Previous Article in Journal
A Luminol-Based, Peroxide-Free Fenton Chemiluminescence System Driven by Cu(I)-Polyethylenimine-Lipoic Acid Nanoflowers for Ultrasensitive SARS-CoV-2 Immunoassay
Previous Article in Special Issue
Development of CPE/ssDNA-Based Electrochemical Sensor for the Detection of Leucine to Assess Soil Health
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Two-Dimensional Carbon-Based Electrochemical Sensors for Pesticide Detection: Recent Advances and Environmental Monitoring Applications

1
Department of Chemistry, Madanapalle Institute of Technology & Science (MITS), Deemed to be University, Madanapalle 517325, India
2
Department of Chemistry, Soonchunhyang University, Asan 31538, Republic of Korea
3
Department of Chemistry, Presidency University, Bengaluru 560064, India
4
Department of Chemistry, Sri Venkateswara College, University of Delhi, New Delhi 110021, India
5
Polymer Electronics Lab, Department of Bioelectronics and Biosensors, Alagappa University, Karaikudi 630003, India
*
Authors to whom correspondence should be addressed.
Biosensors 2026, 16(1), 62; https://doi.org/10.3390/bios16010062
Submission received: 4 December 2025 / Revised: 8 January 2026 / Accepted: 12 January 2026 / Published: 14 January 2026

Abstract

Pesticides have been widely applied in agricultural practices over the past decades to protect crops from pests and other harmful organisms. However, their extensive use results in the contamination of soil, water, and agricultural products, posing significant risks to human and environmental health. Exposure to pesticides can lead to skin irritation, respiratory disorders, and various chronic health problems. Moreover, pesticides frequently enter surface water bodies such as rivers and lakes through agricultural runoff and leaching processes. Therefore, developing effective analytical methods for the rapid and sensitive detection of pesticides in food and water is of great importance. Electrochemical sensing techniques have shown remarkable progress in pesticide analysis due to their high sensitivity, simplicity, and potential for on-site monitoring. Two-dimensional (2D) carbon nanomaterials have emerged as efficient electrocatalysts for the precise and selective detection of pesticides, owing to their large surface area, excellent electrical conductivity, and unique structural features. In this review, we summarize recent advancements in the electrochemical detection of pesticides using 2D carbon-based materials. Comprehensive information on electrode fabrication, sensing mechanisms, analytical performance—including sensing range and limit of detection—and the versatility of 2D carbon composites for pesticide detection is provided. Challenges and future perspectives in developing highly sensitive and selective electrochemical sensing platforms are also discussed, highlighting their potential for simultaneous pesticide monitoring in food and environmental samples. Carbon-based electrochemical sensors have been the subject of many investigations, but their practical application in actual environmental and food samples is still restricted because of matrix effects, operational instability, and repeatability issues. In order to close the gap between laboratory research and real-world applications, this review critically examines sensor performance in real-sample conditions and offers innovative approaches for in situ pesticide monitoring.

1. Introduction

The accelerating growth of the global population leads to significant demand for food supply, making it difficult to achieve greater yields in agriculture due to problems such as poor soil health, lack of fertilizers, unpredictable weather conditions, irrigation issues, crop infections, harmful parasites, and plant-eating pests. One notable advancement is the emergence of precision agriculture or smart agriculture, which is garnering significant attention for its ability to optimize resource use while enhancing productivity. As a critical driver of increasing agricultural production, pesticides are widely applied to crops across the globe due to their effectiveness in controlling plant pests, thereby minimizing the need for manual labour [1]. Pesticides encompass various substances, including herbicides used to terminate weeds and other undesirable vegetation; insecticides for controlling a broad array of insect species; fungicides used to prevent the growth of moulds and mildew; disinfectants to curb the spread of harmful bacteria; and rodenticides for controlling the population of mice and rats [2]. The toxic nature of pesticides affects not only the intended pest species but also disrupts ecosystems by damaging the habitats of beneficial organisms. Additionally, many pesticides exhibit environmental persistence, allowing them to remain active for prolonged periods. These properties trigger their accumulation in the tissues of organisms and subsequent transfer through trophic levels, leading to bioaccumulation and biomagnification. Furthermore, pesticides can significantly affect water purity and quality by dispersing into adjacent ecosystems, leading to the contamination of aquatic life and fertile soil. Prolonged and intensive use of pesticides can substantially heighten environmental risks by accelerating resistance development in target pests and promoting the emergence of secondary pest infestations [3]. India ranks fourth in agrochemical production, yet its pesticide usage remains relatively low on a per-hectare basis compared to countries such as Japan, South Korea, and the United States. Soil is the main reservoir for pesticides applied in agricultural practices. While only around 10% of these chemicals effectively reach their intended targets, a significant portion, ranging from 20% to 70%, along with their breakdown products, end up in the soil, where they can persist as long-lasting residues. Based on their predominant chemical constituents, pesticides are systematically classified into organochlorines, organophosphates, and carbamates [4,5,6]. Based on the exposure (amount and location), certain pesticides are toxic to other animals and humans if they enter our food chain through sources such as air pollution, groundwater, food storage, or harvest. The primary cause of illnesses that seriously impair our health is contaminated food [2]. The recent investigation suggested that edible plants are contaminated by approximately 30% with pesticide residues [7]. Regulation (EC) No 396/2005 establishes maximum residue levels (MRLs) of pesticide residues in food and feed of plant and animal origin intended for human or animal consumption at 0.01 ppm [1]. The high exposure of plant tissues to pesticides can lead to a phytotoxic effect, oxidative damage and alteration of enzyme systems impacting the physiological activities of plants [8]. The prolonged exposure to pesticides in humans leads to the development of neurological diseases due to their inhibitory action on acetylcholinesterase (AChE) [9]. Nerve damage in both humans and animals causes neurological problems that can affect mobility, sensation, gland secretion, or organ function. Moreover, the affected nerves may cause other serious health problems. Several environmental protection groups have classified pesticides as a priority hazardous substance. A few of the toxic pesticide components are shown in Figure 1. The routine techniques, including mass spectrometry, liquid–gas chromatography, Raman, fluorescence spectroscopy, etc., are frequently used to detect pesticides [10,11,12]. The speed and accessibility of analysis techniques, high selectivity with sensitivity, interfering molecules, and high-cost levels are only a few of the numerous obstacles and demands that still need to be addressed. Electrochemical sensing mechanisms are crucial in many applications because they employ a range of transduction strategies to detect analytes with high sensitivity and selectivity. The primary techniques are impedimetric, potentiometric, amperometric, and transistor-based sensing; each offers unique advantages for certain applications [2]. An interesting area of research in electrochemical detection is electrochemical (EC) sensors, a potential alternative sensor platform for detecting chemical pollutants, pesticides, etc., [13]. These sensors can be seamlessly integrated into compact, portable devices for rapid and real-time analysis in various application areas, including agro-environmental monitoring, food quality, and safety [14]. When optimally engineered for pesticide detection, EC sensors exhibit exceptional sensitivity and selectivity, even in complex sample matrices. In electrochemical detection, functional materials can be used to modify the working electrode, enabling it to detect the target pesticide with high sensitivity and selectivity while responding much more quickly. The EC sensors work on the following fundamental mechanisms: electrocatalysis, the signal transduction process, and the use of functional materials [15]. These sensors convert chemical or biological interactions into electrical signals via transducers, which can be electrodes or other materials. The performance of these sensors is influenced by the immobilization of biomolecules and the electrode’s surface properties. Moreover, enhanced electron-transfer processes increase sensitivity and specificity by boosting electron transfer mechanisms. Electrocatalytic efficiency is strongly influenced by concentration and kinetic factors. Additionally, the use of advanced materials like nanomaterials and metal–organic frameworks improves sensor performance and contributes to stability and selectivity under varied conditions by increasing mass transport and surface area [16]. Even though electrochemical sensors have a lot of potential, concerns like stability in many environmental scenarios and the requirement for more breakthroughs in sensor design continue to be crucial for this field’s future developments.
Several sensing techniques are used in the electrochemical detection of pesticides utilizing 2D carbon-based materials, depending on the electrode architecture and pesticide class. Analyte accumulation on graphene or g-C3N4 surfaces is facilitated by strong π–π interactions and hydrogen bonding, which are characteristic adsorption-controlled processes for organophosphates and carbamates. In sensors with metal nanoparticles or heterostructures, electrocatalytic mechanisms predominate, allowing for higher sensitivity and enhanced redox reactions. Inhibition-based sensing, which is especially pertinent to organophosphate pesticides, makes use of the reduction in enzyme activity, such as acetylcholinesterase, to produce detectable changes in current. Affinity-based techniques that use aptamer-functionalized electrodes or molecularly imprinted polymers also provide great selectivity toward certain pesticide compounds, which makes them appropriate for complicated sample matrices.
While a lot of work has gone into creating extremely sensitive electrochemical sensors for pesticide detection, sensitivity by itself does not guarantee dependable operation in real-world scenarios. Complex matrices comprising organic matter, inorganic ions, and coexisting pollutants can significantly affect electrochemical responses in real environmental and food samples, frequently resulting in signal suppression, amplification, or electrode fouling. The translation of laboratory-scale sensor performance to practical applications is severely hampered by these matrix effects. Therefore, a vital understanding of matrix effects and how they affect electrochemical sensing is necessary for the logical design of sensors that can detect pesticides in complex samples with accuracy and repeatability. The following section examines the causes, implications, and mitigation techniques for matrix effects in electrochemical pesticide sensing. In this context, the present review focused on the electrochemical detection of several kinds of pesticides, whose chemical structures are summarized in Figure 2.

2. Sensor Design Strategies Based on Pesticide Chemical Classes

2.1. Organophosphates

Among the most popular insecticides, organophosphate pesticides are distinguished by the presence of a functional group containing phosphorus. Their great capacity to inhibit acetylcholinesterase (AChE) via phosphorylating the enzyme’s active site accounts for their principal toxicological and electrochemical relevance [17]. Many electrochemical sensing methods are based on this inhibitory mechanism. Since organophosphates usually show little direct redox activity from an electrochemical perspective, enzyme inhibition-based methods are frequently used to detect them indirectly. In such systems, the decrease in enzyme activity (for example, reduced thiocholine oxidation) is electrochemically measured and associated with the concentration of organophosphate. Surface-modified electrodes, such as those with nanostructured carbon materials, metal nanoparticles, or conducting polymers, are commonly used to improve sensitivity and selectivity [18].

2.2. Carbamates

A class of nitrogen-containing chemicals called carbamate insecticides is frequently employed in the management of agricultural pests. They differ from organophosphate pesticides in that their main biochemical effect is the reversible suppression of acetylcholinesterase via carbamylation of the enzyme active site [18]. From an electrochemical perspective, carbamates frequently show substantial intrinsic redox activity at electrode surfaces instead of adsorption-dominated behaviour. As a result, the design of electrochemical sensors for carbamate detection places a strong emphasis on the selection of suitable electrode materials, surface functionalization to increase adsorption efficiency, and signal amplification techniques utilizing nanostructured materials to boost analytical performance and sensitivity.

2.3. Organochlorine Pesticides

Organochlorine pesticides are distinguished by a high level of chlorination, which provides great hydrophobicity and chemical durability. Because of their low inherent electrochemical redox activity, their detection is mostly determined by adsorption-controlled processes and hydrophobic interactions on the electrode surface [19]. In order to improve adsorption efficiency and analyte preconcentration, electrochemical sensor design strategies prioritize surface-modified electrodes that incorporate functional nanomaterials. Furthermore, tailored sensing interfaces and affinity-based recognition elements are used to enhance analytical performance and selectivity in complicated environmental matrices.

2.4. Pyrethroid Pesticides

Pyrethroid pesticides are synthetic derivatives of natural pyrethrins that are commonly utilized due to their great insecticidal efficiency and minimal mammalian toxicity. Pyrethroids often show limited direct redox activity electrochemically and are mostly identified by indirect sensing mechanisms or adsorption-controlled procedures [20]. In order to improve adsorption efficiency and signal transduction, surface-modified electrodes, functional nanomaterials, and affinity-based interfaces are prioritized in electrochemical sensor design for pyrethroid detection. To increase sensitivity and selectivity, customized sensing methods are frequently used, such as molecular recognition layers and electrodes modified with nanocomposite [20].

3. Matrix Effect in Electrochemical Detection of Pesticides

Matrix effect is a significant challenge in electrochemical analysis, as it is caused by the presence of interfering compounds in real matrices that can either suppress or enhance the electrochemical signal. In electrochemical pesticide sensing, the analytical response can be significantly influenced by coexisting species present in environmental matrices, such as soil, food extracts, wastewater, agricultural runoff water, and river and lake water. The signal accuracy is altered by interfering substances, which produce overlapping redox signals, alter electron transfer kinetics, compete for adsorption sites and foul the electrode surface. The inorganic ions Ni2+, Co2+, Fe3+, and Cu2+ can primarily form complexes with certain pesticides [21]. Nitroaromatic substances, which closely resemble the electrochemical behaviour of nitro-containing pesticides, can also interfere with detection. Phenolic compounds mainly interfere with carbon-based electrodes through π-π adsorption, thereby limiting the availability of active sites [22]. Some biomolecules, such as dopamine, uric acid and ascorbic acid, can induce surface fouling or contribute substantial background currents, while humic substances, surfactants, and saccharides (sucrose and glucose, etc.) block catalytic pores and hamper diffusion. Hence, a selective electrochemical sensing platform requires minimizing the matrix-derived artefacts either through an antifouling coating or preferential adsorption of the catalyst, or proper host–guest recognition. In this context, researchers have concentrated on the synthesis and practical applications of 2D-carbon derived materials, which have been applied in the analysis of various real samples.

4. Overview of 2D Nanomaterials Towards Sensing of Pesticides

4.1. Graphene

Graphene is a vital two-dimensional, sp2 hybridized polycyclic aromatic carbon nanomaterial that appears to be a skeletal honeycomb and can be further functionalized or modified better with various functional groups (-OH, -O-, -COOH and -NH2, etc.), polymers, metal, metal oxide and other substances. Graphene exhibits low magnetic properties due to short-term electron–electron interactions and long-term coulomb interactions, making it a low magnetic noise, having high carrier mobility and high surface sensitivity [23,24]. Two-dimensional graphene (2D-G) has high conductivity in terms of both thermal and electrical aspects due to its resistance to moisture and high electron mobility. The 2D-G conductivity and optical transparency have led to its application as a transparent conductive layer in advanced photonic devices [25]. The 2D-G exhibits major properties that position it as a key material for transformative innovations in various scientific domains. This makes 2D-G a crucial material for application in batteries and cells as anodes, and solar panels, supercapacitors, fuel cells, anticorrosion coatings, DNA sequencing, drug delivery, biosensors, and as a significant physical surface support for desired biomolecules to transport into the solution media in electrochemical sensor applications. The 2D-G, characterized by its highly conjugated π-electron network and an exceptional specific surface approaching nearly 2600 m2 g−1, demonstrates superior adsorption efficacy towards organic molecules, owing to its electronic structure and expansive interfacial domain [26,27,28,29]. These unique, complex, and multifaceted properties of 2D-G, along with its high surface-to-volume ratio and porosity, are poised to advance future developments by enhancing the interaction between the analyte and the electrode surface, thereby promoting improved sensitivity and detection limits, highlighting the critical role of the electrode in real-time detection and monitoring of pesticides [30,31,32,33,34,35,36].

