 / Published: 26 April 2022

Abstract

:
Cancer is by far the most common cause of death worldwide. There are more than 200 types of cancer known hitherto depending upon the origin and type. Early diagnosis of cancer provides better disease prognosis and the best chance for a cure. This fact prompts world-leading scientists and clinicians to develop techniques for the early detection of cancer. Thus, less morbidity and lower mortality rates are envisioned. The latest advancements in the diagnosis of cancer utilizing nanotechnology have manifested encouraging results. Cancerous cells are well known for their substantial amounts of hydrogen peroxide (H2O2). The common methods for the detection of H2O2 include colorimetry, titration, chromatography, spectrophotometry, fluorimetry, and chemiluminescence. These methods commonly lack selectivity, sensitivity, and reproducibility and have prolonged analytical time. New biosensors are reported to circumvent these obstacles. The production of detectable amounts of H2O2 by cancerous cells has promoted the use of bio- and electrochemical sensors because of their high sensitivity, selectivity, robustness, and miniaturized point-of-care cancer diagnostics. Thus, this review will emphasize the principles, analytical parameters, advantages, and disadvantages of the latest electrochemical biosensors in the detection of H2O2. It will provide a summary of the latest technological advancements of biosensors based on potentiometric, impedimetric, amperometric, and voltammetric H2O2 detection. Moreover, it will critically describe the classification of biosensors based on the material, nature, conjugation, and carbon-nanocomposite electrodes for rapid and effective detection of H2O2, which can be useful in the early detection of cancerous cells.

Graphical Abstract

1. Introduction

Biosensors are simple devices that are small and are generally used in the field of medicine, pharmaceutical industries, environmental technology, and food industry. They are used for the measurement of many biological and chemical substances [1]. Owing to the advancement of science and technology, the research involved in biosensors has successfully made the biosensing devices small and efficient [2]. The use of the latest novel techniques and availability of a new biomaterial have made the biosensors efficient and have extended their use in multiple industries such as pharmaceutical, environmental, agriculture, and industrial laboratories [3]. Biosensors are of many types, and electrobiochemical biosensors have been commonly used for over 20 years in the field of diagnostics to detect biochemicals like glucose, lactate, cholesterol, urea, creatinine, DNA, antigens, antibodies, and cancer markers [4]. Electrobiochemical biosensors are also useful in the analysis of food materials and drinks and are extensively utilized in environmental and pharmaceutical laboratories [5,6]. Cancer is one of the most fatal diseases, and every year, more than 10 million new cases and 6 million deaths are reported worldwide [7,8]. Cancer is linked with high rates of morbidities and mortalities and more than 8.7 million deaths worldwide in 2015 [9]. The cancer incidence in high- and low-income countries is similar to the trend of increase in lower–middle-income countries because of the increase in risk factors associated with cancer [10]. In the United States of America (USA), cancer is the second leading cause of mortality, with heart disease being the first, and in a study published in 2017, it was estimated that more than 0.6 million people will die annually from cancer [8]. The survival rate of cancer patients drastically increases if the cancer is detected in earlier stages. The appearance of alarming systems in cancer patients is usually after cancer has spread in the body to multiple locations or has metastasized in different locations and organs, which characterizes an advanced stage of cancer. Most of the people are diagnosed with cancer at an advanced stage, which causes a high risk for mortality. To have a better disease prognosis, it is imperative that new research should focus on the early detection of cancerous cells in the body.
The use of biosensing devices, which have been designed to detect biochemicals, holds vast potential in the early diagnosis of cancers. Biosensors work by detecting a biological moiety or analyte and then converting it into an electrical signal, which can be detected and analyzed by the biosensor device. Many cancerous cells release specific chemicals called biomarkers, which can be detected using a biosensor device. The specific biomarker levels can also help in analyzing the effectiveness of anticancer therapy. The use of biosensor devices is a promising technique, which can help in early and accurate detection, imaging of cancerous cells, monitoring of angiogenesis, detection of proliferation, and tracking of metastatic changes and the efficacy of anticancer therapeutic regimens [11]. The latest research in biosensor devices using the latest techniques, such as nanotechnology-empowered diagnostics, can help in the identification of specific cancer biomarkers, which can help in the detection of cancer, disease progression, disease remission, and further proliferation. The biomarkers of cancer can be overexpressed proteins, surface antigens, active or inactive metabolites, miRNA, or the cancerous cells themselves. Many biosensors are excellent for use as an effective analytical device because of their capability to detect specific cancer biomarkers due to their highly sensitive, selective, robust, and miniaturized point-of-care cancer diagnostic capability [12]. Hydrogen peroxide (H2O2) is normally present inside the body and is vital in initiating and performing many important physiological processes. H2O2 is a by-product of respiratory chain and enzymes oxidases (glucose oxidase, cholesterol oxidase, glutamate oxidase, etc.) [13,14,15,16,17]. Hydrogen peroxide is a reactive oxygen species (ROS), which helps in regulating normal body functions such as cell growth, activation of the immune system, and programmed apoptotic changes [13,14,15,16,17]. The body system normally functions in homeostasis, and an increased level of H2O2 due to increased production can cause harm to the body. Increased levels of H2O2 can cause damage to normal cells [18], increase inflammatory responses [19], and cause cancer [20]. H2O2 regulates cancer cell characteristics, including invasion, proliferation, migration, apoptosis, and angiogenesis. Oxidative stress is associated with high levels of ROS, common in many types of cancerous cells. H2O2 has a specific role as a second messenger in pro-tumorigenic signaling pathways of cancerous cells [21,22]. GPX2 regulates cancer progression by regulating the hydrogen peroxide level in the cells, so when the level of H2O2 is downregulated to a normal level and the oxidative stress is relieved, it can help in dysregulating cancer cell homeostasis [23]. H2O2 has recently been a prime focus of research because of its high biological significance. When studied within living systems, it is noteworthy to check the concentration of H2O2 in mammals. The cellular compartment concentration has to be in the physiological range of 1 nM to 0.5 µM [24,25]. The latest research in the field of cancer diagnosis via biomarker-based techniques has been evaluated as successful because the process ensures high-precision, reliable, and sensitive data. The processes of biomarker-based cancer diagnostic are simple, which makes it a popular choice. Recent studies are focusing on profiling the cell functions with the efflux of endogenous H2O2 as potential biomarkers for diagnosing various cancers by measuring them using conventional biological assays [26]. However, before moving to the practical implications of using H2O2 as a potential biomarker for cancer diagnostics in living systems, detection of increased oxidative stress, prediction of neurodegenerative diseases, and detection of tumor growth inside living organisms, it is imperative to develop methods and techniques that can precisely detect and measure the level of H2O2 inside the cellular compartments [27].
The current problem with detecting H2O2 in cellular compartments is its low concentration in the body, as well as reactivity, which makes it difficult to separate its normal physiological concentration in a healthy organism from the concentration in a diseased or high-risk state. Therefore, scientists and researchers are focusing on developing sensors that can detect and quantify H2O2 in different systems and physiological conditions. Currently, many analytical techniques such as titrimetry, spectrophotometry, chemiluminescence, chromatography, fluorescence, and phosphorescence are measuring and determining the concentration of H2O2 in cellular compartments to develop consistent, precise, sensitive, fast, efficient, and low-cost methods. Currently, many methods investigated based on these analytical techniques have methodological disadvantages such as small sensitivity and selectivity, time-consuming and complex process, susceptibility to interference, and high-cost running instruments [28]. Alternatively, the advantages of “electrochemical sensors” are that they are highly sensitive and selective, reliable, quick, less costly, simple, and practical, and therefore, they are an optimum solution for exact and sensitive H2O2 detection [29]. The advancements of science and the latest research are looking for potential solutions to detect cancer at early stages and provide individualized therapeutic regimens. However, these techniques and methods still have many restrictions and limitations in the context of clinical examinations, histopathological analysis, imaging mammography, and chemotherapeutic adverse drug effects, as well as their high running costs [30,31]. In the detection of breast cancer, the patients are first exposed to a high amount of radiation in mammography because of its inadequate test sensitivity. Detection and disease progression are usually confirmed using a biopsy, an invasive procedure used to conduct histopathology of the disease [32]. These commonly used procedures are highly risky, uncomfortable, invasive, and costly. Therefore, the latest research is currently focused on developing noninvasive, inexpensive, highly selective screening, diagnostic, and therapeutic approaches to improve the disease prognosis. Several reviews regarding the importance of the carbon nanomaterial in the electrochemical sensing of H2O2 have been published. Out of those reviews, Yang et al. and Wang et al. conducted their reviews by comparing different nanomaterials for the creation of electrochemical biosensors and their applications for the detection of biomolecules [33,34]. Other materials have also been used in creating biosensors such as graphene electrochemical biosensors [35,36,37] and chemical sensors [38,39], which have been reviewed by Kuila and coworkers. Depending on the aim of the existing reviews, they provide an overview of biosensors that have been used to detect biological analytes. Similarly, Ping and coworkers [40] reviewed the strengths, advantages, and existing applications of 2D graphene-based aptasensors, whereas Chen and coworkers [13] focused their review on the carbon nanomaterial and transition metals in the electrocatalytic reduction of H2O2 in different samples. Similarly, Zhang et al. aimed to review the role of carbon materials in improving the sensitivity of H2O2 biosensors [41]. Regardless of these existing review articles, a comprehensive overview on the carbon-based nanomaterials and their composite with metal NPs, metal oxides, and biomolecules for the electrochemical detection of H2O2 secreted from cancerous cells is still missing in the literature.
This review highlights the recent development in the application of carbon nanomaterials and metal nanoparticles for H2O2 detection. A subdivision of the sensors has been made depending on the nanomaterial used: (i) metal nanoparticles, (ii) graphene modified with metal or metal oxide nanoparticles to form “graphene nanocomposites,” (iii) enzyme-loaded graphene-like 2D nanomaterials, and (iv) carbon nanotubes modified with metal or metal oxide nanoparticles. Finally, the current potential and challenges of using carbon nanomaterials for H2O2 detection are outlined. Relative electrochemical properties such as limit of detection (LOD), sensitivity, and stability are reported for each sensor, and a critical comparison between the results has been carried out by summarizing the strengths and weaknesses of the various sensors found in the review.

2. Classical Methods for H2O2 Detection

2.1. Electrochemical Systems

Electrobiochemical biosensors are built on the principle that they have a biorecognition component providing an electroactive constituent after reacting with an analyte that is transformed into an electrical signal that is measured using a physiochemical transducer as shown in Figure 1. The electrochemical transducers help in the detection and monitoring of the changes in the electrical current and potential. The commonly used easily detected biorecognition elements are enzymes. Antibodies, complete cells, and DNA can also be conventionally used for the construction of electrochemical biosensors as biorecognizable elements [1,42] by loading on a metal surface or carbon electrodes [43]. The most used electrochemical transducers are amperometric, potentiometric, conductometric, and impedimetric. The advantage of using electrochemical biosensors in analytical techniques is that they can be effectively used to reduce the size of the device. Additional advantages are that these sensors are low cost, highly sensitive, stable, and reproducible; show a linear response; can detect turbid samples; and are efficient due to using low sample volumes and chemical consumption [5,6,44,45].

2.2. Potentiometric

Biosensors act on the principle of measuring the electric potential generated by an electrode in the absence of a substantial amount of current via a reference and functional electrode that has been adjusted to sense selective analytes, with membranes that cover the electrode surface. When a target analyte interacts with the membrane covering the electrode surface, a subsequent change in electric potential is detected and measured by the electrode [47]. Potentiometric biosensors work by measuring the electric potential at zero current, which helps in differentiating the reference and functional electrodes. Potentiometric sensors measure the generated electric potential of ion-selective electrodes during biological reactions with target-specific ions. In potentiometric biosensors, the enzyme is attached on the surface of the electrode by glutaraldehyde crosslinking or an adsorption process. The probe of pH meter is covered by the membrane, where biological reaction either produces or absorbs hydrogen ions. The alteration in hydrogen ions causes a change in pH, which is a measure of the concentration of the analyte [48]. For instance, two example of potentiometric biosensors depending upon the type of electrode used are as follows: (a) Potentiometric biosensors use a Nafion membrane/Pt electrode for H2O2 determination, with an additional advantage of a perm-selective barrier [49]. Ascorbate and redox-active species reduce the overall electrode response, which further potentiates coupling between the redox potential on the Pt electrode and Donnan potential and increases sensitivity in detecting H2O2. The present potentiometric biosensor has a sensitivity of 125.1 ± 5.9 mV/decade, linear range of 10–1000 µM, and LOD of 10 µM [49]. (b) Zheng et al. used an MnO2/CPE to detect H2O2 [50]. The biosensor shows sensitivity of 19.4–121 mV/decade, with a wide linear range of 0.3–363 µM and LOD of 0.12 µM. The analytical parameters of the MnO2 doped/CPE biosensor were superior to those of the up-to-date potentiometric biosensors, i.e., Nafion membrane/Pt electrode [49], because of the enhanced electrode surface area, linear range, and LOD, except for the sensitivity. The coupling between the redox potential on the Pt electrode and Donnan potential made the Nafion membrane/Pt electrode superior to the MnO2 doped/CPE in terms of sensitivity [50].

2.3. Amperometric Biosensor

Zhao et al. developed an amperometric biosensor by immobilizing HRP on a silica sol–gel matrix on CPE for the determination of extracellular H2O2 excreted from breast cancer cells. The amperometric biosensor detected H2O2 via a sequence-specific peptide immobilized on the electrode surface and explicitly bound with horseradish peroxidase (HRP) in an auspicious orientation. The composed biosensors showed a linear range from 1.0 × 10−7 to 1.0 × 10−4 M, with a detection limit (LOD) of 3.0 × 10−8 M [51]. Zhao et al. showed a linear calibration of H2O2, i.e., 2 × 10−5 to 2.6 × 10−3 M, under optimum conditions [52]. In another strategy, a hydrogen peroxide biosensor was fabricated by coating a sol–gel–horseradish peroxidase LSPR layer onto a Nafion-methylene green modified electrode to develop a probe for H2O2 detection. The developed electrode exhibited sensitivity of 13.5 μA mM−1, with a detection limit of 1.0 × 10−7 M and response to 95% of the steady-state current in <20 s [53]. To detect H2O2, Tripathi et al. developed a novel biosensor by entrapping HRP in a new Ormosil composite doped with ferrocene monocarboxylic acid–bovine serum albumin conjugate and multiwalled carbon nanotubes (MWNTs). The developed biosensor showed a linear range of 0.02–4.0 mM, with a LOD of 5.0 μM (S/N = 3) [54]. In another study, a titania sol–gel matrix entrapping hemoglobin (Hb) was used as a peroxidase mimetic to sense H2O2 with a linear range from 5.0 × 10−7 to 5.4 × 10−5 M and a detection limit of 1.2 × 10−7 M [55]. Povedano et al. used His-Tag–Zinc finger commercial (His–Tag–ZFP) protein. The His–Tag–ZFP prefers to bind with RNA hybrids over ssRNAs, ssDNAs, and dsDNAs. These were further conjugated with streptavidin–HRP (Strep–HRP) in order to detect H2O2 with a LOD of 0.91 nM [56]. Reduced graphene oxide wrapped ZnMn2O4 microspheres (ZnMn2O4@rGO)-modified glassy carbon electrode (GCE) (ZnMn2O4@rGO/GCE) was used to make amperometric biosensors to detect H2O2. The resultant electrode showed a linear detection with a wide concentration range of 0.03–6000 μM and was used to detect H2O2 released from human breast carcinoma cells (MCF-7) as low as 0.012 μM [57]. Dong et al. reported a high-performance sensor using high-index facets of Au–Pd nanocubes loaded on large surface reduced grapheme oxide (rGO), and GCEs were modified by physical adsorption of both nanocomposites. The synthesized biosensor with three-dimensional nanocomposites possessed high sensitivity toward H2O2, with a minimum LOD of 4 nM, wide linear range from 0.005 μM to 3.5 mM, and swift response time [58]. Later, Jia and his coworker developed a nonenzymatic biosensor composed of poly(diallyldimethylammonium chloride) (PDDA)-capped rGO nanosheets loaded with a trimetallic AuPtAg nanoalloy. This biosensor detects H2O2 released from carcinoma cells with a LOD of 1.2 nM and a wide linear range from 0.05 μM to 5.5 mM [59]. In another study, hierarchical Mo2C@MoS2 consisting of interlayer-expanded MoS2 nanosheets wrapped on Mo2C nanorods was built as a highly sensitive, bifunctional electrochemical biosensor to detect H2O2 produced by cancerous cells, with sensitivity of 1080 μA mM−1 cm−2 and LOD of 0.2 μM [60]. Thiruppatthi et al. reported a simple stimulus responding aminophenol, pre-anodized screen-printed carbon electrode (SPCE*/AP) that could detect NADH and H2O2. The electrode was built by adsorbing aminophenol on the surface of the electrode, prepared from aminophenylboronic acid via boronic acid deprotection with H2O2. The resulting biosensor displayed linear ranges from 50 to 500 µM and from 200 µM to 2 mM, with detection limits (S/N = 3) of 4.2 and 28.9 µM for NADH and H2O2, respectively [61]. Maji et al. used cetyltrimethylammonium bromide-loaded gold nanorods (AuNRs) immobilized on a GC electrode to construct an amperometric biosensor (AuNRs/GC for the electrocatalytic detection of H2O2 under localized surface Plasmon resonance (LSPR) excitation (808 nm, 2 W cm−2). This biosensor showed an exaggerated improvement in its biosensing properties (~2–4-fold), with a wide linear range from 5.0 μM to 5.0 mM, LOD of 1.8 μM, and sensitivity of 1.6 μA mM−1 cm−2 [62]. In another study, self-supported MoS2 nanosheet arrays were built, and they showed highly potent electrocatalytic performance, with a LOD of 1.0 μM (S/N = 3) and high sensitivity of 5.3 mA mM−1 cm−2. This biosensor with self-supported MoS2 nanosheet arrays successfully detected trace amounts of H2O2 released from live A549 cancer cells [63].

2.4. Calorimetric Biosensors

A new area of nanotechnology and its integration with biosensors has introduced the concept of calorimetric biosensors for cancerous cell diagnosis and detection. Li et al. used a microfluidic paper-based analytical device (μ-PAD) for the synchronous sensitive and visual detection of H2O2 released from cancer cells. μ-PAD construction was done using a layer-by-layer modification of concanavalin A, graphene quantum dots (GQDs)-labeled flower-like Au@Pd alloy nanoparticles (NPs) probe, and vertical alignment of cancerous cells on the surface of ZnO [64]. In the study by Zhang et al. porous platinum NPs on graphene oxide (Pt-NPs/GO) were used in building a calorimetric biosensor. The resultant nanocomposite functioned as a peroxidase mimetic, which could catalyze peroxidase substrate reaction in the presence of H2O2. Building on this phenomenon, Pt-NPs/GO acts as a signal transducer in developing a colorimetric assay for cancerous cell detection [65]. Additionally, porous, alloy-structured PtPd nanorods (PtPd PNRs) were used as a peroxidase mimetic for H2O2 detection. The PtPd PNRs were found to have a detection limit of 8.6 nM and a linear range from 20 nM to 50 mM and were used as a signal transducer to develop an innovative detection method for studying the flux of H2O2 released from cells [66]. Folate and iron-substituted polyoxometalate [(FeOH2)2SiW10O36] provided a novel method for the detection of H2O2 with good sensitivity, with a linear range of 1.34 × 10−7 to 6.7 × 10−5 mol L−1, and low detection limit (1 × 10−7 mol L−1) and swift response toward H2O2. Ye et al. showed a new analysis method based on calorimetric analysis. The colorimetric biosensing strategy was based on iodide-responsive Cu–Au nanoparticles (Cu–Au NPs) combined with the iodide-catalyzed H2O2–3,3,5,5-tetramethylbenzidine (TMB) reaction system. The bimetallic Cu–Au NPs absorbed iodide, thus indirectly inducing the colorimetric signal variation of the H2O2–TMB system. The results demonstrated economically effective, simple, label-free visualization of H2O2 from cancerous cells with high selectivity and sensitivity. The resultant calorimetric biosensor operates with a linear range from 50 to 500 cells/mL and a LOD of 5 cells in 100 μL [67]. Calorimetric biosensors can be designed using single wall nano tube (SWNT), which is subsequently embedded within a collagen matrix. When there is an angiogenic stimulation of human umbilical vein endothelial cells (HUVECs), H2O2 molecules are released, which can be detected using this SWNT sensor. The constructed calorimetric biosensor shows calibration from 12.5 to 400 nM and can measure H2O2 at a nanomolar concentration in HUVEC from humans, with 1 s temporal and 300 nm spatial resolutions [68]. Other biosensors based on the principle of calorimetry include a biosensor with an electrode made of a 2D hybrid material (RGO–PMS@AuNPs). This biosensor displayed remarkable electrochemical performance and possessed high sensitivity and high selectivity in detecting H2O2 in 0.1 M phosphate-buffered saline as compared to enzymatic biosensors. The developed biosensor has an additional advantage over other sensors because it is nontoxic and can detect H2O2 without any intrusion by common interfering agents, with high sensitivity of 39.2 μA mM−1 cm−2, broad detection ranges from 0.5 μM to 50 mM, and a LOD of 60 nM. The sensor has high efficiency and can detect H2O2 in trace amounts, i.e., as low as nanomolar, secreted from living HeLa and HepG2 tumor cells [69].

2.5. Chemiluminescence Material for the Detection of H2O2

For the early diagnosis and detection of cancerous cells, it is important that molecules that indicate changes or biomarkers should be efficiently imaged and sensed. These parameters are important especially in studies that are evaluating the clinical mechanisms and designing effective chemotherapeutic agents [70,71]. The diagnostic and therapeutic methods for multiple detections are slow and need repetitive sampling, which results in low sensitivity and accuracy, because of heterogeneous sampling for separate detections [72]. The multiple fluorescence (FL) technique has promising results when used in in situ concurrent detection of multiple biomolecules. This technique has certain limitations such as weak compatibility with different biological systems, toxicity to living cells, and necessity for specialized synthesis and preparation [70]. Additionally, the FL signals generated using this technique faced interference, changes from background effects, and photobleaching while operating. Thus, it is highly desirable that in situ sequential detection of multiple biomolecules using within a complex biological sample is greatly desirable, the FL technique should be researched in cancer diagnostics without the current limitations [73]. The chemiluminescence (CL) technique is based on the principle that light is generated because of the energy released during a chemical reaction due to the de-excitation of the high energy moieties to the ground state or through energy transfer to luminophore molecules as shown in Figure 2 [74,75,76]. CL methods have gained popularity because these techniques are highly sensitive, are free of interference, phototoxicity, and photobleaching, and show no changes from background effects. The combination of CL methods with enzymes and analytes such as firefly luciferase (FFLuc) for 5′-triphosphate disodium salt (ATP) [77,78] and HRP [79,80] may result in a highly sensitive and competent method for H2O2 detection. In the recent advancement in the field of NPs, the multifunctional NPs with shell-like structures in their core have promising results in simultaneous diagnosis and treatment in living systems [81,82,83,84]. These multifunctional NPs were synthesized in the study by Ren et al. where dual functioning NPs were developed by combining HRPSiO2@FFLuc NPs with the enzyme-based core–shell structures, where the enzymes HRP and FFLuc were the main components of the core and shell of the NPs. They used the dual functioning NPs for the simultaneous in situ sequential detections and imaging of two biomolecules, namely, ATP and H2O2, in the same biological system. The surroundings of tumor cells or tissues are slightly acidic, and SiO2 is sensitive to an acidic environment, which causes the breakage of the SiO2 layer/component and exposes FFLuc and HRP (outside) and the SiO2 core (inside) to catalytic reactions. This results in the emission of two separate but simultaneous chemiluminescence signals for the sequential detection of ATP and H2O2, which avoids the signal interference between each other [73]. In another study by Lee et al., a novel contrasting agent was successfully synthesized, which was highly sensitive and specific and could image H2O2 in living systems [85]. The authors used peroxalate NPs to image H2O2 by inducing a chemiluminescent reaction using three components: H2O2, peroxalate esters, and fluorescent dyes. The peroxalate NPs were coated with peroxalate esters (hydrophobic polymer in its matrix). These NPs image H2O2 via a dual step process. Firstly, H2O2 diffusion occurs in the NPs, which then causes a reaction with the peroxalate ester groups and generates dioxetanedione, creating high energy inside the NPs [86,87], which subsequently then chemically excites the encapsulated fluorescent dyes, via a chemically initiated electron-exchange luminescence mechanism [88,89], leading to CL from the NPs and allows imaging of H2O2. Additionally, Lee et al. developed a method to synthesize peroxalate micelles, with a composition of amphiphilic peroxalate-based copolymers, rubrene (fluorescent dye), and a “stealth” polyethylene glycol (PEG) molecule to evade macrophage phagocytosis, which could successfully detect H2O2 through CL. These peroxalate-loaded micelles detected H2O2 within nanomolar concentrations (>50 nM) and were highly sensitive in detecting H2O2 in low physiological concentrations inside living systems [90].
Another study found that using peroxyoxalate chemiluminescent (POCL) NPs, H2O2 could be detected in trace amounts within living systems (in vivo) using optimized CL techniques in the near-infrared (NIR) wavelength. The detection of H2O2 using NIR is efficient in living systems because the penetration power of these NIR rays is higher because of the reduced photon-limiting interferences (scattering and absorption) happening within biological mediums [91,92,93]. CL using luminol was synthesized using o-benzyl alcohol-decorated block poly(carbonate)s copolymer, viz., PMPC–ONA, giving the resultant micelles a high H2O2 detection ability. In these micelles, luminol, fluorophore, and hemin were wrapped, forming an L/H/S@PMPC–ONA nanoprobe. These micelles work based on the principle that in the presence of H2O2 in the system, H2O2 diffuses within NPs, reacts with the hemin, and generates high energy reactive oxygen. The high energy reactive oxygen then chemically excites the luminol, activating the CL to expose nitrosobenzaldehyde recognition sites. This process destabilizes the micelles and releases the fluorescent indicator (fluorophore), which helps in imaging H2O2 [94]. Lee et al. additionally synthesized a nanoprobe using multiple molecule integration, i.e., dye/peroxalate NPs, which exhibited more enhanced and controlled CL, and hence displayed widespread applications in biomedical imaging of H2O2. This new enhanced nanoprobe was synthesized using nanoscopic coaggregation of a dye, which exhibited the aggregation-enhanced fluorescence phenomenon with a peroxalate, which had a high response to H2O2, which converted the energy generated from the chemical reaction to electronic excitation [95]. Additionally, Lee et al. successfully detected and imaged H2O2 via CL resonance energy transfer in the NIR wavelength using quantum dots functionalized with a luminol derivative [96]. Geng et al. devised a method to detect H2O2 via aggregation-induced emission fluorogen using 2,3-bis(4-(phenyl(4-(1,2,2-triphenylvinyl)-phenyl)amino)phenyl)-fumaronitrile (TPETPAFN), resulting in dye-encapsulated NPs [97]. A polyoxometalate (POM), vanadomolybdophosphoric heteropoly acid (H5PMo10V2O40, PMoV2), shows similar activity like peroxidases and functions by catalyzing the luminol/H2O2 reaction to generate CL. This phenomenon was shown in the study by Jia et al. where the study results showed an enzyme-free luminol/H2O2/PMoV2 CL system, which could be utilized for its high sensitivity in detecting H2O2. This enzyme-free luminol/H2O2/PMoV2 CL system exhibited good linear dependence with respect to H2O2 concentration within a wide range of up to 5 to 5000 nM (LOD) [98].

2.6. Titrimetry

The titrimetry technique can be used to analyze an unknown amount of H2O2 in a known sample concentration. The titrimetric technique uses iodometry, permanganate, and cerium (IV) in an acidic medium. In the study by Klassen et al., the concentration of H2O2 was assessed at 300 µM using the I3 method after the calibration with permanganate. ε max measurement was made at 351 nm as 25,800 M−1 cm−1 using the calibration plot of the I3 method titrated against potassium dichromate (KMnO4) [99]. In the study by Murty et al., the concentration of H2O2 was measured potentiometrically in an acidic medium using 8–11 M phosphoric acid [47]. Kieber and Helz synthesized a method for the detection of H2O2 by modifying the iodometric titration method using water matrices, where iodine was liberated as follows:
H2O2 + 2H+ + 2I → I2 + 2H2O
I2 + 2H2O + C6H5AsO = 2I + 2H+ + C6H5AsO(OH)2
The I2 produced was consumed by adding an excess of phenylarsine oxide. The end result was declared by titrating the remaining amount of phenylarsine oxide with iodine [100] when the intense blue color of the starch–iodine complex disappeared. The LOD was 0.02 µM. In another study, a two-step absorbance, microtiter plate method was developed by titrating an acidified H2O2 solution with standard cerium (IV) sulfate. In the second step, cerium (IV) sulfate was converted into cerium (III) sulfate, and potassium iodide was converted into iodine [101]. This process is commonly used and possesses additional advantages over the other methods because of its simplicity and low running costs, but its limitation is its inaccuracy at lower concentrations. Additionally, the other limitations of the method are that it consumes more time and requires skilled personnel to perform the calibration of the instrument.

2.7. Spectroscopy

One of the most common, convenient, and extensively used methods for determining and measuring H2O2 is spectroscopy. This method is based on the principle that colored compounds are formed with respect to absorbance measurements comparative method of methyl blue and toluidine. A method comparing the reaction of methyl blue and toluidine blue with iodine solution was introduced for determination of H2O2 based on the following reaction:
H2O2 + 2KI + 2HCl → I2(aq) + 2KCl + 2H2O.
In the comparison, methyl blue when reacted with iodine gave a single-peak visible spectrum with a higher extinction coefficient (=49,100 M−1 cm−1) [102]. In another study by Matsubara et al., a method using a mixture of titanium IV and 2,4((5-bromopyridyl)azo)5-(N-propyl-N-sulfopropyl amino) phenol disodium for determining H2O2 [103] was demonstrated. Molar absorptivity was found to be 5.7104 M−1 cm−1 at 539 nm. In the study by Clapp et al., the measurement of H2O2 was done using an aqueous solution with titanium (IV) sulfate. This method yielded a yellow peroxotitanium species at a wavelength of 407 nm [104]. An in vitro method for the detection of H2O2 was developed using the 1,10-phenanthroline method. The advantages of this method are its short processing time, increased sensitivity, and high reproducibility [105]. In another study, catalytic decomposition of H2O2 was demonstrated by monomeric molybdenum (VI) by mixing hydroquinone, ammonium molybdate, and anilinium sulfate with varying H2O2 concentrations and determining the absorbance at 550 nm [106]. Zhang and Wong demonstrated a method for the estimation of the concentration of H2O2 in marine water at acidic pH of 4 in the presence of HRP at 592 nm using leuco crystal violet oxidation. The LOD for H2O2 was found to be 20 nM with ±1% accuracy [107]. In the study by Huang et al., a fast, reproducible, and reliable method for the detection and measurement of H2O2 was demonstrated. This method used 4AAP-DEA-βCD-hemin, and the LOD was 8.4 × 10−5, with a molar absorption coefficient of 1.65 × 104 mol/L/cm [108]. Zhang et al. showed the determination of H2O2 in pulp bleaching effluents. The study shows that H2O2, in the presence of sulfuric acid solution, chemically reacted with vanadium pentoxide and formed a peroxovanadate complex that is reddish-brown [109].

2.8. Colorimetry

The method of determining H2O2 using iodide and starch was first developed in 1943 by Eisenberg. The H2O2 samples were treated with a titanium sulfate reagent, and the changes in color were quantified with the presence of H2O2. The chemical reaction of H2O2 with the titanium sulfate reagent is shown as follows:
Ti+4 + H2O2 + 2H2O = H2TiO4 + 4H+.
The formation of a yellow compound called pertitanic acid determined the H2O2 concentration within a range of 0.2–3.0 mg/100 mL [110]. Another study showed a more sensitive method using colorimetry. The study showed that the oxidation of iodide takes place in the presence of (NH4)2MoO4 (ammonium molybdate), which helps in determining the concentration of H2O2 even in micromolar quantities. The study determined the molar absorptivity of the starch–iodine complex (intense blue) at a value of 39.45 mmol−1 cm per liter at a wavelength of 570 nm [111]. The colorimetric method based on enzymes using plant extracts was developed by Fernando et al., where a sharp pink quinoneimine dye was formed. The pink dye formation took place when H2O2 reacted with phenol, 4-aminoantipyrine, and HRP in 0.4 M phosphate buffer with pH of 7.0 [112]. The assay results were considered optimum when the assay conditions were maintained at pH 7.0, temperature of 37 °C, 0.7 mM H2O2 concentration, and 1 U/mL enzyme concentration within 30 min. The optimum assay resulted in a limit of quantitation and LOD of 411 and 136 mM, respectively. Another simple method to detect the H2O2 released by cells within a tissue culture was based on the principle that phenol red oxidizes in the presence of H2O2. The study results showed a direct linear relationship between the concentration of H2O2 and absorbance, which had a range of 1 to 60 nmol/mL. The absorbance was measured at 520 nm [113]. Another fast and reliable method for determining H2O2 was developed using a colorimetry technique. In the method, 4-nitrophenylboronic acid was utilized for determining the concentration of H2O2 in an aqueous medium, where nitrophenylboronic acid reacted with H2O2 and produced 4-nitrophenol. The LOD was found to be ~1.0 μM [114]. Nitinaivinij et al. used the principle of colorimetry and demonstrated the determination of H2O2 in a very low concentration. The method utilized the technique of chromaticity analysis of silver nanoprisms (AgNPrs). The AgNPrs decomposed in the presence of H2O2, producing yellow color, and showed the H2O2 concentration at 1.57 mM with high accuracy and sensitivity [115]. The advantage of this method is that the determination of H2O2 can be carried out using a simple apparatus, but this method could give false-positive readings, and the results were not applicable to determine H2O2 within turbid samples.

2.9. Chromatography

Chromatographic techniques are commonly used for separation. High-performance liquid chromatography (HPLC) is an analytical technique used for the detection and separation of different moieties. In the study by Takahashi, separation of H2O2 was achieved using an electrochemical detector and a cation-exchange resin gel column of sulfonated styrene-divinylbenzene copolymer. This method was found to have a linearity of 0.9984. The LOD was measured at 0.2 pmol [116]. In another method by Wada et al., H2O2 separation was achieved using an octadecylsilyl column, and the LOD was measured at 1.1 µM [117]. In another study, H2O2 was determined using gas chromatography in the presence of oxidized butyric acid, and its absorbance was found at a wavelength of 517 nm [118]. Another method in a study used a ligand exchange-type column for the separation of H2O2. The column was packed using a sulfonated polystyrene/divinylbenzene cation-exchange [119]. Steinberg et al. used the principle of reverse-phase chromatographic techniques in HPLC to determine H2O2. Iodovanillic acid was formed and was detected using UV absorption at 280 nm with a LOD of ~0.1 μM [120]. The advantages of chromatographic techniques in H2O2 determination are that these methods are relatively simple, have low operational costs, and use a wide range of stationary phases and columns. The limitations of this technique are its costly overall equipment, its long operational time, interferences, and the necessity for a specialized operator to run the machine.

2.10. Fluorescence

Another common method to detect H2O2 that has wide applications is based on the principle of fluorescent signal detection. In fluorescence sensors, the excitation of electrons is achieved from an external photon source, in contrast to CL, where light is generated via a chemical reaction [121]. Many fluorescent probes have been constructed using different materials. The probes include naphthofluorescein disulfonate [122], homovanillic acid [123], peroxyfluor-1 [124], peroxyresorufin-1 [124], single-walled carbon nanotubes [125], peroxyxanthone-1 [124], and phosphine-based fluorescent reagents [126]. In one study, the fluorescent biosensors helped in the detection of intracellular H2O2 in mice peritoneal macrophages [122]. In the study by Miller et al., three fluorescent probes that were detectable via confocal and two-photon spectroscopic methods from the peroxysensor family were successfully developed. Each fluorescent probe emitted at a different wavelength from the other, which allowed these probes to be used in various applications with respect to specific emitting wavelengths [124]. Recently, intracellular H2O2 concentration can be measured using HyPer, a genetically encoded H2O2 biosensor (Figure 3) [127]. HyPer is a chimeric protein [128] composed from the permuted yellow fluorescent protein (cpYFP) and H2O2-sensitive domain of the bacterial transcription factor OxyR, which is responsible for sensing H2O2 [129]. In the study by Belousov et al., an H2O2 sensor named HyPer was developed and studied. The HyPer sensor was successful in detecting an increase in H2O2 levels in HeLa cells during Apo2L/TRAIL protein-induced apoptosis (programmed cell death). This sensor also detected increased levels of H2O2 in cells taken from rat adrenal medulla (PC-12) that had been previously exposed to nerve growth factor [128]. The HyPer family includes five probes: HyPer [128], HyPer2 [130], HyPer3 [131], HyPer7 [132], and HyPerRed [133]. GEFIs of this family consist of a circularly permuted fluorescent protein (cpYFP for the numbered HyPers or cpmApple for HyPerRed) integrated via short peptide linkers into the bacterial transcription factor OxyR lacking a DNA-binding domain. Upon oxidation by H2O2, OxyR forms an intramolecular disulfide bond [134] that elicits conformational rearrangements. These rearrangements are then transmitted into the chromophore center of a fluorescent moiety of a GEFI, causing fluorescence alterations that can be subsequently detected. HyPer and its improved derivates, HyPer2 and HyPer3, contain cpYFP. cpYFP has two excitation peaks at 420 and 500 nm and a single emission peak at 516 nm. When the OxyR domain is oxidized by H2O2, the intensity of fluorescence excited at approximately 420 nm (F420) decreases, whereas the intensity of fluorescence excited at approximately 500 nm (F500) increases proportionally. A sensor readout is generated as a F500/F420 ratio [128].
Another study by Xu et al. showed a specific probe called Mito-H2O2, which is used to detect mitochondrial-associated H2O2 levels in HeLa cells. The study further showed that Mito-H2O2 was an effective, sensitive, and quick mitochondrial-targeted sensor [135]. Xiao et al. also developed another fluorescent probe called ER-H2O2 specifically for targeting the endoplasmic reticulum, which was equally effective, sensitive, and quick in the detection of H2O2. Xiao et al. induced apoptosis in both the organelles using L-buthionine sulfoximine, and both of these probes were tested for H2O2 specificity and selectivity [136]. Shen et al. developed a microfluidic method, which had droplets in combination with gold nanoclusters. This method was demonstrated to have high sensitivity for the detection of H2O2 secreted by a single cell. When a single cell was isolated using a microdroplet (with a volume of 4.2 nL), it can secrete H2O2, which causes florescent changes in HRP-gold nanoclusters with high specificity and high sensitivity of 200–400 attomole. The high throughput performance (~103 single-cell encapsulated microdroplets per minute) of the resultant microfluidic device makes it a powerful tool to investigate cell-to-cell heterogeneity in releasing H2O2 at the large scale, promising revelation of new knowledge to understand the biological role of H2O2 in tumor cells [137]. Moreover, Wang et al. fabricated a Ce6@Lum-AuNPs nanoprobe using green syntheses methods. They successfully loaded luminol-gold NPs with the fluorescent receptor Chlorin e6 (Ce6). The resultant fluorescent Ce6@Lum-AuNPs proved successful towards fluorescent bioimaging of cancerous cells [138].

3. Recent Advances

3.1. Current Approaches in the Construction of Biosensors

Over the past 200 years, the use of enzymes was common because of specific substrate sensitivity. However, enzymes are highly unstable and sensitive and are prone to denaturation caused by environmental changes such as pH and temperature. Therefore, recent studies have focused on using an artificial pseudo-catalyst instead of enzymes to overcome the drawbacks [139,140,141]. Denaturation of enzymes is common in enzyme groups such as peroxidases, catalases, monoamine oxidase, choline oxidase, uricase, and ascorbate oxidase. Peroxidases, also known as heme proteins, constitute the prosthetic group, i.e., ferriprotoporphyrin, and are usually found to have a molecular weight of 30 to 150 kDa [142,143]. Peroxidases are oxidoreductases and are produced by many animals, plants, and microorganisms. Peroxidases reduce H2O2 and help in the oxidation of aromatic amines, phenols, and organic and inorganic substrates [144] and are extensively utilized in biochemistry, enzyme immunoassays, wastewater treatment plants containing phenol compounds, synthesis of aromatic compounds, and removing H2O2 from food materials [143]. The application of peroxidase enzymes is extensive, and they are commonly used in analytical techniques for the detection of glucose [145], cholesterol [146], uric acid [147], H2O2 [148], alcohols [149], and phenols [144]. Peroxidase enzymes are also used in the pharmaceutical industry for the construction of biosensors for the detection of different drugs in the body. As previously mentioned, enzymes are prone to degradation; hence, the latest research involves the replacement of enzymes with pseudo-catalysts, i.e., inorganic/organic. These materials are chosen because of their low cost, stability, and convenience [150,151,152].

3.2. Electrochemical Sensing of H2O2 via Metal Nanoparticles

Nanotechnology has advanced, and there are many types of NPs available nowadays. NPs can be classified based on the nanomaterial used to synthesize them. Commonly available nanomaterials include metal NPs [153,154], carbon nanomaterials [155,156], and metallic oxide nanostructures [157]. Nowadays, NPs are used in manufacturing H2O2 electrochemical sensors, exhibiting distinctive electrical and catalytic properties toward the reduction or oxidization of H2O2 and having a broad range of stability based on the nanomaterial used in them. However, until now, most of the studies reporting cost-effective H2O2 detecting electrochemical sensors have a detection limit of the sub-micromolar level [38,158]. To detect H2O2 in cellular matrices, the electrobiochemical sensors should be sensitive enough to sense H2O2 concentration in nanomoles. Currently, those biosensors that are highly sensitive and have optimum H2O2 detection limits have been developed using HRP and metal nanoparticles [159,160], which decreases the long-term operational stability and increases the operational costs. In the study by Wang et al., real-time electrochemical detection of H2O2 via small MoS2 NPs in Raw 264.7 cancerous cells was performed. The resulting device had a detection limit lower than 2.5 nM and a wide linear range of up to five orders of magnitude [161]. In vivo monitoring of H2O2 secreted from living cells is essential in understanding cellular signaling pathways. The release of H2O2 from living cells is very low because the selective detection of H2O2 at a low level is challenging. To overcome this difficulty of detecting endogenous H2O2 from live cells, Dou et al. synthesized three hybrid metal nanoflower sensors for the detection and monitoring of H2O2 concealed from living MCF-7 cancerous cells. The three-hybrid metal Au–Pd–Pt nanoflower-decorated MoS2 nanosheet-modified sensors were developed using simple wet chemistry. The three-hybrid metal nanoflower sensors (Au–Pd–Pt/MoS2) show a synergetic increase in the electrocatalytic reduction of H2O2 with an ultrasmall detection limit as low as the sub-nanomolar level. Immobilization of aminin glycoproteins on the nanocomposite surface will result in an increase of its biocompatibility, which, in turn, enhances composite adherence to cells. This property of nanocomposites can be effectively used in future applications directed toward monitoring the secretion of H2O2 from living cells and cellular apparatus and may be utilized in developing highly efficient and sensitive cancer diagnostics sensors [162]. Sun et al. synthesized a dumbbell-shaped PtxPd100–x–Fe3O4 NP composite, which could effectively determine the secretion of H2O2 from Raw 264.7 cells with a detection limit of 5 nM [160]. Chang et al. developed a sensitive fluorescent assay to determine H2O2 with a wide linear range of 1 to 100 μM and detection limit of 0.8 μM. A fluorescent biosensor based on the inner filter effect (IFE) was manufactured using poly (vinyl pyrrolidone)-protected gold nanoparticles (PVP–AuNPs) and fluorescent BSA-protected gold nanoclusters (BSA–AuNCs). The BSA–AuNCs acted as an IFE fluorophore pair. The high extinction coefficient of PVP–AuNPs served as a dominant absorber and influenced the emission of the fluorophore in the BSA–AuNCs assay. The surface Plasmon resonance (SPR) of PVP–AuNPs was significantly enhanced with an increase in H2O2 concentration. The increased H2O2 then caused the significant induction of the fluorescent quenching effect of BSA–AuNCs [163]. Cui et al. showed a fast, simple, and reagent-free method for H2O2 detection. The study used luminol-reduced Au NPs for the determination of H2O2. The resulting biosensor had the electrochemiluminescence application in effectively determining the concentration of H2O2 within limits of 3 × 10−7–1.0 × 10−3 mol L−1 with a low detection limit of 1.0 × 10−7 mol L−1 (S/N = 3) [164].
Liu et al. synthesized porphyrin functionalized ceria (Por-Ceria) uniform nanoparticles as a calorimetric probe for H2O2 detection [165]. A nickel phosphide nanosheet array on a titanium mesh (Ni2P NA/TM) possesses superior analytical performance with a rapid retort time of <5 s. Manufactured biosensors showed high selectivity and stability, with a wide linear range of 0.001–20 mM, ultrasmall LOD of 0.2 μM (S/N = 3), and high sensitivity of 690.7 μA mM−1 cm−2 [166]. Small (10–30 nm) platinum nanoparticles (Pt-NPs) were prepared via protein-directed one-pot reduction. The resultant BSA/Pt-NPs composite shows colorimetric determination of H2O2 with a linear range from 50 μM to 3.0 mM, LOD of 7.9 μM, and visually detected lowest concentration of 200 μM [167].
Ultrathin silver nanosheets that can detect H2O2 with a LOD of 0.17 µM, linear range of 5–6000 μM, and fast response time <2 s were synthesized by Ma et al. The synthesized biosensors showed real-time determination of H2O2 released from living HeLa and SH-SY5Y cells, with high sensitivity of 320.3 µA mM−1 cm−2 [168].
The synergistic combination of p-type semiconductive channels of layered double hydroxides (LDHs) exhibited multifunctional properties, a distinctive morphology, and abundant surface active sites. The Fe3O4@CuAl NSs modified electrode exhibited excellent electrocatalytic activity toward H2O2 reduction. The projected biosensor revealed prominent electrochemical sensing of H2O2 with an extensive linear range of eight orders of magnitude and a low detection limit of 1 nM (S/N = 3) [169]. Copper(I) phosphide nanowires on 3D porous copper foam (Cu3P NWs/CF) were fabricated via electrochemical anodized Cu(OH)2 NWs to manufacture noble metal-free electrocatalysts. The Cu3P NWs/CF-based sensor exhibited first-rate electrocatalytic reduction of H2O2 with a detection limit as low as that achieved by noble metal-free electrocatalysts, i.e., 2 nM. The developed sensor assured sensitive and consistent determination of H2O2 excretion from living tumorigenic cells [170]. Xiong et al. developed a nickel phosphide nanosheet array on a titanium mesh (Ni2P NA/TM) using an economical and effective metal toward electrocatalytic H2O2 reduction. Ni2P NA/TM, being a nonenzymatic H2O2 sensor, presented superior analytical performance, with a swift response time <5 s and wide linear range of 0.001–20 mM. The resultant electrode exhibited high sensitivity of 690.7 μA mM−1 cm−2 and ultrasmall detection limit of 0.2 μM (S/N = 3) [166]. A Prussian blue nanocube-decorated molybdenum disulfide (MoS2-PBNCs) nanocomposite was designed for the electrochemical sensing of H2O2. Interestingly, a sensor for label-free sensing of carcinoembryonic antigen (CEA) can be constructed by using MoS2-PBNCs nanocomposites. The electrochemical response of the MoS2-based immunosensor was linear, with a CEA concentration range from 0.005 to 10 ng mL−1 and minimum recognition limit of 0.54 pg mL−1 [171].

3.3. H2O2 Detection Using Enzymatic Biosensors

Various analytical techniques, i.e., chemiluminescence [95], fluorescence [172], and electrochemistry [80,160], have been employed for the analysis of H2O2 at the cellular level. Among them, electrochemical sensors are an area of high interest and provide fast, economically effective, and real-time determination via a simple mechanism with ultrahigh sensitivity and selectivity. Electrochemical detection is considered a powerful tool for the determination of other electroactive metabolites such as glucose [173], dopamine [174], and O2 [175] secreted from live cells. The high selectivity and sensitivity of enzymes made them valuable for the electrochemical biosensing of H2O2. Horseradish peroxidase (HRP) enzymes draw considerable attention for the construction of electrochemical biosensors because of their efficient catalysis of H2O2 [176,177]. Wang et al. reported a highly sensitive sequence-selective DNA sensor composed of an HRP-labeled probe. The proposed biosensor successfully detected the K-ras gene, which is associated with colorectal cancer. Thiol (–SH) modified capture probe adsorbed chemically on the gold electrode via self-assembly and exhibiting a detection limit of 5.85 × 10−12 mol L−1, hybridization of nucleic acid (target DNA:K-ras gene), and a HRP labeled oligonucleotide detection probe can be achieved using the sandwich way method. Wang et al. developed an extremely sensitive sequence-selective DNA sensor on an HRP-labeled probe to detect the specific K-ras gene which is associated with colorectal cancer. At first, the capture probe modified with –SH was chemically adsorbed on the gold electrode by self assembly. Then, a complementary nucleic acid (target DNA:K-ras gene) was hybridized with an HRP labeled oligonucleotide detection probe in a sandwich way with a detection limit of 5.85 × 10−12 mol L−1 [178]. Bruno et al. developed horseradish peroxidase conjugated gold nano biosensors for detection of H2O2 released by prostate cancerous cells. The proposed biosensor can detect hydrogen peroxide (H2O2) in a wide linear range from 2 to 100 μM with a low detection limit of 0.01 μM [179]. A Cyt c loaded nanostructured TiO2 film was successfully prepared by Luo, which exhibits natural enzymatic activity toward H2O2, redox formal potential (E0′) of 108.0 ± 1.9 mV versus Ag|AgCl, and an heterogeneous electron transfer rate constant (ks) of 13.8 ± 2.1 s−1 [157]. To stabilize the enzyme model, Zhou et al. used an enzyme cytochrome c (Cyt c), to facilitate the transfer of electrons between the redox enzyme and electrode. Cyt c was immobilized stably into the molecular hydrogel to maintain its innate bioactivity toward H2O2. The use of Cyt c is a consistent methodology to regulate H2O2 at an optimized potential with high selectivity over other ROS, oxygen, metal ions, and ascorbic acid. The in vivo sensing of H2O2 from living cells, a small molecular hydrogel provides long-lasting stability and good reproducibility [180].

3.4. Carbon-Based Material for H2O2 Sensing

The success of graphene has boosted great research in the synthesis and characterization of graphene-like 2D materials, single and few-atom-thick layers of van der Waals materials, which show fascinating and useful properties. The single atom layer of C is the most transparent, strongest, and thinnest material and exhibits electrical conductance much better than Cu, with the ability to endure a current density that is six orders of magnitude [36,181,182]. The structure of some of the carbon-based material is shown in Figure 4. Recently, graphene has attracted great interest in the development of biosensors, i.e., optical and electrochemical, with improved performance owing to its integration with different nanomaterials (metals and metal oxides) and quantum dots [183,184,185]. Researchers have developed great interest in the emerging class of carbon-based 2D materials (graphene) because of their distinctive properties with applications in sensing and biosensing, electronics, catalysis, composites, and coatings. The excellent optical and electrical properties of carbon-based 2D materials made their use emergent in sensing and biosensing and showed real-time application in the field of biochemistry and nanomedicines [186,187]. Graphene-like 2D layered nanomaterials boron nitride (BN), transition metal dichalcogenides, graphite–carbon nitride (gC3N4), graphenes, and transition metal oxides have been investigated broadly [188,189]. Boron nitride nanosheets contain alternate nitrogen and boron atoms in a honeycomb lattice structure with extensive band gap, and BN is an insulator [190]. Instead of various transduction techniques, electrochemical methods are well known for analytical biomarker detection via graphene 2D-based sensors [191].