4.2. Derivatives of 2D-G

The enhanced structural versatility of graphene-based derivatives accelerates fast electron transfer kinetics, efficient mass transport of analytes, and permits selective functionalization with suitable moieties, thereby enabling selective and sensitive analysis of selected analytes in electrochemical sensing [37]. The popular 2D-G derivatives, such as graphene oxide, reduced graphene oxide, and carbon nanowalls, are significant functionalized materials. Two-dimensional graphene oxide (2D-GO) material is composed of oxygen-rich functional groups, which exhibit varied electronic character, chemical stability, lower conductivity, and high reactivity compared to 2D-G. The 2D-GO is prepared by the oxidation of graphene, which produces a dispersible aqueous solution. The plane structure of 2D-GO, proven to be primarily composed of epoxy (-O-) and hydroxyl (-OH) groups, is very appropriate for non-carbon atom doping or polymer bonding applications, making it accessible as an electrode modification material [36]. The two-dimensional reduced graphene oxide (2D-rGO) is achieved by decreasing or reducing the oxygen functional groups from 2D-GO. By approaching or restoring the sp2 carbon network of 2D-G, 2D-rGO enhances its interaction capabilities with analytes through various underlying mechanisms such as π-π interactions, hydrogen bonding, and electrostatic forces [38,39,40]. The 2D-rGO goes a step forward compared with 2D-GO, because of its different flexibilities, simple preparation method, ease of dispersion in solvents, electrical performance control and other physical properties like exceptional mechanical performance, electrical charge transfer ability, biological compatibility, and superior light transmittance, making it a potential candidate in biosensing and other sensor materials [41,42]. The further improvement of sensing or redox performance on the electrode surface can be greatly achieved by modifying 2D-rGO with suitable conducting materials [43].

4.3. Graphitic Carbon Nitride

Two-dimensional graphitic carbon nitride (2D-g-C3N4) sp2-hybridized carbon and nitrogen material, resembling a graphene-like polymeric structure, exhibits distinctive physical and chemical properties [44]. The material exhibits remarkable performance due to several inherent characteristics. It can be synthesized through environmentally sustainable and economical routes using precursors such as melamine, urea, or thiourea via a straightforward thermal polymerization process. Its tri-s-triazine ring structure and high degree of polycondensation impart exceptional physicochemical stability. The material also shows great potential in energy generation and storage applications, along with distinct fluorescence behaviour. Possessing a bandgap of about 2.7 eV, it effectively absorbs blue light with wavelengths shorter than 475 nm. Its outstanding optical properties contribute to excellent photo and electrocatalytic activity. Moreover, the electronic structure can be finely tuned, enabling modulation of charge carrier distribution owing to its intrinsic polarity and semiconducting nature. Finally, the porous architecture enhances its surface reactivity, making it a highly promising candidate for environmental remediation applications [45,46,47,48,49,50]. The foremost benefit of 2D-g-C3N4 over graphene is its bandgap, which features outstanding capabilities in absorbing light, thereby enhancing its potential in water oxidation and reduction processes, photocatalysis, optoelectronics, and nano-sensing [51]. It has attracted considerable scholarly attention in drug delivery, tissue engineering, and wound dressing and is a very promising material in cancer predictions, therapy because of its nontoxicity, biocompatibility, and its capability to penetrate the natural tissues [52]. The morphology and electrochemical sensing properties of 2D-g-C3N4 are significantly improved in terms of electronic states, creating new active sites, compressed band structure and elevating charge transport rate by the addition of non-metal (nitrogen, sulphur, oxygen, carbon, and phosphorus), monodoping with germanium, metal, metal oxides, conductive nano carbon substances (like graphene and carbon nanotube) and MXene (transition metal carbides, nitrides) substances [53,54,55,56,57,58,59,60]. On the other hand, owing to the high surface area and porous nature of 2D-g-C3N4 nanosheets, it is highly appropriate for environmental stressor (pesticide residues) sensing applications.

4.4. Graphdiyne

Li et al. prepared thin films of two-dimensional graphdiyne (2D-GDY) through a cross-coupling reaction, which has become a prominent member of 2D carbon-based materials [61]. The 2D-GDY has a unique pattern of linear diacetylene linkages between benzene rings, creating spatial variations in charge density and enabling enhanced electron mobility, an extended π-system, and a consistent pore architecture that enables efficient mass transport across both in-plane and out-of-plane directions [62,63,64]. Diverging from 2D-G, the 2D-GDY lattice predominantly exhibits a finite band gap, imparting semiconducting electronic properties. Furthermore, unlike the brittle fracture behaviour commonly associated with 2D-G, 2D-GDY nanosheets demonstrate polymer-like stretchability, offering enhanced mechanical flexibility and low density [65,66]. Owing to the intrinsic properties of its sp-hybridized carbon atoms, along with the highly reactive C-C triple bonds and fully conjugated architecture, 2D-GDY demonstrates a pronounced capacity for electronic doping and chemical modifications. This opens avenues in scientific and biomedical fields, including optical properties, bio-imaging, sensing technology, targeted obtained drug delivery, cancer therapy and antimicrobial applications. The 2D-GDY exhibits a high extinction coefficient across broad wave lengths and strong NIR absorption with a large surface area, resulting in outstanding photo thermal conversion efficiency [67]. The 2D-GDY is actively involved in removal and separation technology, such as desalination or water treatment, gas separation and storage, organic separation, and heavy metal removal through their high pore density, π-conjugation, polarity and large surface area [68,69,70,71,72]. The planar and Dirac cone structure of 2D-GDY exhibits high carrier mobility, and surface area represents the enhanced charge transfer and surface interactions on modified electrodes, facilitating the enrichment of target analytes during electrochemical detection. In addition, 2D-GDY has a tunable bandgap, can be functionalized by hetero atom or groups to get 2D-GDY functionalized composites, which are potentially used as electrode materials for enabling outstanding conductivity and enhanced electrochemical reactivity with numerous active sites [73,74,75,76].
In electrochemical sensing mechanisms, 2D-G and 2D-g-C3N4 play complementary but distinct roles. The 2D-G derivatives exhibit outstanding electrical conductivity, a large electroactive surface area, and fast electron-transfer kinetics. Hence, sharp voltammetric peaks, high amperometric currents, low charge-transfer resistance, and ultralow detection limits could be achieved. However, their sensitivity and adsorption are relatively limited, and these shortcomings can be addressed through functionalization strategies such as 2D-GO/2D-rGO modification, defect engineering, or decoration with metal nanoparticles. Even after functionalization, the graphene derivative retains metal-like conductivity and faster heterogeneous electron-transfer kinetics, leading to higher peak currents, sharper and better-resolved voltammetric peaks, and lower detection limits for electroactive pesticides (organophosphates, nitro-pesticides, carbamates). In contrast, functionalized 2D-g-C3N4 shows improved adsorption and selectivity due to its N-rich surface. But its semiconducting framework limits charge transport, resulting in comparatively broader and lower-intensity peaks. Therefore, 2D-G functionalized composites outperform 2D-g-C3N4-based ones in peak quality and sensitivity, while 2D-g-C3N4 is better used as a complementary component to enhance selectivity rather than as the primary peak-generating material.

5. Synthesis of 2D Carbon-Based Materials

5.1. The 2D-GO

The 2D-GO is one of the most widely studied derivatives of graphite oxide because of its adjustable properties, various functionalities and scalable synthesis. The 2D-GO is generally synthesized from the graphite through a chemical oxidation process, which introduces high-density oxygen-containing functional groups, for example, hydroxyl group, carboxylic group and epoxy group in the graphene. Because of these, it enhances the dispersibility in water and also provides active sites for further modification. For these reasons, GO is often used as a precursor to graphene-based nanocomposites. The most widely used path for synthesizing GO is the Hummers’ method, first established in 1958. This method uses potassium permanganate (KMnO4) and sodium nitrate (NaNO3) in a strong sulfuric acid (H2SO4) medium for the conversion of graphite into GO [77]. It became the benchmark method owing to its relatively reduced reaction time and higher reproducibility compared to earlier methods. However, the traditional Hummers’ process generates hazardous nitrogen oxides and manganese-containing waste as a by-product, which raises considerable environmental and safety concerns [78]. To mitigate these limitations, numerous modified Hummers’ methods have been proposed. For example, Marcano et al. (2010) eliminated NaNO3 from the reaction and introduced a H2SO4/H3PO4 mixture, which increased oxidation efficiency, yielded higher quality 2D-GO and lowered toxic emissions. Such modifications not only enhanced safety but also facilitated more environmentally friendly large-scale synthesis [78]. Recently, investigations have increasingly focused on eco-friendly oxidation routes, utilizing mild oxidants and greener solvents to reduce hazardous waste. For example, hydrogen peroxide (H2O2) has been successfully applied as the only oxidizing agent in a dilute H2SO4 system, with the acid acting primarily as a medium and reusable control agent. These green synthesis techniques significantly minimize the consumption of harsh chemicals yet maintaining efficiency, making 2D-GO production safer, quicker, and more viable for large-scale and biomedical purposes [79]. Therefore, 2D-GO synthesis has progressed from the initial chemical oxidation routes to contemporary, environmentally friendly approaches. Although the Hummers’ method and its modification remain the most widely utilized, continued efforts to develop safer, sustainable, rapid and large-scale methods demonstrate the advancement in this field.

5.2. The 2D-rGO

The 2D-rGO is typically produced from 2D-GO via processes that remove oxygen-containing functional groups, which helps to restore the conjugated sp2 carbon lattice and electrical conductivity. Among the current strategies, chemical, thermal, and electrochemical reduction are the most studied approaches. Every technique offers unique benefits and challenges depending on its intended use.

5.2.1. Chemical Reduction

Chemical reduction is the most employed method for synthesizing 2D-rGO, in which GO dispersions are processed with reducing agents that eliminate oxygenated functional groups and partially recover the conjugated sp2 carbon network. During this process, the GO powder is usually suspended in water or another solvent to create a homogeneous suspension, followed by the gradual addition of the reducing agent under agitation, heating, or controlled pH conditions to promote oxygen removal. The reduced product is then obtained by centrifugation, rinsed several times to remove residual reagents, and dried for subsequent use. Hydrazine hydrate was one of the initial reductants utilized, enabling efficient elimination of oxygen groups and producing conductive graphene-like nanosheets. However, its high toxicity, instability, and environmental risks significantly limit its large-scale production [80]. Sodium borohydride (NaBH4) is another powerful reductant that ensures efficient reduction, yielding rGO with enhanced conductivity. The process generally involves dispersing GO in an aqueous medium, subsequently, the dropwise addition of NaBH4 under continuous agitation at room temperature or moderate temperatures. Although effective, NaBH4 is prone to instability and rapid reaction, limiting its feasibility for industrial scale [81]. Greener alternatives have been designed to replace hazardous chemicals. For example, ascorbic acid (vitamin C) has been identified to be an ideal replacement due to its low toxicity, natural availability, and reducing power. Fernández-Merino et al. (2010) demonstrated that the use of vitamin C for the reduction in GO dispersions produces high-quality rGO with conductivity nearly equivalent to hydrazine-treated samples [82]. In 2015, Abdolhosseinzadeh et al. introduced a rapid and scalable vitamin C-based method in which GO is dispersed in an aqueous medium, incorporating vitamin C as the reductant, and the mixture is warmed under mild thermal conditions to generate rGO for large-scale applications [83]. Other mild reductants have also been investigated. Hydroquinone has served as a facile reductant for the conversion of GO into rGO, usually by mixing hydroquinone with aqueous GO dispersions under controlled pH and thermal conditions, leading to conductive graphene nanosheets [84]. In addition, gaseous hydrogen reduction performed at high temperatures following the thermal expansion of GO provides a route to high-quality graphene sheets with controllable thickness, although it requires specific high-temperature environments [85].

5.2.2. Thermal Reduction

Thermal reduction is one of the easiest approaches to produce rGO from GO. The process involves thermal treatment of GO at elevated temperatures under inert gas (Ar, N2) or reducing (H2/Ar) environments, leading to the decomposition of oxygen-containing moieties (hydroxyl, epoxy, and carboxyl) into gaseous species such as CO2, CO, and H2O. Such annealing restores the extended sp2 carbon lattice and significantly improves the electrical conductivity of the obtained material. Larciprete et al. (2011) contributed the mechanical insights into thermal reduction, highlighting a dual-pathway process. They demonstrated that the reduction pathway highly depends on oxygen coverage: at low oxygen density, oxygen groups are eliminated gradually (mainly hydroxyl and epoxy), in contrast, at high oxygen concentration, abrupt CO/CO2 release takes place, resulting in higher defects in the carbon lattice [86]. Building on this, Sengupta et al. (2018) conducted a temperature-dependent analysis from 300 to 450 °C. They showed that ~350 °C is an optimal temperature for annealing, at which a high degree of deoxygenation is obtained while preserving structural integrity. This finding emphasized the practical balance between conductivity restoration and structural stability [87]. Klemeyer et al. (2020) further highlighted the importance of material geometry in the thermal reduction process. They demonstrated that thin GO films reduce faster and more effectively compared to bulk powders due to improved heat transfer and gas diffusion. In contrast, bulk samples need higher temperatures to achieve an equivalent reduction level. This finding underlined the importance of tailoring thermal treatment conditions according to the morphology of the GO starting material [88]. In a recent study, Valentini et al. (2023) have shown the feasibility of thermal reduction (<300 °C) for flexible substrates at low temperature. While this method yielded rGO with lower conductivity compared to those obtained via >1000 °C treatments, it maintained film flexibility, making it highly suitable for portable and flexible electronics [89].