3.4.1. Graphene-Based Metal-Free Electrocatalysts

The application of carbon materials in analytical and industrial electrochemistry is well known owing to their low cost and electrocatalytic potential in a number of redox reactions [192]. Recently, groups of researchers showed that surface functionalization of graphene materials results in diverse behavior, which made them benevolent in sensing in contrast to intrinsic graphene. Zhou et al. showed the chemical reduction of graphene oxide into chemically reduced graphene oxide (CR-GO) via hydrazine, and the resultant GCE constructed from the obtained CR-GO showed excellent sensing capability for H2O2 detection. The synthesized electrochemical biosensor exhibited a lower detection limit of 0.05 μM and wide linear range from 0.05 to 1500 μM, which precedes the use of functionalized carbon materials in electrochemical sensing [155].
In another work, Takahashi et al. reported rGO modified GCE via electrodeposition. The electrochemical studies showed an enhancement in the sensing performance of the rGO modified electrode that was considerably better than the original electrode for hydrogen peroxide detection. Some studies showed a high electron density on the defective sites (edges) of modified graphene oxide, which made it a potential candidate for the electrocatalytic reduction of H2O2 [193]. The synthesis of novel quality graphene is important in exploiting graphene application for electrochemical sensing. Chemical and physical (thermal method) reduction of GO (hydrophilic GO to hydrophobic graphene) is the most effective method to manufacture graphene on a large scale. During chemical and physical reduction, exfoliated graphene becomes disorderly aggregated, which results in the decrease in their disperse behavior in water and limits their practical applications [194]. Later on, some researchers fixed this problem using various dispersants, i.e., sodium dodecyl sulfate, cetyltrimethyl ammonium bromide (CTAB), and DNA. These dispersants enhanced the disperse behavior and stability of graphene in an aqueous environment. Lv et al. simply introduced DNA molecules on the graphene surface using the self-assembly method and formed graphene–DNA hybrids (GN/DNA). DNA–graphene was found to show a physical interaction, i.e., π–π stacking via aromatic rings of graphene and N-containing functional moiety in DNA, which results in a strong interaction between graphene and DNA. Stacking DNA on the graphene surface not only enhanced graphene dispersion in aqueous media but also imparted an electron-rich character to graphene by forming a GN/DNA composite. Comparative studies showed that the GN/DNA modified electrode exhibited higher sensitivity, wide detection range, and swift response time in contrast to the GN-modified electrode for the electrochemical sensing of H2O2 [195]. Woo et al. fabricated a multiwalled carbon nanotube–graphene composite (MWCNT–graphene) via a direct in situ chemical reduction of graphene oxide and pre-treated MWCNT mixture. The prepared component showed a uniform network of ultrathin graphene sheets stuck between nanotube bundles. Structural analysis showed that the morphology of graphene present between nanotube bundles was comparatively higher than pure graphene, which showed wrinkled and aggregated morphology. The electrochemical sensor constructed from the resultant MWCNT–graphene exhibited a wide detection range from 20 μM to 2.1 mM and low detection limit of 9.4 μM. Synergic increase in the electrochemical performance of the MWCNT–graphene composite is attributed to high electrical conductivity of MWCNTs [196]. Recently, metal-free electrocatalysts, heteroatom-doped graphene, play a crucial role in H2O2 detection. The electronic properties of graphene can be altered drastically by doping graphene with N, S, and B, which play a crucial role in operating the electronic properties. Wang and coworkers used the nitrogen plasma treatment strategy to produce N-doped graphene from reduced graphene oxide as a starting material. Spectral studies of N-graphene showed that the nitrogen atom was substituted into graphene sheets with three different nitrogens, including graphitic N, pyridinic N, and pyrrolic N. The concentration of nitrogen in graphene sheets was optimized by monitoring the plasma exposure time, and the resultant N-doped graphene showed improved electrocatalytic performance as compared to pristine graphene in electrochemical sensing [197].
Wu et al. reported the synthesis of N-doped graphene using hydrazine as a nitrogen source, with a 4.5% N/C atomic ratio, and reducing agent. Structural studies of N-doped graphene were made via XPS measurements. Structural analyses showed 28% pyridinic N, 49% pyrrolic N, 19% graphitic N, and 4% oxidized N [198]. Increased sensitivity, a wide linear range, and a low detection limit were achieved using N-doped graphene as compared to pristine graphene. In addition to N, Yeh et al. successfully synthesized boron-doped graphene nanosheets (BGNs) using B2O3 and graphene nanosheets through an atmospheric-pressure carbothermal reaction. Boron doping on the graphene surface created defects in nearby sites and uneven charge separation, which, in turn, facilitated the charge transfer to neighbor atoms. The resultant BGN-doped graphene showed a wide linearity range from 1.0 to 20.0 mM, detection limit of 3.8 μM, and much higher sensitivity (266.7 μA mM−1 cm−2) compared with undoped GNs [199]. Recently, the electrochemical performance of the detection of H2O2 was further improved using co-doped graphene with two elements. Yang et al. synthesized N and B co-doped graphene (NB-G) using a microwave-activated chemical–thermal treatment strategy. In this strategy, they first developed N-graphene using GO and cyanamide as a precursor, followed by microwave treatment. The boron atom was doped on N-modified graphene via the pyrolysis of the N-G and B2O3 mixture at 900 °C for 0.5 h in an Ar atmosphere to obtain BN-G [200]. Electrochemical studies of NB-G were made using ferric/ferrous coupling of K3[Fe(CN)6]/K4[Fe(CN)6]. The prepared electrode exhibited outstanding electrocatalytic reduction of H2O2 and a rapid response time, with a linear range from 0.5 μM to 5 mM and detection limit as low as 0.05 μM. The excellent electrochemical performance of the NB-G electrode is attributed to the novel structural network, with high charge transfer and large surface area, and the synergistic effect between the two heteroatoms of B and N [200]. Table 1 shows electrochemical performance of non-enzymatic metal free H2O2 sensors based on graphene.

3.4.2. Carbon Composite with Enzymes for H2O2 Detection

Noble metals, nonnoble metal oxide, and sulfide-modified graphene composites are used to immobilize HRP for the construction of enzymatic H2O2 biosensors [212,213]. Song at al. [214] reported MoS2–graphene (MoS2-Gr)-based biocompatible biosensors for the ultrasensitive detection of H2O2. MoS2-Gr nanosheets were prepared using GO and NaMoO4 as precursors using the solvothermal method, and a change in solution color from reddish-brown (GO) to black confirmed the dispersion of dark flower-like MoS2 nanoparticles on the Gr surface. Structural analyses were made using XRD results, which confirmed the formation of MoS2-Gr composites. Electrostatic interaction arose between negatively charged MoS2-Gr nanosheets and positively charged HRP and resulted in the formation of the HRP-MoS2-Gr composite. The appearance of the peak in the UV–Vis spectra at 402 nm confirmed the immobilization of HRP on MoS2-Gr, whereas no peak was noticed in case of MoS2-Gr nanosheets. The HRP-MoS2-Gr fabricated biosensor showed excellent stability and enhanced electrocatalytic performance for H2O2 detection. The resultant biosensors exhibited a low detection limit of 0.049 μM and broad linear range from 0.2 μM to 1.103 mM.
Later, Yu et al. immobilized horseradish peroxidase (HRP) on Au-decorated graphene oxide. The fabricated biosensors showed a fast response with remarkable performance, such as low detection limit (7.5 × 10−9 M) and real-time measurement of cellular H2O2 in living cells [27]. Liu et al. used horseradish peroxidase (HRP) immobilized on 3D porous graphene (PGN) to develop a real-time biosensor for the detection of H2O2 from living cells. Nanoporous graphene plays a significant role in the excess absorption of HRP, accelerates the diffusion rate, and shows excellent electrochemical performance toward H2O2 with a LOD of 0.0267 nM and wide linear range of seven orders of magnitude [215]. Enzymatic biosensors suffer from two major problems, namely, enzymatic loss and inactivation, which greatly affect biosensor performance. Fan and his coworker overcame this problem by encapsulating horseradish peroxidase on biomimetic graphene capsules (GRCAPS) using CaCO3 as a porous sacrificial template to mimic the existence form of bioenzymes in organisms as shown in Figure 5. As a result, the synthesized biosensor showed a low detection limit of 3.3 mmol L−1 and wide linear range of 0.01–12 mmol L−1 [216]. Wu et al. used another strategy to construct horseradish peroxidase–attapulgite nanohybrids on glassy carbon to fabricated biosensors. The prepared biosensor showed a rapid response, high sensitivity, and a low detection limit with a wide linear range for the detection of hydrogen peroxide released from RAW 264.7 macrophage cells [217]. Table 2 shows electrochemical performance of enzymatic H2O2 biosensor loaded on graphene.

3.4.3. Graphene Composite with Metal Nanoparticles for H2O2 Detection

Dai et al. prepared heterogeneous Co3O4 dodecahedrons that contain carbon, and encapsulated Au nanoparticles (Au@C-Co3O4) were proposed via the pyrolysis of Au nanoparticle-encapsulated zeolitic imidazolate framework-67 (Au@ZIF-67). A remarkable increase in electrocatalytic performance with ultrahigh sensitivity of 7553 μA mM−1 cm−2 and with a detection limit of 19 nM was observed using the electrode fabricated from the porous Au@C-Co3O4 even with the 0.85% Au content in the composite. The synthesized biosensors were applicable for monitoring H2O2 concentration, which will be helpful in identifying cancerous cells [237]. A metal organic framework consisting of porphyrinand iron metal decorated on well-ordered mesoporous carbon (OMC) for hydrogen peroxide (H2O2) secreted from viable cells. Porphyrinic iron metal-organic framework (pFeMOF)-decorated ordered mesoporous carbon (OMC) was developed to detect hydrogen peroxide (H2O2) released from viable cells. Increased stability and electrical conduction were noticed with the introduction of OMC. Electrocatalytic reduction of H2O2 was observed at two different linear ranges, i.e., from 70.5 to 1830.5 μM and from 0.5 to 70.5 μM, with high sensitivity of 67.54 μA mM−1 at a low concentration and 22.29 μA mM−1 at a high concentration and with a detection limit (LOD) as low as 0.45 μM [238]. A nonenzymatic H2O2 electrochemical sensor was developed by immobilizing 2D ultrathin MnO2 nanosheets onto glassy carbon electrodes (GCE) with a Nafion film. The amperometric study showed an excellent increase in electrocatalytic reduction of H2O2 with an extreme low detection limit (5 nM), wide linear range (25 nM−2 μM and 10–454 μM), and high sensitivity of 3261 mA M−1 cm−2 via the immobilization of the MnO2 nanosheets. The constructed biosensors were efficaciously employed for real-time monitoring of H2O2 released from SP2/0 cells in trace amounts [239].
The functionalized hollow-structured nanospheres (HNSs) centered on Pd nanoparticles (NPs) adorned double shell-structured N-doped graphene quantum dots (N-GQDs)@N-doped carbon (NC) HNSs, with ultrafine Pd NPs and “nanozyme” N-GQDs as dual signal-amplifying nanoprobes, act as an exceedingly effective electrochemical sensor for the detection of H2O2 released from cancer cells. The hybrid HNS material-based synthesized electrochemical biosensors demonstrate excellent performance, which involves an ultrasmall detection limit as low as nanomolar and a rapid response time. The extra sensitivity, selectivity, and reproducibility of the synthesized biosensors make them valuable for real-time tracking of H2O2 released from different living cancer cells in a normal state and treated with chemotherapy and radiotherapy [26]. Heteroatom-doped graphene (N and B) exhibits multidimensional electron transport pathways, which make their use valuable in electrocatalytic sensing of H2O2 with excellent stability and response time. Table 3 and Table 4 show electrochemical performance of graphene-supported non-noble metal and noble metal nanoparticles.
The resultant N and B co-doped graphene (NB-G)-based electrochemical sensor showed a linear response from 0.5 μM to 5 mM with a LOD of 0.05 μM (S/N = 3). This increase in sensitivity with an ultralow detection limit to microlevel attributed to the NB-G constructive structure and special effects arose from the co-doping of N and boron in graphene [200]. CuFe2O4 nanoparticle-doped reduced graphene oxide based on a CPE was used as a voltammetric sensor for hydrogen peroxide (H2O2) sensing. The synthesized sensors showed a rapid amperometric response in less than 2 s and wide linear range of 2 to 200 μM with a low detection limit of 0.52 μM under optimum conditions (pH 5) [264]. Xi et al. synthesized N and S dual-doped graphene (NSG) co-doped carbocatalyst via one-pot syntheses. The NSG-modified electrode showed superior catalytic activity toward sensing, including a linear range up to 1.7 mM. ZnMn2O4-wrapped reduced graphene oxide microspheres (ZnMn2O4@rGO) act as an excellent electrocatalyst for H2O2 reduction. The ZnMn2O4@rGO-pasted glassy carbon electrode (ZnMn2O4@rGO/GCE) displayed a linear detection range of 0.03–6000 μM with a detection limit of 0.012 μM. The resultant biosensor showed promising results in physiology and diagnostics and was applicable in the determination of H2O2 secreted from human breast cancer cells (MCF-7) [57]. An AuNPs-NH2/Cu-MOF composite was prepared via ammoniation of Au NPs, anchored with a Cu-based metal organic framework (Cu-MOF). The synthesized AuNPs-NH2/Cu-MOF composite was further modified with a GCE to prepare an AuNPs-NH2/Cu-MOF/GCE electrode. The synthesized AuNPs-NH2/Cu-MOF/GCE composites possessed high sensitivity and selectivity, and they can be used as an electrochemical enzyme-free sensor for the quantitative detection of H2O2. Instead of quantitative H2O2 detection, the synthesized electrochemical sensor showed a wide linear response toward H2O2 concentrations ranging 5–850 μM with a LOD down to 1.2 μM [265]. Wang et al. improved the sensitivity of the electrode using hemin-capped biomineralized gold nanoparticles (Hem@AuNPs)-doped reduced graphene oxide (rGO), followed by coating with chitosan (CS). The resultant electrode from the prepared nanohybrids showed excellent electrocatalytic reduction of H2O2 with superior sensitivity, stability, and response time of few seconds. The most important feature of the synthesized electrode from the resultant nanohybrid is its lower detection limit of 9.3 nM and linear range of five orders of magnitude. Such characteristics enable this biosensor to detect H2O2 releasing from living Hela cells accurately and make this biosensor valuable for ultrasmall detection of H2O2 from living HeLa cells precisely [266]. Sun and his coworkers designed a novel nonenzymatic hydrogen peroxide sensor using intermetallic PtPb nanoplates (PtPb/G) supported on graphene with enhanced electrochemical performance for H2O2 detection in neutral solution and H2O2 released from the cells. The nanocomposite exhibited excellent electrocatalytic activity for the electrochemical reduction of H2O2 in half-cell test and with wide linear detection range of 2 nM to 2.5 mM and ultralow detection limit of 2 nM. An experiment further showed that the sensitivity of intermetallic PtPb nanoplates is 12.7 times higher than that of a commercial Pt/C electrode for the detection of H2O2 released from Raw 264.7 cells [267]. A graphene/Nafion/azure/I/Au nanoparticle composites modified glass carbon electrode (graphene/Nafion/AzI/AuNPs/GCE) was used for the construction of a nonenzymatic H2O2 sensor. The performance of the synthesized biosensors was recorded under optimum conditions, i.e., pH of 4.0 and potential of −0.2 V, upon the addition of H2O2. A stable current was obtained in less than 3 s, with a detection limit of 10 μM (S/N = 3) and a linear range of 30 μM to 5 mM [268]. Ju et al. reported a green and simple strategy for the in situ growth of surfactant-free Au nanoparticles (Au NPs) on nitrogen-doped graphene quantum dots (Au NPs–N-GQDs). The reported strategy showed the in situ formation of the Au NPs–N-GQDs hybrid by simple mixing of N-GQDs and HAuCl4·4H2O without any reductant and surfactant. The prepared nanocomposite (Au NPs–N-GQDs) exhibited a low detection limit of 0.12 μM and sensitivity of 186.22 μA/mM cm2 for the electrochemical detection of hydrogen peroxide (H2O2) [79]. Another research group developed a microelectrode with high sensitivity, a wide linear range, and good selectivity for the detection of H2O2 released from female cancer cells. The synthesized hierarchical nanohybrid microelectrode was composed of 3D porous graphene enfolded activated carbon fiber (ACF). This technique, i.e., green ionic liquid (IL), plays a crucial role in the simultaneous superficial and effective electrodeposition and electrochemical reduction of GO nanosheets on ACF to form a 3D porous ionic liquid functionalized electrochemically reduced GO (ERGO)-wrapped ACF (IL–ERGO/ACF) [269].
Table 4. Graphene supported noble metal nanoparticles for electrochemical detection of H2O2.
Table 4. Graphene supported noble metal nanoparticles for electrochemical detection of H2O2.
Graphene Based MaterialSensitivity
μA mM−1 cm−2
Linear Range
(μM)
Detection Limit
(μM)
Ref.
Au-PEI/GO 460.0 0.5–1680 0.2[270]
AgNPs-MWCNT-rGO 0.833 100–100,0000.9[271]
RGO-Au-PTBO 63.395.0–1077.1 0.2[272]
Ag-MnOOH-GO 59.14 0.5–17,8000.2 [273]
Au NPs@POM-G 58.87 5.0–18,000 1.54[274]
AgNPs-TWEEN-GO 0.7459 20–23,100 8.7[41]
GO-ATP-Pd 504.85 0.1–10,0000.016[275]
GN/Au-NPs-0.5–500 0.22[276]
GN-Pt0.01 2–710 0.5[277]
Ag NWs-graphene12.3710.0–34,300 1.0[278]
GR-AuNRs389.2 30–500010[279]
Au@C-Co3O47553 -0.019[237]
Au NPs-N-GQDs186.22 0.25–13,327 0.12[79]
AuNPs-NH2/Cu-MOF/GCE 1.71 5–850 1.2[265]
GO/Au@Pt@Au -0.05–17,5000.02 [280]
NG-hAuPd 5095.5 0.1–200.02[281]
PDA-RGO/Ag NP 0.0111 0.5–80002.07[282]
AgNPs/GN -100–100,0000.5[283]
Ag/SG -100–136,500 0.14[284]
Pt/PG341.14 1–14770.5[285]
PDDA-RGO/MnO2/AuNPs 1132.8 5.0–5000.6[286]
AgNP/rGO-100–80,0007.1[287]
AgNPs-GO -10–20,0000.5[288]
RGO-AuNP 5.3 250–22,500 6.2[289]
GNPs/SGS 27.7 20–15,000 0.2[290]
AgNPs-CNT-rGO-10–10,0001.0[291]
PpyNFs/AgNPs-rGO 0.7367 100–50001.099[292]
polystyrene@RGO-Pt 0.0675 0.5–80000.1[41]
Graphene/Nafion/Azl/AuNPs -30–5000 10[268]
Pt/GN 0.0204 2.5–66500.8[41]
RGO-AuNPs (B) 9.5 25–41,5005.0[273]
PtAu/G-CNTs 313.4 2.0–8561 0.6[293]
PtAuNPs-CTAB-GR 0.1654 0.005–4.8 0.0017[294]
PtAu/RGO 4.105 0.015–8.730.008[295]
Pt/graphene-CNT paper 1.410–25.0 0.01[296]
pFeMOF/OMC67.54 70.5–1830.50.45 [238]
Pd-PEI/GO -0.5–4590.2[297]
Pd-NPs/GN 0.019 0.001–20000.0002[298]
PdNPGNs 2.75 0.1–1000 0.05[299]
RGO-PMS@AuNPs 39.2 0.5–50,000 0.06[69]
2Au1Ag-PDA/CFME12966 0–55 0.12[300]
TiO2NTs/r-GO/AgNPs 1152 15,500–50,000 2.2[301]
PtPb/G-2–2.5 0.02[267]
3DGA-AuNPs/cytc/GCE351.57 - [302]
PdPt NCs@SGN/GCE-1–300 0.3 [303]
AuNFs/Fe3O4@ZIF-8-MoS2-5–120 0.9 [304]

3.4.4. Graphene-Loaded Biomolecules for Selective Detection of H2O2

Recently, graphene-based heme protein electrodes have gained wide attention for H2O2 detection. These graphene-based materials offer an appropriate microenvironment to maintain the redox bioactivity of proteins and make the transfer of electrons feasible between redox proteins (active centers) and the principal electrode [232]. A mixture of a strong acid and an oxidizing agent is used for the synthesis of graphene oxide from graphite [289]. GO serves as a precursor of graphene and as a sensing element. Several proteins, including cytochrome c, horseradish peroxidase (HRP), and myoglobin (Mb), were incubated. Zuo et al. [227] fabricated a heme proteins-modified GO electrode from GO suspension. Immobilization of protein on a GO sheet is associated with strong hydrophobic and electrostatic interactions between proteins and GO. The innate characteristics of the proteins remain unaltered in the presence of GO, which offers an appropriate microenvironment for the immobilization of protein with an intact structure. Studies revealed that the protein-based GO modified electrodes have an advantage over the featureless voltammograms because of the emittance of redox peaks from proteins on these electrodes, which stipulate an efficient electrical wiring of the redox centers of proteins to the surface of the electrode in the presence of GO. Importantly, the proteins retained their intrinsic peroxidase activity upon forming mixtures with GO and the catalytic properties provide a high sensing performance for H2O2 detection with low detection limit and wide detection range. Furthermore, Mani and coworkers improved the performance of Mb-based H2O2 biosensors using an RGO-MWCNT-Pt/Mb electrode [233]. The RGO-MWCNT-Pt composite was prepared using the wet chemical method, which provides good affinity and a large surface area for the accumulation of excess Mb. The Pt nanoparticles in the RGO-MWCNT-Pt/Mb composite showed excellent electrocatalytic activity and efficiently prohibited the accumulation and restacking of graphene sheets and CNTs. The resultant electrode (RGO-MWCNT-Pt/Mb) showed an excellent wide linear range from 10 pM to 0.19 nM with a detection limit of 6 pM and much higher sensitivity (1.99 μA pM−1 cm−2) compared to other biosensors.
Additionally, HRP-fabricated H2O2 electrochemical biosensors were prepared using nano-graphene for the direct electron transfer from HRP to the substrates (electrode) [305]. These HRP-anchored graphene-based materials determine H2O2 with higher selectivity and sensitivity [224,306]. Zhang et al. reported immobilization of HRP and lysozyme enzymes on graphene oxide sheets in phosphate buffer solution by incubating GO with enzymes at 4 °C. The immobilized enzyme molecules were studied in situ using AFM, which clearly disclosed HRP molecules (bright spots) on the surface of GO. Strong hydrogen bonding and electrostatic interaction play a key role in loading enzymes (HRP and lysozyme) on graphene oxide, which was much higher than that on previously reported studies and was found to be the optimum solid substrate for the immobilization of the enzyme [307]. Moreover, Fan and coworkers applied the same method to generate graphene-poly (sodium 4-styrenesulfonate)/biomimetic graphene capsules (GS-PSS/GRCAPS) nanocomposites for direct electrochemical sensing of H2O2. Initially, porous CaCO3 was used as a support for HRP encapsulation in GRCAPS. Afterward, a GS-PSS/GRCAPS composite was synthesized via layer-by-layer electrostatic self-assembly, in which negatively charged GS-PSS electrostatically interact with positively charged PEI@GRCAPS [261]. GRCAPS was revealed to mimic the existing enzymes in living cells and provide a satisfactory microenvironment for HRP to realize direct electron transfer at the modified electrode. The resultant electrochemical biosensor exhibited long-term stability, low detection limit, extensive linear range, and an excellent anti-interference ability. A nonenzymatic and highly electrocatalytic H2O2 biosensor was proposed using a novel electrode composed of hemin-capped biomineralized gold nanoparticles (Hem@AuNPs), rGO, and chitosan (CS). The excellent rGO conductivity and outstanding electrocatalytic performance of Hem@AuNPs make them suitable for developing ultrasensitive biosensors for real-time determination of H2O2. Taking advantages of the peroxidase-like activities of nanohybrids, the resultant electrode demonstrated a highly selective and outstanding electrochemical performance toward H2O2 with fast response, improved sensitivity, and stability. More significantly, the lower determination limit of 9.3 nM and wider linear ranges of five orders of magnitude enable this biosensor to accurately detect H2O2 released from living HeLa cells [266]. Jiao et al. reported nonenzymatic biosensors for dynamic, most significant ROS. Intracellular nonenzymatic monitoring of H2O2 was achieved via loading of AuPtAg nanoalloy on rGO capped with poly (diallyldimethylammonium chloride). The constructed biosensor showed rapid and precise measurement of H2O2 released from cancerous cells. The precise and accurate detection of H2O2 is due to the remarkable rGO and PDDA conductivity with outstanding synergistic electrocatalytic performance of ternary alloys. The remarkable electrochemical performance of the resultant biosensor, with a low detection limit (1.2 nM) and wide linear range (from 0.05 μM to 5.5 mM), is due to the peroxidase-like activity of the AuPtAg nanoalloy [59]. In another study, a cytochrome c (Cyt c)-immobilized Au nanoparticle-loaded 3D graphene aerogel (3DGA) was synthesized for the detection of H2O2. Morphological and surface study of the 3DGA-AuNPs revealed efficacious formation of 3D-networked assembly, which helps in enhancing conductivity and effective enzyme immobilization. The large surface of 3DGA and biocompatibility of AuNPs help in enabling direct electron transfer between the electrode and Cyt c. The as-prepared 3DGA-AuNPs/Cyt c/GCE exhibited a pair of well-defined redox peaks of a FeIII/II redox couple of Cyt c and revealed excellent electrocatalytic potential toward H2O2 with high sensitivity of 351.57 μA mM−1 cm−2 [302].

3.5. Carbon Nanotubes (CNTs)

CNTs, an allotropic form of carbon, are composed of a graphene sheet packed in a tube constituting a cylinder (single-walled CNTs (SWCNTs)) or concentric and closed tubules (multiwalled CNTs (MWCNTs)) [308,309,310]. The combination of CNTs in biosensors offers numerous advantages, including increased surface area, smooth charge transfer, stacking of various biomolecules, and improved conductivity of the resulting platform as shown in Table 5 [310,311,312].

3.5.1. H2O2 Electrochemical Sensors Based on the Association of CNTs and Hemoproteins

Direct electrochemical assignment of proteins in biosensors is an area of high interest. However, direct electron transfer between proteins and electrodes is faced with major problems, i.e., the distance between the redox center and electrode and protein denaturation. Different methods, including polymer adsorption, covalent binding, and layer-by-layer film assembly, are well known for the deposition of protein molecules on the electrode surface. The excellent electrocatalytic properties of CNTs make them valuable in loading biomolecules and for use as biosensors. CNTs function as nanowires and boost the electron transfer from the protein’s redox center to the electrode. The heme-containing proteins (Hb, Mb, Cyt c, and HRP) are the most common analytes for protein detection [313]. Heme proteins are the center of several biological redox reactions. Therefore, several studies reported the efficacy of these proteins as a biosensor for H2O2, nitrite, or hydrogen sulfide detection. Yang et al. have developed a method to directly bind hemoglobin to a vertically aligned CNT surface. They modified the nanotubes so they could use diazonium chemistry to directly bind hemoglobin. In amperometric detection of H2O2, an Hb-ACNT electrode exhibits a wide concentration range (40 μM to 3 mM), LOD of 5.4 M, high sensitivity, and long-term stability [314]. This aligned NT forest shows accumulation of Hb on a large area rather than immobilization in unsystematic tangled webs of CNT. Furthermore, Esplandiu et al. immobilized Mb to detect H2O2, studied direct electron transfer kinetic, and showed that vertically aligned NT forests possess better kinetics compared to the epoxy incorporated SWCNT/Mb sensor [315]. In addition, their LOD was 50 nM for H2O2, superior to other random and aligned NT methods. The release of H2O2 from living HepG2 cancer cells was studied by Zhang and coworkers, who constructed an enzyme-based biosensor with a LOD of 0.23 μM using SWCNTs as a robust scaffold for Hb immobilization. The constructed biosensor was also used for the quantification of H2O2 released from HepG2 cells via in situ biosynthesis of ZnO quantum dots, which was further confirmed by fluorescence staining [316]. Wang et al. applied a simple dispersion method to coat a GCEs with SWCNT and heme proteins in the presence of cetylrimethylammonium bromide (CTAB) [317]. CTAB-suspended NTs facilitate the immobilization of Mb, Cyt c, and HRP on the electrode surface. Redox chemistry of heme was studied in the presence of SWCNTs. The developed electrode was precisely used for nitrite and H2O2 detection, which gave rise to a new-fangled peak in cyclic voltammograms with decreased in heme reduction peak. The results indicate that the electrode exhibited a response time of only 4 s, LOD of 3.6 M, and less sensitivity for H2O2 detection. Several researchers have also demonstrated the efficacy of the immobilization of heme proteins on polymers. For instance, Hb was immobilized on polyelectrolyte surfactant polymers [318], where Hb retained a secondary structure, thus reducing the effect of protein denaturation in polymers. Moreover, the addition of SWCNT to the nanocomposite enhanced the reaction kinetics, and H2O2 was sensed with a LOD of 0.8 M. Likewise, MWCNT and Mb were immobilized on the collagen polymers [319], where H2O2 was measured with a linear range from 0.6 to 39 M. Nagaraju et al. used self-assembled monolayers of 4-aminothiophenol on gold electrodes with immobilized Cyt c for H2O2 detection [320]. Three orders of magnitude of faster electron transfer kinetics were observed with SWCNT in these monolayers as compared to the non-SWCNT monolayer. The results confirmed that NTs increased the direct electron transfer. However, large step changes (3.8 mM) in H2O2 were used, and no LOD was calculated. Thus, the sensitivity was not very significant. A layer-by-layer approach was also implemented to immobilize proteins rather than simultaneous deposition of all components, e.g., chitosan-stabilized NTs were placed on GCE, followed by the accumulation of gold NPs on chitosan, and subsequently, Hb was bound to the gold surface [321]. This method is beneficial in retaining the bioactivity of Hb and increases the amount of enzyme activity. The method showed a LOD of 0.2 M for H2O2 detection. The heme-based biosensors showed rapid and fast detection of changes in H2O2. In general, HRP is the most widely used in biosensing as compared to other heme-containing proteins, which shows the best results compared to others. The lowest limits of detection for H2O2 were achieved using an aligned NT geometry, which supports the accumulation of the heme protein and which, in turn, leads to fast electron transfer from proteins to electrodes. Future studies are required to address the reproducibility of electrode fabrication, practical geometries, and uses for real samples. SWCNTs, HRP, and 1-butyl-3-methylimidazolium tetrafluoroborate (BMIM·BF4) were combined to construct a cellular H2O2 sensor. At a working potential of −0.35V, HRP-BMIM·BF4/SWCNTs/CFUME showed a dynamic range of ~10.2 μM, with a low detection limit of 0.13 μM (S/N = 3) and high sensitivity of 4.25 A/M cm2. Due to its small dimension and low working potential, HRP-BMIM·BF4/SWCNTs/CFUME allowed direct amperometric real-time monitoring of H2O2 in HeLa cells treated with camptothecin (an anticancer drug) without complex data processing and extra surface coatings to prevent interference. Thus, its testing evidently demonstrated a significantly high level of H2O2 in HeLa cells under camptothecin stress [322]. A schematic presentation of enzyme loaded CNTs for the detection of H2O2 in living cells is shown in Figure 6.
Table 5. CNTs-supported metal nanoparticles biosensors for electrochemical sensing of H2O2 secreted from cancerous cells.
Table 5. CNTs-supported metal nanoparticles biosensors for electrochemical sensing of H2O2 secreted from cancerous cells.
CNTs H2O2 BiosensorsSensitivity
μA mM−1 cm−2
Linear Range
(μM)
Detection Limit
(μM)
Ref.
ZnO/COOH-MWNTs-1–21 --[323]
GCE/MWCNTs-CDs0.039 3.5–300 0.25 [324]
((APy)6[H2W12O40])/(SWCNT-COOH)--0.4[325]
GCE/CNTs-PAMAM DENs-PtNCs987.53–4000.8[326]
GCE/C60-MWCNTs
CS-IL/MB/CuNP
0.0243 2–40.055[327]
3D PB NPs/G-CNTs0.11343 1–31610.095[328]
CF@N-CNTAs–AuNPs1421–43000.05[329]
CDs/MWCNTs/GCE-----[324]
OECT/PET/CE-CNTs/PtNPs-0.5–1000.2[330]
ZNBs/fMWCNTs-0.049–22 0.035[331]
ZnONPs/MWCNTs-1000–200,000-[332]
N-CNTs30 -0.5[333]
(GC) (BG-CNPs/GC)-- [334]
GCE/rGONRs/MnO20.01420.25–2245 0.071[335]

3.5.2. H2O2 Electrochemical Sensors Based on the Association of Metallic Nanoparticles and CNTs

In the last few years, the connotation of metal nanoparticles with CNTs has been considered a valuable alternate for the development of highly sensitive and selective sensors for H2O2 detection shown in Figure 7 [41,337,338]. Zhang et al. developed a remarkable stretchy nanohybrid microelectrode using carbon fibers [329]. The constructed microelectrode reproduced a remarkable analytical signal at 0.300 V, with an ultrasmall LOD as low as nanomolar. Rapid and ultrasmall sensing of H2O2 excreted from MCF-7 and MDA-MB-231 cells was achieved because of the synergistic catalytic activity of the N-CNTA-decorated AuNPs. Table 6 showed non-enzymatic CNTs biosensors for H2O2 sensing.
Bai et al. developed an electrode with high electrocatalytic activity using carbon dots (CDs) and oxidized MWCNTs modified GCEs [324]. The excellent biosensing capability of the MWCNT/CD/GCEs composite is associated with the large surface area and electron acceptor characteristics of MWCNTs and with the excellent donor capacity of the CDs. The analytical performance of the sensor is highly dependent on the MWCNT:CD ratio, so 10:1 was selected for electrode manufacturing. The developed biosensors successfully quantified the H2O2 secreted from HeLa cells with a linear range of 3.5 × 10−6 and 3 × 10−4 M and a LOD of 0.25 mM.
Liu and Ding used a Pt-encapsulated poly(amidoamine) dendrimer with amine terminations (G6-NH2 PAMAM dendrimer) covalently attached to a carboxylated CNT composite [326]. Elaborated architectures showed a rapid, reproducible, and steady response at 0.150 V, with a linear range between 3 and 400 mM. The biosensor successfully detected H2O2 in MCF-7 cells with an LOD of 0.8 mM. Liu at al. modified GCEs by dendrimer-encapsulated Pt nanoclusters and carbon nanotubes (Pt-DENs/CNTs) to detect extracellular H2O2 excreted from live cells. Those Pt-DENs/CNTs nanocomposites were characterized by UV-Vis spectra, SEM, energy-dispersive X-ray spectroscopy, and TEM. The nonenzymatic sensor displayed exceptional catalytic activity in H2O2 reduction. The effective nonenzymatic sensing capability of H2O2 reduction revealed that the Pt-DENs/CNTs sensor has potential application in screening H2O2 in cellular processes [326].
Real-time monitoring of H2O2 secreted from living cells is important to understand the occurrence of diseases and searching of new therapeutic strategies. Zhao et al. successfully synthesized three-dimensional carbon nanotubes spaced graphene aerogel decorated with Prussian blue nanoparticles (3D PB NPs/G-CNTs) by one-step mild temperature treatment, in which the PB NPs acquire intrinsic peroxidase-like activity. The 3D porous structure of G-CNTs with large surface and high electrical conductivity can efficiently enhance catalytic performance and help in real-time detection of H2O2 released from living cells. The composite exhibited good catalytic performance toward H2O2 reduction with sensitivity of 134.3 μA mM−1 cm−2, LOD of 95 nM, and wide linear range of 1–3161 μM [328].
Zhang et al. integrated Fe3O4 and Cu nanoparticles (NPs) into the NCNTs to produce N-doped carbon@Fe3O4-Cu nanotubes (NCNTs@Fe3O4@Cu) through a one-pot high-temperature decomposition. Then, Au NPs were assembled on the magnetic NCNTs to obtain an NCNTs@Fe3O4@Au composite by galvanic replacement with Cu NPs. The resultant composite provided a suitable platform for the immobilization of the enzyme to fabricate biosensors for H2O2 monitoring. After the cytochrome c (Cyt c) was accumulated by the NCNTs@Fe3O4@Au composite, the Cyt c/NCNTs@Fe3O4@Au gathered to the surface of the electrode with an external magnet [336]. Tabrizi et al. developed a flow injection amperometric sandwich-type aptasensor for the detection of human leukemic lymphoblasts (CCRF-CEM). An amperometric biosensor was synthesized by decorating nanogold on poly(3,4-ethylenedioxythiophene) (PEDOT-Aunano). PEDOT-Aunano acts as a nano-platform for immobilizing a thiolated sgc8c aptamer and MWCNTs loaded PdNPs/3,4,9,10-perylene tetracarboxylic acid (MWCNTs-Pdnano/PTCA) to assemble a catalytic labeled aptamer. In this strategy, the CCRF-CEM cancer cells were sandwiched between the immobilized sgc8c aptamer on PEDOT-Aunano (electrode) and sgc8c aptamer MWCNTs-Pdnano/PTCA/aptamer (catalytic site). The resultant sandwich-type aptasensor determined the CCRF-CEM cancer cell concentration using 0.1 mM H2O2 (electroactive component). The MWCNTs-Pdnano nanocomposites enhanced the electrocatalytic reduction of H2O2, which further boost sensor sensitivity toward CCRF-CEM cancer cells. The proposed sandwich-type aptasensor displayed outstanding analytical performance for real-time determination of CCRF-CEM cancer cells with high selectivity, ranging from 1.0 × 101 to 5.0 × 105 cells mL−1 with an LOD of 8 cells mL−1 [339].

3.5.3. In Vivo Sensing of H2O2 Release from Carcinoma Cells

Considering the significance of cellular H2O2 in cell pharmacology and pathophysiology, accurate and reliable in vivo detection of cellular H2O2 is sorely needed. Sensing H2O2 at cellular level is constrained by several factors, including small cell size, low concentration of cellular H2O2, and interferences in the culture medium [215,340]. Such in situ monitoring of the cellular release of H2O2 provides a new in vitro drug screening platform for personalized medicine and cancer therapy. Enzyme-based electrochemical sensing is efficient for continuous in situ monitoring of H2O2 because of its high sensitivity, rapid response, and selectivity [27,267]. In vivo monitoring of H2O2 secreted from living cells is essential in understanding cellular signaling pathways. The release of H2O2 from living cells is very low because the selective detection of H2O2 at a low level is challenging.
Chen et al. used a flow-through mode sensing strategy based on cell-in-lumen configuration for ultra-small detection of H2O2 secreted from the H1299 carcinoma cell. The current strategy involved the growth of cells on the inner surface of a porous hollow fiber (PHF), while a sensing layer comprised of multi-walled carbon nanotubes, gold nanoparticles (AuNPs), and enzymes accumulated on the outer surface of the PHF. The porous structure of the resultant electrode proved beneficial in the exchange of H2O2 from cell to sensing layer in a short time span. The resultant electrode exhibits ultra-small sensitivity to detect H2O2 at the nanomolar level having a detection limit of 6 nM with a wide linear range of 0.01–5 [341]. Ye et al. fabricated a PdPt NCs@SGN/GCE non enzymatic electrochemical biosensor comprised of Pd-Pt nanocages and SnO2/graphene nanosheets. The resultant electrode displayed excellent catalytic activity toward H2O2 in situ secreted from human cervical cancer cells (Hela cells) with high selectivity and sensitivity, a low detection limit of 0.3 mM, and a large linear range from 1 mM to 300 mM [303]. Fe3O4 quantum dot was decorated on three-dimensional graphene nanocomposites (Fe3O4/3DG NCs) for real time in-situ monitoring of H2O2 released from living cancer cells. The fabricated electrochemical sensor mimics peroxidase-like activity with high sensitivity of 274.15 mA M−1 cm−2, a low detection limit (78 nM), fast response (2.8 s), and outstanding reproducibility [342].
A nonenzymatic electrochemical sensor was constructed by immobilizing 2D ultrathin MnO2 nanosheets onto glassy carbon electrodes (GCE) with a Nafion film for real-time monitoring of H2O2 released from SP2/0 cells in trace amounts. The amperometric study showed an excellent increase in electrocatalytic reduction of H2O2 with an extreme low detection limit (5 nM), wide linear range (25 nM−2 μM and 10–454 μM), and high sensitivity of 3261 mA M−1 cm−2 via the immobilization of the MnO2 nanosheets [239]. Xi et al. synthesized N and S dual-doped graphene (NSG) co-doped carbocatalyst via one-pot syntheses. The NSG-modified electrode showed superior catalytic activity toward sensing, including a linear range up to 1.7 mM. The prepared electrode showed high sensitivity of 0.266 mA cm−2 mM−1 with a detection limit as low as 1 μM (S/N = 3), with good discernment, reproducibility, stability, and biocompatibility with real-time determination of H2O2 secreted from live cancerous cells [343]. Later on, Zhao et al. used a well-controlled strategy for the syntheses of the graphene fiber microelectrode via MnO2 nanowire (MnO2-NWs) assembly (MnO2-NWs@Au-NPs). The prepared microelectrode showed proficient catalytic performance toward the redox reaction of H2O2. The nanohybrid microelectrode showed in vivo real-time detection of H2O2 released from human breast cancer cells [344]. Recently, Chen and his co-worker established high-index facets of an Au-Pd nanocubes loaded rGO composite. The resultant electrode comprised of three-dimensional nanocomposites showed a detection limit of 4 nM, a wide linear range from 0.005 μM to 3.5 mM, and real time monitoring of endogenous H2O2 in human serum samples released from a living breast cancer cell [345].

3.6. MXenes Materials

So far, various nanomaterials have been reported and utilized for the development of incrementally efficient biosensors. Among the most recently reported nanomaterials available for biosensors, MXenes have attracted much attention for their huge potential in biosensor development because of their unique characteristics [346]. MXenes are two-dimensional inorganic compounds with a thickness of a few atomic layers and they are composed of transition metal carbides, nitrides, or carbonitrides such as titanium carbide (Ti3C2) and titanium carbonitride (Ti2CN), which confers them with exceptional characteristics, including high conductivity and superior fluorescent, optical, and plasmonic properties [347,348]. Moreover, the biocompatible property of MXenes enables their biomedical application [349,350]. Since they were first reported in 2011, MXenes have been used to develop various types of advanced biosensors, including electrochemical, fluorescent/optical, and surface-enhanced Raman spectroscopy (SERS) biosensors, by augmenting MXene characteristics to make them suitable for specific types of biosensors or by combining them with other nanomaterials [351,352]. Recent studies on the development of highly effective MXene biosensors show that this novel nanomaterial is the most ideal candidate for biosensor development at present. So far, no considerable development was seen in MXenes-based biosensors for detection of H2O2 released from a cancer cell. However, we foresee MXenes as having outstanding potential for detection of H2O2 at an ultra-low level with durable stability and long working hours.

4. Conclusions and Future Perspectives

Carbon nanomaterials have gained prodigious attention over the last two decades because of their higher applicability in electrochemical sensors. This review shed light on the application of carbon nanomaterials and their composite with metal, metal oxides, and biomolecules for the fabrication of electrochemical sensors for real-time monitoring of hydrogen peroxide. Initially, we discussed the recent advancement in the development of heme protein biosensors with carbon nanomaterials as immobilization matrix and their application in the detection of H2O2. Subsequently, the synthesis and application of graphene-supported nano-catalysts (metal-free, noble metals, and nonnoble metals) was discussed in detail for the construction of nonenzymatic H2O2 electrochemical sensors. Despite the extensive advancement in the design and application of carbon nanomaterials for the electrocatalytic determination of H2O2, it is crucially important to develop new techniques and methods for the synthesis of carbon-based electrocatalysts with a novel structure and extraordinary activity. Some of the most highly ultra-sensitive biosensors for detecting H2O2 at an ultra-low level are displayed in Table 6. Furthermore, the comprehensive understanding and exploration of the structure–property relationship of carbon nanomaterials and their extensive use in H2O2 sensors require more efforts and research. Particularly, its excellent electrical conductivity, electron mobility, small band gap, and ultrahigh surface area make it widely applicable in biosensors. These advantages of graphene would bestow good conductivity to the capsule film, further facilitate fast electron transfer between enzyme and basal electrode, and enhance the sensitivity and detection limit of biosensors. We prophesy excellent biosensing potential of new MXenes materials and carbon-based material for detection of H2O2 released from cancer cells at an ultra-low level with remarkable stability and selectivity.

Author Contributions

Conceptualization, A.K., A.A.-H. and S.E.D.; methodology, T.A. and A.I.; formal analysis, S.A.H. and A.K.; investigation, A.K. and S.E.D.; resources, A.K. and A.A.-H.; data curation, T.A., A.I. and S.A.H.; writing—original draft preparation, T.A., A.I. and A.K.; writing—review and editing, S.A.H., J.U., A.K. and A.A.-H.; supervision, A.K., A.A.-H. and S.E.D.; project administration, A.K., A.A.-H. and S.E.D.; funding acquisition, S.E.D. All authors have read and agreed to the published version of the manuscript.

Funding

The project was supported by grant from the Oman Research Council (TRC) through the funded project (BFP/RGP/HSS/19/198).

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are available from the corresponding author upon request.