5.2.3. Electrochemical Reduction

Electrochemical reduction has been recognized as a clean, chemical-free, and precisely controllable method to synthesize rGO. Unlike conventional chemical and thermal methods, this method eliminates the need for harmful reagents or high-energy treatments. Instead, an imposed potential directly delivers electrons to GO films or suspensions, which reduce oxygen-containing functional groups as well as partially restore the sp2-conjugated carbon framework. Consequently, this method is particularly attractive for sensor fabrication and device integration, since it can be applied directly on electrode surfaces. The process usually involves coating GO onto a conductive substrate, for example, glassy carbon electrodes, metal electrodes or indium tin oxide (ITO), followed by immersion in an electrolyte solution (for instance, phosphate-buffered saline PBS, KCl). A fixed potential or cyclic voltammetry (CV) is then applied, leading to gradual deoxygenation. Reduction level can be adjusted by controlling the applied potential, scan rate, and electrolyte concentration. In most cases, the reduced material is directly suitable for electrochemical applications without further treatment. Wang et al. (2009) reported one of the early studies, in which GO-coated electrodes were reduced electrochemically in electrolyte solution, resulting in rGO with remarkably improved conductivity, which is highly effective for electrochemical sensors [90]. Pei and Cheng (2012) showed its benefits over chemical reduction compared to the chemical process, pointing out that electrochemical reduction methods allow fine control amount of oxygen while reducing chemical waste [91]. Furthermore, previous studies have concentrated on electrolyte engineering. For example, Stankovich et al. (2013) demonstrated that changing the ion composition of the supporting electrolyte markedly influences reduction effectiveness, electrochemical capacity, and long-term stability of rGO films [80].

5.3. Graphdiyne

Graphdiyne (GDY), a two-dimensional carbon allotrope characterized by sp-sp2 hybridization and evenly distributed acetylenic linkages, is typically synthesized through Glaser–Hay oxidative coupling of terminal alkynes in the presence of copper-based catalysts. Generally, it relies on the deprotection of hexakis[(trimethylsilyl)-ethynyl]benzene (HEB-TMS) to obtain the reactive monomer hexaethynylbenzene (HEB), then subjected to Cu-catalyzed homocoupling to form extended carbon lattices. A typical method involves solution-phase polymerization using organic solvents. For example, deprotected HEB was introduced into a pyridine solution which contained Cu(OAc)2 and kept under argon at room temperature for 24 h. The obtained material was purified by consecutive washing with pyridine, DMF, HCl and water and freeze-drying to obtain high-purity GDY [92]. Likewise, CuCl-catalyzed polymerization of HEB-TMS in DMF at 60 °C for 24 h generated a CuO/GDY composite, which was subsequently purified by acid treatment to eliminate copper species, resulting in black-brown GDY nanosheets [93]. This DMF-CuCl approach has been widely popular because of its reproducibility and controllable synthesis. Scale-up methods have been demonstrated in which gram-level synthesis was obtained by the reactivation of several grams of HEB–TMS with CuCl in DMF under the identical conditions [94]. A related CuCl-assisted method involved heating HEB-TMS and CuCl in DMF at 60 °C, 24 h, resulting in CuO/GDY, which, after several solvent washes and HCl treatment, was obtained as pure GDY [95]. Despite these advances in synthesis, 2D-GDY remains rarely employed as a flat-form electrode material in electrochemical pesticide sensing, mainly due to several practical and application-oriented limitations. The synthesis routes are still relatively complex, low-yield, and often suffer from limited reproducibility, particularly when translated to device-level fabrication. Moreover, the diacetylene linkages in GDY are chemically reactive and can undergo degradation under harsh electrochemical environments, such as acidic or alkaline media and repeated redox cycling. In addition, structural defects and poor intersheet contact frequently lead to lower effective electrical conductivity, while the lack of abundant functional groups or heteroatoms limits strong adsorption and selective recognition of pesticide molecules. Consequently, well-established two-dimensional materials, including graphene derivatives and 2D-g-C3N4, currently offer ultralow detection limits, superior stability, and proven performance in real-sample analysis, leaving limited practical incentive to adopt 2D-GDY for electrochemical pesticide sensing at its present stage of development.

5.4. The 2D-g-C3N4

The 2D-g-C3N4 is generally produced by pyrolytic condensation of nitrogen-rich, oxygen-deficient compounds containing C-N frameworks, for example, as triazine or heptazine. However, these compounds are often thermally unstable, costly, or reactive, limiting the direct synthesis of single-phase sp3 carbon nitrides difficult because of their poor thermodynamic characteristics [96]. Among various approaches, the most widely established method is thermal polycondensation of simple precursors (e.g., cyanamide, dicyandiamide, urea, and melamine). Cyanamide converts into 2D-g-C3N4 at ~550 °C, evolving NH3, whereas NaOH pre-treatment reduces the condensation temperature at ~500 °C, producing a higher surface area and lower aggregation [97]. However, cyanamide is high-cost and reactive, which limits its large-scale use. To address these issues, Xu et al. introduced guanidine hydrochloride, which is a water-soluble precursor and eco-friendly, following a similar condensation route to synthesize g-C3N4 at 550 °C in Ar [98]. Another low-cost precursor is urea, pyrolysis between 450 and 600 °C to generate 2D-g-C3N4 with its tunable surface areas (12–83 m2/g) varying with calcination temperature [99]. Bulk g-C3N4 produced by direct condensation typically exhibits very low surface area (~10 m2 g−1), limiting catalytic efficiency [100]. To overcome these issues, template strategies have been widely employed. Hard templates, for example, SBA-15 or silica, direct the formation of porous structures. On the other hand, surfactants or ionic liquids generate soft templates, such as the incorporation of Fe-g-C3N4 onto SBA-15, increased surface area (from 8 to 506 m2 g−1) and improved photocatalytic benzene oxidation [101]. Similarly, ordered mesoporous SBA-15 replicas produced well-defined, highly porous g-C3N4 [102]. Nevertheless, template removal steps may introduce impurities and complicate the synthesis process. Recent studies have optimized melamine-based methods, such as direct pyrolysis at 600 °C generate bulk g-C3N4. Then, post-annealing at 500–550 °C for varying durations fine-tunes porosity, crystallinity, and surface activity. Beyond bulk preparation, exfoliation strategies can now enable to formation of a nanosheet [101]. A recent 2023 study outlined multiple exfoliation approaches to enhance the properties of melamine-based g-C3N4. The one-step thermal route generates bulk g-C3N4 through calcining melamine, whereas a three-step alkali-assisted strategy delivers edge defects by KOH treatment, resulting in highly reactive nanosheets. Ultrasonic exfoliation in ethanol-water gives an easy, environment-friendly route to generate nanosheets, though precise size control remained a challenge. Thermal exfoliation through re-calcination modestly improved surface area, while chemical exfoliation with H2SO4 produced highly porous, defective frameworks, albeit requiring careful concerns and neutralization [102]. These methods produce thinner g-C3N4 nanosheets with high surface areas, adjustable band structures, and enhanced catalytic performance.

6. Synthesis Strategies and Their Influence on Electrochemical Sensing Performance

The synthesis procedure of 2D carbon-based materials greatly influences their electrochemical sensing capabilities by regulating essential material properties such as defect density, surface functional groups, conductivity, and active site availability. This section focuses on how various synthesis methodologies affect sensor-relevant features that are essential for pesticide detection rather than just preparation techniques. Commonly employed chemical and electrochemical exfoliation techniques for graphene derivatives generate structural flaws and oxygen-containing functional groups that improve analyte adsorption and enable hydrogen-bonding and π–π interactions with pesticide molecules. However, excessive defects could impair electrical conductivity, necessitating post-treatment or optimization techniques. On the other hand, controlled functionalization and heteroatom doping are made possible by hydrothermal and solvothermal synthesis methods, which enhance charge-transfer kinetics and electrocatalytic activity, especially when metal nanoparticles or heterostructures are included. Thermal polymerization of nitrogen-rich precursors for g-C3N4 produces a stable framework with numerous nitrogen functionalities that facilitate specific interactions with carbamate and organophosphate insecticides. However, in order to improve electron transport, its very low inherent conductivity requires composite synthesis with graphene or carbon nanotubes. Overall, maximizing electrochemical sensitivity, selectivity, and repeatability in pesticide sensing applications requires adjusting synthesis conditions to balance conductivity, surface chemistry, and defect density.

7. Recent Advances in Electrochemical Detection of Pesticides Using 2D Carbon Materials

Pesticides pose a major threat to both human health and the environment, with excessive residues in agricultural products severely compromising food safety. As a result, the sensitive and reliable detection of pesticide residues in vegetables, fruits, and water bodies is essential for ensuring food security and environmental protection. In recent years, 2D carbon materials, including rGO and g-C3N4, have attracted considerable attention for electrochemical pesticide sensing owing to their large surface area, excellent electrical conductivity, abundant surface functional groups, and strong affinity toward organic contaminants.

7.1. Two-Dimensional Carbon Derivative Aptamer-Based Sensing

Beyond catalytic enhancement, the integration of molecular recognition elements, particularly aptamers, with 2D carbon materials has enabled highly selective pesticide sensing. For example, MnMoO4/rGO nanocomposites were successfully employed in an aptamer-based electrochemical biosensor for fenitrothion (FNT) detection [103]. As illustrated in Figure 3A,B, the optimized MnMoO4:rGO (2:1) nanocomposite synthesized via a citric-acid-assisted hydrothermal route exhibited a planar hexagonal morphology with superior electrochemical properties. The corresponding DPV responses (Figure 3C) revealed concentration-dependent signal suppression upon FNT binding, while the calibration plot (Figure 3D) demonstrated an ultralow detection limit with excellent linearity. The strong interfacial coupling between MnMoO4 and rGO enabled high selectivity against structurally similar pesticides and reliable performance in wastewater, tap water, and rice extract matrices.
In another study, a 2D-rGO-based electrochemical aptasensor was developed by Asma Zaid Al Menhail et al. [104] to detect three hazardous neonicotinoid pesticides: imidacloprid, thiamethoxam, and clothianidin. The 2D-rGO surface was then functionalized with 1-pyrenebutyric acid (py), followed by carbodiimide hydrochloride/N-hydroxy succinimide activation to enable covalent bonding with amine-labelled aptamers (Figure 4). The development of the aptasensor was eventually completed by immobilizing aptamers specific to imidacloprid, thiamethoxam, and clothianidin and blocking any unreacted sites with ethanolamine. Tomato and rice samples were homogenized, extracted with water and acetonitrile, vortexed, centrifuged, evaporated, and then reconstituted in binding buffer. The extracts were filtered and diluted 1:10 before being spiked with known concentrations (10, 50, and 100 ng mL−1) of imidacloprid, thiamethoxam, and clothianidin. The spiked extracts were analyzed using the fabricated multiplexed aptasensor. DPV was used to record the current response of each aptamer-functionalized electrode upon exposure to its specific neonicotinoid target. The aptasensor exhibited excellent selectivity and high sensitivity, achieving a detection limit as low as 0.01 ng mL−1 across a broad linear range of 0.01–100 ng mL−1. The proposed method offers a low-cost, portable, and reliable tool for on-site pesticide detection, thereby contributing to better environmental protection and improved food safety.

7.2. Two-Dimensional-Carbon Derivatives with Metal Oxide-Based Sensing

Hybridization of redox-active metal oxides with 2D carbon materials has proven to be an effective strategy for enhancing electrochemical sensing performance. In such architectures, the 2D carbon component serves as a conductive scaffold that facilitates rapid electron transfer, suppresses nanoparticle aggregation, and enhances analyte adsorption through π-π interactions. In that connection, Vinitha M et al. [105] provided a schematic preparation of a metal oxide with 2D carbon nanocomposite GdPO4/RGO. To synthesize a nanocomposite, Gd(NO3)3⋅6H2O was taken in deionized water and stirred continuously for 10 min. Citric acid and NaH2PO4 were then added one after the other. After adding the appropriate amount of 2D-GO, the mixture was continuously stirred for an hour. The entire mixture was then transferred to a Teflon autoclave, which was maintained at 180 °C temperature for 12 h. Thereafter, to remove impurities in the GdPO4/RGO, several washes with DI water and ethanol were carried out. Furthermore, EIS and CV studies of bare GCE and modified GCE with GdPO4/RGO, which reveal that the bare GCE have limited electron transfer, resulting in a high Rct (886 Ω) and the fabricated GdPO4/RGO/GCE exhibited a remarkably low Rct (101 Ω). The DPV curve recorded on the GdPO4/RGO/GCE, exhibiting a rapid and sensitive response to FNT additions. Distinct and well-resolved reduction peaks of FNT were observed with broad linear ranges. The reproducibility and spiked real samples, such as river and tap water, were analyzed, demonstrating that the GdPO4/RGO/GCE holds potential and is significant for applications in the electrochemical sensing of agricultural pollutants.
Graphitic carbon nitride (g-C3N4) has emerged as a promising 2D platform for electrochemical sensing owing to its chemical stability, nitrogen-rich framework, and ability to form heterojunctions with metal oxides. Aravind Radha et al. [106] developed a sensor, CaZrO3@g-C3N4 modified glassy carbon electrode (CaZrO3@g-C3N4/GCE) for the detection of diethofencarb (DFC). Firstly, they were synthesized CaZrO3 by mixing Ca(NO3)2 and ZrOCl2 in 0.07 L of DI water and then adding KOH solution dropwise while stirring. The complete homogeneous solution was transferred to an autoclave and subjected to maintain temperature up to 100 °C for 24 h. The resulting product, known as CaZrO3, was calcined for 3 h at a temperature of around 1000 °C. CaZrO3 and g-C3N4 were combined in a 2:1 ratio and treated under ultrasonication. The resulting final composite (CaZrO3@g-C3N4) was washed and dried at 80 °C. EIS study of CaZrO3@g-C3N4 had higher conductivity and better electron transport performance, as evidenced by a significantly lower Rct (53.4 Ω cm2) than bare GCE (1085 Ω cm2). The CaZrO3@g-C3N4 electrode electrochemical activity was examined via the CV method. CV curve explores the high electrocatalytic activity of CaZrO3@g-C3N4@ observed by the high Ipa = 20.70 μA (Epa = 0.37 V) and low performance of bare GCE Ipa = 7.01 μA (Epa = 0.41 V). Furthermore, the influence of pH on DFC oxidation revealed optimal performance at neutral pH. DPV measurements demonstrated a wide linear range and low detection limit, with excellent selectivity, repeatability, and reproducibility in real food samples.
To detect malathion in vegetable samples, Waribam S. D. et al. [107] constructed a g-C3N4-supported CuO-derived biochar (B-CuO/g-C3N4). Initially, 10 g of melamine monomer was heated to 550 °C to produce a pale-yellow g-C3N4. Then, the CuO material was synthesized by combining 0.5 g of CuCl2·2H2O with 0.05 g of urea, followed by the addition of NH4OH and sonication. After that, the resulting mixture was then subjected to hydrothermal treatment for 12 h at 150 °C. Finally, the B-CuO/g-C3N4 composite was then mechanically blended with biochar and heated to 600 °C for 4 h under an inert nitrogen atmosphere. The electrocatalytic performance of B-CuO/g-C3N4/GCE was analyzed, revealing a lower peak potential separation (ΔEp = 0.324 V) compared to g-C3N4/GCE, which exhibited a higher peak separation (ΔEp = 0.366 V). Square wave anodic stripping voltammetry (SWASV) analysis showed that the malathion peak current increased proportionally with concentration, exhibiting good linearity and a low detection limit. The reproducibility of the SWASV measurement was experimentally assessed by modifying the electrode 5 times under identical working conditions, with an RSD of approximately 9.69 ± 0.32% (n = 5), demonstrating sensor accuracy. To assess the practical feasibility, the sensor was applied for the detection of malathion in tomato and apple extracts, yielding recovery percentages ranging from 87.64 to 120.59%.
In an investigation by Sladana Durdic et al. [108], they developed Co-doped CeO2 nanoparticles decorated on the g-C3N4 surface, which was applied to find out fenitrothion (FNT) in real samples. The authors first synthesized g-C3N4 by melamine polymerization at 550 °C. For nanoparticle preparation via the hydrothermal method, CoCl2·6H2O and Ce(NO3)3·6H2O were dissolved in 100 mL distilled water with 3% NH3(aq) added dropwise under stirring for 1 h, then combined with a 50 mL g-C3N4 suspension (0.03 g mL−1) and subjected to magnetic stirring followed by 2 h of sonication [108]. CV analysis of GCp/g-C3N4@Co-doped CeO2 revealed a superior Ipa compared to other modified and bare electrodes. EIS study, GCp/g-C3N4@Co-doped CeO2 demonstrated superior electron transport capability, with a significantly lower electron-transfer resistance (Rct) (2874 Ω) compared to bare GCE (6081 Ω). The bare GCp electrode exhibited poor electrochemical performance, with a ΔEp of 531 mV and low intensity peak currents (Ipa = 34.6 μA, Ipc = −28.8 μA). On the other hand, the GCp/g-C3N4@Co-doped CeO2 sensor showed ΔEp of 252 mV, and the superior current intensity of both Ipa and Ipc peaks were 42.6 μA and −38.4 μA, respectively. SWV responses demonstrate that the electrode GCp/g-C3N4@Co-doped CeO2 successfully detects the FNT pesticide in an extensive concentration range of 0.01–13.7 μM, and determined LOQ and LOD were 0.0097 μM and 0.0032 μM, respectively. The suggested method’s RSD for FNT concentrations of 0.08 μM, 0.5 μM, and 4.5 μM were 2.16%, 0.83%, and 0.85% respectively. Additionally, the standard addition method was used for effective sensing of FNT in real samples, such as apple and tap water. These studies collectively demonstrate that synergistic interactions between metal oxides and 2D carbon materials, rather than the intrinsic activity of individual components, govern sensor performance.