Acknowledgments

The authors would like to thank the University of Nizwa for the generous support of this project. The authors also extend their appreciation to the Deanship of scientific research at King Khalid University for funding this work through the research groups program under Grant No. RGP.1/259/42.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Martinkova, P.; Kostelnik, A.; Válek, T.; Pohanka, M. Main streams in the construction of biosensors and their applications. Int. J. Electrochem. Sci. 2017, 12, 7386–7403. [Google Scholar] [CrossRef]
  2. Li, P.; Lee, G.-H.; Kim, S.Y.; Kwon, S.Y.; Kim, H.-R.; Park, S. From diagnosis to treatment: Recent advances in patient-friendly biosensors and implantable devices. ACS Nano 2021, 15, 1960–2004. [Google Scholar] [CrossRef] [PubMed]
  3. Huang, Y.; Xu, J.; Liu, J.; Wang, X.; Chen, B. Disease-related detection with electrochemical biosensors: A review. Sensors 2017, 17, 2375. [Google Scholar] [CrossRef]
  4. Monošík, R.; Stred'anský, M.; Šturdík, E. Application of electrochemical biosensors in clinical diagnosis. J. Clin. Lab. Anal. 2012, 26, 22–34. [Google Scholar] [CrossRef] [PubMed]
  5. Monosik, R.; Stredansky, M.; Tkac, J.; Sturdik, E. Application of enzyme biosensors in analysis of food and beverages. Food Anal. Methods 2012, 5, 40–53. [Google Scholar] [CrossRef]
  6. Faridbod, F.; Gupta, V.K.; Zamani, H.A. Electrochemical sensors and biosensors. Int. J. Electrochem. 2011, 2011. [Google Scholar] [CrossRef] [Green Version]
  7. World Health Organization. The World Health Report: 2004: Changing History; World Health Organization: Geneva, Switzerland, 2004. [Google Scholar]
  8. Siegel, R.L.; Miller, K.D.; Fedewa, S.A.; Ahnen, D.J.; Meester, R.G.; Barzi, A.; Jemal, A. Colorectal cancer statistics, 2017. CA Cancer J. Clin. 2017, 67, 177–193. [Google Scholar] [CrossRef] [PubMed]
  9. Fitzmaurice, C.; Allen, C.; Barber, R.M.; Barregard, L.; Bhutta, Z.A.; Brenner, H.; Dicker, D.J.; Chimed-Orchir, O.; Dandona, R.; Dandona, L. Global, regional, and national cancer incidence, mortality, years of life lost, years lived with disability, and disability-adjusted life-years for 32 cancer groups, 1990 to 2015: A systematic analysis for the global burden of disease study. JAMA Oncol. 2017, 3, 524–548. [Google Scholar] [PubMed]
  10. Steward, B.; Kleihues, P. Colorectal Cancer; World Cancer Report; IACR Press: Lyon, France, 2003. [Google Scholar]
  11. Iannazzo, D.; Espro, C.; Celesti, C.; Ferlazzo, A.; Neri, G. Smart biosensors for cancer diagnosis based on graphene quantum dots. Cancers 2021, 13, 3194. [Google Scholar] [CrossRef]
  12. Mahato, K.; Kumar, A.; Maurya, P.K.; Chandra, P. Shifting paradigm of cancer diagnoses in clinically relevant samples based on miniaturized electrochemical nanobiosensors and microfluidic devices. Biosens. Bioelectron. 2018, 100, 411–428. [Google Scholar] [CrossRef]
  13. Chen, W.; Cai, S.; Ren, Q.-Q.; Wen, W.; Zhao, Y.-D. Recent advances in electrochemical sensing for hydrogen peroxide: A review. Analyst 2012, 137, 49–58. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, X.; Martindale, J.L.; Liu, Y.; Holbrook, N.J. The cellular response to oxidative stress: Influences of mitogen-activated protein kinase signalling pathways on cell survival. Biochem. J. 1998, 333, 291–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Schreck, R.; Rieber, P.; Baeuerle, P.A. Reactive oxygen intermediates as apparently widely used messengers in the activation of the NF-kappa B transcription factor and HIV-1. EMBO J. 1991, 10, 2247–2258. [Google Scholar] [CrossRef] [PubMed]
  16. Abe, J.-I.; Berk, B.C. Fyn and JAK2 mediate Ras activation by reactive oxygen species. J. Biol. Chem. 1999, 274, 21003–21010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Elias, H.; Vayssié, S. Reactive peroxo compounds generated in situ from hydrogen peroxide: Kinetics and catalytic application in oxidation processes. Peroxide Chem. Mech. Prep. Asp. Oxyg. Transf. 2000, 128–138. [Google Scholar] [CrossRef]
  18. Imlay, J.A.; Linn, S. Mutagenesis and stress responses induced in Escherichia coli by hydrogen peroxide. J. Bacteriol. 1987, 169, 2967–2976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Mittal, M.; Siddiqui, M.R.; Tran, K.; Reddy, S.P.; Malik, A.B. Reactive oxygen species in inflammation and tissue injury. Antioxid. Redox Signal. 2014, 20, 1126–1167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Sen, S.; Chakraborty, R.; Sridhar, C.; Reddy, Y.; De, B. Free radicals, antioxidants, diseases and phytomedicines: Current status and future prospect. Int. J. Pharm. Sci. Rev. Res. 2010, 3, 91–100. [Google Scholar]
  21. Nogueira, V.; Hay, N. Molecular pathways: Reactive oxygen species homeostasis in cancer cells and implications for cancer therapy. Clin. Cancer Res. 2013, 19, 4309–4314. [Google Scholar] [CrossRef] [Green Version]
  22. Martinez-Outschoorn, U.E.; Balliet, R.M.; Lin, Z.; Whitaker-Menezes, D.; Howell, A.; Sotgia, F.; Lisanti, M.P. Hereditary ovarian cancer and two-compartment tumor metabolism: Epithelial loss of BRCA1 induces hydrogen peroxide production, driving oxidative stress and NFκB activation in the tumor stroma. Cell Cycle 2012, 11, 4152–4166. [Google Scholar] [CrossRef] [Green Version]
  23. Brigelius-Flohe, R.; Kipp, A. Glutathione peroxidases in different stages of carcinogenesis. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2009, 1790, 1555–1568. [Google Scholar] [CrossRef] [PubMed]
  24. Weinstain, R.; Savariar, E.N.; Felsen, C.N.; Tsien, R.Y. In vivo targeting of hydrogen peroxide by activatable cell-penetrating peptides. J. Am. Chem. Soc. 2014, 136, 874–877. [Google Scholar] [CrossRef]
  25. Zhu, L.; Zhang, Y.; Xu, P.; Wen, W.; Li, X.; Xu, J. PtW/MoS2 hybrid nanocomposite for electrochemical sensing of H2O2 released from living cells. Biosens. Bioelectron. 2016, 80, 601–606. [Google Scholar] [CrossRef] [PubMed]
  26. Xi, J.; Xie, C.; Zhang, Y.; Wang, L.; Xiao, J.; Duan, X.; Ren, J.; Xiao, F.; Wang, S. Pd nanoparticles decorated N-doped graphene quantum dots@ N-doped carbon hollow nanospheres with high electrochemical sensing performance in cancer detection. ACS Appl. Mater. Interfaces 2016, 8, 22563–22573. [Google Scholar] [CrossRef] [PubMed]
  27. Yu, C.; Wang, L.; Li, W.; Zhu, C.; Bao, N.; Gu, H. Detection of cellular H2O2 in living cells based on horseradish peroxidase at the interface of Au nanoparticles decorated graphene oxide. Sens. Actuators B Chem. 2015, 211, 17–24. [Google Scholar] [CrossRef]
  28. Razmi, H.; Mohammad-Rezaei, R.; Heidari, H. Self-assembled Prussian blue nanoparticles based electrochemical sensor for high sensitive determination of H2O2 in acidic media. Electroanalysis 2009, 21, 2355–2362. [Google Scholar] [CrossRef]
  29. Liu, Y.; Wang, D.; Xu, L.; Hou, H.; You, T. A novel and simple route to prepare a Pt nanoparticle-loaded carbon nanofiber electrode for hydrogen peroxide sensing. Biosens. Bioelectron. 2011, 26, 4585–4590. [Google Scholar] [CrossRef]
  30. Liu, C.-J.; Yu, S.-L.; Liu, Y.-P.; Dai, X.-J.; Wu, Y.; Li, R.-J.; Tao, J.-C. Synthesis, cytotoxic activity evaluation and HQSAR study of novel isosteviol derivatives as potential anticancer agents. Eur. J. Med. Chem. 2016, 115, 26–40. [Google Scholar] [CrossRef]
  31. Cardoso, A.R.; Moreira, F.T.; Fernandes, R.; Sales, M.G.F. Novel and simple electrochemical biosensor monitoring attomolar levels of miRNA-155 in breast cancer. Biosens. Bioelectron. 2016, 80, 621–630. [Google Scholar] [CrossRef] [PubMed]
  32. Yahalom, G.; Weiss, D.; Novikov, I.; Bevers, T.B.; Radvanyi, L.G.; Liu, M.; Piura, B.; Iacobelli, S.; Sandri, M.T.; Cassano, E. An antibody-based blood test utilizing a panel of biomarkers as a new method for improved breast cancer diagnosis. Biomark. Cancer 2013, 5, 71–80. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, C.; Denno, M.E.; Pyakurel, P.; Venton, B.J. Recent trends in carbon nanomaterial-based electrochemical sensors for biomolecules: A review. Anal. Chim. Acta 2015, 887, 17–37. [Google Scholar] [CrossRef] [Green Version]
  34. Wang, Z.; Dai, Z. Carbon nanomaterial-based electrochemical biosensors: An overview. Nanoscale 2015, 7, 6420–6431. [Google Scholar] [CrossRef] [PubMed]
  35. Vashist, S.K.; Luong, J.H. Recent advances in electrochemical biosensing schemes using graphene and graphene-based nanocomposites. Carbon 2015, 84, 519–550. [Google Scholar] [CrossRef]
  36. Lawal, A.T. Synthesis and utilisation of graphene for fabrication of electrochemical sensors. Talanta 2015, 131, 424–443. [Google Scholar] [CrossRef] [PubMed]
  37. Kuila, T.; Bose, S.; Khanra, P.; Mishra, A.K.; Kim, N.H.; Lee, J.H. Recent advances in graphene-based biosensors. Biosens. Bioelectron. 2011, 26, 4637–4648. [Google Scholar] [CrossRef]
  38. Liu, Y.; Dong, X.; Chen, P. Biological and chemical sensors based on graphene materials. Chem. Soc. Rev. 2012, 41, 2283–2307. [Google Scholar] [CrossRef]
  39. Wu, S.; He, Q.; Tan, C.; Wang, Y.; Zhang, H. Graphene-based electrochemical sensors. Small 2013, 9, 1160–1172. [Google Scholar] [CrossRef]
  40. Ping, J.; Wang, Y.; Fan, K.; Wu, J.; Ying, Y. Direct electrochemical reduction of graphene oxide on ionic liquid doped screen-printed electrode and its electrochemical biosensing application. Biosens. Bioelectron. 2011, 28, 204–209. [Google Scholar] [CrossRef] [PubMed]
  41. 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]
  42. Pohanka, M.; Skládal, P. Electrochemical biosensors—Principles and applications. J. Appl. Biomed. 2008, 6, 57–64. [Google Scholar] [CrossRef] [Green Version]
  43. Vigneshvar, S.; Sudhakumari, C.; Senthilkumaran, B.; Prakash, H. Recent advances in biosensor technology for potential applications–An overview. Front. Bioeng. Biotechnol. 2016, 4, 11. [Google Scholar] [CrossRef] [Green Version]
  44. Lazcka, O.; del Campo, F.J.; Munoz, F.X. Pathogen detection: A perspective of traditional methods and biosensors. Biosens. Bioelectron. 2007, 22, 1205–1217. [Google Scholar] [CrossRef] [PubMed]
  45. He, F. Development of Capillary-Driven Microfludic Biosensors for Food Safety and Quality Assurance. Ph.D. Thesis, University of Massachusetts Amherst, Amherst, MA, USA, 2014. [Google Scholar]
  46. Zhu, C.; Yang, G.; Li, H.; Du, D.; Lin, Y. Electrochemical sensors and biosensors based on nanomaterials and nanostructures. Anal. Chem. 2015, 87, 230–249. [Google Scholar] [CrossRef]
  47. Pundir, C.S.; Deswal, R.; Narwal, V. Quantitative analysis of hydrogen peroxide with special emphasis on biosensors. Bioprocess Biosyst. Eng. 2018, 41, 313–329. [Google Scholar] [CrossRef]
  48. Yunus, S.; Jonas, A.M.; Lakard, B. Potentiometric biosensors. In Encyclopedia of Biophysics; Roberts, G.C.K., Ed.; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  49. Parrilla, M.; Cánovas, R.; Andrade, F.J. Enhanced potentiometric detection of hydrogen peroxide using a platinum electrode coated with nafion. Electroanalysis 2017, 29, 223–230. [Google Scholar] [CrossRef]
  50. Zheng, X.; Guo, Z. Potentiometric determination of hydrogen peroxide at MnO2-doped carbon paste electrode. Talanta 2000, 50, 1157–1162. [Google Scholar] [CrossRef]
  51. Zhao, J.; Yan, Y.; Zhu, L.; Li, X.; Li, G. An amperometric biosensor for the detection of hydrogen peroxide released from human breast cancer cells. Biosens. Bioelectron. 2013, 41, 815–819. [Google Scholar] [CrossRef]
  52. Li, J.; Tan, S.N.; Ge, H. Silica sol-gel immobilized amperometric biosensor for hydrogen peroxide. Anal. Chim. Acta 1996, 335, 137–145. [Google Scholar] [CrossRef]
  53. Wang, B.; Dong, S. Sol–gel-derived amperometric biosensor for hydrogen peroxide based on methylene green incorporated in Nafion film. Talanta 2000, 51, 565–572. [Google Scholar] [CrossRef]
  54. Tripathi, V.S.; Kandimalla, V.B.; Ju, H. Amperometric biosensor for hydrogen peroxide based on ferrocene-bovine serum albumin and multiwall carbon nanotube modified ormosil composite. Biosens. Bioelectron. 2006, 21, 1529–1535. [Google Scholar] [CrossRef]
  55. Yu, J.; Ju, H. Amperometric biosensor for hydrogen peroxide based on hemoglobin entrapped in titania sol–gel film. Anal. Chim. Acta 2003, 486, 209–216. [Google Scholar] [CrossRef]
  56. Povedano, E.; Montiel, V.R.-V.; Gamella, M.; Serafín, V.; Pedrero, M.; Moranova, L.; Bartosik, M.; Montoya, J.J.; Yáñez-Sedeño, P.; Campuzano, S. A novel zinc finger protein–based amperometric biosensor for miRNA determination. Anal. Bioanal. Chem. 2019, 412, 5031–5041. [Google Scholar] [CrossRef] [PubMed]
  57. Li, Y.; Huan, K.; Deng, D.; Tang, L.; Wang, J.; Luo, L. Facile synthesis of ZnMn2O4@ rGO microspheres for ultrasensitive electrochemical detection of hydrogen peroxide from human breast cancer cells. ACS Appl. Mater. Interfaces 2019, 12, 3430–3437. [Google Scholar] [CrossRef] [PubMed]
  58. Dong, W.; Ren, Y.; Bai, Z.; Yang, Y.; Chen, Q. Fabrication of hexahedral Au-Pd/graphene nanocomposites biosensor and its application in cancer cell H2O2 detection. Bioelectrochemistry 2019, 128, 274–282. [Google Scholar] [CrossRef] [PubMed]
  59. Jiao, J.; Pan, M.; Liu, X.; Li, B.; Liu, J.; Chen, Q. A non-enzymatic sensor based on trimetallic nanoalloy with poly (diallyldimethylammonium chloride)-capped reduced graphene oxide for dynamic monitoring hydrogen peroxide production by cancerous cells. Sensors 2020, 20, 71. [Google Scholar] [CrossRef] [Green Version]
  60. Shu, Y.; Zhang, L.; Cai, H.; Yang, Y.; Zeng, J.; Ma, D.; Gao, Q. Hierarchical mo2c@ MoS2 nanorods as electrochemical sensors for highly sensitive detection of hydrogen peroxide and cancer cells. Sens. Actuators B Chem. 2020, 127863. [Google Scholar] [CrossRef]
  61. Thiruppathi, M.; Lin, P.-Y.; Chou, Y.-T.; Ho, H.-Y.; Wu, L.-C.; Ho, J.-A.A. Simple aminophenol-based electrochemical probes for non-enzymatic, dual amperometric detection of NADH and hydrogen peroxide. Talanta 2019, 200, 450–457. [Google Scholar] [CrossRef]
  62. Maji, S.K. Plasmon-enhanced electrochemical biosensing of hydrogen peroxide from cancer cells by gold nanorods. ACS Appl. Nano Mater. 2019, 2, 7162–7169. [Google Scholar] [CrossRef]
  63. Du, H.; Zhang, X.; Liu, Z.; Qu, F. A supersensitive biosensor based on MoS2 nanosheet arrays for the real-time detection of H2O2 secreted from living cells. Chem. Commun. 2019, 55, 9653–9656. [Google Scholar] [CrossRef] [PubMed]
  64. Li, L.; Zhang, Y.; Zhang, L.; Ge, S.; Liu, H.; Ren, N.; Yan, M.; Yu, J. Based device for colorimetric and photoelectrochemical quantification of the flux of H2O2 releasing from MCF-7 cancer cells. Anal. Chem. 2016, 88, 5369–5377. [Google Scholar] [CrossRef]
  65. Zhang, L.-N.; Deng, H.-H.; Lin, F.-L.; Xu, X.-W.; Weng, S.-H.; Liu, A.-L.; Lin, X.-H.; Xia, X.-H.; Chen, W. In situ growth of porous platinum nanoparticles on graphene oxide for colorimetric detection of cancer cells. Anal. Chem. 2014, 86, 2711–2718. [Google Scholar] [CrossRef]
  66. Ge, S.; Liu, W.; Liu, H.; Liu, F.; Yu, J.; Yan, M.; Huang, J. Colorimetric detection of the flux of hydrogen peroxide released from living cells based on the high peroxidase-like catalytic performance of porous PtPd nanorods. Biosens. Bioelectron. 2015, 71, 456–462. [Google Scholar] [CrossRef] [PubMed]
  67. Ye, X.; Shi, H.; He, X.; Wang, K.; He, D.; Yan, L.A.; Xu, F.; Lei, Y.; Tang, J.; Yu, Y. Iodide-responsive Cu–Au nanoparticle-based colorimetric platform for ultrasensitive detection of target cancer cells. Anal. Chem. 2015, 87, 7141–7147. [Google Scholar] [CrossRef] [PubMed]
  68. Kim, J.-H.; Patra, C.R.; Arkalgud, J.R.; Boghossian, A.A.; Zhang, J.; Han, J.-H.; Reuel, N.F.; Ahn, J.-H.; Mukhopadhyay, D.; Strano, M.S. Single-molecule detection of H2O2 mediating angiogenic redox signaling on fluorescent single-walled carbon nanotube array. ACS Nano 2011, 5, 7848–7857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Maji, S.K.; Sreejith, S.; Mandal, A.K.; Ma, X.; Zhao, Y. Immobilizing gold nanoparticles in mesoporous silica covered reduced graphene oxide: A hybrid material for cancer cell detection through hydrogen peroxide sensing. ACS Appl. Mater. Interfaces 2014, 6, 13648–13656. [Google Scholar] [CrossRef] [PubMed]
  70. Wang, Y.; Tang, L.; Li, Z.; Lin, Y.; Li, J. In situ simultaneous monitoring of ATP and GTP using a graphene oxide nanosheet–based sensing platform in living cells. Nat. Protoc. 2014, 9, 1944. [Google Scholar] [CrossRef]
  71. McKibbin, P.L.; Kobori, A.; Taniguchi, Y.; Kool, E.T.; David, S.S. Surprising repair activities of nonpolar analogs of 8-oxoG expose features of recognition and catalysis by base excision repair glycosylases. J. Am. Chem. Soc. 2012, 134, 1653–1661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Li, L.; Lin, H.; Lei, C.; Nie, Z.; Huang, Y.; Yao, S. Label-free fluorescence assay for thrombin based on unmodified quantum dots. Biosens. Bioelectron. 2014, 54, 42–47. [Google Scholar] [CrossRef] [PubMed]
  73. Ren, D.; Wong, N.T.; Handoko, A.D.; Huang, Y.; Yeo, B.S. Mechanistic insights into the enhanced activity and stability of agglomerated Cu nanocrystals for the electrochemical reduction of carbon dioxide to n-propanol. J. Phys. Chem. Lett. 2016, 7, 20–24. [Google Scholar] [CrossRef]
  74. Zhou, Y.; Zhang, Y.; Lau, C.; Lu, J. Sequential determination of two proteins by temperature-triggered homogeneous chemiluminescent immunoassay. Anal. Chem. 2006, 78, 5920–5924. [Google Scholar] [CrossRef] [PubMed]
  75. Liu, M.; Lin, Z.; Lin, J.-M. A review on applications of chemiluminescence detection in food analysis. Anal. Chim. Acta 2010, 670, 1–10. [Google Scholar] [CrossRef]
  76. Kong, H.; Liu, D.; Zhang, S.; Zhang, X. Protein sensing and cell discrimination using a sensor array based on nanomaterial-assisted chemiluminescence. Anal. Chem. 2011, 83, 1867–1870. [Google Scholar] [CrossRef] [PubMed]
  77. Ji, D.; Mohsen, M.G.; Harcourt, E.M.; Kool, E.T. ATP-Releasing Nucleotides: Linking DNA Synthesis to Luciferase Signaling. Angew. Chem. Int. Ed. 2016, 55, 2087–2091. [Google Scholar] [CrossRef] [Green Version]
  78. Liu, B.-F.; Ozaki, M.; Hisamoto, H.; Luo, Q.; Utsumi, Y.; Hattori, T.; Terabe, S. Microfluidic chip toward cellular ATP and ATP-conjugated metabolic analysis with bioluminescence detection. Anal. Chem. 2005, 77, 573–578. [Google Scholar] [CrossRef] [PubMed]
  79. 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. 2015, 87, 1903–1910. [Google Scholar] [CrossRef] [PubMed]
  80. Bai, J.; Jiang, X. A facile one-pot synthesis of copper sulfide-decorated reduced graphene oxide composites for enhanced detecting of H2O2 in biological environments. Anal. Chem. 2013, 85, 8095–8101. [Google Scholar] [CrossRef]
  81. Lu, Y.; Liu, Y.; Zhang, S.; Wang, S.; Zhang, S.; Zhang, X. Aptamer-Based Plasmonic Sensor Array for Discrimination of Proteins and Cells with the Naked Eye. Anal. Chem. 2013, 85, 6571–6574. [Google Scholar] [CrossRef]
  82. Gong, Y.; Chen, X.; Lu, Y.; Yang, W. Self-assembled dipeptide–gold nanoparticle hybrid spheres for highly sensitive amperometric hydrogen peroxide biosensors. Biosens. Bioelectron. 2015, 66, 392–398. [Google Scholar] [CrossRef] [PubMed]
  83. Wang, Y.; Gu, H. Core-Shell-Type Magnetic Mesoporous Silica Nanocomposites for Bioimaging and Therapeutic Agent Delivery. Adv. Mater. 2015, 27, 576–585. [Google Scholar] [CrossRef]
  84. Yang, J.; Shen, D.; Zhou, L.; Li, W.; Li, X.; Yao, C.; Wang, R.; El-Toni, A.M.; Zhang, F.; Zhao, D. Spatially confined fabrication of core–shell gold nanocages@mesoporous silica for near-infrared controlled photothermal drug release. Chem. Mater. 2013, 25, 3030–3037. [Google Scholar] [CrossRef]
  85. Lee, D.; Khaja, S.D.; Velasquez-Castano, J.C.; Dasari, M.; Sun, C.; A Petros, J.; Taylor, W.R.; Murthy, N. In vivo imaging of hydrogen peroxide with chemiluminescent nanoparticles. Nat. Mater. 2007, 6, 765–769. [Google Scholar] [CrossRef] [PubMed]
  86. Arnous, A.; Petrakis, C.; Makris, D.P.; Kefalas, P. A peroxyoxalate chemiluminescence-based assay for the evaluation of hydrogen peroxide scavenging activity employing 9,10-diphenylanthracene as the fluorophore. J. Pharmacol. Toxicol. Methods 2002, 48, 171–177. [Google Scholar] [CrossRef]
  87. Koike, R.; Kato, Y.; Motoyoshiya, J.; Nishii, Y.; Aoyama, H. Unprecedented chemiluminescence behaviour during peroxyoxalate chemiluminescence of oxalates with fluorescent or electron-donating aryloxy groups. Luminescence 2006, 21, 164–173. [Google Scholar] [CrossRef]
  88. Stevani, C.V.; Silva, S.M.; Baader, W.J. Studies on the Mechanism of the Excitation Step in Peroxyoxalate Chemiluminescence. Eur. J. Org. Chem. 2000, 2000, 4037–4046. [Google Scholar] [CrossRef]
  89. Matsumoto, M. Advanced chemistry of dioxetane-based chemiluminescent substrates originating from bioluminescence. J. Photochem. Photobiol. C 2004, 5, 27–53. [Google Scholar] [CrossRef]
  90. Lee, D.; Dasari, M.; Erigala, V.; Murthy, N.; Yu, J.; Dickson, R. Detection of hydrogen peroxide with chemiluminescent micelles. Int. J. Nanomed. 2008, 3, 471–476. [Google Scholar] [CrossRef] [Green Version]
  91. Weissleder, R.; Pittet, M.J. Imaging in the era of molecular oncology. Nature 2008, 452, 580–589. [Google Scholar] [CrossRef] [Green Version]
  92. Zhang, K.; Kaufman, R.J. From endoplasmic-reticulum stress to the inflammatory response. Nature 2008, 454, 455–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Peer, D.; Karp, J.M.; Hong, S.; Farokhzad, O.C.; Margalit, R.; Langer, R. Nanocarriers as an emerging platform for cancer therapy. Nat. Nanotechnol. 2007, 2, 751–760. [Google Scholar] [CrossRef] [PubMed]
  94. Nan, Y.; Zhao, W.; Li, N.; Liang, Z.; Xu, X. Chemiluminescence-triggered fluorophore release: Approach for in vivo fluorescence imaging of hydrogen peroxide. Sens. Actuators B Chem. 2019, 281, 296–302. [Google Scholar] [CrossRef]
  95. Lee, Y.-D.; Lim, C.-K.; Singh, A.; Koh, J.; Kim, J.; Kwon, I.C.; Kim, S. Dye/Peroxalate Aggregated Nanoparticles with Enhanced and Tunable Chemiluminescence for Biomedical Imaging of Hydrogen Peroxide. ACS Nano 2012, 6, 6759–6766. [Google Scholar] [CrossRef] [PubMed]
  96. Lee, E.S.; Deepagan, V.G.; Gil You, D.; Jeon, J.; Yi, G.-R.; Lee, J.Y.; Lee, D.S.; Suh, Y.D.; Park, J.H. Nanoparticles based on quantum dots and a luminol derivative: Implications for in vivo imaging of hydrogen peroxide by chemiluminescence resonance energy transfer. Chem. Commun. 2016, 52, 4132–4135. [Google Scholar] [CrossRef] [PubMed]
  97. Geng, J.; Li, K.; Qin, W.; Tang, B.Z.; Liu, B. Red-Emissive Chemiluminescent Nanoparticles with Aggregation-Induced Emission Characteristics for In Vivo Hydrogen Peroxide Imaging. Part. Part. Syst. Charact. 2014, 31, 1238–1243. [Google Scholar] [CrossRef]
  98. Jia, Y.; Sun, S.; Cui, X.; Wang, X.; Yang, L. Enzyme-like catalysis of polyoxometalates for chemiluminescence: Application in ultrasensitive detection of H2O2 and blood glucose. Talanta 2019, 205, 120139. [Google Scholar] [CrossRef] [PubMed]
  99. Klassen, N.V.; Marchington, D.; McGowan, H.C. H2O2 Determination by the I3 Method and by KMnO4 Titration. Anal. Chem. 1994, 66, 2921–2925. [Google Scholar] [CrossRef]
  100. Kieber, R.J.; Helz, G.R. Two-method verification of hydrogen peroxide determinations in natural waters. Anal. Chem. 1986, 58, 2312–2315. [Google Scholar] [CrossRef]
  101. Putt, K.S.; Pugh, R.B. A High-throughput microtiter plate based method for the determination of peracetic acid and hydrogen peroxide. PLoS ONE 2013, 8, e79218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Zaribafan, A.; Haghbeen, K.; Fazli, M.; Akhondali, A. Spectrophotometric method for hydrogen peroxide determination through oxidation of organic dyes. Environ. Stud. Persian Gulf 2014, 1, 93–101. [Google Scholar]
  103. Matsubara, C.; Kudo, K.; Kawashita, T.; Takamura, K. Spectrophotometric determination of hydrogen peroxide with titanium 2-((5-bromopyridyl)azo)-5-(N-propyl-N-sulfopropylamino)phenol reagent and its application to the determination of serum glucose using glucose oxidase. Anal. Chem. 1985, 57, 1107–1109. [Google Scholar] [CrossRef]