7.3. Two-Dimensional-Carbon Derivatives with Molecularly Imprinted Polymer-Based Sensing

Molecularly imprinted polymers (MIPs) integrated with 2D-Carbon derivatives provide outstanding electrochemical benefits for pesticide sensing, primarily due to their synergistic effects on sensitivity, selectivity, and practical applicability. The 2D-Carbon with MIPs creates highly specific recognition cavities that match the perfect size, shape, and active functional groups of pesticide molecules, enabling selective targeting even among structurally similar agrochemicals. In this context, a CdMoO4/g-C3N4 based molecularly imprinted electrochemical sensor was developed by Mehmet Lutfi Yola [109] through a combination of material synthesis, electrode fabrication, and analytical application to achieve ultrasensitive carbendazim detection. 2D-g-C3N4 was first synthesized by thermal condensation of melamine, after which CdMoO4 microspheres were uniformly grown on 2D-g-C3N4 sheets using a one-pot in situ hydrothermal method, forming a heterojunction nanocomposite with enhanced charge-transfer properties. The nanocomposite was drop-cast onto a polished glassy carbon electrode, followed by electropolymerization of pyrrole in the presence of carbendazim as a template to form a molecularly imprinted polymer layer; subsequent template removal generated selective recognition cavities on the electrode surface. Electrochemical characterization revealed that the CdMoO4/g-C3N4 modification significantly reduced charge-transfer resistance and improved electron-transfer kinetics, resulting in a wide linear detection range 1.0 × 10−5–1.0 × 10−3 µM and an ultralow detection limit of 2.5 × 10−6 µM. The sensor exhibited excellent selectivity against interfering pesticides, high stability, reproducibility, and reusability, and was successfully applied to real apple and orange juice samples with recoveries close to 100%, confirming minimal matrix effects. Hence, the synergistic integration of CdMoO4 with 2D-g-C3N4 and the molecular imprinting strategy enabled the construction of a robust, sensitive, and selective electrochemical platform, proving strong potential for practical pesticide monitoring and food safety applications.
Similarly, Ayman H. Kamel et al. [110] synthesized 2D-rGO with MIPs using an emulsion photopolymerization method with imidacloprid as the template, methacrylic acid as the functional monomer, ethylene glycol dimethacrylate as the cross-linker, and benzoyl peroxide as the initiator towards the detection of imidacloprid. Soxhlet extraction was then used to remove the template and create specific recognition cavities, and a non-imprinted polymers were prepared for comparison. A glassy carbon electrode was first modified with 2D-rGO to serve as a solid-contact ion-to-electron transducer, and then a PVC-based sensing membrane containing the MIP beads, plasticizer, and ionic additive was cast onto the 2D-rGO layer to fabricate an all-solid-state potentiometric sensor. The incorporation of 2D-rGO significantly enhanced potential stability, minimized water-layer formation, reduced signal drift, and improved durability, while the MIP layer provided high molecular selectivity toward imidacloprid. The resulting sensor exhibited a wide linear detection range from 0.5 to 1000 µM, and a low detection limit of 0.2 µM. The sensor demonstrated strong selectivity against structurally related neonicotinoids and common interferents, as well as high repeatability and long-term stability. Furthermore, the sensor was successfully utilized to the direct measurement of imidacloprid in commercial pesticide formulations, with recoveries ranging from 94.5% to 106.4%, demonstrating good agreement with HPLC results. A synergistic combination of MIP-based molecular recognition and 2D-rGO-based solid-contact transduction resulted in the development of a cost-effective, sensitive, selective, and stable potentiometric sensing platform appropriate for routine agro-industrial pesticide monitoring and quality control.

7.4. Two-Dimensional-Carbon Derivatives with Enzyme-Based Biosensing

Enzyme-based biosensing platforms integrated with 2D-carbon derivatives offer significant electrochemical benefits for pesticide detection. Enzymes such as acetylcholinesterase, organophosphorus hydrolase, and laccase exhibit high biochemical specificity toward pesticides. In contrast, 2D-Carbon derivatives serve as excellent conductive matrices with large electroactive surface areas, promoting rapid electron transfer between the enzyme active sites and the electrode. This results in enhanced current response and lower detection limits. Especially in 2D-g-C3N4, being nitrogen-rich semiconducting materials, supply abundant surface functional groups that facilitate stable enzyme immobilization through electrostatic interactions and hydrogen bonding, preserve enzymatic activity, and enable strong interaction of pesticide molecules.
Kai Zhang et al. [111] synthesized reduced graphene oxide-chitosan (rGO-CHI) by chemically reducing graphene oxide in an acidic chitosan medium with ascorbic acid to obtain a conductive and biocompatible composite, onto which methamidophos was immobilized as a recognition molecule due to its strong binding affinity toward AChE. A polished glassy carbon electrode was modified by drop-casting the rGO-CHI composite, followed by methamidophos immobilization and bovine serum albumin blocking to prevent nonspecific adsorption, resulting in an electrochemical biosensor based on the AChE inhibition principle. The addition of rGO-CHI greatly improved electron transfer, but the interaction between AChE and organophosphorus pesticides increased charge-transfer resistance, resulting in sensitive current fluctuations that were measured by DPV. The biosensor demonstrated outstanding electrochemical performance, with low detection limits ranging from 0.05 to 0.52 ng mL−1, remarkable reproducibility and stability, and the ability to detect various organophosphorus and carbamate pesticides across a wide concentration range. Therefore, the synergistic combination of rGO conductivity, chitosan biocompatibility, and methamidophos-mediated AChE inhibition enabled the development of a highly sensitive, stable, and broad-spectrum electrochemical biosensing platform suitable for rapid and practical pesticide monitoring in food safety applications.
Similarly, Sehrish Bilal et al. [112] prepared a NiCr2O4/g-C3N4-based electrochemical AChE biosensor heterostructured composite through an in situ chemical reaction. NiCr2O4 particles were uniformly deposited onto g-C3N4 sheets, yielding a high-surface area, conductive, and biocompatible transducer material. The composite was immobilized onto pencil graphite electrodes, functionalized with carboxyl groups, and activated using EDC/NHS chemistry to enable covalent attachment of AChE enzymes extracted from three insect species (Apis mellifera, Tribolium castaneum, and Zootermopsis nevadensis). Electrochemical characterization confirmed enhanced electron-transfer kinetics after composite modification and successful enzyme immobilization, while the resulting bioelectrodes exhibited strong electrocatalytic oxidation of thiocholine and sensitive inhibition responses toward malathion. The biosensors showed species-dependent analytical performance with wide linear ranges and nanomolar detection limits, along with good stability, reproducibility, and selectivity. Practical applicability was demonstrated through the successful detection of malathion in spiked wheat flour samples using the Apis mellifera AChE-based biosensor, which achieved high recovery values comparable to HPLC analysis. The synergistic combination of NiCr2O4 and g-C3N4 provided an efficient enzyme-immobilization matrix and signal-amplifying platform. This highlights the strong potential of such metal oxide g-C3N4 composites for food safety monitoring, sensitive pesticide toxicity assessment, and environmental analysis.

7.5. Structure–Property–Performance Relationships

Across the reported studies, clear structure–property–performance relationships can be identified. Extended 2D architecture gives abundant electroactive sites and facilitates rapid electron transport, leading to lower LOD. Heterojunction formation between 2D carbons and metal oxides reduces interfacial resistance and enhances signal stability. Metal doping and composite engineering introduce additional redox-active sites, improving sensitivity and widening the linear detection range. Meanwhile, surface modification with aptamers enables highly selective biosensing platforms with least interference effects.

7.6. Practical Applicability and Outlook

A key advantage of 2D carbon-based electrochemical sensors lies in their demonstrated applicability to real-world samples. Reliable detection of pesticide residues has been achieved in wastewater, tap and river water, and food matrices such as tomato, rice, apple, grape, and leafy vegetables, with acceptable recovery values and low relative standard deviation, as summarized in previous publications [99,100,101,102,103,104,105,106,107]. The malathion sensor based on B-CuO/g- C3N4 further highlights this applicability, where the synthesis route, electrochemical behaviour, and SWASV responses collectively demonstrate high sensitivity and real-sample feasibility [107].
Similarly, 2D carbon materials have become appealing options for direct electrochemical sensing and reliable templates for the production of 2D carbon-based electrochemical sensing materials. These materials have the potential to improve the sensitivity of electroanalytical procedures for sensing pharmaceuticals, environmental contaminants, and electrochemical energy storage systems. Since the existence of these hazardous pesticide compounds poses exposure and potential bioaccumulation dangers in organisms, it is crucial to develop innovative ways for monitoring and controlling their concentrations. As a result, establishing a sensitive and efficient pesticide detection method is crucial for environmental monitoring and improvement (recent improvements were noted in Table 1). Furthermore, platforms such as g-C3N4 nanosheet modified electrodes have made it possible to detect different residues in environmental and food matrices quickly, reliably, and without interference, which supports more efficient environmental monitoring and food safety (Table 2). The exceptional analytical capability demonstrated by these 2D carbon-driven electrochemical sensors is significantly relevant to today’s environmental and food safety assessment. The intrinsic advantages of 2D materials, particularly their large surface area, abundant functional active sites, rapid charge transport, and customizable surface chemistry support the fabrication of sensors with remarkably low LOD, often reaching the ng mL−1 level [104]. Such sensitivity is essential for monitoring pesticide residues in water as well as agricultural products, since regulatory policies limits require trace-level detection to ensure public and environmental health. Several recent reviews have emphasized that graphene, rGO, g-C3N4, and related 2D materials possess remarkable electrocatalytic performance and fast electron-transfer kinetics, making them ideal for monitoring pollutants such as neonicotinoids, organophosphates, carbamates, heavy metals, and antibiotics in a wide range of environmental matrices [113,114,115]. Additionally, the use of electrochemical techniques such as differential pulse voltammetry (DPV), square wave voltammetry (SWV), and square wave anodic stripping voltammetry (SWASV) facilitates the development of cost-effective, miniaturized, and portable sensing devices that demand minimal instrumentation is an essential feature for real-time and field-deployable analysis. The successful application of these 2D carbon-based sensors in complex real-world matrices, including wastewater, tomato, rice, apple, and tap/river water extracts (Refs. [97,99,100,102]), further demonstrates their stability, excellent recovery rates, and low RSD values. These findings confirm their ability to function efficiently without requiring extensive laboratory-based preprocessing. Overall, the ease of portability, analytical reliability, and strong real-sample compatibility of 2D-carbon-based sensors make them as powerful tools for quick hazard assessment, environmental monitoring, and ensuring food safety, thus significantly contributing to global health as well as environmental protection efforts.