  104. Clapp, P.A.; Evans, D.F.; Sheriff, T.S. Spectrophotometric determination of hydrogen peroxide after extraction with ethyl acetate. Anal. Chim. Acta 1989, 218, 331–334. [Google Scholar] [CrossRef]
  105. Mukhopadhyay, D.; Dasgupta, P.; Roy, D.S.; Palchoudhuri, S.; Chatterjee, I.; Ali, S.; Dastidar, S.G. A Sensitive In vitro Spectrophotometric Hydrogen Peroxide Scavenging Assay using 1,10-Phenanthroline. Free Radic. Antioxid. 2016, 6, 124–132. [Google Scholar] [CrossRef]
  106. Elnemma, E.M. Spectrophotometric Determination of Hydrogen Peroxide by a Hydroquinone-Aniline System Catalyzed by Molybdate. Bull. Korean Chem. Soc. 2004, 25, 127–129. [Google Scholar]
  107. Zhang, L.-S.; Wong, G.T. Spectrophotometric determination of H2O2 in marine waters with leuco crystal violet. Talanta 1994, 41, 2137–2145. [Google Scholar] [CrossRef]
  108. Huang, Y.; Cai, R.; Mao, L.; LIU, Z.; HUANG, H. Spectrophotometric determination of hydrogen peroxide using β-CD-Hemin as a mimetic enzyme of peroxidase. Anal. Sci. 1999, 15, 889–894. [Google Scholar] [CrossRef] [Green Version]
  109. Zhang, Q.; Fu, S.; Li, H.; Liu, Y. A novel method for the determination of hydrogen peroxide in bleaching effluents by spectroscopy. BioResources 2013, 8, 3699–3705. [Google Scholar] [CrossRef] [Green Version]
  110. Eisenberg, G. Industrial and engineering chemistry. Ind. Eng. Chem. Anal. Ed. 1943, 15, 327–328. [Google Scholar] [CrossRef]
  111. Graf, E.; Penniston, J.T. Method for determination of hydrogen peroxide, with its application illustrated by glucose assay. Clin. Chem. 1980, 26, 658–660. [Google Scholar] [CrossRef]
  112. Pick, E.; Keisari, Y. A simple colorimetric method for the measurement of H2O2 produced by cells in culture J. Immunol. Methods 1980, 38, 161–170. [Google Scholar] [CrossRef]
  113. Fernando, C.D.; Soysa, P. Optimized enzymatic colorimetric assay for determination of hydrogen peroxide (H2O2) scavenging activity of plant extracts. MethodsX 2015, 2, 283–291. [Google Scholar] [CrossRef] [PubMed]
  114. Su, G.; Wei, Y.; Guo, M. Direct Colorimetric Detection of Hydrogen Peroxide Using 4-Nitrophenyl Boronic Acid or Its Pinacol Ester. Am. J. Anal. Chem. 2011, 2, 879–884. [Google Scholar] [CrossRef] [Green Version]
  115. Nitinaivinij, K.; Parnklang, T.; Thammacharoen, C.; Ekgasit, S.; Wongravee, K. Colorimetric determination of hydrogen peroxide by morphological decomposition of silver nanoprisms coupled with chromaticity analysis. Anal. Methods 2014, 6, 9816–9824. [Google Scholar] [CrossRef]
  116. Takahashi, A.; Hashimoto, K.; Kumazawa, S.; Nakayama, T. Determination of Hydrogen Peroxide by High-Performance Liquid Chromatography with a Cation-Exchange Resin Gel Column and Electrochemical Detector. Anal. Sci. 1999, 15, 481–483. [Google Scholar] [CrossRef] [Green Version]
  117. Wada, M.; Inoue, K.; Ihara, A.; Kishikawa, N.; Nakashima, K.; Kuroda, N. Determination of organic peroxides by liquid chromatography with on-line post-column ultraviolet irradiation and peroxyoxalate chemiluminescence detection. J. Chromatogr. A 2003, 987, 189–195. [Google Scholar] [CrossRef]
  118. Nepomnyashchikh, Y.V.; Borkina, G.G.; Karavaeva, A.V.; Perkel’, A.L. Photometric and Gas-Chromatographic Determination of Hydrogen Peroxide and Peroxybutanoic Acid in Oxidized Butanoic Acid. J. Anal. Chem. 2005, 60, 1024–1028. [Google Scholar] [CrossRef]
  119. Magara, K.; Ikeda, T.; Sugimoto, T.; Hosoya, S. Quantitative Analysis of Hydrogen Peroxide by High Performance Liquid Chromatography. Jpn. TAPPI J. 2007, 61, 1481–1493. [Google Scholar] [CrossRef] [Green Version]
  120. Tarno, H.; Qi, H.; Endoh, R.; Kobayashi, M.; Goto, H.; Futai, K. Types of frass produced by the ambrosia beetle Platypus quercivorus during gallery construction, and host suitability of five tree species for the beetle. J. For. Res. 2011, 16, 68–75. [Google Scholar] [CrossRef]
  121. Wielandt, H. On the eigenvalues of A + B and AB. J. Res. Natl. Bur. Stand. Sect. B Math. Sci. 1973, 77B, 61. [Google Scholar] [CrossRef]
  122. Xu, K.; Tang, B.; Huang, H.; Yang, G.; Chen, Z.; Li, P.; An, L. Strong red fluorescent probes suitable for detecting hydrogen peroxide generated by mice peritoneal macrophages. Chem. Commun. 2005, 48, 5974–5976. [Google Scholar] [CrossRef]
  123. Paździoch-Czochra, M.; Wideńska, A. Spectrofluorimetric determination of hydrogen peroxide scavenging activity. Anal. Chim. Acta 2002, 452, 177–184. [Google Scholar] [CrossRef]
  124. Miller, E.W.; Albers, A.E.; Pralle, A.; Isacoff, E.Y.; Chang, C.J. Boronate-Based Fluorescent Probes for Imaging Cellular Hydrogen Peroxide. J. Am. Chem. Soc. 2005, 127, 16652–16659. [Google Scholar] [CrossRef] [Green Version]
  125. Qian, P.; Qin, Y.; Lyu, Y.; Li, Y.; Wang, L.; Wang, S.; Liu, Y. A hierarchical cobalt/carbon nanotube hybrid nanocomplex-based ratiometric fluorescent nanosensor for ultrasensitive detection of hydrogen peroxide and glucose in human serum. Anal. Bioanal. Chem. 2019, 411, 1517–1524. [Google Scholar] [CrossRef] [PubMed]
  126. Onoda, M.; Uchiyama, T.; Mawatari, K.-I.; Kaneko, K.; Nakagomi, K. Simple and Rapid Determination of Hydrogen Peroxide Using Phosphine-based Fluorescent Reagents with Sodium Tungstate Dihydrate. Anal. Sci. 2006, 22, 815–817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Lyublinskaya, O.; Antunes, F. Measuring intracellular concentration of hydrogen peroxide with the use of genetically encoded H2O2 biosensor HyPer. Redox Biol. 2019, 24, 101200. [Google Scholar] [CrossRef]
  128. Belousov, V.V.; Fradkov, A.F.; Lukyanov, K.; Staroverov, D.; Shakhbazov, K.S.; Terskikh, A.V.; Lukyanov, S. Genetically encoded fluorescent indicator for intracellular hydrogen peroxide. Nat. Methods 2006, 3, 281–286. [Google Scholar] [CrossRef]
  129. Zheng, M.; Åslund, F.; Storz, G. Activation of the OxyR Transcription Factor by Reversible Disulfide Bond Formation. Science 1998, 279, 1718–1722. [Google Scholar] [CrossRef]
  130. Markvicheva, K.N.; Bilan, D.; Mishina, N.; Gorokhovatsky, A.Y.; Vinokurov, L.M.; Lukyanov, S.; Belousov, V.V. A genetically encoded sensor for H2O2 with expanded dynamic range. Bioorganic Med. Chem. 2011, 19, 1079–1084. [Google Scholar] [CrossRef]
  131. Bilan, D.S.; Pase, L.; Joosen, L.; Gorokhovatsky, A.Y.; Ermakova, Y.G.; Gadella, T.W.J.; Grabher, C.; Schultz, C.; Lukyanov, S.; Belousov, V.V. HyPer-3: A Genetically Encoded H2O2 Probe with Improved Performance for Ratiometric and Fluorescence Lifetime Imaging. ACS Chem. Biol. 2013, 8, 535–542. [Google Scholar] [CrossRef] [Green Version]
  132. Pak, V.V.; Ezerina, D.; Lyublinskaya, O.; Pedre, B.; Tyurin-Kuzmin, P.A.; Mishina, N.M.; Thauvin, M.; Young, D.; Wahni, K.; Gache, S.A.M.; et al. Ultrasensitive Genetically Encoded Indicator for Hydrogen Peroxide Identifies Roles for the Oxidant in Cell Migration and Mitochondrial Function. Cell Metab. 2020, 31, 642–653.e6. [Google Scholar] [CrossRef] [PubMed]
  133. Ermakova, Y.; Bilan, D.; Matlashov, M.; Mishina, N.; Markvicheva, K.N.; Subach, O.M.; Subach, F.V.; Bogeski, I.; Hoth, M.; Enikolopov, G.; et al. Red fluorescent genetically encoded indicator for intracellular hydrogen peroxide. Nat. Commun. 2014, 5, 5222. [Google Scholar] [CrossRef]
  134. Choi, H.-J.; Kim, S.-J.; Mukhopadhyay, P.; Cho, S.; Woo, J.-R.; Storz, G.; Ryu, S.-E. Structural Basis of the Redox Switch in the OxyR Transcription Factor. Cell 2001, 105, 103–113. [Google Scholar] [CrossRef] [Green Version]
  135. Xu, J.; Zhang, Y.; Yu, H.; Gao, X.; Shao, S. Mitochondria-Targeted Fluorescent Probe for Imaging Hydrogen Peroxide in Living Cells. Anal. Chem. 2015, 88, 1455–1461. [Google Scholar] [CrossRef] [PubMed]
  136. Xiao, H.; Li, P.; Hu, X.; Shi, X.; Zhang, W.; Tang, B. Simultaneous fluorescence imaging of hydrogen peroxide in mitochondria and endoplasmic reticulum during apoptosis. Chem. Sci. 2016, 7, 6153–6159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Shen, R.; Liu, P.; Zhang, Y.; Yu, Z.; Chen, X.; Zhou, L.; Nie, B.; Żaczek, A.; Chen, J.; Liu, J. Sensitive Detection of Single-Cell Secreted H2O2 by Integrating a Microfluidic Droplet Sensor and Au Nanoclusters. Anal. Chem. 2018, 90, 4478–4484. [Google Scholar] [CrossRef] [PubMed]
  138. Chen, Y.; Ye, J.; Lv, G.; Liu, W.; Jiang, H.; Liu, X.; Wang, X. Hydrogen Peroxide and Hypochlorite Responsive Fluorescent Nanoprobes for Sensitive Cancer Cell Imaging. Biosensors 2022, 12, 111. [Google Scholar] [CrossRef]
  139. Hu, L.; Yuan, Y.; Zhang, L.; Zhao, J.; Majeed, S.; Xu, G. Copper nanoclusters as peroxidase mimetics and their applications to H2O2 and glucose detection. Anal. Chim. Acta 2013, 762, 83–86. [Google Scholar] [CrossRef]
  140. Liu, H.; Gu, C.; Xiong, W.; Zhang, M. A sensitive hydrogen peroxide biosensor using ultra-small CuInS2 nanocrystals as peroxidase mimics. Sens. Actuators B Chem. 2015, 209, 670–676. [Google Scholar] [CrossRef]
  141. Su, L.; Qin, W.; Zhang, H.; Rahman, Z.U.; Ren, C.; Ma, S.; Chen, X. The peroxidase/catalase-like activities of MFe2O4 (M = Mg, Ni, Cu) MNPs and their application in colorimetric biosensing of glucose. Biosens. Bioelectron. 2015, 63, 384–391. [Google Scholar] [CrossRef]
  142. Regalado, C.; García-Almendárez, B.E.; Duarte-Vázquez, M.A. Biotechnological applications of peroxidases. Phytochem. Rev. 2004, 3, 243–256. [Google Scholar] [CrossRef]
  143. Hamid, M. Potential applications of peroxidases. Food Chem. 2009, 115, 1177–1186. [Google Scholar] [CrossRef]
  144. Chekin, F.; Gorton, L.; Tapsobea, I. Direct and mediated electrochemistry of peroxidase and its electrocatalysis on a variety of screen-printed carbon electrodes: Amperometric hydrogen peroxide and phenols biosensor. Anal. Bioanal. Chem. 2014, 407, 439–446. [Google Scholar] [CrossRef]
  145. Yang, H.; Liu, B.; Ding, Y.; Li, L.; Ouyang, X. Fabrication of cuprous oxide nanoparticles-graphene nanocomposite for determination of acetaminophen. J. Electroanal. Chem. 2015, 757, 88–93. [Google Scholar] [CrossRef]
  146. Chulkova, I.; Derina, K.; Taishibekova, Y. The modified electrode for the determination of cholesterol. In Proceedings of the Chemistry and Chemical Technology in the XXI Century: Materials of the XVI International Scientific-Practical Conference of Students and Young Scientists Dedicated to the 115th Anniversary of Professor L.P. Kuleva, Tomsk, Russia, 25–29 May 2015; pp. 195–197. [Google Scholar]
  147. Jelikić-Stankov, M.D.; Djurdjevic, P.; Stankov, D. Determination of uric acid in human serum by an enzymatic method using N-methyl-N-(4-aminophenyl)-3-methoxyaniline reagent. J. Serb. Chem. Soc. 2003, 68, 691–698. [Google Scholar] [CrossRef]
  148. Zhou, B.; Wang, J.; Guo, Z.; Tan, H.; Zhu, X. A simple colorimetric method for determination of hydrogen peroxide in plant tissues. Plant Growth Regul. 2006, 49, 113–118. [Google Scholar] [CrossRef]
  149. Chinnadayyala, S.R.; Kakoti, A.; Santhosh, M.; Goswami, P. A novel amperometric alcohol biosensor developed in a 3rd generation bioelectrode platform using peroxidase coupled ferrocene activated alcohol oxidase as biorecognition system. Biosens. Bioelectron. 2014, 55, 120–126. [Google Scholar] [CrossRef]
  150. Yu, F.; Huang, Y.; Cole, A.J.; Yang, V.C. The artificial peroxidase activity of magnetic iron oxide nanoparticles and its application to glucose detection. Biomaterials 2009, 30, 4716–4722. [Google Scholar] [CrossRef] [Green Version]
  151. Mu, J.; Zhang, L.; Zhao, M.; Wang, Y. Co3O4 nanoparticles as an efficient catalase mimic: Properties, mechanism and its electrocatalytic sensing application for hydrogen peroxide. J. Mol. Catal. A Chem. 2013, 378, 30–37. [Google Scholar] [CrossRef]
  152. Yoon, J.; Lee, T.; Bapurao G., B.; Jo, J.; Oh, B.-K.; Choi, J.-W. Electrochemical H2O2 biosensor composed of myoglobin on MoS2 nanoparticle-graphene oxide hybrid structure. Biosens. Bioelectron. 2017, 93, 14–20. [Google Scholar] [CrossRef]
  153. Wen, Z.; Ci, S.; Li, J. Pt Nanoparticles Inserting in Carbon Nanotube Arrays: Nanocomposites for Glucose Biosensors. J. Phys. Chem. C 2009, 113, 13482–13487. [Google Scholar] [CrossRef]
  154. Pingarrón, J.M.; Yáñez-Sedeño, P.; González-Cortés, A. Gold nanoparticle-based electrochemical biosensors. Electrochim. Acta 2008, 53, 5848–5866. [Google Scholar] [CrossRef]
  155. Zhou, M.; Zhai, Y.M.; Dong, S.J. Electrochemical sensing and biosensing platform based on chemically reduced graphene oxide. Anal. Chem. 2009, 81, 5603–5613. [Google Scholar] [CrossRef]
  156. Xu, X.; Jiang, S.; Hu, Z.; Liu, S. Nitrogen-doped carbon nanotubes: High electrocatalytic activity toward the oxidation of hydrogen peroxide and its application for biosensing. ACS Nano 2010, 4, 4292–4298. [Google Scholar] [CrossRef]
  157. Luo, Y.; Liu, H.; Rui, Q.; Tian, Y. Detection of Extracellular H2O2 Released from Human Liver Cancer Cells Based on TiO2 Nanoneedles with Enhanced Electron Transfer of Cytochrome c. Anal. Chem. 2009, 81, 3035–3041. [Google Scholar] [CrossRef] [PubMed]
  158. Shao, Y.; Wang, J.; Wu, H.; Liu, J.; Aksay, I.A.; Lin, Y. Graphene based electrochemical sensors and biosensors: A review. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 2010, 22, 1027–1036. [Google Scholar] [CrossRef]
  159. Kim, G.; Lee, Y.-E.K.; Xu, H.; Philbert, M.A.; Kopelman, R. Nanoencapsulation method for high selectivity sensing of hydrogen peroxide inside live cells. Anal. Chem. 2010, 82, 2165–2169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Sun, X.; Guo, S.; Liu, Y.; Sun, S. Dumbbell-like PtPd–Fe3O4 nanoparticles for enhanced electrochemical detection of H2O2. Nano Lett. 2012, 12, 4859–4863. [Google Scholar] [CrossRef] [PubMed]
  161. Wang, T.; Zhu, H.; Zhuo, J.; Zhu, Z.; Papakonstantinou, P.; Lubarsky, G.; Lin, J.; Li, M. Biosensor Based on Ultrasmall MoS2 Nanoparticles for Electrochemical Detection of H2O2 Released by Cells at the Nanomolar Level. Anal. Chem. 2013, 85, 10289–10295. [Google Scholar] [CrossRef] [PubMed]
  162. Dou, B.; Yang, J.; Yuan, R.; Xiang, Y. Trimetallic Hybrid Nanoflower-Decorated MoS2 Nanosheet Sensor for Direct in Situ Monitoring of H2O2 Secreted from Live Cancer Cells. Anal. Chem. 2018, 90, 5945–5950. [Google Scholar] [CrossRef] [PubMed]
  163. Chang, H.-C.; Ho, J.-A.A. Gold nanocluster-assisted fluorescent detection for hydrogen peroxide and cholesterol based on the inner filter effect of gold nanoparticles. Anal. Chem. 2015, 87, 10362–10367. [Google Scholar] [CrossRef] [PubMed]
  164. Cui, H.; Wang, W.; Duan, C.-F.; Dong, Y.-P.; Guo, J.-Z. Synthesis, characterization, and electrochemiluminescence of luminol-reduced gold nanoparticles and their application in a hydrogen peroxide sensor. Chem.–A Eur. J. 2007, 13, 6975–6984. [Google Scholar] [CrossRef] [PubMed]
  165. Liu, Q.; Yang, Y.; Lv, X.; Ding, Y.; Zhang, Y.; Jing, J.; Xu, C. One-step synthesis of uniform nanoparticles of porphyrin functionalized ceria with promising peroxidase mimetics for H2O2 and glucose colorimetric detection. Sens. Actuators B Chem. 2017, 240, 726–734. [Google Scholar] [CrossRef]
  166. Xiong, X.; You, C.; Cao, X.; Pang, L.; Kong, R.; Sun, X. Ni2P nanosheets array as a novel electrochemical catalyst electrode for non-enzymatic H2O2 sensing. Electrochim. Acta 2017, 253, 517–521. [Google Scholar] [CrossRef]
  167. Chen, L.; Wang, N.; Wang, X.; Ai, S. Protein-directed in situ synthesis of platinum nanoparticles with superior peroxidase-like activity, and their use for photometric determination of hydrogen peroxide. Mikrochim. Acta 2013, 180, 1517–1522. [Google Scholar] [CrossRef]
  168. Ma, B.; Kong, C.; Hu, X.; Liu, K.; Huang, Q.; Lv, J.; Lu, W.; Zhang, X.; Yang, Z.; Yang, S. A sensitive electrochemical nonenzymatic biosensor for the detection of H2O2 released from living cells based on ultrathin concave Ag nanosheets. Biosens. Bioelectron. 2018, 106, 29–36. [Google Scholar] [CrossRef]
  169. Asif, M.; Liu, H.; Aziz, A.; Wang, H.; Wang, Z.; Ajmal, M.; Xiao, F.; Liu, H. Core-shell iron oxide-layered double hydroxide: High electrochemical sensing performance of H2O2 biomarker in live cancer cells with plasma therapeutics. Biosens. Bioelectron. 2017, 97, 352–359. [Google Scholar] [CrossRef]
  170. Li, Z.; Xin, Y.; Wu, W.; Fu, B.; Zhang, Z. Topotactic Conversion of Copper(I) Phosphide Nanowires for Sensitive Electrochemical Detection of H2O2 Release from Living Cells. Anal. Chem. 2016, 88, 7724–7729. [Google Scholar] [CrossRef]
  171. Su, S.; Han, X.; Lu, Z.; Liu, W.; Zhu, D.; Chao, J.; Fan, C.; Wang, L.; Song, S.; Weng, L.; et al. Facile Synthesis of a MoS2–Prussian Blue Nanocube Nanohybrid-Based Electrochemical Sensing Platform for Hydrogen Peroxide and Carcinoembryonic Antigen Detection. ACS Appl. Mater. Interfaces 2017, 9, 12773–12781. [Google Scholar] [CrossRef] [PubMed]
  172. Yuan, L.; Lin, W.; Xie, Y.; Chen, B.; Zhu, S. Single fluorescent probe responds to H2O2, NO, and H2O2/NO with three different sets of fluorescence signals. J. Am. Chem. Soc. 2012, 134, 1305–1315. [Google Scholar] [CrossRef] [PubMed]
  173. Boero, C.; Casulli, M.A.; Olivo, J.; Foglia, L.; Orso, E.; Mazza, M.; Carrara, S.; De Micheli, G. Design, development, and validation of an in-situ biosensor array for metabolite monitoring of cell cultures. Biosens. Bioelectron. 2014, 61, 251–259. [Google Scholar] [CrossRef] [Green Version]
  174. Shi, B.-X.; Wang, Y.; Zhang, K.; Lam, T.-L.; Chan, H.L.-W. Monitoring of dopamine release in single cell using ultrasensitive ITO microsensors modified with carbon nanotubes. Biosens. Bioelectron. 2011, 26, 2917–2921. [Google Scholar] [CrossRef] [PubMed]
  175. Li, D.-W.; Qin, L.-X.; Li, Y.; Nia, R.P.; Long, Y.-T.; Chen, H.-Y. CdSe/ZnS quantum dot–Cytochrome c bioconjugates for selective intracellular O 2–Sensing. Chem. Commun. 2011, 47, 8539–8541. [Google Scholar] [CrossRef] [PubMed]
  176. Han, M.; Liu, S.; Bao, J.; Dai, Z. Pd nanoparticle assemblies—As the substitute of HRP, in their biosensing applications for H2O2 and glucose. Biosens. Bioelectron. 2012, 31, 151–156. [Google Scholar] [CrossRef]
  177. Wang, Y.; Hasebe, Y. Carbon felt-based bioelectrocatalytic flow-through detectors: Highly sensitive amperometric determination of H2O2 based on a direct electrochemistry of covalently modified horseradish peroxidase using cyanuric chloride as a linking agent. Sens. Actuators B Chem. 2011, 155, 722–729. [Google Scholar] [CrossRef]
  178. Wang, Z.; Yang, Y.; Leng, K.; Li, J.; Zheng, F.; Shen, G.; Yu, R. A Sequence-Selective Electrochemical DNA Biosensor Based on HRP-Labeled Probe for Colorectal Cancer DNA Detection. Anal. Lett. 2008, 41, 24–35. [Google Scholar] [CrossRef]
  179. Crulhas, B.P.; Ramos, N.P.; Castro, G.R.; Pedrosa, V.A. Detection of hydrogen peroxide releasing from prostate cancer cell using a biosensor. J. Solid State Electrochem. 2016, 20, 2427–2433. [Google Scholar] [CrossRef] [Green Version]
  180. Zhou, J.; Liao, C.; Zhang, L.; Wang, Q.; Tian, Y. Molecular Hydrogel-Stabilized Enzyme with Facilitated Electron Transfer for Determination of H2O2 Released from Live Cells. Anal. Chem. 2014, 86, 4395–4401. [Google Scholar] [CrossRef]
  181. Goenka, S.; Sant, V.; Sant, S. Graphene-based nanomaterials for drug delivery and tissue engineering. J. Control. Release 2014, 173, 75–88. [Google Scholar] [CrossRef]
  182. Pumera, M.; Ambrosi, A.; Bonanni, A.; Chng, E.L.K.; Poh, H.L. Graphene for electrochemical sensing and biosensing. TrAC Trends Anal. Chem. 2010, 29, 954–965. [Google Scholar] [CrossRef]
  183. Gao, H.; Duan, H. 2D and 3D graphene materials: Preparation and bioelectrochemical applications. Biosens. Bioelectron. 2015, 65, 404–419. [Google Scholar] [CrossRef] [PubMed]
  184. Favero, G.; Fusco, G.; Mazzei, F.; Tasca, F.; Antiochia, R. Electrochemical Characterization of Graphene and MWCNT Screen-Printed Electrodes Modified with AuNPs for Laccase Biosensor Development. Nanomaterials 2015, 5, 1995–2006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Song, Y.; Luo, Y.; Zhu, C.; Li, H.; Du, D.; Lin, Y. Recent advances in electrochemical biosensors based on graphene two-dimensional nanomaterials. Biosens. Bioelectron. 2016, 76, 195–212. [Google Scholar] [CrossRef] [PubMed]
  186. Chen, Y.; Tan, C.; Zhang, H.; Wang, L. Two-dimensional graphene analogues for biomedical applications. Chem. Soc. Rev. 2015, 44, 2681–2701. [Google Scholar] [CrossRef]
  187. Tan, S.M.; Sofer, Z.; Pumera, M. Biomarkers Detection on Hydrogenated Graphene Surfaces: Towards Applications of Graphane in Biosensing. Electroanalysis 2013, 25, 703–705. [Google Scholar] [CrossRef]
  188. Yang, G.; Zhu, C.; Du, D.; Zhu, J.; Lin, Y. Graphene-like two-dimensional layered nanomaterials: Applications in biosensors and nanomedicine. Nanoscale 2015, 7, 14217–14231. [Google Scholar] [CrossRef] [PubMed]
  189. Gupta, A.; Sakthivel, T.; Seal, S. Recent development in 2D materials beyond graphene. Prog. Mater. Sci. 2015, 73, 44–126. [Google Scholar] [CrossRef]
  190. Lin, Y.; Connell, J.W. Advances in 2D boron nitride nanostructures: Nanosheets, nanoribbons, nanomeshes, and hybrids with graphene. Nanoscale 2012, 4, 6908–6939. [Google Scholar] [CrossRef] [PubMed]
  191. Zhu, C.; Du, D.; Lin, Y. Graphene and graphene-like 2D materials for optical biosensing and bioimaging: A review. 2D Mater. 2015, 2, 032004. [Google Scholar] [CrossRef]
  192. McCreery, R.L. Advanced Carbon Electrode Materials for Molecular Electrochemistry. Chem. Rev. 2008, 108, 2646–2687. [Google Scholar] [CrossRef] [PubMed]
  193. Takahashi, S.; Abiko, N.; Anzai, J.-I. Redox Response of Reduced Graphene Oxide-Modified Glassy Carbon Electrodes to Hydrogen Peroxide and Hydrazine. Materials 2013, 6, 1840–1850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Hamilton, C.E.; Lomeda, J.R.; Sun, Z.; Tour, J.M.; Barron, A.R. High-Yield Organic Dispersions of Unfunctionalized Graphene. Nano Lett. 2009, 9, 3460–3462. [Google Scholar] [CrossRef]
  195. Lv, W.; Guo, M.; Liang, M.-H.; Jin, F.-M.; Cui, L.; Zhi, L.; Yang, Q.-H. Graphene-DNA hybrids: Self-assembly and electrochemical detection performance. J. Mater. Chem. 2010, 20, 6668–6673. [Google Scholar] [CrossRef]
  196. Woo, S.; Kim, Y.-R.; Chung, T.D.; Piao, Y.; Kim, H. Synthesis of a graphene–carbon nanotube composite and its electrochemical sensing of hydrogen peroxide. Electrochim. Acta 2012, 59, 509–514. [Google Scholar] [CrossRef]
  197. Wang, Y.; Shao, Y.; Matson, D.W.; Li, J.; Lin, Y. Nitrogen-Doped Graphene and Its Application in Electrochemical Biosensing. ACS Nano 2010, 4, 1790–1798. [Google Scholar] [CrossRef]
  198. Li, M.; Wu, Z.-S.; Ren, W.; Cheng, H.-M.; Tang, N.; Wu, W.; Zhong, W.; Du, Y. The doping of reduced graphene oxide with nitrogen and its effect on the quenching of the material’s photoluminescence. Carbon 2012, 50, 5286–5291. [Google Scholar] [CrossRef]
  199. Yeh, M.-H.; Li, Y.-S.; Chen, G.-L.; Lin, L.-Y.; Li, T.-J.; Chuang, H.-M.; Hsieh, C.-Y.; Lo, S.-C.; Chiang, W.-H.; Ho, K.-C. Facile Synthesis of Boron-doped Graphene Nanosheets with Hierarchical Microstructure at Atmosphere Pressure for Metal-free Electrochemical Detection of Hydrogen Peroxide. Electrochim. Acta 2015, 172, 52–60. [Google Scholar] [CrossRef]
  200. Yang, G.-H.; Zhou, Y.-H.; Wu, J.-J.; Cao, J.-T.; Li, L.-L.; Liu, H.-Y.; Zhu, J.-J. Microwave-assisted synthesis of nitrogen and boron co-doped graphene and its application for enhanced electrochemical detection of hydrogen peroxide. RSC Adv. 2013, 3, 22597–22604. [Google Scholar] [CrossRef]
  201. Zor, E.; Saglam, M.E.; Akin, I.; Saf, A.O.; Bingol, H.; Ersoz, M. Green synthesis of reduced graphene oxide/nanopolypyrrole composite: Characterization and H2O2 determination in urine. RSC Adv. 2014, 4, 12457–12466. [Google Scholar] [CrossRef]
  202. Luo, J.; Chen, Y.; Ma, Q.; Liu, R.; Liu, X. Layer-by-layer assembled ionic-liquid functionalized graphene–polyaniline nanocomposite with enhanced electrochemical sensing properties. J. Mater. Chem. C 2014, 2, 4818–4827. [Google Scholar] [CrossRef]
  203. Wang, Q.; Li, M.; Szunerits, S.; Boukherroub, R. Environmentally Friendly Reduction of Graphene Oxide Using Tyrosine for Nonenzymatic Amperometric H2O2 Detection. Electroanalysis 2014, 26, 156–163. [Google Scholar] [CrossRef]
  204. Nguyen, V.H.; Tran, T.H.; Shim, J.-J. Glassy carbon electrode modified with a graphene oxide/poly(o-phenylenediamine) composite for the chemical detection of hydrogen peroxide. Mater. Sci. Eng. C 2014, 44, 144–150. [Google Scholar] [CrossRef] [PubMed]
  205. Huang, Y.; Li, S.F.Y. Electrocatalytic performance of silica nanoparticles on graphene oxide sheets for hydrogen peroxide sensing. J. Electroanal. Chem. 2013, 690, 8–12. [Google Scholar] [CrossRef]
  206. Kong, F.-Y.; Li, W.-W.; Wang, J.-Y.; Fang, H.-L.; Fan, D.-H.; Wang, W. Direct electrolytic exfoliation of graphite with hemin and single-walled carbon nanotube: Creating functional hybrid nanomaterial for hydrogen peroxide detection. Anal. Chim. Acta 2015, 884, 37–43. [Google Scholar] [CrossRef] [PubMed]
  207. Lei, W.; Wu, L.; Huang, W.; Hao, Q.; Zhang, Y.; Xia, X. Microwave-assisted synthesis of hemin–graphene/poly(3,4-ethylenedioxythiophene) nanocomposite for a biomimetic hydrogen peroxide biosensor. J. Mater. Chem. B 2014, 2, 4324–4330. [Google Scholar] [CrossRef]
  208. Zhang, T.; Gu, Y.; Li, C.; Yan, X.; Lu, N.; Liu, H.; Zhang, Z.; Zhang, H. Fabrication of Novel Electrochemical Biosensor Based on Graphene Nanohybrid to Detect H2O2 Released from Living Cells with Ultrahigh Performance. ACS Appl. Mater. Interfaces 2017, 9, 37991–37999. [Google Scholar] [CrossRef]
  209. Xi, F.; Zhao, D.; Wang, X.; Chen, P. Non-enzymatic detection of hydrogen peroxide using a functionalized three-dimensional graphene electrode. Electrochem. Commun. 2013, 26, 81–84. [Google Scholar] [CrossRef]
  210. Zhang, J.; Zhao, M.; Yang, J.; Wu, G.; Wu, H.; Chen, C.; Liu, A. Metal-free rGO/GO hybrid microelectrode array for sensitive and in-situ hydrogen peroxide sensing. Electrochim. Acta 2019, 326, 134967. [Google Scholar] [CrossRef]
  211. Tian, Y.; Wei, Z.; Zhang, K.; Peng, S.; Zhang, X.; Liu, W.; Chu, K. Three-dimensional phosphorus-doped graphene as an efficient metal-free electrocatalyst for electrochemical sensing. Sens. Actuators B Chem. 2017, 241, 584–591. [Google Scholar] [CrossRef] [Green Version]
  212. Radhakrishnan, S.; Kim, S.J. An enzymatic biosensor for hydrogen peroxide based on one-pot preparation of CeO2-reduced graphene oxide nanocomposite. RSC Adv. 2015, 5, 12937–12943. [Google Scholar] [CrossRef]
  213. Wang, S.; Zhu, Y.; Yang, X.; Li, C. Photoelectrochemical detection of H2O2 based on flower-like CuInS2-graphene hybrid. Electroanalysis 2014, 26, 573–580. [Google Scholar] [CrossRef]
  214. Song, H.; Ni, Y.; Kokot, S. Investigations of an electrochemical platform based on the layered MoS2–graphene and horseradish peroxidase nanocomposite for direct electrochemistry and electrocatalysis. Biosens. Bioelectron. 2014, 56, 137–143. [Google Scholar] [CrossRef]
  215. Liu, Y.; Liu, X.; Guo, Z.; Hu, Z.; Xue, Z.; Lu, X. Horseradish peroxidase supported on porous graphene as a novel sensing platform for detection of hydrogen peroxide in living cells sensitively. Biosens. Bioelectron. 2017, 87, 101–107. [Google Scholar] [CrossRef]
  216. Fan, Z.; Lin, Q.; Gong, P.; Liu, B.; Wang, J.; Yang, S. A new enzymatic immobilization carrier based on graphene capsule for hydrogen peroxide biosensors. Electrochim. Acta 2015, 151, 186–194. [Google Scholar] [CrossRef]
  217. Wu, P.; Cai, Z.; Chen, J.; Zhang, H.; Cai, C. Electrochemical measurement of the flux of hydrogen peroxide releasing from RAW 264.7 macrophage cells based on enzyme-attapulgite clay nanohybrids. Biosens. Bioelectron. 2011, 26, 4012–4017. [Google Scholar] [CrossRef] [PubMed]
  218. Wang, Y.; Zhang, H.; Yao, D.; Pu, J.; Zhang, Y.; Gao, X.; Sun, Y. Direct electrochemistry of hemoglobin on graphene/Fe3O4 nanocomposite-modified glass carbon electrode and its sensitive detection for hydrogen peroxide. J. Solid State Electrochem. 2013, 17, 881–887. [Google Scholar] [CrossRef] [Green Version]
  219. Cheng, Y.; Feng, B.; Yang, X.; Yang, P.; Ding, Y.; Chen, Y.; Fei, J. Electrochemical biosensing platform based on carboxymethyl cellulose functionalized reduced graphene oxide and hemoglobin hybrid nanocomposite film. Sens. Actuators B Chem. 2013, 182, 288–293. [Google Scholar] [CrossRef]
  220. Xie, L.; Xu, Y.; Cao, X. Hydrogen peroxide biosensor based on hemoglobin immobilized at graphene, flower-like zinc oxide, and gold nanoparticles nanocomposite modified glassy carbon electrode. Colloids Surf. B Biointerfaces 2013, 107, 245–250. [Google Scholar] [CrossRef] [PubMed]
  221. Li, M.; Xu, S.; Tang, M.; Liu, L.; Gao, F.; Wang, Y. Direct electrochemistry of horseradish peroxidase on graphene-modified electrode for electrocatalytic reduction towards H2O2. Electrochim. Acta 2011, 56, 1144–1149. [Google Scholar] [CrossRef]
  222. Zhang, L.; Han, G.; Liu, Y.; Tang, J.; Tang, W. Immobilizing haemoglobin on gold/graphene–chitosan nanocomposite as efficient hydrogen peroxide biosensor. Sens. Actuators B Chem. 2014, 197, 164–171. [Google Scholar] [CrossRef]
  223. Liu, H.; Su, X.; Duan, C.; Dong, X.; Zhou, S.; Zhu, Z. Microwave-assisted hydrothermal synthesis of Au NPs–Graphene composites for H2O2 detection. J. Electroanal. Chem. 2014, 731, 36–42. [Google Scholar] [CrossRef]
  224. Vilian, A.T.E.; Chen, S.-M. Simple approach for the immobilization of horseradish peroxidase on poly-l-histidine modified reduced graphene oxide for amperometric determination of dopamine and H2O2. RSC Adv. 2014, 4, 55867–55876. [Google Scholar] [CrossRef]
  225. Xiong, W.; Qu, Q.; Liu, S. Self-assembly of ultra-small gold nanoparticles on an indium tin oxide electrode for the enzyme-free detection of hydrogen peroxide. Mikrochim. Acta 2014, 181, 983–989. [Google Scholar] [CrossRef]
  226. Sheng, Q.; Wang, M.; Zheng, J. A novel hydrogen peroxide biosensor based on enzymatically induced deposition of polyaniline on the functionalized graphene–carbon nanotube hybrid materials. Sens. Actuators B Chem. 2011, 160, 1070–1077. [Google Scholar] [CrossRef]
  227. 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 2010, 55, 3055–3060. [Google Scholar] [CrossRef]
  228. Wang, T.; Liu, J.; Ren, J.; Wang, J.; Wang, E. Mimetic biomembrane–AuNPs–graphene hybrid as matrix for enzyme immobilization and bioelectrocatalysis study. Talanta 2015, 143, 438–441. [Google Scholar] [CrossRef] [PubMed]
  229. Nandini, S.; Manjunatha, R.; Shanmugam, S.; Melo, J.S.; Suresh, G.S. Electrochemical biosensor for the selective determination of hydrogen peroxide based on the co-deposition of palladium, horseradish peroxidase on functionalized-graphene modified graphite electrode as composite. J. Electroanal. Chem. 2013, 689, 233–242. [Google Scholar] [CrossRef]
  230. Nalini, S.; Shanmugam, S.; Neelagund, S.E.; Melo, J.S.; Suresh, G.S.; Nandini, S. Amperometric hydrogen peroxide and cholesterol biosensors designed by using hierarchical curtailed silver flowers functionalized graphene and enzymes deposits. J. Solid State Electrochem. 2014, 18, 685–701. [Google Scholar] [CrossRef]
  231. Huang, K.-J.; Niu, D.-J.; Liu, X.; Wu, Z.-W.; Fan, Y.; Chang, Y.-F.; Wu, Y.-Y. Direct electrochemistry of catalase at amine-functionalized graphene/gold nanoparticles composite film for hydrogen peroxide sensor. Electrochim. Acta 2011, 56, 2947–2953. [Google Scholar] [CrossRef]
  232. Dinesh, B.; Mani, V.; Saraswathi, R.; Chen, S.-M. Direct electrochemistry of cytochrome c immobilized on a graphene oxide–carbon nanotube composite for picomolar detection of hydrogen peroxide. RSC Adv. 2014, 4, 28229–28237. [Google Scholar] [CrossRef]
  233. Mani, V.; Dinesh, B.; Chen, S.-M.; Saraswathi, R. Direct electrochemistry of myoglobin at reduced graphene oxide-multiwalled carbon nanotubes-platinum nanoparticles nanocomposite and biosensing towards hydrogen peroxide and nitrite. Biosens. Bioelectron. 2014, 53, 420–427. [Google Scholar] [CrossRef]
  234. Liu, F.; Xu, Q.; Huang, W.; Zhang, Z.; Xiang, G.; Zhang, C.; Liang, C.; Lian, H.; Peng, J. Green synthesis of porous graphene and its application for sensitive detection of hydrogen peroxide and 2,4-dichlorophenoxyacetic acid. Electrochim. Acta 2019, 295, 615–623. [Google Scholar] [CrossRef]
  235. Ebrahimi, A.; Zhang, K.; Dong, C.; Subramanian, S.; Butler, D.; Bolotsky, A.; Goodnight, L.; Cheng, Y.; Robinson, J.A. FeSx-graphene heterostructures: Nanofabrication-compatible catalysts for ultra-sensitive electrochemical detection of hydrogen peroxide. Sens. Actuators B Chem. 2019, 285, 631–638. [Google Scholar] [CrossRef]
  236. Hu, Z.; Dai, Z.; Hu, X.; Yang, B.; Liu, Q.; Gao, C.; Zheng, X.; Yu, Y. A facile preparation of FePt-loaded few-layer MoS2 nanosheets nanocomposites (F-MoS2-FePt NCs) and their application for colorimetric detection of H2O2 in living cells. J. Nanobiotechnol. 2019, 17, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Dai, H.; Chen, Y.; Niu, X.; Pan, C.; Chen, H.; Chen, X. High-performance electrochemical biosensor for nonenzymatic H2O2 sensing based on Au@C-Co3O4 heterostructures. Biosens. Bioelectron. 2018, 118, 36–43. [Google Scholar] [CrossRef]
  238. Liu, J.; Bo, X.; Yang, J.; Yin, D.; Guo, L. One-step synthesis of porphyrinic iron-based metal-organic framework/ordered mesoporous carbon for electrochemical detection of hydrogen peroxide in living cells. Sens. Actuators B Chem. 2017, 248, 207–213. [Google Scholar] [CrossRef]
  239. Shu, Y.; Xu, J.; Chen, J.; Xu, Q.; Xiao, X.; Jin, D.; Pang, H.; Hu, X. Ultrasensitive electrochemical detection of H2O2 in living cells based on ultrathin MnO2 nanosheets. Sens. Actuators B Chem. 2017, 252, 72–78. [Google Scholar] [CrossRef]
  240. Ensafi, A.A.; Jafari–Asl, M.; Rezaei, B. A novel enzyme-free amperometric sensor for hydrogen peroxide based on Nafion/exfoliated graphene oxide–Co3O4 nanocomposite. Talanta 2013, 103, 322–329. [Google Scholar] [CrossRef]
  241. Sarkar, A.; Ghosh, A.B.; Saha, N.; Bhadu, G.R.; Adhikary, B. Newly Designed Amperometric Biosensor for Hydrogen Peroxide and Glucose Based on Vanadium Sulfide Nanoparticles. ACS Appl. Nano Mater. 2018, 1, 1339–1347. [Google Scholar] [CrossRef]
  242. 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 H2O2. Mikrochim. Acta 2014, 181, 631–638. [Google Scholar] [CrossRef]
  243. Zheng, L.; Ye, D.; Xiong, L.; Xu, J.; Tao, K.; Zou, Z.; Huang, D.; Kang, X.; Yang, S.; Xia, J. Preparation of cobalt-tetraphenylporphyrin/reduced graphene oxide nanocomposite and its application on hydrogen peroxide biosensor. Anal. Chim. Acta 2013, 768, 69–75. [Google Scholar] [CrossRef]
  244. Hosu, I.S.; Wang, Q.; Vasilescu, A.; Peteu, S.F.; Raditoiu, V.; Railian, S.; Zaitsev, V.; Turcheniuk, K.; Wang, Q.; Li, M.; et al. Cobalt phthalocyanine tetracarboxylic acid modified reduced graphene oxide: A sensitive matrix for the electrocatalytic detection of peroxynitrite and hydrogen peroxide. RSC Adv. 2015, 5, 1474–1484. [Google Scholar] [CrossRef]
  245. Liu, X.; Zhu, H.; Yang, X. An amperometric hydrogen peroxide chemical sensor based on graphene-Fe3O4 multilayer films modified ITO electrode. Talanta 2011, 87, 243–248. [Google Scholar] [CrossRef] [PubMed]
  246. Yang, S.; Li, G.; Wang, G.; Zhao, J.; Hu, M.; Qu, L. A novel nonenzymatic H2O2 sensor based on cobalt hexacyanoferrate nanoparticles and graphene composite modified electrode. Sens. Actuators B Chem. 2015, 208, 593–599. [Google Scholar] [CrossRef]
  247. Kubendhiran, S.; Thirumalraj, B.; Chen, S.-M.; Karuppiah, C. Electrochemical co-preparation of cobalt sulfide/reduced graphene oxide composite for electrocatalytic activity and determination of H2O2 in biological samples. J. Colloid Interface Sci. 2018, 509, 153–162. [Google Scholar] [CrossRef]
  248. Karimi, M.A.; Banifatemeh, F.; Hatefi-Mehrjardi, A.; Tavallali, H.; Eshaghia, Z.; Deilamy-Rad, G. A novel rapid synthesis of Fe2O3/graphene nanocomposite using ferrate (VI) and its application as a new kind of nanocomposite modified electrode as electrochemical sensor. Mater. Res. Bull. 2015, 70, 856–864. [Google Scholar] [CrossRef]
  249. Li, Z.; Zheng, X.; Zheng, J. A non-enzymatic sensor based on Au@Ag nanoparticles with good stability for sensitive detection of H2O2. New J. Chem. 2016, 40, 2115–2120. [Google Scholar] [CrossRef]
  250. Zhu, S.; Guo, J.; Dong, J.; Cui, Z.; Lu, T.; Zhu, C.; Zhang, D.; Ma, J. Sonochemical fabrication of Fe3O4 nanoparticles on reduced graphene oxide for biosensors. Ultrason. Sonochem. 2013, 20, 872–880. [Google Scholar] [CrossRef]
  251. Zhang, P.; Huang, Y.; Lu, X.; Zhang, S.; Li, J.; Wei, G.; Su, Z. One-Step Synthesis of Large-Scale Graphene Film Doped with Gold Nanoparticles at Liquid–Air Interface for Electrochemistry and Raman Detection Applications. Langmuir 2014, 30, 8980–8989. [Google Scholar] [CrossRef] [PubMed]
  252. Ye, Y.; Kong, T.; Yu, X.; Wu, Y.; Zhang, K.; Wang, X. Enhanced nonenzymatic hydrogen peroxide sensing with reduced graphene oxide/ferroferric oxide nanocomposites. Talanta 2012, 89, 417–421. [Google Scholar] [CrossRef]
  253. Yang, X.; Wang, L.; Zhou, G.; Sui, N.; Gu, Y.; Wan, J. Electrochemical Detection of H2O2 Based on Fe3O4 Nanoparticles with Graphene Oxide and Polyamidoamine Dendrimer. J. Clust. Sci. 2014, 26, 789–798. [Google Scholar] [CrossRef]
  254. Zhu, M.; Li, N.; Ye, J. Sensitive and Selective Sensing of Hydrogen Peroxide with Iron-Tetrasulfophthalocyanine-Graphene-Nafion Modified Screen-Printed Electrode. Electroanalysis 2012, 24, 1212–1219. [Google Scholar] [CrossRef]
  255. Palanisamy, S.; Chen, S.-M.; Sarawathi, R. A novel nonenzymatic hydrogen peroxide sensor based on reduced graphene oxide/ZnO composite modified electrode. Sens. Actuators B Chem. 2012, 166–167, 372–377. [Google Scholar] [CrossRef]
  256. Jiang, B.-B.; Wei, X.-W.; Wu, F.-H.; Wu, K.-L.; Chen, L.; Yuan, G.-Z.; Dong, C.; Ye, Y. A non-enzymatic hydrogen peroxide sensor based on a glassy carbon electrode modified with cuprous oxide and nitrogen-doped graphene in a nafion matrix. Mikrochim. Acta 2014, 181, 1463–1470. [Google Scholar] [CrossRef]
  257. Xu, F.; Deng, M.; Li, G.; Chen, S.; Wang, L. Electrochemical behavior of cuprous oxide–reduced graphene oxide nanocomposites and their application in nonenzymatic hydrogen peroxide sensing. Electrochim. Acta 2013, 88, 59–65. [Google Scholar] [CrossRef]
  258. Liu, M.; Liu, R.; Chen, W. Graphene wrapped Cu2O nanocubes: Non-enzymatic electrochemical sensors for the detection of glucose and hydrogen peroxide with enhanced stability. Biosens. Bioelectron. 2013, 45, 206–212. [Google Scholar] [CrossRef] [PubMed]
  259. Li, L.; Du, Z.; Liu, S.; Hao, Q.; Wang, Y.; Li, Q.; Wang, T. A novel nonenzymatic hydrogen peroxide sensor based on MnO2/graphene oxide nanocomposite. Talanta 2010, 82, 1637–1641. [Google Scholar] [CrossRef] [PubMed]
  260. Dong, S.; Xi, J.; Wu, Y.; Liu, H.; Fu, C.; Liu, H.; Xiao, F. High loading MnO2 nanowires on graphene paper: Facile electrochemical synthesis and use as flexible electrode for tracking hydrogen peroxide secretion in live cells. Anal. Chim. Acta 2015, 853, 200–206. [Google Scholar] [CrossRef] [PubMed]
  261. Feng, X.; Zhang, Y.; Song, J.; Chen, N.; Zhou, J.; Huang, Z.; Ma, Y.; Zhang, L.; Wang, L. MnO2/graphene nanocomposites for nonenzymatic electrochemical detection of hydrogen peroxide. Electroanalysis 2015, 27, 353–359. [Google Scholar] [CrossRef]
  262. Hassan, M.; Jiang, Y.; Bo, X.; Zhou, M. Sensitive nonenzymatic detection of hydrogen peroxide at nitrogen-doped graphene supported-CoFe nanoparticles. Talanta 2018, 188, 339–348. [Google Scholar] [CrossRef] [PubMed]
  263. Lu, N.; Zheng, B.; Gu, Y.; Yan, X.; Zhang, T.; Liu, H.; Xu, H.; Xu, Z.; Li, X.; Zhang, Z. Fabrication of CoNPs-embedded porous carbon composites based on morphochemical imprinting strategy for detection of H2O2 released from living cells. Electrochim. Acta 2019, 321, 134717. [Google Scholar] [CrossRef]
  264. Benvidi, A.; Nafar, M.T.; Jahanbani, S.; Tezerjani, M.D.; Rezaeinasab, M.; Dalirnasab, S. Developing an electrochemical sensor based on a carbon paste electrode modified with nano-composite of reduced graphene oxide and CuFe2O4 nanoparticles for determination of hydrogen peroxide. Mater. Sci. Eng. C 2017, 75, 1435–1447. [Google Scholar] [CrossRef] [PubMed]
  265. Dang, W.; Sun, Y.; Jiao, H.; Xu, L.; Lin, M. AuNPs-NH2/Cu-MOF modified glassy carbon electrode as enzyme-free electrochemical sensor detecting H2O2. J. Electroanal. Chem. 2020, 856, 113592. [Google Scholar] [CrossRef]
  266. Wang, W.; Tang, H.; Wu, Y.; Zhang, Y.; Li, Z. Highly electrocatalytic biosensor based on Hemin@ AuNPs/reduced graphene oxide/chitosan nanohybrids for non-enzymatic ultrasensitive detection of hydrogen peroxide in living cells. Biosens. Bioelectron. 2019, 132, 217–223. [Google Scholar] [CrossRef]
  267. Sun, Y.; Luo, M.; Meng, X.; Xiang, J.; Wang, L.; Ren, Q.; Guo, S. Graphene/Intermetallic PtPb Nanoplates Composites for Boosting Electrochemical Detection of H2O2 Released from Cells. Anal. Chem. 2017, 89, 3761–3767. [Google Scholar] [CrossRef] [PubMed]
  268. Zhang, Y.; Liu, Y.; He, J.; Pang, P.; Gao, Y.; Hu, Q. Electrochemical behavior of graphene/Nafion/Azure I/Au nanoparticles composites modified glass carbon electrode and its application as nonenzymatic hydrogen peroxide sensor. Electrochim. Acta 2013, 90, 550–555. [Google Scholar] [CrossRef]
  269. Wang, L.; Dong, Y.; Zhang, Y.; Zhang, Z.; Chi, K.; Yuan, H.; Zhao, A.; Ren, J.; Xiao, F.; Wang, S. PtAu alloy nanoflowers on 3D porous ionic liquid functionalized graphene-wrapped activated carbon fiber as a flexible microelectrode for near-cell detection of cancer. NPG Asia Mater. 2016, 8, e337. [Google Scholar] [CrossRef] [Green Version]
  270. Yuan, B.; Xu, C.; Liu, L.; Shi, Y.; Li, S.; Zhang, R.; Zhang, D. Polyethylenimine-bridged graphene oxide–gold film on glassy carbon electrode and its electrocatalytic activity toward nitrite and hydrogen peroxide. Sens. Actuators B 2014, 198, 55–61. [Google Scholar] [CrossRef]
  271. Lorestani, F.; Shahnavaz, Z.; Mn, P.; Alias, Y.; Manan, N.S.A. One-step hydrothermal green synthesis of silver nanoparticle-carbon nanotube reduced-graphene oxide composite and its application as hydrogen peroxide sensor. Sens. Actuators B Chem. 2015, 208, 389–398. [Google Scholar] [CrossRef]
  272. Chang, H.-C.; Wang, X.; Shiu, K.-K.; Zhu, Y.; Wang, J.; Li, Q.; Chen, B.; Jiang, H. Layer-by-layer assembly of graphene, Au and poly (toluidine blue O) films sensor for evaluation of oxidative stress of tumor cells elicited by hydrogen peroxide. Biosens. Bioelectron. 2013, 41, 789–794. [Google Scholar] [CrossRef] [PubMed]
  273. Bai, X.; Shiu, K.-K. Investigation of the optimal weight contents of reduced graphene oxide–gold nanoparticles composites and theirs application in electrochemical biosensors. J. Electroanal. Chem. 2014, 720–721, 84–91. [Google Scholar] [CrossRef]
  274. Liu, R.; Li, S.; Zhang, G.; Dolbecq, A.; Mialane, P.; Keita, B. Polyoxometalate-Mediated Green Synthesis of Graphene and Metal Nanohybrids: High-Performance Electrocatalysts. J. Clust. Sci. 2014, 25, 711–740. [Google Scholar] [CrossRef]
  275. Yu, B.; Feng, J.; Liu, S.; Zhang, T. Preparation of reduced graphene oxide decorated with high density Ag nanorods for non-enzymatic hydrogen peroxide detection. RSC Adv. 2013, 3, 14303–14307. [Google Scholar] [CrossRef]
  276. Fang, Y.; Guo, S.; Zhu, C.; Zhai, Y.; Wang, E. Self-Assembly of Cationic Polyelectrolyte-Functionalized Graphene Nanosheets and Gold Nanoparticles: A Two-Dimensional Heterostructure for Hydrogen Peroxide Sensing. Langmuir 2010, 26, 11277–11282. [Google Scholar] [CrossRef] [PubMed]
  277. Xu, F.; Sun, Y.; Zhang, Y.; Shi, Y.; Wen, Z.; Li, Z. Graphene–Pt nanocomposite for nonenzymatic detection of hydrogen peroxide with enhanced sensitivity. Electrochem. Commun. 2011, 13, 1131–1134. [Google Scholar] [CrossRef]
  278. Gao, C.-H.; Zhu, X.-Z.; Zhang, L.; Zhou, D.-Y.; Wang, Z.-K.; Liao, L.-S. Comparative studies on the inorganic and organic p-type dopants in organic light-emitting diodes with enhanced hole injection. Appl. Phys. Lett. 2013, 102, 153301. [Google Scholar] [CrossRef]
  279. Pang, P.; Yang, Z.; Xiao, S.; Xie, J.; Zhang, Y.; Gao, Y. Nonenzymatic amperometric determination of hydrogen peroxide by graphene and gold nanorods nanocomposite modified electrode. J. Electroanal. Chem. 2014, 727, 27–33. [Google Scholar] [CrossRef]
  280. Li, X.-R.; Xu, M.-C.; Chen, H.-Y.; Xu, J.-J. Bimetallic Au@Pt@Au core–shell nanoparticles on graphene oxide nanosheets for high-performance H2O2 bi-directional sensing. J. Mater. Chem. B 2015, 3, 4355–4362. [Google Scholar] [CrossRef]
  281. Yao, L.; Yan, Y.; Lee, J.-M. Synthesis of Porous Pd Nanostructure and Its Application in Enzyme-Free Sensor of Hydrogen Peroxide. ACS Sustain. Chem. Eng. 2017, 5, 1248–1252. [Google Scholar] [CrossRef]
  282. Fu, L.; Lai, G.; Jia, B.; Yu, A. Preparation and Electrocatalytic Properties of Polydopamine Functionalized Reduced Graphene Oxide-Silver Nanocomposites. Electrocatalysis 2014, 6, 72–76. [Google Scholar] [CrossRef]
  283. Liu, S.; Tian, J.; Wang, L.; Sun, X. Microwave-assisted rapid synthesis of Ag nanoparticles/graphene nanosheet composites and their application for hydrogen peroxide detection. J. Nanopart. Res. 2011, 13, 4539–4548. [Google Scholar] [CrossRef]
  284. Shan, L.; Liu, H.; Wang, G. Preparation of tungsten-doped BiVO4 and enhanced photocatalytic activity. J. Nanopart. Res. 2015, 17, 181. [Google Scholar] [CrossRef]
  285. Liu, J.; Bo, X.; Zhao, Z.; Guo, L. Highly exposed Pt nanoparticles supported on porous graphene for electrochemical detection of hydrogen peroxide in living cells. Biosens. Bioelectron. 2015, 74, 71–77. [Google Scholar] [CrossRef]
  286. Zhang, C.; Zhang, Y.; Miao, Z.; Ma, M.; Du, X.; Lin, J.; Han, B.; Takahashi, S.; Anzai, J.-I.; Chen, Q. Dual-function amperometric sensors based on poly (diallydimethylammoniun chloride)-functionalized reduced graphene oxide/manganese dioxide/gold nanoparticles nanocomposite. Sens. Actuators B Chem. 2016, 222, 663–673. [Google Scholar] [CrossRef]
  287. Liu, S.; Wang, L.; Tian, J.; Luo, Y.; Zhang, X.; Sun, X. Aniline as a dispersing and stabilizing agent for reduced graphene oxide and its subsequent decoration with Ag nanoparticles for enzymeless hydrogen peroxide detection. J. Colloid Interface Sci. 2011, 363, 615–619. [Google Scholar] [CrossRef] [PubMed]
  288. Zhu, J.; Kim, K.; Liu, Z.; Feng, H.; Hou, S. Electroless Deposition of Silver Nanoparticles on Graphene Oxide Surface and Its Applications for the Detection of Hydrogen Peroxide. Electroanalysis 2014, 26, 2513–2519. [Google Scholar] [CrossRef]
  289. Zhang, P.; Zhang, X.; Zhang, S.; Lu, X.; Li, Q.; Su, Z.; Wei, G. One-pot green synthesis, characterizations, and biosensor application of self-assembled reduced graphene oxide–gold nanoparticle hybrid membranes. J. Mater. Chem. B 2013, 1, 6525–6531. [Google Scholar] [CrossRef]
  290. Li, S.-J.; Shi, Y.-F.; Liu, L.; Song, L.-X.; Pang, H.; Du, J.-M. Electrostatic self-assembly for preparation of sulfonated graphene/gold nanoparticle hybrids and their application for hydrogen peroxide sensing. Electrochim. Acta 2012, 85, 628–635. [Google Scholar] [CrossRef]
  291. Zhang, Y.; Wang, Z.; Ji, Y.; Liu, S. Synthesis of Ag nanoparticle–carbon nanotube–reduced graphene oxide hybrids for highly sensitive non-enzymatic hydrogen peroxide detection. RSC Adv. 2015, 5, 39037–39041. [Google Scholar] [CrossRef]
  292. Nia, P.M.; Lorestani, F.; Meng, W.P.; Alias, Y. A novel non-enzymatic H2O2 sensor based on polypyrrole nanofibers–silver nanoparticles decorated reduced graphene oxide nano composites. Appl. Surf. Sci. 2015, 332, 648–656. [Google Scholar]
  293. Lu, D.; Zhang, Y.; Lin, S.; Wang, L.; Wang, C. Synthesis of PtAu bimetallic nanoparticles on graphene–carbon nanotube hybrid nanomaterials for nonenzymatic hydrogen peroxide sensor. Talanta 2013, 112, 111–116. [Google Scholar] [CrossRef] [PubMed]