8. Conclusions and Future Perspective

The determination of pesticides is essential because their consumption by humans creates serious health issues, such as neurological disorders and cancer, and also pollutes soil, water, and air. Two-dimensional carbon materials, including graphene, rGO, graphdiyne, and g-C3N4, as well as their mixed nanocomposites, have shown remarkable electrocatalytic properties in the electrochemical detection of hazardous contaminants, biomolecules, and various heavy metals. The significant sensing ability of 2D carbon materials is due to their high sensitivity, selectivity, vast surface area, adjustable defects, excellent electrical conductivity, and rapid analysis. Hence, these materials meet the demands in food safety and environmental monitoring applications. In addition, 2D carbon-based materials also show remarkable detecting efficiency towards pesticides, with a dynamic range and low limits of detection supported by their distinctive physicochemical characteristics.
Here, we thoroughly reviewed the recent developments in the electrochemical determination of pesticides using 2D-carbon materials as an electrocatalyst. We also discussed the toxicity of various pesticides. The synthesis, properties, and advantages of 2D carbon materials are discussed in detail. The fabrication of various electrochemical strategies towards pesticide detection is elaborately addressed. The electrochemical sensing methodology, the versatility of 2D carbon materials, selectivity, and real sample analysis of pesticides are properly discussed. Although electrochemical sensors have demonstrated a great deal of promise for pesticide detection, their in situ environmental application is still difficult. Analytical dependability is generally limited by complex sample matrices in natural streams and agricultural runoff, which frequently result in interference, electrode fouling, and decreased selectivity. Furthermore, in variable field circumstances like temperature, pH, and ionic strength, many sensors have poor long-term stability and signal drift. Even though reports of ultra-low detection limits are common, on-site applications still struggle to strike a balance between high sensitivity and real-time monitoring. Large-scale implementation is further hindered by problems with reproducibility and calibration, and continuous monitoring is hampered by limited integration with portable and wireless systems. Therefore, the development of strong antifouling electrode materials, highly selective recognition techniques, standardized fabrication procedures, and clever data-processing methods should be the main emphasis of future research. Transitioning electrochemical sensors from lab research to useful in situ pesticide monitoring will require a focus on sensor miniaturization, field validation, and comparison with traditional analytical methods.
Future research should focus on the simultaneous detection of multiple pesticides using a single electrochemical technology. It will be a challenging task to find a remarkable 2D carbon-based electrocatalyst that can detect multiple target molecules simultaneously. Furthermore, it would be essential to identify the effect of environmental conditions, including temperature, pH, and humidity, on the reliability and accuracy of the sensors. To create reliable, field-ready sensors, it is important to study functionalization techniques, composite creation, and device engineering. In addition, enhancing selectivity, developing economical and sustainable synthesis techniques, integrating sensors with digital platforms for real-time monitoring, and encouraging extensive testing in environmental and agricultural settings are some future prospects.

Author Contributions

K.I.: Conceptualization, Data curation, Investigation, Writing—original draft. A.A.: Writing—original draft, Visualization. G.V.P.: Writing—original draft, Investigation, Writing—review and editing. Y.V.M.R.: Formal analysis, Visualization. L.I.G.: Formal analysis, Visualization. J.W.: Formal analysis, Visualization. T.H.K.: Supervision, Resources, Funding acquisition, Project administration, Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Research Foundation of Korea (NRF) grant funded by the Korea Government (MOE and MSIT) (RS-2021-NR060121 and RS-2025–00559158). This research was supported by Korea Basic Science Institute (National research Facilities and Equipment Center) grant funded by the Ministry of Education (RS-2022-NF000922). This research was supported by the ‘regional innovation mega project’ program through the Korea Innovation Foundation funded by the Ministry of Science and ICT (2023-DD-UP-0007). This work was also supported by the Soonchunhyang University research fund.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

During the preparation of this manuscript, the authors used ChatGPT (OpenAI, https://openai.com/) and Gemini (Google, https://gemini.google.com/app) for language editing and improving the clarity of the text. The authors reviewed and edited the output and take full responsibility for the content of this publication.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miglione, A.; Raucci, A.; Mancini, M.; Gioia, V.; Frugis, A.; Cinti, S. An electrochemical biosensor for on-glove application: Organophosphorus pesticide detection directly on fruit peels. Talanta 2025, 283, 127093. [Google Scholar] [CrossRef]
  2. Ejigu, A.; Tefera, M.; Guadie, A. Electrochemical detection of pesticides: A comprehensive review on voltammetric determination of malathion, 2,4-D, carbaryl, and glyphosate. Electrochem. Commun. 2024, 169, 107839. [Google Scholar] [CrossRef]
  3. Bilge, S.; Aftab, S.; Donar, Y.O.; Özoylumlu, B.; Sınağ, A. Waste biomass derived carbon-based materials for electrochemical detection of pesticides from real samples. Inorg. Chem. Commun. 2025, 171, 113649. [Google Scholar] [CrossRef]
  4. Tanabe, S.; Madhusree, B.; Öztürk, A.A.; Tatsukawa, R.; Miyazaki, N.; Özdamar, E.; Aral, O.; Samsun, O.; Öztürk, B. Persistent organochlorine residues in harbour porpoise (Phocoena phocoena) from the Black Sea. Mar. Pollut. Bull. 1997, 34, 338–347. [Google Scholar] [CrossRef]
  5. Devi, P.I.; Thomas, J.; Raju, R.K. Pesticide consumption in India: A spatiotemporal analysis. Agric. Econ. Res. Rev. 2017, 30, 163. [Google Scholar] [CrossRef]
  6. Rajan, S.; Parween, M.; Raju, N.J. Pesticides in the hydrogeo-environment: A review of contaminant prevalence, source and mobilisation in India. Environ. Geochem. Health 2023, 45, 5481–5513. [Google Scholar] [CrossRef]
  7. P, S.; Thasale, R.; Kumar, D.; Mehta, T.G.; Limbachiya, R. Human health risk assessment of pesticide residues in vegetable and fruit samples in Gujarat State, India. Heliyon 2022, 8, e10876. [Google Scholar] [CrossRef]
  8. Nath, U.; Puzari, A.; Jamir, T. Toxicological assessment of synthetic pesticides on physiology of Phaseolus vulgaris L. and Pisum sativum L. along with their correlation to health hazards: A case study in south-west Nagaland, India. J. Saudi Soc. Agric. Sci. 2024, 23, 300–311. [Google Scholar] [CrossRef]
  9. Chen, Y.; Yang, Z.; Nian, B.; Yu, C.; Maimaiti, D.; Chai, M.; Yang, X.; Zang, X.; Xu, D. Mechanisms of Neurotoxicity of Organophosphate Pesticides and Their Relation to Neurological Disorders. Neuropsychiatr. Dis. Treat. 2024, 20, 2237–2254. [Google Scholar] [CrossRef]
  10. Botitsi, H.V.; Garbis, S.D.; Economou, A.; Tsipi, D.F. Current mass spectrometry strategies for the analysis of pesticides and their metabolites in food and water matrices. Mass Spectrom. Rev. 2011, 30, 907–939. [Google Scholar] [CrossRef]
  11. Parzanese, C.; Lancioni, C.; Castells, C. Gas-liquid chromatography as a tool to determine vapor pressure of low-volatility pesticides: A critical study. Anal. Chim. Acta 2025, 1354, 343932. [Google Scholar] [CrossRef]
  12. Vodova, M.; Nejdl, L.; Pavelicova, K.; Zemankova, K.; Rrypar, T.; Skopalova Sterbova, D.; Bezdekova, J.; Nuchtavorn, N.; Macka, M.; Adam, V.; et al. Detection of pesticides in food products using paper-based devices by UV-induced fluorescence spectroscopy combined with molecularly imprinted polymers. Food Chem. 2022, 380, 132141. [Google Scholar] [CrossRef]
  13. Piskin, E.; Alakus, Z.; Budak, F.; Cetinkaya, A.; Ozkan, S.A. Nanoparticle-supported electrochemical sensors for pesticide analysis in fruit juices. J. Pharm. Biomed. Anal. Open 2025, 5, 100056. [Google Scholar] [CrossRef]
  14. Ding, L.; Guo, J.; Chen, S.; Wang, Y. Electrochemical sensing mechanisms of neonicotinoid pesticides and recent progress in utilizing functional materials for electrochemical detection platforms. Talanta 2024, 273, 125937. [Google Scholar] [CrossRef]
  15. Wu, Y.; Ali, S.; White, R.J. Electrocatalytic Mechanism for Improving Sensitivity and Specificity of Electrochemical Nucleic Acid-Based Sensors with Covalent Redox Tags—Part I. ACS Sens. 2020, 5, 3833–3841. [Google Scholar] [CrossRef]
  16. Yuan, H.; Li, N.; Fan, W.; Cai, H.; Zhao, D. Metal-Organic Framework Based Gas Sensors. Adv. Sci. 2022, 9, 2104374. [Google Scholar] [CrossRef]
  17. Nunes, E.W.; Silva, M.K.L.; Rascon, J.; Leiva-Tafur, D.; Lapa, R.M.L.; Cesarino, I. Acetylcholinesterase Biosensor Based on Functionalized Renewable Carbon Platform for Detection of Carbaryl in Food. Biosensors 2022, 12, 486. [Google Scholar] [CrossRef]
  18. Saqib, M.; Solomonenko, A.N.; Barek, J.; Dorozhko, E.V.; Korotkova, E.I.; Aljasar, S.A. Graphene derivatives-based electrodes for the electrochemical determination of carbamate pesticides in food products: A review. Anal. Chim. Acta 2023, 1272, 341449. [Google Scholar] [CrossRef]
  19. Vrabelj, T.; Finšgar, M. Recent Progress in Non-Enzymatic Electroanalytical Detection of Pesticides Based on the Use of Functional Nanomaterials as Electrode Modifiers. Biosensors 2022, 12, 263. [Google Scholar] [CrossRef]
  20. Zhang, L.; Zhao, M.; Xiao, M.; Im, M.H.; El-Aty, A.M.A.; Shao, H.; She, Y. Recent Advances in the Recognition Elements of Sensors to Detect Pyrethroids in Food: A Review. Biosensors 2022, 12, 402. [Google Scholar] [CrossRef]
  21. Zhou, Z.-D.; Li, S.-Q.; Liu, Y.; Du, B.; Shen, Y.-Y.; Yu, B.-Y.; Wang, C.-C. Two bis-ligand-coordinated Zn(ii)-MOFs for luminescent sensing of ions, antibiotics and pesticides in aqueous solutions. RSC Adv. 2022, 12, 7780–7788. [Google Scholar] [CrossRef]
  22. Wu, Y.; Zhang, X.; Liu, C.; Tian, L.; Zhang, Y.; Zhu, M.; Qiao, W.; Wu, J.; Yan, S.; Zhang, H.; et al. Adsorption Behaviors and Mechanism of Phenol and Catechol in Wastewater by Magnetic Graphene Oxides: A Comprehensive Study Based on Adsorption Experiments, Mathematical Models, and Molecular Simulations. ACS Omega 2024, 9, 15101–15113. [Google Scholar] [CrossRef]
  23. Allen, M.J.; Tung, V.C.; Kaner, R.B. Honeycomb Carbon: A Review of Graphene. Chem. Rev. 2010, 110, 132–145. [Google Scholar] [CrossRef]
  24. Zhao, Y.; Li, X.; Zhou, X.; Zhang, Y. Review on the graphene based optical fiber chemical and biological sensors. Sens. Actuators B Chem. 2016, 231, 324–340. [Google Scholar] [CrossRef]
  25. Urade, A.R.; Lahiri, I.; Suresh, K.S. Graphene Properties, Synthesis and Applications: A Review. JOM 2023, 75, 614–630. [Google Scholar] [CrossRef]
  26. Zafar, M.A.; Liu, Y.; Hernandez, F.C.R.; Varghese, O.K.; Jacob, M.V. Plasma-Based Synthesis of Freestanding Graphene from a Natural Resource for Sensing Application. Adv. Mater. Interfaces 2023, 10, 2202399. [Google Scholar] [CrossRef]
  27. Gao, N.; Tan, R.; Cai, Z.; Zhao, H.; Chang, G.; He, Y. A novel electrochemical sensor via Zr-based metal organic framework–graphene for pesticide detection. J. Mater. Sci. 2021, 56, 19060–19074. [Google Scholar] [CrossRef]
  28. Karadurmus, L.; Cetinkaya, A.; Kaya, S.I.; Ozkan, S.A. Recent trends on electrochemical carbon-based nanosensors for sensitive assay of pesticides. Trends Environ. Anal. Chem. 2022, 34, e00158. [Google Scholar] [CrossRef]
  29. Amin, A.; Prasad, G.V.; Vinothkumar, V.; Jang, S.J.; Oh, D.E.; Kim, T.H. Reduced Graphene Oxide/β-Cyclodextrin Nanocomposite for the Electrochemical Detection of Nitrofurantoin. Chemosensors 2025, 13, 247. [Google Scholar] [CrossRef]
  30. Liu, L.; Zhang, J.; Zhao, J.; Liu, F. Mechanical properties of graphene oxides. Nanoscale 2012, 4, 5910. [Google Scholar] [CrossRef]
  31. Ozcelikay, G.; Karadurmus, L.; Bilge, S.; Sınağ, A.; Ozkan, S.A. New analytical strategies Amplified with 2D carbon nanomaterials for electrochemical sensing of food pollutants in water and soils sources. Chemosphere 2022, 296, 133974. [Google Scholar] [CrossRef]
  32. Makhsin, S.R. Physical Properties of Graphene. JMechE 2023, 12, 225–267. [Google Scholar] [CrossRef]
  33. Novoselov, K.S.; Fal′ko, V.I.; Colombo, L.; Gellert, P.R.; Schwab, M.G.; Kim, K. A roadmap for graphene. Nature 2012, 490, 192–200. [Google Scholar] [CrossRef]
  34. Nurunnabi, M.; Parvez, K.; Nafiujjaman, M.; Revuri, V.; Khan, H.A.; Feng, X.; Lee, Y. Bioapplication of graphene oxide derivatives: Drug/gene delivery, imaging, polymeric modification, toxicology, therapeutics and challenges. RSC Adv. 2015, 5, 42141–42161. [Google Scholar] [CrossRef]
  35. Nag, A.; Mitra, A.; Mukhopadhyay, S.C. Graphene and its sensor-based applications: A review. Sens. Actuators A Phys. 2018, 270, 177–194. [Google Scholar] [CrossRef]
  36. Zafar, M.A.; Waligo, D.; Varghese, O.K.; Jacob, M.V. Advances in graphene-based electrochemical biosensors for on-site pesticide detection. Front. Carbon 2023, 2, 1325970. [Google Scholar] [CrossRef]
  37. Ragumoorthy, C.; Nataraj, N.; Chen, S.-M.; Kiruthiga, G. Yttrium iron oxide embedded reduced graphene oxide: A trace level detection platform for carbamate pesticide in agricultural products. Process Saf. Environ. Prot. 2025, 196, 106875. [Google Scholar] [CrossRef]
  38. Kherra, N.; Mosbah, A.; Baaziz, H.; Zerriouh, A.; Amusa, H.K.; Darwish, A.S.; Lemaoui, T.; Banat, F.; AlNashef, I.M. Machine learning and quantum-driven multiscale design of graphene-based 2D adsorbents for enhanced aqueous pesticide removal. J. Environ. Chem. Eng. 2025, 13, 118336. [Google Scholar] [CrossRef]
  39. Chen, D.; Bai, W.; Deng, H.; Li, Z.; Chen, L.; Liu, S.; Zhang, K.; Wang, C.; Zheng, S.; Wang, S. Graphene oxide-based dual-color thin-film fluorescent nanolabels for enhanced immunochromatographic detection of pesticides. Anal. Chim. Acta 2025, 1372, 344466. [Google Scholar] [CrossRef]
  40. Venkata Prasad, G.; Vinothkumar, V.; Joo Jang, S.; Eun Oh, D.; Hyun Kim, T. Multi-walled carbon nanotube/graphene oxide/poly(threonine) composite electrode for boosting electrochemical detection of paracetamol in biological samples. Microchem. J. 2023, 184, 108205. [Google Scholar] [CrossRef]
  41. Razaq, A.; Bibi, F.; Zheng, X.; Papadakis, R.; Jafri, S.H.M.; Li, H. Review on Graphene-, Graphene Oxide-, Reduced Graphene Oxide-Based Flexible Composites: From Fabrication to Applications. Materials 2022, 15, 1012. [Google Scholar] [CrossRef]
  42. Mirres, A.C.D.M.; Silva, B.E.P.D.M.D.; Tessaro, L.; Galvan, D.; Andrade, J.C.D.; Aquino, A.; Joshi, N.; Conte-Junior, C.A. Recent Advances in Nanomaterial-Based Biosensors for Pesticide Detection in Foods. Biosensors 2022, 12, 572. [Google Scholar] [CrossRef]
  43. Wu, S.; Hao, J.; Yang, S.; Sun, Y.; Wang, Y.; Zhang, W.; Mao, H.; Song, X.-M. Layer-by-layer self-assembly film of PEI-reduced graphene oxide composites and cholesterol oxidase for ultrasensitive cholesterol biosensing. Sens. Actuators B Chem. 2019, 298, 126856. [Google Scholar] [CrossRef]
  44. Alonso, A.V.; Murugesan, A.; Gogoi, R.; Chandrabose, S.; Abass, K.S.; Sharma, V.; Kandhavelu, M. Graphitic carbon nitride nanoparticle: G-C3N4 synthesis, characterization, and its biological activity against glioblastoma. Eur. J. Pharmacol. 2025, 1003, 177999. [Google Scholar] [CrossRef]
  45. Elhaddad, E.; Al-fawwaz, A.T.; Rehan, M. An effective stannic oxide/graphitic carbon nitride (SnO2 NPs@g-C3N4) nanocomposite photocatalyst for organic pollutants degradation under visible-light. J. Saudi Chem. Soc. 2023, 27, 101677. [Google Scholar] [CrossRef]
  46. Ye, M.; Li, B.; Zhang, K.; Yan, R.; Dai, L.; Zhang, S.; Li, Y.; Hu, J.; Yang, B.; Yao, Y. Application of graphitic carbon nitride (g-C3N4) in solid polymer electrolytes: A mini review. Electrochem. Commun. 2025, 176, 107939. [Google Scholar] [CrossRef]
  47. Wudil, Y.S.; Ahmad, U.F.; Gondal, M.A.; Al-Osta, M.A.; Almohammedi, A.; Sa’id, R.S.; Hrahsheh, F.; Haruna, K.; Mohamed, M.J.S. Tuning of graphitic carbon nitride (g-C3N4) for photocatalysis: A critical review. Arab. J. Chem. 2023, 16, 104542. [Google Scholar] [CrossRef]
  48. Basivi, P.K.; Selvaraj, Y.; Perumal, S.; Bojarajan, A.K.; Lin, X.; Girirajan, M.; Kim, C.W.; Sangaraju, S. Graphitic carbon nitride (g–C3N4)–Based Z-scheme photocatalysts: Innovations for energy and environmental applications. Mater. Today Sustain. 2025, 29, 101069. [Google Scholar] [CrossRef]
  49. Qamar, M.A.; Javed, M.; Shahid, S.; Shariq, M.; Fadhali, M.M.; Ali, S.K.; Shak, M. Synthesis and applications of graphitic carbon nitride (g-C3N4) based membranes for wastewater treatment: A critical review. Heliyon 2023, 9, e12685. [Google Scholar] [CrossRef]
  50. Klini, A.; Kokkotos, S.; Vasilaki, E.; Vamvakaki, M. Graphitic Carbon Nitride (G-C3n4) Optical Sensors for Environmental Monitoring. Diam. Relat. Mater. 2025, 155, 112301. [Google Scholar] [CrossRef]
  51. Mortazavi, B.; Shahrokhi, M.; Shojaei, F.; Rabczuk, T.; Zhuang, X. A machine learning-assisted exploration of the structural stability, electronic, optical, heat conduction and mechanical properties of C3N4 graphitic carbon nitride monolayers. Comput. Mater. Today 2025, 5, 100024. [Google Scholar] [CrossRef]
  52. Pourmadadi, M.; Rahmani, E.; Eshaghi, M.M.; Shamsabadipour, A.; Ghotekar, S.; Rahdar, A.; Romanholo Ferreira, L.F. Graphitic carbon nitride (g-C3N4) synthesis methods, surface functionalization, and drug delivery applications: A review. J. Drug Deliv. Sci. Technol. 2023, 79, 104001. [Google Scholar] [CrossRef]
  53. Gajarajan, S.; Kuchi, C.; Reddy, A.O.; Krishna, K.G.; Kumar, K.S.; Naresh, B.; Sreedhara Reddy, P. Graphitic carbon nitride (g-C3N4) decorated on CuCo2O4 nanosphere composites for enhanced electrochemical performance for energy storage applications. Mater. Today Commun. 2024, 39, 108688. [Google Scholar] [CrossRef]
  54. Zia, R.; Nazir, M.S.; Rouf, H.; Ahmad, N.; Ahmad, F.; Ali, Z.; Hassan, S.U. Modified lignin etched sulphur-doped graphitic carbon nitride for electrochemical sensing of cypermethrin in drinking water: Experimental and theoretical insights. Int. J. Biol. Macromol. 2025, 321, 146503. [Google Scholar] [CrossRef]
  55. Ilager, D.; Shetti, N.P.; Foucaud, Y.; Badawi, M.; Aminabhavi, T.M. Graphene/g-carbon nitride (GO/g-C3N4) nanohybrids as a sensor material for the detection of methyl parathion and carbendazim. Chemosphere 2022, 292, 133450. [Google Scholar] [CrossRef] [PubMed]
  56. Kumar, J.V.; Rhim, J. Fluorescence Probe for Detection of Malathion Using Sulfur Quantum Dots@Graphitic-Carbon Nitride Nanocomposite. Luminescence 2025, 40, e70112. [Google Scholar] [CrossRef]