  294. Liu, P.; Li, J.; Liu, X.; Li, M.; Lu, X. One-pot synthesis of highly dispersed PtAu nanoparticles–CTAB–graphene nanocomposites for nonenzyme hydrogen peroxide sensor. J. Electroanal. Chem. 2015, 751, 1–6. [Google Scholar] [CrossRef]
  295. Cui, X.; Wu, S.; Li, Y.; Wan, G. Sensing hydrogen peroxide using a glassy carbon electrode modified with in-situ electrodeposited platinum-gold bimetallic nanoclusters on a graphene surface. Mikrochim. Acta 2014, 182, 265–272. [Google Scholar] [CrossRef]
  296. Sun, Y.; He, K.; Zhang, Z.; Zhou, A.; Duan, H. Real-time electrochemical detection of hydrogen peroxide secretion in live cells by Pt nanoparticles decorated graphene–carbon nanotube hybrid paper electrode. Biosens. Bioelectron. 2015, 68, 358–364. [Google Scholar] [CrossRef]
  297. Xu, C.; Zhang, L.; Liu, L.; Shi, Y.; Wang, H.; Wang, X.; Wang, F.; Yuan, B.; Zhang, D. A novel enzyme-free hydrogen peroxide sensor based on polyethylenimine-grafted graphene oxide-Pd particles modified electrode. J. Electroanal. Chem. 2014, 731, 67–71. [Google Scholar] [CrossRef]
  298. Liu, H.; Chen, X.; Huang, L.; Wang, J.; Pan, H. Palladium Nanoparticles Embedded into Graphene Nanosheets: Preparation, Characterization, and Nonenzymatic Electrochemical Detection of H2O2. Electroanalysis 2014, 26, 556–564. [Google Scholar] [CrossRef]
  299. Chen, X.-M.; Cai, Z.-X.; Huang, Z.-Y.; Oyama, M.; Jiang, Y.-Q.; Chen, X. Ultrafine palladium nanoparticles grown on graphene nanosheets for enhanced electrochemical sensing of hydrogen peroxide. Electrochim. Acta 2013, 97, 398–403. [Google Scholar] [CrossRef]
  300. Sun, W.; Cai, X.; Wang, Z.; Zhao, H.; Lan, M. A novel modification method via in-situ reduction of AuAg bimetallic nanoparticles by polydopamine on carbon fiber microelectrode for H2O2 detection. Microchem. J. 2020, 154, 104595. [Google Scholar] [CrossRef]
  301. Xie, Y.; Zhan, Y. Electrochemical capacitance of porous reduced graphene oxide/nickel foam. J. Porous Mater. 2015, 22, 403–412. [Google Scholar] [CrossRef]
  302. Zhao, Y.; Hu, Y.; Hou, J.; Jia, Z.; Zhong, D.; Zhou, S.; Huo, D.; Yang, M.; Hou, C. Electrochemical biointerface based on electrodeposition AuNPs on 3D graphene aerogel: Direct electron transfer of Cytochrome c and hydrogen peroxide sensing. J. Electroanal. Chem. 2019, 842, 16–23. [Google Scholar] [CrossRef]
  303. Fu, Y.; Huang, D.; Li, C.; Zou, L.; Ye, B. Graphene blended with SnO2 and Pd-Pt nanocages for sensitive non-enzymatic electrochemical detection of H2O2 released from living cells. Anal. Chim. Acta 2018, 1014, 10–18. [Google Scholar] [CrossRef] [PubMed]
  304. Lu, J.; Hu, Y.; Wang, P.; Liu, P.; Chen, Z.; Sun, D. Electrochemical biosensor based on gold nanoflowers-encapsulated magnetic metal-organic framework nanozymes for drug evaluation with in-situ monitoring of H2O2 released from H9C2 cardiac cells. Sens. Actuators B Chem. 2020, 311, 127909. [Google Scholar] [CrossRef]
  305. Hu, J.; Wisetsuwannaphum, S.; Foord, J.S. Glutamate biosensors based on diamond and graphene platforms. Faraday Discuss. 2014, 172, 457–472. [Google Scholar] [CrossRef] [PubMed]
  306. Chang, Q.; Tang, H. Optical determination of glucose and hydrogen peroxide using a nanocomposite prepared from glucose oxidase and magnetite nanoparticles immobilized on graphene oxide. Mikrochim. Acta 2014, 181, 527–534. [Google Scholar] [CrossRef]
  307. Zhang, J.; Zhang, F.; Yang, H.; Huang, X.; Liu, H.; Zhang, J.; Guo, S. Graphene Oxide as a Matrix for Enzyme Immobilization. Langmuir 2010, 26, 6083–6085. [Google Scholar] [CrossRef]
  308. Dresselhaus, M.; Dresselhaus, G.; Jorio, A. Unusual properties and structure of carbon nanotubes. Annu. Rev. Mater. Res. 2004, 34, 247–278. [Google Scholar] [CrossRef]
  309. Primo, E.N.; Gutiérrez, F.; Rubianes, M.D.; Ferreyra, N.F.; Rodríguez, M.C.; Pedano, M.L.; Gasnier, A.; Gutierrez, A.; Eguílaz, M.; Dalmasso, P.; et al. Electrochemistry in one dimension: Applications of carbon nanotubes. Electrochem. Carbon Electrodes 2015, 83–120. [Google Scholar] [CrossRef]
  310. Yu, D.; Wang, P.; Zhao, Y.; Fan, A. Iodophenol blue-enhanced luminol chemiluminescence and its application to hydrogen peroxide and glucose detection. Talanta 2016, 146, 655–661. [Google Scholar] [CrossRef] [PubMed]
  311. Georgakilas, V.; Perman, J.A.; Tucek, J.; Zboril, R. Broad Family of Carbon Nanoallotropes: Classification, Chemistry, and Applications of Fullerenes, Carbon Dots, Nanotubes, Graphene, Nanodiamonds, and Combined Superstructures. Chem. Rev. 2015, 115, 4744–4822. [Google Scholar] [CrossRef]
  312. Yáñez-Sedeño, P.; González-Cortés, A.; Agüí, L.; Pingarrón, J.M. Uncommon carbon nanostructures for the preparation of electrochemical immunosensors. Electroanalysis 2016, 28, 1679–1691. [Google Scholar] [CrossRef]
  313. Jacobs, C.B.; Peairs, M.J.; Venton, B.J. Review: Carbon nanotube based electrochemical sensors for biomolecules. Anal. Chim. Acta 2010, 662, 105–127. [Google Scholar] [CrossRef] [PubMed]
  314. Yang, J.; Xu, Y.; Zhang, R.; Wang, Y.; He, P.; Fang, Y. Direct Electrochemistry and Electrocatalysis of the Hemoglobin Immobilized on Diazonium-Functionalized Aligned Carbon Nanotubes Electrode. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 2009, 21, 1672–1677. [Google Scholar] [CrossRef]
  315. Esplandiu, M.J.; Pacios, M.; Cyganek, L.; Bartroli, J.; Del Valle, M. Enhancing the electrochemical response of myoglobin with carbon nanotube electrodes. Nanotechnology 2009, 20, 355502. [Google Scholar] [CrossRef]
  316. Zhang, H.; Ruan, J.; Liu, W.; Jiang, X.; Du, T.; Jiang, H.; Alberto, P.; Gottschalk, K.-E.; Wang, X. Monitoring dynamic release of intracellular hydrogen peroxide through a microelectrode based enzymatic biosensor. Anal. Bioanal. Chem. 2018, 410, 4509–4517. [Google Scholar] [CrossRef] [PubMed]
  317. Wang, S.; Xie, F.; Liu, G. Direct electrochemistry and electrocatalysis of heme proteins on SWCNTs-CTAB modified electrodes. Talanta 2009, 77, 1343–1350. [Google Scholar] [CrossRef]
  318. Chen, L.; Lu, G. Novel amperometric biosensor based on composite film assembled by polyelectrolyte-surfactant polymer, carbon nanotubes and hemoglobin. Sens. Actuators B Chem. 2007, 121, 423–429. [Google Scholar] [CrossRef]
  319. Zong, S.; Cao, Y.; Ju, H. Amperometric biosensor for hydrogen peroxide based on myoglobin doped multiwalled carbon nanotube enhanced grafted collagen matrix. Anal. Lett. 2007, 40, 1556–1568. [Google Scholar] [CrossRef]
  320. Nagaraju, D.; Pandey, R.K.; Lakshminarayanan, V. Electrocatalytic studies of Cytochrome c functionalized single walled carbon nanotubes on self-assembled monolayer of 4-ATP on gold. J. Electroanal. Chem. 2009, 627, 63–68. [Google Scholar] [CrossRef]
  321. Chen, S.; Yuan, R.; Chai, Y.; Yin, B.; Xu, Y. Multilayer Assembly of Hemoglobin and Colloidal Gold Nanoparticles on Multiwall Carbon Nanotubes/Chitosan Composite for Detecting Hydrogen Peroxide. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 2008, 20, 2141–2147. [Google Scholar] [CrossRef]
  322. Ren, Q.-Q.; Wu, J.; Zhang, W.-C.; Wang, C.; Qin, X.; Liu, G.-C.; Li, Z.-X.; Yu, Y. Real-time in vitro detection of cellular H2O2 under camptothecin stress using horseradish peroxidase, ionic liquid, and carbon nanotube-modified carbon fiber ultramicroelectrode. Sens. Actuators B Chem. 2017, 245, 615–621. [Google Scholar] [CrossRef]
  323. Pandey, R.R.; Guo, Y.; Gao, Y.; Chusuei, C.C. A Prussian blue ZnO carbon nanotube composite for chronoamperometrically assaying H2O2 in BT20 and 4T1 breast cancer cells. Anal. Chem. 2019, 91, 10573–10581. [Google Scholar] [CrossRef] [PubMed]
  324. Bai, J.; Sun, C.; Jiang, X. Carbon dots-decorated multiwalled carbon nanotubes nanocomposites as a high-performance electrochemical sensor for detection of H2O2 in living cells. Anal. Bioanal. Chem. 2016, 408, 4705–4714. [Google Scholar] [CrossRef]
  325. Sahraoui, Y.; Chaliaa, S.; Maaref, A.; Haddad, A.; Bessueille, F.; Jaffrezic-Renault, N. Synergistic effect of polyoxometalate and single walled carbon nanotubes on peroxidase-like mimics and highly sensitive electrochemical detection of hydrogen peroxide. Electroanalysis 2020, 32, 683–689. [Google Scholar] [CrossRef]
  326. Liu, J.-X.; Ding, S.-N. Non-enzymatic amperometric determination of cellular hydrogen peroxide using dendrimer-encapsulated Pt nanoclusters/carbon nanotubes hybrid composites modified glassy carbon electrode. Sens. Actuators B Chem. 2017, 251, 200–207. [Google Scholar] [CrossRef]
  327. Roushani, M.; Bakyas, K.; Dizajdizi, B.Z. Development of sensitive amperometric hydrogen peroxide sensor using a CuNPs/MB/MWCNT-C60-Cs-IL nanocomposite modified glassy carbon electrode. Mater. Sci. Eng. C 2016, 64, 54–60. [Google Scholar] [CrossRef] [PubMed]
  328. Zhao, P.; Zhao, Y.; Jiang, L.; Chen, S.; Ji, Z.; Hou, C.; Huo, D.; Yang, M. 3D carbon nanotubes spaced graphene aerogel incorporated with prussian blue nanoparticles for real-time detection of H2O2 released from living cells. J. Electrochem. Soc. 2020, 167, 047511. [Google Scholar] [CrossRef]
  329. Zhang, Y.; Xiao, J.; Sun, Y.; Wang, L.; Dong, X.; Ren, J.; He, W.; Xiao, F. Flexible nanohybrid microelectrode based on carbon fiber wrapped by gold nanoparticles decorated nitrogen doped carbon nanotube arrays: In situ electrochemical detection in live cancer cells. Biosens. Bioelectron. 2018, 100, 453–461. [Google Scholar] [CrossRef] [PubMed]
  330. Guo, X.; Cao, Q.; Liu, Y.; He, T.; Liu, J.; Huang, S.; Tang, H.; Ma, M. Organic electrochemical transistor for in situ detection of H2O2 released from adherent cells and its application in evaluating the in vitro cytotoxicity of nanomaterial. Anal. Chem. 2019, 92, 908–915. [Google Scholar] [CrossRef]
  331. Tavakkoli, H.; Akhond, M.; Ghorbankhani, G.A.; Absalan, G. Electrochemical sensing of hydrogen peroxide using a glassy carbon electrode modified with multiwalled carbon nanotubes and zein nanoparticle composites: Application to HepG2 cancer cell detection. Mikrochim. Acta 2020, 187, 105. [Google Scholar] [CrossRef]
  332. Wayu, M.B.; Spidle, R.T.; Devkota, T.; Deb, A.K.; Delong, R.K.; Ghosh, K.C.; Wanekaya, A.K.; Chusuei, C.C. Morphology of hydrothermally synthesized ZnO nanoparticles tethered to carbon nanotubes affects electrocatalytic activity for H2O2 detection. Electrochim. Acta 2013, 97, 99–104. [Google Scholar] [CrossRef] [Green Version]
  333. Goran, J.M.; Phan, E.N.H.; Favela, C.A.; Stevenson, K.J. H2O2 Detection at carbon nanotubes and nitrogen-doped carbon nanotubes: Oxidation, reduction, or disproportionation? Anal. Chem. 2015, 87, 5989–5996. [Google Scholar] [CrossRef] [PubMed]
  334. Li, X.; Li, H.; Liu, T.; Hei, Y.; Hassan, M.; Zhang, S.; Lin, J.; Wang, T.; Bo, X.; Wang, H.-L.; et al. The biomass of ground cherry husks derived carbon nanoplates for electrochemical sensing. Sens. Actuators B Chem. 2018, 255, 3248–3256. [Google Scholar] [CrossRef]
  335. Wu, Z.-L.; Li, C.-K.; Yu, J.-G.; Chen, X.-Q. MnO2/reduced graphene oxide nanoribbons: Facile hydrothermal preparation and their application in amperometric detection of hydrogen peroxide. Sens. Actuators B Chem. 2017, 239, 544–552. [Google Scholar] [CrossRef]
  336. Zhang, M.; Zheng, J.; Wang, J.; Xu, J.; Hayat, T.; Alharbi, N.S. Direct electrochemistry of cytochrome c immobilized on one dimensional Au nanoparticles functionalized magnetic N-doped carbon nanotubes and its application for the detection of H2O2. Sens. Actuators B Chem. 2019, 282, 85–95. [Google Scholar] [CrossRef]
  337. Chen, S.; Yuan, R.; Chai, Y.; Hu, F. Electrochemical sensing of hydrogen peroxide using metal nanoparticles: A review. Mikrochim. Acta 2013, 180, 15–32. [Google Scholar] [CrossRef]
  338. Chou, T.-C.; Wu, K.-Y.; Hsu, F.-X.; Lee, C.-K. Pt-MWCNT modified carbon electrode strip for rapid and quantitative detection of H2O2 in food. J. Food Drug Anal. 2018, 26, 662–669. [Google Scholar] [CrossRef] [Green Version]
  339. Tabrizi, M.A.; Shamsipur, M.; Saber, R.; Sarkar, S. Flow injection amperometric sandwich-type aptasensor for the determination of human leukemic lymphoblast cancer cells using MWCNTs-Pdnano/PTCA/aptamer as labeled aptamer for the signal amplification. Anal. Chim. Acta 2017, 985, 61–68. [Google Scholar] [CrossRef]
  340. Werner, E. Determination of cellular H2O2 production. Sci. Signal. 2003, 2003, PL3. [Google Scholar] [CrossRef]
  341. Ma, Z.; Jiang, M.; Zhu, Q.; Luo, Y.; Chen, G.; Pan, M.; Xie, T.; Huang, X.; Chen, D. A porous hollow fiber sensor for detection of cellular hydrogen peroxide release based on cell-in-lumen configuration. Sens. Actuators B Chem. 2020, 321, 128516. [Google Scholar] [CrossRef]
  342. Zhao, Y.; Huo, D.; Bao, J.; Yang, M.; Chen, M.; Hou, J.; Fa, H.; Hou, C. Biosensor based on 3D graphene-supported Fe3O4 quantum dots as biomimetic enzyme for in situ detection of H2O2 released from living cells. Sens. Actuators B Chem. 2017, 244, 1037–1044. [Google Scholar] [CrossRef]
  343. Xi, J.; Zhang, Y.; Wang, Q.; Xiao, J.; Chi, K.; Duan, X.; Chen, J.; Tang, C.; Sun, Y.; Xiao, F.; et al. Multi-element doping design of high-efficient carbocatalyst for electrochemical sensing of cancer cells. Sens. Actuators B Chem. 2018, 273, 108–117. [Google Scholar] [CrossRef]
  344. Zhao, A.; She, J.; Manoj, D.; Wang, T.; Sun, Y.; Zhang, Y.; Xiao, F. Functionalized graphene fiber modified by dual nanoenzyme: Towards high-performance flexible nanohybrid microelectrode for electrochemical sensing in live cancer cells. Sens. Actuators B Chem. 2020, 310, 127861. [Google Scholar] [CrossRef]
  345. Yang, Y.; Ohnoutek, L.; Ajmal, S.; Zheng, X.; Feng, Y.; Li, K.; Wang, T.; Deng, Y.; Liu, Y.; Xu, D.; et al. “Hot edges” in an inverse opal structure enable efficient CO2 electrochemical reduction and sensitive in situ Raman characterization. J. Mater. Chem. A 2019, 7, 11836–11846. [Google Scholar] [CrossRef]
  346. Sinha, A.; Dhanjai; Zhao, H.; Huang, Y.; Lu, X.; Chen, J.; Jain, R. MXene: An emerging material for sensing and biosensing. TrAC Trends Anal. Chem. 2018, 105, 424–435. [Google Scholar] [CrossRef]
  347. Chen, J.; Tong, P.; Huang, L.; Yu, Z.; Tang, D. Ti3C2 MXene nanosheet-based capacitance immunoassay with tyramine-enzyme repeats to detect prostate-specific antigen on interdigitated micro-comb electrode. Electrochim. Acta 2019, 319, 375–381. [Google Scholar] [CrossRef]
  348. Chen, X.; Sun, X.; Xu, W.; Pan, G.; Zhou, D.; Zhu, J.; Wang, H.; Bai, X.; Dong, B.; Song, H. Ratiometric photoluminescence sensing based on Ti3C2MXene quantum dots as an intracellular pH sensor. Nanoscale 2018, 10, 1111–1118. [Google Scholar] [CrossRef]
  349. Lin, H.; Chen, Y.; Shi, J. Insights into 2D MXenes for versatile biomedical applications: Current advances and challenges ahead. Adv. Sci. 2018, 5, 1800518. [Google Scholar] [CrossRef] [Green Version]
  350. Dai, C.; Lin, H.; Xu, G.; Liu, Z.; Wu, R.; Chen, Y. Biocompatible 2D titanium carbide (MXenes) composite nanosheets for pH-responsive MRI-guided tumor hyperthermia. Chem. Mater. 2017, 29, 8637–8652. [Google Scholar] [CrossRef]
  351. Liu, J.; Jiang, X.; Zhang, R.; Zhang, Y.; Wu, L.; Lu, W.; Li, J.; Li, Y.; Zhang, H. MXene-Enabled Electrochemical Microfluidic Biosensor: Applications toward Multicomponent Continuous Monitoring in Whole Blood. Adv. Funct. Mater. 2019, 29, 1807326. [Google Scholar] [CrossRef]
  352. Guan, Q.; Ma, J.; Yang, W.; Zhang, R.; Zhang, X.; Dong, X.; Fan, Y.; Cai, L.; Cao, Y.; Zhang, Y.; et al. Highly fluorescent Ti3C2 MXene quantum dots for macrophage labeling and Cu2+ ion sensing. Nanoscale 2019, 11, 14123–14133. [Google Scholar] [CrossRef]
Figure 1. Distinct strategies for the electrochemical detection of H2O2 including Cyto-biosensors, Immuno-biosensors, Enzymatic and non-enzymatic biosensors. Figure reproduced with permission from [46]. Copyright 2014. American Chemical Society (Washington, DC, USA).
Figure 1. Distinct strategies for the electrochemical detection of H2O2 including Cyto-biosensors, Immuno-biosensors, Enzymatic and non-enzymatic biosensors. Figure reproduced with permission from [46]. Copyright 2014. American Chemical Society (Washington, DC, USA).
Nanomaterials 12 01475 g001
Figure 2. Mechanism of the chemiluminescent material for the detection of H2O2 released in cancer cells. Excitation and de-excitation of chemiluminescence materials can be seen during chemical reaction.
Figure 2. Mechanism of the chemiluminescent material for the detection of H2O2 released in cancer cells. Excitation and de-excitation of chemiluminescence materials can be seen during chemical reaction.
Nanomaterials 12 01475 g002
Figure 3. Analysis of HyPer fluorescence in K562 cells exposed to extracellular H2O2. (a) Scheme demonstrating the changes in the excitation spectrum of HyPer upon oxidation. (b) Flow cytometry histograms of K562 cells measured after two-minute exposure to different concentrations of H2O2. Reproduced with permission from [127]. Copyright 2019, Science Direct (Amsterdam, The Netherlands).
Figure 3. Analysis of HyPer fluorescence in K562 cells exposed to extracellular H2O2. (a) Scheme demonstrating the changes in the excitation spectrum of HyPer upon oxidation. (b) Flow cytometry histograms of K562 cells measured after two-minute exposure to different concentrations of H2O2. Reproduced with permission from [127]. Copyright 2019, Science Direct (Amsterdam, The Netherlands).
Nanomaterials 12 01475 g003
Figure 4. Different types of carbon-based materials, i.e., (a) Graphene, (b) Carbon nanotube, and (c) reduced grapheme oxide, used in electrochemical sensing of H2O2.
Figure 4. Different types of carbon-based materials, i.e., (a) Graphene, (b) Carbon nanotube, and (c) reduced grapheme oxide, used in electrochemical sensing of H2O2.
Nanomaterials 12 01475 g004
Figure 5. Mechanism of the synthesis of the graphene enzyme composite for the electrochemical sensing of H2O2. Reproduced with permission from [216]. Copyright 2015, Science Direct.
Figure 5. Mechanism of the synthesis of the graphene enzyme composite for the electrochemical sensing of H2O2. Reproduced with permission from [216]. Copyright 2015, Science Direct.
Nanomaterials 12 01475 g005
Figure 6. Enzyme-loaded CNTs for the detection of H2O2 in living cells. Reproduced with permission from [336]. Copyright 2019, Elsevier Ltd (Amsterdam, The Netherlands).
Figure 6. Enzyme-loaded CNTs for the detection of H2O2 in living cells. Reproduced with permission from [336]. Copyright 2019, Elsevier Ltd (Amsterdam, The Netherlands).
Nanomaterials 12 01475 g006
Figure 7. Mechanism of metal nanoparticle loaded CNTs for real-time analyses of H2O2 secreted from live cells. Reproduced with permission from [324], Copyright 2016 Spinger Ltd (Berlin/Heidelberg, Germany).
Figure 7. Mechanism of metal nanoparticle loaded CNTs for real-time analyses of H2O2 secreted from live cells. Reproduced with permission from [324], Copyright 2016 Spinger Ltd (Berlin/Heidelberg, Germany).
Nanomaterials 12 01475 g007
Table 1. Non-enzymatic metal free H2O2 electrochemical sensors based on graphene.
Table 1. Non-enzymatic metal free H2O2 electrochemical sensors based on graphene.
Carbon MaterialSensitivity
μA mM−1 cm−2
Liner Range
(μM)
Detection Limit
(μM)
Ref.
CR-GO-0.05–15000.05[155]
Graphene-MWCNT32.91 20–21009.4[196]
rGO/nPPy47.69 0.1–40.034[201]
IL-GR-s-PANI280.00.5–20000.06[202]
rGO/Tyrosine69.07 100–210080[203]
Poly(o-Phenylenediamine)/GO16.2 2.5–250.84[204]
BGNs266.71000–20,0003.8[199]
NB-G-0.5–50000.05[200]
GSnano/CS18.785.22–10,4302.6[205]
GN-HN-SWCNT0.0150.2–4000.05[206]
H-GNs/PEDOT2350.1–100.08[207]
NS-GQD/G-0.4–330.026[208]
Functionalized 3D Graphene169.70.4–6600.08[209]
rGO/GO hybrid MEA-0.18–9.6-[210]
3D-G/GCE-0.2–41,200 0.17 [211]
Table 2. Graphene-based enzymatic biosensors for H2O2 detection in Cancerous cells.
Table 2. Graphene-based enzymatic biosensors for H2O2 detection in Cancerous cells.
Graphene-Based MaterialsSensitivity
μA mM−1 cm−2
Linear Range
(μM)
Detection Limit
(μM)
Ref.
GE/Fe3O4/Hb GCE0.3837100–1700 6.00[218]
rGO-CMC/Hb-0.083–13.940.08[219]
Hb/AuNPs/ZnO/Gr-6.0–11300.8[220]
HRP/graphene-0.33–14.00.11[221]
Hb/Au/GR-CS3.47 × 105 2.0–935 0.35[222]
Hb/Au NPs-Gr -0.1–70 0.03[223]
HRP/P-L-His-rGO 2.6 × 1050.2–5000 0.05[224]
GS-PSS/GRCAPS -10–12,000 3.3[216]
HRP/AuNP/ThGP 0.0860.5–1800 0.01[225]
PANI/HRP/GE-CNT/AuPt NPs 370 0.5–100 0.17[226]
Au/graphene/HRP/CS -5.0–5130 1.7[227]
MP11/DMPG-AuNPs/PDDA-G243.7 20–280 2.6[228]
(HRP-Pd)/f-graphene 92.8225–35000.05[229]
HRP-f-graphene-Ag 143.525–19,350 5.0[230]
HRP/CeO2-rGO 4.650.1–500 0.021[212]
HRP-MoS2-Gr 679.70.2–1103 0.049[214]
Catalase/AuNPs/graphene-NH2 13.4 0.3–600 0.05[231]
Cyt c/GO-MWCNT/Au NP0.5331 × 105–1.4 × 10−4 27.7 × 10−6[232]
RGO-MWCNT-Pt/Mb 1.9901 × 10−5–1.9 × 10−416 × 10−6[233]
PPY-He-RGO -0.1–10 0.13[41]
HRP/PGN/GCE-8.0 ×10−11–6.64 × 10−72.6 × 10−5[215]
PGR/catalase/GCE-1.0 × 10−7–7.7 × 10−6 1.5 × 10−3[234]
FeSx/graphene--5 × 10−4[235]
F-MoS2-FePt NCs-8–300 2.24[236]
Table 3. Graphene-supported non-Noble metal nanoparticles for electrochemical detection of H2O2.
Table 3. Graphene-supported non-Noble metal nanoparticles for electrochemical detection of H2O2.
Graphene-Based MaterialsSensitivity
μA mM−1 cm−2
Linear Range
(μM)
Detection Limit
(μM)
Ref.
Nafion/EGO/Co3O4 5601–100 0.3[240]
CoHCFNPs/GR 0.00070.6–379.50.1[41]
VS2 NPs/GCE41.96 20.5–2.50.224[241]
CoOxNPs/ERGO 148.65–10000.2[242]
CoTPP/RGO 0.00130.1–46000.02[243]
rGO/CoPc-COOH 14.5100–12,000 60[244]
(PDDA-G/Fe3O4)n 61.220–6250 2.5[245]
Fe3O4/GO-PAMAM 1.385 20–1000 2.0[246]
CoS/RGO2.519 0.1 to 2542.40.042[247]
rGO-Fe2O3 0.08550–9000 6.0[248]
Fe3O4/rGO 387.6 1–20,000 0.17[249]
Ni2P NA/TM690.7 0.001–20 0.2 [166]
Fe3O4/RGO 22.27 0.5–3000 0.18[41]
Fe3O4/RGO 0.0468 4.0–1000 2.0[250]
PB/TiO2-GR 480.97 0.04–2000 0.0086[251]
RGO/Fe3O4 688.0 100–6000 3.2[252]
Cu-MOF-GN 57.73 10–11,180 2.0[33]
PFECS/rGO 117.142 10–190 1.253[253]
FeTSPc-GR-Nafion 36.93 0.2–5000 0.08[254]
RGO/ZnO 13.49 0.02–22.48 0.02[255]
Cu2O/N-graphene 26.67 5.0–3570 0.8[256]
Cu2O-rGO0.020730–12,800 21.7[257]
CuS/RGO 0.035 5–1500 0.27[80]
Cu2O/GNs -300–7800 20.8[258]
GO/MnO238.2 5.0–600 0.8[259]
MnO2-ERGO59.0 100–45,400 10[260]
MnO2 nanosheet/graphene -10–9002.0[261]
CoFe/NGR435.7-0.28[262]
Co@NCS-0.5–75000.08[263]
Table 6. Ultra-sensitive electrochemical biosensors for detection of H2O2.
Table 6. Ultra-sensitive electrochemical biosensors for detection of H2O2.
H2O2 BiosensorsSensitivity
μA mM−1 cm−2
Linear Range (μM)Detection Limit (μM)Ref.
GN-HN-SWCNT0.0150.2–4000.05[206]
Cyt c/GO-MWCNT/Au NP0.5331 × 105–1.4 × 10−427.7 × 10−6[232]
HRP/P-L-His-rGO2.6 × 1050.2–50000.05[224]
HRP/AuNP/ThGP0.0860.5–18000.01[225]
RGO-MWCNT-Pt/Mb1.9901 × 10−5–1.9 × 10−416 × 10−6[233]
CoTPP/RGO0.00130.1–46000.02[243]
GN-Pt0.012–7100.5[277]
PDA-RGO/Ag NP0.01110.5–80002.07[282]
Pt/GN0.02042.5–66500.8[41]
PtAuNPs-CTAB-GR0.16540.005–4.80.0017[294]
Pd-NPs/GN0.0190.001–20000.0002[298]
GCE/MWCNTs-CDs0.0393.5–3000.25[324]
GCE/C60-MWCNTs
CS-IL/MB/CuNP
0.02432–40.055[327]
GCE/rGONRs/MnO20.01420.25–22450.071[335]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ahmad, T.; Iqbal, A.; Halim, S.A.; Uddin, J.; Khan, A.; El Deeb, S.; Al-Harrasi, A. Recent Advances in Electrochemical Sensing of Hydrogen Peroxide (H2O2) Released from Cancer Cells. Nanomaterials 2022, 12, 1475. https://doi.org/10.3390/nano12091475

AMA Style

Ahmad T, Iqbal A, Halim SA, Uddin J, Khan A, El Deeb S, Al-Harrasi A. Recent Advances in Electrochemical Sensing of Hydrogen Peroxide (H2O2) Released from Cancer Cells. Nanomaterials. 2022; 12(9):1475. https://doi.org/10.3390/nano12091475

Chicago/Turabian Style

Ahmad, Touqeer, Ayesha Iqbal, Sobia Ahsan Halim, Jalal Uddin, Ajmal Khan, Sami El Deeb, and Ahmed Al-Harrasi. 2022. "Recent Advances in Electrochemical Sensing of Hydrogen Peroxide (H2O2) Released from Cancer Cells" Nanomaterials 12, no. 9: 1475. https://doi.org/10.3390/nano12091475

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