  57. Ahmad, N.; Kumar, A.; Rachchh, N.; Jyothi, S.R.; Bhanot, D.; Kumari, B.; Kumar, A.; Abosaoda, M.K. Developing a highly sensitive electrochemical sensor for malathion detection based on green g-C3N4 @LiCoO2 nanocomposites. RSC Adv. 2025, 15, 3378–3388. [Google Scholar] [CrossRef]
  58. Razzaque, S.; Abubakar, M.; Farid, M.A.; Zia, R.; Nazir, S.; Razzaque, H.; Ali, A.; Ali, Z.; Mahmood, A.; Al-Masry, W.; et al. Detection of toxic cypermethrin pesticides in drinking water by simple graphitic electrode modified with Kraft lignin@Ni@g-C3 N4 nano-composite. J. Mater. Chem. B 2024, 12, 9364–9374. [Google Scholar] [CrossRef]
  59. Shi, Y.; Niu, Z.; Rodas-Gonzalez, A.; Wang, S.; Wang, Y.; Yang, L. An electrochemical adsorption-catalysis sensor based on 2D MXene/g-C3N4 doped with Au-COOH for selective detection of GMP in meat. Chem. Eng. J. 2024, 500, 155893. [Google Scholar] [CrossRef]
  60. Otoh, E.F.; Odey, M.O.; Martin, O.I.; Agurokpon, D.C. In silico engineering of graphitic carbon nitride nanostructures through germanium mono-doping and codoping with transition metals (Ni, Pd, Pt) as sensors for diazinon organophosphorus pesticide pollutants. BMC Chem. 2025, 19, 78. [Google Scholar] [CrossRef]
  61. Li, G.; Li, Y.; Liu, H.; Guo, Y.; Li, Y.; Zhu, D. Architecture of graphdiyne nanoscale films. Chem. Commun. 2010, 46, 3256. [Google Scholar] [CrossRef]
  62. Nidhi, H.V.; Koppad, V.S.; Babu, A.M.; Varghese, A. Properties, Synthesis and Emerging Applications of Graphdiyne: A Journey Through Recent Advancements. Top. Curr. Chem. 2024, 382, 19. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, T.; Jin, Z. Application of graphdiyne for promote efficient photocatalytic hydrogen evolution. Results Surf. Interfaces 2024, 14, 100188. [Google Scholar] [CrossRef]
  64. Chen, S.; Xue, Y.; Li, Y. 2D graphdiyne, what’s next? Next Mater. 2023, 1, 100031. [Google Scholar] [CrossRef]
  65. Mortazavi, B.; Shojaei, F.; Zhuang, X. A first-principles investigation on the structural, stability and mechanical properties of novel Ag, Au and Cu-graphdiyne magnetic semiconducting monolayers. Nano Trends 2023, 4, 100021. [Google Scholar] [CrossRef]
  66. Mortazavi, B.; Shahrokhi, M.; Zhuang, X.; Rabczuk, T. Boron–graphdiyne: A superstretchable semiconductor with low thermal conductivity and ultrahigh capacity for Li, Na and Ca ion storage. J. Mater. Chem. A 2018, 6, 11022–11036. [Google Scholar] [CrossRef]
  67. Liu, R.; Liu, H.; Li, Y.; Yi, Y.; Shang, X.; Zhang, S.; Yu, X.; Zhang, S.; Cao, H.; Zhang, G. Nitrogen-doped graphdiyne as a metal-free catalyst for high-performance oxygen reduction reactions. Nanoscale 2014, 6, 11336–11343. [Google Scholar] [CrossRef]
  68. Wang, Z.; Wang, H.; Zhang, X.; Yuan, Y.; Wang, L.; Liu, J.; Chen, C. Graphdiyne-based nanomaterials: Synthesis, properties, and biomedical applications. ChemPhysMater 2025, 4, 207–233. [Google Scholar] [CrossRef]
  69. Li, S.; Zhao, Y.; Wang, D. Progress in graphdiyne-based membrane for gas separation and water purification. ChemPhysMater 2025, 4, 234–250. [Google Scholar] [CrossRef]
  70. Li, Y. Special Issue on “Fundament and Application of 2D Graphdiyne”. ChemPhysMater 2025, 4, 205–206. [Google Scholar] [CrossRef]
  71. Gao, X.; Liu, H.; Wang, D.; Zhang, J. Graphdiyne: Synthesis, properties, and applications. Chem. Soc. Rev. 2019, 48, 908–936. [Google Scholar] [CrossRef]
  72. Zhao, L.-X.; Fan, Y.-G.; Zhang, X.; Li, C.; Cheng, X.-Y.; Guo, F.; Wang, Z.-Y. Graphdiyne biomaterials: From characterization to properties and applications. J. Nanobiotechnol. 2025, 23, 169. [Google Scholar] [CrossRef]
  73. Vaidyanathan, A.; Wagh, V.; Chakraborty, B. Enhanced hydrogen storage in graphdiyne through compressive strain: Insights from density functional theory simulations. J. Energy Storage 2025, 117, 116153. [Google Scholar] [CrossRef]
  74. Ai, Y.; Yan, L.; Zhang, S.; Ye, X.; Xuan, Y.; He, S.; Wang, X.; Sun, W. Ultra-sensitive simultaneous electrochemical detection of Zn(II), Cd(II) and Pb(II) based on the bismuth and graphdiyne film modified electrode. Microchem. J. 2023, 184, 108186. [Google Scholar] [CrossRef]
  75. Song, D.; Xu, X.; Huang, X.; Li, G.; Wang, X.; Zhao, Y.; Gao, F. Nitrogen and fluoride co-doped graphdiyne with metal-organic framework (MOF)-derived NiCo2O4–Co3O4 nanocages as sensing layers for ultra-sensitive pesticide detection. Anal. Chim. Acta 2023, 1252, 341012. [Google Scholar] [CrossRef]
  76. Shi, J.; Liu, S.; Li, P.; Lin, Y.; Luo, H.; Wu, Y.; Yan, J.; Huang, K.-J.; Tan, X. Self-powered dual-mode sensing strategy based on graphdiyne and DNA nanoring for sensitive detection of tumor biomarker. Biosens. Bioelectron. 2023, 237, 115557. [Google Scholar] [CrossRef] [PubMed]
  77. Hummers, W.S.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  78. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved Synthesis of Graphene Oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  79. Cao, K.; Tian, Z.; Zhang, X.; Wang, Y.; Zhu, Q. Green preparation of graphene oxide nanosheets as adsorbent. Sci. Rep. 2023, 13, 9314. [Google Scholar] [CrossRef]
  80. Stankovich, S.; Dikin, D.A.; Piner, R.D.; Kohlhaas, K.A.; Kleinhammes, A.; Jia, Y.; Wu, Y.; Nguyen, S.T.; Ruoff, R.S. Synthesis of graphene-based nanosheets via chemical reduction of exfoliated graphite oxide. Carbon 2007, 45, 1558–1565. [Google Scholar] [CrossRef]
  81. Shin, H.J.; Kim, K.K.; Benayad, A.; Yoon, S.; Park, H.K.; Jung, I.; Jin, M.H.; Jeong, H.; Kim, J.M.; Choi, J.; et al. Efficient Reduction of Graphite Oxide by Sodium Borohydride and Its Effect on Electrical Conductance. Adv. Funct. Mater. 2009, 19, 1987–1992. [Google Scholar] [CrossRef]
  82. Fernández-Merino, M.J.; Guardia, L.; Paredes, J.I.; Villar-Rodil, S.; Solís-Fernández, P.; Martínez-Alonso, A.; Tascón, J.M.D. Vitamin C Is an Ideal Substitute for Hydrazine in the Reduction of Graphene Oxide Suspensions. J. Phys. Chem. C 2010, 114, 6426–6432. [Google Scholar] [CrossRef]
  83. Abdolhosseinzadeh, S.; Asgharzadeh, H.; Seop Kim, H. Fast and fully-scalable synthesis of reduced graphene oxide. Sci. Rep. 2015, 5, 10160. [Google Scholar] [CrossRef]
  84. Wang, G.; Yang, J.; Park, J.; Gou, X.; Wang, B.; Liu, H.; Yao, J. Facile Synthesis and Characterization of Graphene Nanosheets. J. Phys. Chem. C 2008, 112, 8192–8195. [Google Scholar] [CrossRef]
  85. Wu, Z.-S.; Ren, W.; Gao, L.; Liu, B.; Jiang, C.; Cheng, H.-M. Synthesis of high-quality graphene with a pre-determined number of layers. Carbon 2009, 47, 493–499. [Google Scholar] [CrossRef]
  86. Larciprete, R.; Fabris, S.; Sun, T.; Lacovig, P.; Baraldi, A.; Lizzit, S. Dual Path Mechanism in the Thermal Reduction of Graphene Oxide. J. Am. Chem. Soc. 2011, 133, 17315–17321. [Google Scholar] [CrossRef] [PubMed]
  87. Sengupta, I.; Chakraborty, S.; Talukdar, M.; Pal, S.K.; Chakraborty, S. Thermal reduction of graphene oxide: How temperature influences purity. J. Mater. Res. 2018, 33, 4113–4122. [Google Scholar] [CrossRef]
  88. Klemeyer, L.; Park, H.; Huang, J. Geometry-Dependent Thermal Reduction of Graphene Oxide Solid. ACS Mater. Lett. 2021, 3, 511–515. [Google Scholar] [CrossRef]
  89. Valentini, C.; Montes-García, V.; Livio, P.A.; Chudziak, T.; Raya, J.; Ciesielski, A.; Samorì, P. Tuning the electrical properties of graphene oxide through low-temperature thermal annealing. Nanoscale 2023, 15, 5743–5755. [Google Scholar] [CrossRef] [PubMed]
  90. Wei, D.; Liu, Y.; Wang, Y.; Zhang, H.; Huang, L.; Yu, G. Synthesis of N-Doped Graphene by Chemical Vapor Deposition and Its Electrical Properties. Nano Lett. 2009, 9, 1752–1758. [Google Scholar] [CrossRef]
  91. Pei, S.; Cheng, H.-M. The reduction of graphene oxide. Carbon 2012, 50, 3210–3228. [Google Scholar] [CrossRef]
  92. Sun, X.; Duan, M.; Li, R.; Meng, Y.; Bai, Q.; Wang, L.; Liu, M.; Yang, Z.; Zhu, Z.; Sui, N. Ultrathin Graphdiyne/Graphene Heterostructure as a Robust Electrochemical Sensing Platform. Anal. Chem. 2022, 94, 13598–13606. [Google Scholar] [CrossRef] [PubMed]
  93. Zhu, G.; Zhang, C.; Zhang, C.; Yi, Y. Electroless deposition of platinum on graphdiyne for electrochemical sensing. Chem. Eng. J. 2024, 497, 154512. [Google Scholar] [CrossRef]
  94. Xue, Y.; Wang, X.; Sun, B.; Wang, L.; Guo, X. Construction of an electrochemical sensor for the detection of methyl parathion with three-dimensional graphdiyne-carbon nanotubes. Microchim. Acta 2025, 192, 77. [Google Scholar] [CrossRef]
  95. Li, J.; Han, X.; Wang, D.; Zhu, L.; Ha-Thi, M.; Pino, T.; Arbiol, J.; Wu, L.; Nawfal Ghazzal, M. A Deprotection-free Method for High-yield Synthesis of Graphdiyne Powder with In Situ Formed CuO Nanoparticles. Angew. Chem. Int. Ed. 2022, 61, e202210242. [Google Scholar] [CrossRef]
  96. Zhu, J.; Xiao, P.; Li, H.; Carabineiro, S.A.C. Graphitic Carbon Nitride: Synthesis, Properties, and Applications in Catalysis. ACS Appl. Mater. Interfaces 2014, 6, 16449–16465. [Google Scholar] [CrossRef] [PubMed]
  97. Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Müller, J.-O.; Schlögl, R.; Carlsson, J.M. Graphitic carbon nitride materials: Variation of structure and morphology and their use as metal-free catalysts. J. Mater. Chem. 2008, 18, 4893. [Google Scholar] [CrossRef]
  98. Xu, J.; Wu, H.-T.; Wang, X.; Xue, B.; Li, Y.-X.; Cao, Y. A new and environmentally benign precursor for the synthesis of mesoporous g-C3N4 with tunable surface area. Phys. Chem. Chem. Phys. 2013, 15, 4510. [Google Scholar] [CrossRef]
  99. Dong, F.; Wu, L.; Sun, Y.; Fu, M.; Wu, Z.; Lee, S.C. Efficient synthesis of polymeric g-C3N4 layered materials as novel efficient visible light driven photocatalysts. J. Mater. Chem. 2011, 21, 15171. [Google Scholar] [CrossRef]
  100. Chen, X.; Zhang, J.; Fu, X.; Antonietti, M.; Wang, X. Fe-g-C3N4 -Catalyzed Oxidation of Benzene to Phenol Using Hydrogen Peroxide and Visible Light. J. Am. Chem. Soc. 2009, 131, 11658–11659. [Google Scholar] [CrossRef]
  101. Li, Y.; Wang, M.-Q.; Bao, S.-J.; Lu, S.; Xu, M.; Long, D.; Pu, S. Tuning and thermal exfoliation graphene-like carbon nitride nanosheets for superior photocatalytic activity. Ceram. Int. 2016, 42, 18521–18528. [Google Scholar] [CrossRef]
  102. Abega, A.V.; Marchal, C.; Dziurla, M.-A.; Dantio, N.C.B.; Robert, D. Easy three steps gC3N4 exfoliation for excellent photocatalytic activity—An in-depth comparison with conventional approaches. Mater. Chem. Phys. 2023, 304, 127803. [Google Scholar] [CrossRef]
  103. Kumar, J.V.; Elkaffas, R.A.; Eissa, S. Unleashing the power of MnMoO4/rGO nanocomposite towards the electrochemical aptasensing of neurotoxic pesticide fenitrothion. Electrochim. Acta 2025, 536, 146708. [Google Scholar] [CrossRef]
  104. Almenhali, A.Z.; Kanagavalli, P.; Abd-Ellah, M.; Khazaal, S.; El Darra, N.; Eissa, S. Reduced graphene oxide-based electrochemical aptasensor for the multiplexed detection of imidacloprid, thiamethoxam, and clothianidin in food samples. Sci. Rep. 2025, 15, 10329. [Google Scholar] [CrossRef] [PubMed]
  105. Mariyappan, V.; Sundaresan, R.; Chen, S.-M.; Ramachandran, R.; Al-Sehemi, A.G.; Jeevika, A.; Wu, W. Constructing a novel electrochemical sensor for the detection of fenitrothion using rare-earth orthophosphate incorporated reduced graphene oxide composite. Process Saf. Environ. Prot. 2024, 185, 726–738. [Google Scholar] [CrossRef]
  106. Radha, A.; Wang, S.-F. Construction of calcium zirconate nanoparticles confined on graphitic carbon nitride: An electrochemical detection of diethofencarb in food samples. Food Chem. 2025, 465, 141929. [Google Scholar] [CrossRef]
  107. Devi, W.S.; Kaur, R.; Sharma, A.; Thakur, S.; Mehta, S.K.; Raja, V.; Ataya, F.S. Non-enzymatic g-C3N4 supported CuO derived-biochar based electrochemical sensors for trace level detection of malathion. Biosens. Bioelectron. 2025, 267, 116808. [Google Scholar] [CrossRef]
  108. Đurđić, S.; Vlahović, F.; Ognjanović, M.; Gemeiner, P.; Sarakhman, O.; Stanković, V.; Mutić, J.; Stanković, D.; Švorc, Ľ. Nano-size cobalt-doped cerium oxide particles embedded into graphitic carbon nitride for enhanced electrochemical sensing of insecticide fenitrothion in environmental samples: An experimental study with the theoretical elucidation of redox events. Sci. Total Environ. 2024, 909, 168483. [Google Scholar] [CrossRef]
  109. Yola, M.L. Carbendazim imprinted electrochemical sensor based on CdMoO4/g-C3N4 nanocomposite: Application to fruit juice samples. Chemosphere 2022, 301, 134766. [Google Scholar] [CrossRef]
  110. Kamel, A.H.; Abd-Rabboh, H.S.M. Imprinted polymer/reduced graphene oxide-modified glassy carbon electrode-based highly sensitive potentiometric sensing module for imidacloprid detection. Microchem. J. 2024, 197, 109789. [Google Scholar] [CrossRef]
  111. Zhang, K.; Wang, L.; Liu, J.; Zhao, S.; Ding, L.; Zhou, B.; Zhao, B. A Highly Sensitive Electrochemical Methamidophos Immobilized AChE Biosensor for Organophosphorus Pesticides Detection. Int. J. Electrochem. Sci. 2022, 17, 220443. [Google Scholar] [CrossRef]
  112. Bilal, S.; Nasir, M.; Hassan, M.M.; Rehman, M.F.U.; Sami, A.J.; Hayat, A. A novel construct of an electrochemical acetylcholinesterase biosensor for the investigation of malathion sensitivity to three different insect species using a NiCr2 O4/g-C3 N4 composite integrated pencil graphite electrode. RSC Adv. 2022, 12, 16860–16874. [Google Scholar] [CrossRef]
  113. Li, T.; Shang, D.; Gao, S.; Wang, B.; Kong, H.; Yang, G.; Shu, W.; Xu, P.; Wei, G. Two-Dimensional Material-Based Electrochemical Sensors/Biosensors for Food Safety and Biomolecular Detection. Biosensors 2022, 12, 314. [Google Scholar] [CrossRef]
  114. Venegas, C.J.; Bollo, S.; Sierra-Rosales, P. Carbon-Based Electrochemical (Bio)sensors for the Detection of Carbendazim: A Review. Micromachines 2023, 14, 1752. [Google Scholar] [CrossRef] [PubMed]
  115. Lazarević-Pašti, T. Carbon Materials for Organophosphate Pesticide Sensing. Chemosensors 2023, 11, 93. [Google Scholar] [CrossRef]
  116. Nehru, R.; Hsu, Y.-F.; Wang, S.-F.; Dong, C.-D.; Govindasamy, M.; Habila, M.A.; AlMasoud, N. Graphene oxide@Ce-doped TiO2 nanoparticles as electrocatalyst materials for voltammetric detection of hazardous methyl parathion. Microchim. Acta 2021, 188, 216. [Google Scholar] [CrossRef]
  117. Xiao, B. Electrochemical sensor based on Bimetallic phosphosulfide Zn–Ni–P–S Nanocomposite -reduced graphene oxide for determination of Paraoxon Ethyl in agriculture wastewater. Int. J. Electrochem. Sci. 2022, 17, 220672. [Google Scholar] [CrossRef]
  118. Weheabby, S.; Liu, Z.; Pašti, I.A.; Rajić, V.; Vidotti, M.; Kanoun, O. Enhanced electrochemical sensing of methyl parathion using AgNPs@IL/GO nanocomposites in aqueous matrices. Nanoscale Adv. 2025, 7, 2195–2208. [Google Scholar] [CrossRef] [PubMed]
  119. Tan, X.; Liu, Y.; Zhang, T.; Luo, S.; Liu, X.; Tian, H.; Yang, Y.; Chen, C. Ultrasensitive electrochemical detection of methyl parathion pesticide based on cationic water-soluble pillar[5]arene and reduced graphene nanocomposite. RSC Adv. 2019, 9, 345–353. [Google Scholar] [CrossRef] [PubMed]
  120. Ceylan, E.; Ozoglu, O.; Huseyin Ipekci, H.; Tor, A.; Uzunoglu, A. Non-enzymatic detection of methyl parathion in water using CeO2-CuO-decorated reduced graphene oxide. Microchem. J. 2024, 199, 110261. [Google Scholar] [CrossRef]
  121. Alsulami, A.; Kumarswamy, Y.K.; Prashanth, M.K.; Hamzada, S.; Lakshminarayana, P.; Pradeep Kumar, C.B.; Jeon, B.-H.; Raghu, M.S. Fabrication of FeVO4/RGO Nanocomposite: An Amperometric Probe for Sensitive Detection of Methyl Parathion in Green Beans and Solar Light-Induced Degradation. ACS Omega 2022, 7, 45239–45252. [Google Scholar] [CrossRef]
  122. Kaur, R.; Rana, S.; Kaur, K.; Navodita. Ultra-sensitive detection of methylparathion using palladium nanoparticles embedded reduced graphene oxide redox nanoreactors. Microchem. J. 2023, 191, 108871. [Google Scholar] [CrossRef]
  123. Shanmugam, R.; Manavalan, S.; Chen, S.-M.; Keerthi, M.; Lin, L.-H. Methyl Parathion Detection Using SnS2/N, S–Co-Doped Reduced Graphene Oxide Nanocomposite. ACS Sustain. Chem. Eng. 2020, 8, 11194–11203. [Google Scholar] [CrossRef]
  124. Tan, C.; Liu, X.; Yang, K.; Li, J.; Yin, Y.; Zuo, J. A novel electrochemical sensor based on NiO/MoS2/rGO composite material for rapid detection of methyl parathion. Microchim. Acta 2025, 192, 444. [Google Scholar] [CrossRef]
  125. Chen, Z.; Zhang, Y.; Yang, Y.; Shi, X.; Zhang, L.; Jia, G. Hierarchical nitrogen-doped holey graphene as sensitive electrochemical sensor for methyl parathion detection. Sens. Actuators B Chem. 2021, 336, 129721. [Google Scholar] [CrossRef]
  126. Nayem, N.I.; Hossain, M.S.; Rashed, M.A.; Anis-Ul-Haque, K.M.; Ahmed, J.; Faisal, M.; Algethami, J.S.; Harraz, F.A. A sensitive and selective electrochemical detection and kinetic analysis of methyl parathion using Au nanoparticle-decorated rGO/CuO ternary nanocomposite. RSC Adv. 2025, 15, 15348–15365. [Google Scholar] [CrossRef]
  127. Weheabby, S.; You, S.; Pašti, I.A.; Al-Hamry, A.; Kanoun, O. Direct electrochemical detection of pirimicarb using graphene oxide and ionic liquid composite modified by gold nanoparticles. Measurement 2024, 236, 115132. [Google Scholar] [CrossRef]
  128. Tao, T.; Zhou, Y.; Ma, M.; He, H.; Gao, N.; Cai, Z.; Chang, G.; He, Y. Novel graphene electrochemical transistor with ZrO2/rGO nanocomposites functionalized gate electrode for ultrasensitive recognition of methyl parathion. Sens. Actuators B Chem. 2021, 328, 128936. [Google Scholar] [CrossRef]
  129. Sattar, Z.; Fatima, N.; Anjum, S.; Qureshi, N.; Khan, M.I.; Shanableh, A.; Ahmad, F.; Qureshi, H.O.A.; Luque, R. Efficient CeO2/NiO/GO nanocomposites for the detection of toxic carbofuran pesticide in real samples. Results Eng. 2025, 25, 104352. [Google Scholar] [CrossRef]
  130. Anna, M.V.S.S.; Carvalho, S.W.M.M.; Gevaerd, A.; Silva, J.O.S.; Santos, E.; Carregosa, I.S.C.; Wisniewski, A.; Marcolino-Junior, L.H.; Bergamini, M.F.; Sussuchi, E.M. Electrochemical sensor based on biochar and reduced graphene oxide nanocomposite for carbendazim determination. Talanta 2020, 220, 121334. [Google Scholar] [CrossRef]
  131. Srinivasan, S.; Nesakumar, N.; Rayappan, J.B.B.; Kulandaiswamy, A.J. Electrochemical Detection of Imidacloprid Using Cu–rGO Composite Nanofibers Modified Glassy Carbon Electrode. Bull. Environ. Contam. Toxicol. 2020, 104, 449–454. [Google Scholar] [CrossRef]
  132. Luo, J.; Li, S.; Wu, Y.; Pang, C.; Ma, X.; Wang, M.; Zhang, C.; Zhi, X.; Li, B. Electrochemical sensor for imidacloprid detection based on graphene oxide/gold nano/β-cyclodextrin multiple amplification strategy. Microchem. J. 2022, 183, 107979. [Google Scholar] [CrossRef]
  133. Zhao, Y.; Zheng, X.; Wang, Q.; Zhe, T.; Bai, Y.; Bu, T.; Zhang, M.; Wang, L. Electrochemical behavior of reduced graphene oxide/cyclodextrins sensors for ultrasensitive detection of imidacloprid in brown rice. Food Chem. 2020, 333, 127495. [Google Scholar] [CrossRef]
  134. Majidi, M.R.; Ghaderi, S. Facile fabrication and characterization of silver nanodendrimers supported by graphene nanosheets: A sensor for sensitive electrochemical determination of Imidacloprid. J. Electroanal. Chem. 2017, 792, 46–53. [Google Scholar] [CrossRef]
  135. Paula, S.A.; Ferreira, O.A.E.; César, P.A. Determination of Imidacloprid Based on the Development of a Glassy Carbon Electrode Modified with Reduced Graphene Oxide and Manganese (II) Phthalocyanine. Electroanalysis 2020, 32, 86–94. [Google Scholar] [CrossRef]
  136. Pimalai, D.; Putnin, T.; Bamrungsap, S. A highly sensitive electrochemical sensor based on poly(3-aminobenzoic acid)/graphene oxide-gold nanoparticles modified screen printed carbon electrode for paraquat detection. J. Environ. Sci. 2025, 148, 139–150. [Google Scholar] [CrossRef] [PubMed]
  137. Yassin, J.M.; Taddesse, A.M.; Tsegaye, A.A.; Sánchez-Sánchez, M. CdS/g-C3N4/Sm-BDC MOF nanocomposite modified glassy carbon electrodes as a highly sensitive electrochemical sensor for malathion. Appl. Surf. Sci. 2024, 648, 158973. [Google Scholar] [CrossRef]
  138. Kapoor, A.; Rajput, J.K. A prompt electrochemical monitoring platform for sensitive and selective determination of thiamethoxam using Fe2O3@g-C3N4@MSB composite modified glassy carbon electrode. J. Food Compos. Anal. 2023, 115, 105033. [Google Scholar] [CrossRef]
  139. Shanbhag, M.M.; Kalanur, S.S.; Pandiaraj, S.; Alodhayb, A.N.; Pollet, B.G.; Shetti, N.P. Synthesis of graphitic carbon nitride (g-C3N4) by high-temperature condensation for electrochemical evaluation of dichlorophen and thymol in environmental trials. Surf. Interfaces 2024, 50, 104440. [Google Scholar] [CrossRef]
  140. Sakthinathan, S.; Keyan, A.K.; Vasu, D.; Vinothini, S.; Nagaraj, K.; Mangesh, V.L.; Chiu, T.-W. Graphitic Carbon Nitride Incorporated Europium Molybdate Composite as an Enhanced Sensing Platform for Electrochemical Detection of Carbendazim in Agricultural Products. J. Electrochem. Soc. 2022, 169, 127504. [Google Scholar] [CrossRef]
  141. Maji, B.; Achary, L.S.K.; Barik, B.; Jyotsna Sahoo, S.; Mohanty, A.; Dash, P. MnCo2O4 decorated (2D/2D) rGO/g-C3N4-based Non-Enzymatic sensor for highly selective and sensitive detection of Chlorpyrifos in water and food samples. J. Electroanal. Chem. 2022, 909, 116115. [Google Scholar] [CrossRef]
  142. Mansouri, I.; Meribai, A.; Çakar, S.; Ünlü, B.; Özacar, M.; Bessekhouad, Y. Non-enzymatic electrochemical sensor for selective detection of glyphosate in real water samples using SrxZn1−xO/g-C3N4 composite electrode. Mater. Chem. Phys. 2025, 344, 131159. [Google Scholar] [CrossRef]
  143. Zhang, J.; Meng, J.; Li, W.; Gu, K. Investigating the sensitivity and selectivity of electrochemical sensors based on nanomaterials for the detection of carbendazim residue in cherry wine. Food Meas. 2024, 18, 5376–5385. [Google Scholar] [CrossRef]
  144. Dogra, S.; Rajput, J.K. Ultrasensitive electrochemical sensor for nitroaromatics using C3N4-MoS2-Au nanocomposite with pendimethalin as a real sample. J. Environ. Chem. Eng. 2025, 13, 116618. [Google Scholar] [CrossRef]
  145. Xiao, F.; Li, H.; Yan, X.; Yan, L.; Zhang, X.; Wang, M.; Qian, C.; Wang, Y. Graphitic carbon nitride/graphene oxide(g-C3N4/GO) nanocomposites covalently linked with ferrocene containing dendrimer for ultrasensitive detection of pesticide. Anal. Chim. Acta 2020, 1103, 84–96. [Google Scholar] [CrossRef]
Figure 1. A general overview of the classification of some major synthetic organo-pesticides and their typical applications.
Figure 1. A general overview of the classification of some major synthetic organo-pesticides and their typical applications.
Biosensors 16 00062 g001
Figure 2. Chemical structures of several hazardous pesticides.
Figure 2. Chemical structures of several hazardous pesticides.
Biosensors 16 00062 g002
Figure 3. (A) Preparation of MnMO4/rGO nanocomposite via hydrothermal and co-precipitation techniques. (B) Stepwise fabrication of the aptasensor designed for FNT recognition. (C) DPV responses of the modified aptasensor (MnMoO4/rGO-2:1/pyrene/EDC-NHS/Aptamer-modified electrode) after exposure to different FNT concentrations. (D) Corresponding calibration plot for FNT detection (curve of analytical response versus FNT concentration) (Ref. [103]).
Figure 3. (A) Preparation of MnMO4/rGO nanocomposite via hydrothermal and co-precipitation techniques. (B) Stepwise fabrication of the aptasensor designed for FNT recognition. (C) DPV responses of the modified aptasensor (MnMoO4/rGO-2:1/pyrene/EDC-NHS/Aptamer-modified electrode) after exposure to different FNT concentrations. (D) Corresponding calibration plot for FNT detection (curve of analytical response versus FNT concentration) (Ref. [103]).
Biosensors 16 00062 g003
Figure 4. Schematic diagram of the proposed multiplexed aptasensor (Ref. [104]).
Figure 4. Schematic diagram of the proposed multiplexed aptasensor (Ref. [104]).
Biosensors 16 00062 g004
Table 1. Recent applications of 2D carbon material-based electrochemical sensors for pesticide detection.
Table 1. Recent applications of 2D carbon material-based electrochemical sensors for pesticide detection.
2D-Carbon with CompositeTargeted Pesticide/ClassMethodLOD Linear Range Real SampleReference
GO@Ce-doped TiO2methyl parathionDPV0.0016 μM0.002–48.327 μMapple and tomato[116]
Zn-Ni-P-S/GO/GCEparaoxon ethylAMP0.035 μM1–200 μMagriculture wastewater[117]
AgNPs@GO/IL@SPCEmethyl parathionSWV9 μM25–200,000 μMground water and surface water[118]
CP5-rGO/GCEmethyl parathionDPV0.0003 μM0.001–150 μMsoil and tap water[119]
CeO2-CuO/rGOmethyl parathionDPV0.00179 μM0.0038–0.152 μMmineral, drinking, and tap water samples[120]
FeVO4/RGOmethyl parathionAMP0.00070 μM0.001–20 μMgreen beans[121]
RGO-CHIparathion
dichlorvos methamidophos trichlorfon
dimethoate
fenthion methomyl
chlorpyrifos omethoate diazinon
DPV0.05–0.52 ng mL−10–1500 ng mL−1NA[111]
Pd@NrGOmethyl parathionSWASV1.067 ng mL−1
122 ng mL−1
1.8–10 ng mL−1
50–1000 ng mL−1
river water, onion, agricultural run-off, cabbage, lettuce leaves
and tap water
[122]
SnS2/NS-RGOmethyl parathionCV0.00017 μM0.001–176 μMxindian river and black grapes[123]
YFO-rGO/RDEcarbofurani-t0.018 μM0.029–303.5 μMcarrots, cucumber, cabbage, spinach, tomato and potato.[37]
NiO/MoS2/Rgomethyl parathionDPV1.1 ng mL−110–10,000 ng mL−1apple juice[124]
N-HG50/GCEmethyl parathionDPV0.013 nM1 ng mL–150 μg mL−1apple, grape, cucumber and river water[125]
Au@rGO/CuO/GCEmethyl parathionSWV0.045 μM0.4–39.0 μMDhaleswari river[126]
AuNPs@GO/IL@SPCEpirimicarbSWV4.49 μM50–1500 μMgroundwater and surface water[127]
ZrO2/rGOmethyl parathionGECT10 pg mL−11 × 10−5–10 μg L−1Chinese cabbages[128]
CeO2/NiO/GOcarbofuranDPV0.82 μM5–150 μMpotato[129]
CPE-rGOcarbofuranDPV0.0023 μM0.03–0.8 μMdrinking water, lettuce leaves, orange juice, wastewater[130]
Cu-rGOimidaclopridCV0.003247 μMNAsoil sample from paddy field[131]
GO/Au NPs/β-CDimidaclopridDPV0.000133 μM5 × 10−4–0.3 μMChinese cabbage, banana and mango[132]
GCE/rGO/MIPimidaclopridPotentiometry0.8 μM1–1000 μMNA[110]
E-rGO/CDsimidaclopridCV0.02 μM0.5–40 μMbrown rice[133]
AgNDs/GNs/GCEimidaclopridDPV0.814 μM1–100 μMcucumber[134]
GCE/rGO/MnPcimidaclopridCV6.5 μM25–250 μMhoney[135]
SPCE/GO/AuNPs/P3ABAparaquatSWV4.5 × 10−4 μM0.001–100 μMtap water and natural water[136]
MnMoO4/rGOfenitrothionDPV3.0 × 10−4
ng mL−1
1.00 × 10−3–1.00 × 105 ng mL−1wastewater and rice extract{103]
GdPO4/RGOfenitrothionDPV0.007 μM0.01–342 μMriver and tap water[105]
rGO-aptamerimidacloprid thiamethoxam clothianidinDPV6.30 pg mL−1
6.80 pg mL−1
7.10 pg mL−1
0.01–100 ng mL−1
rice and tomato water[104]
Table 2. Recent applications of g-C3N4-based electrochemical sensors for pesticide detection.
Table 2. Recent applications of g-C3N4-based electrochemical sensors for pesticide detection.
2D-Carbon with CompositeTargeted Pesticide/ClassMethodLODLinear RangeReal SampleReference
KL@Ni@g-C3N4CypermethrinCV0.026 μg mL−10.05–0.2 μg mL−1Tap Water[58]
g-C3N4@LiCoO2MalathionDPV0.00438 μM0.005–0.12 μMLettuce[57]
GO/g-C3N4Methyl parathion
carbendazim
SWV8.4 × 10−4 μM
2.0 × 10−6 μM
0.08–100 μM
0.01–250 μM
Water and Soil sample[55]
CdS/g-C3N4/Sm-BDCMalathionDPV0.0074 μM0.03–0.15 µMCabbage[137]
B-CuO/g-C3N4MalathionSWASV1.2 pg mL−10.18–5.66 pg mL−1Soil, Rice, Water and Fruits[107]
g-C3N4@Co-doped CeO2FenitrothionSWV0.0032 μM0.01–13.70 μMTap water and Ethanolic apple extract[108]
CaZrO3@g-C3N4DiethofencarbDPV0.0018 μM0.01–230.04 µMStrawberry, Grapes, Spinach, and Apple[106]
MIP/CdMoO4/g- C3N4CarbendazimCV2.5 × 10−6 µM1.0 × 10−5–1.0 × 10−3 µMFruit juice[109]
Fe2O3@g-C3N4 @MSBThiamethoxamLSV 0.137 µM0.01–200 µMPotato, Rice and River water[138]
g-C3N4/CPEDichlorophen ThymolSWV0.012 µM0.05–100 µMSoil and Water[139]
g-C3N4/EuMoO4CarbendazimDPV0.04 µM50–400 µMApple and Tomato[140]
rGO-g-C3N4-MnCo2O4ChlorpyrifosDPV0.32 × 10−6 μg mL−11.5 × 10−5–7.0 µg mL−1Groundwater, Tap water and Pomegranate sample[141]
KL@Ni@S-g-C3N4CypermethrinCV0.05 µg mL−10.1–1.0 µg mL−1Tap water[54]
NiCr2O4/g-C3N4MalathionCV0.0023 µM0.002–0.10 µMWheat flour[112]
SrxZn1-xO/g-C3N4GlyphosateDPV1.4 × 10−4 µM0.002–0.1 µMLake Sapanca, Tap water and Natural Spring water.[142]
PTH-g-C3N4CarbendazimDPV3.7 × 10−4 µM0.1–85 µMCherry Wine[143]
C3N4-MoS2-AuPendimethalinLSV0.219 µM
0.615 µM 0.479 µM
0–1000 µM
0–500 µM
0–500 µM
Tap water and Distilled water[144]
g-C3N4/GO/Fc-TEDMetolcarbDPV0.0083 µM0.045–213 µMVegetable Spinach[145]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Imran, K.; Amin, A.; Venkata Prasad, G.; Veera Manohara Reddy, Y.; Gita, L.I.; Wilson, J.; Kim, T.H. Two-Dimensional Carbon-Based Electrochemical Sensors for Pesticide Detection: Recent Advances and Environmental Monitoring Applications. Biosensors 2026, 16, 62. https://doi.org/10.3390/bios16010062

AMA Style

Imran K, Amin A, Venkata Prasad G, Veera Manohara Reddy Y, Gita LI, Wilson J, Kim TH. Two-Dimensional Carbon-Based Electrochemical Sensors for Pesticide Detection: Recent Advances and Environmental Monitoring Applications. Biosensors. 2026; 16(1):62. https://doi.org/10.3390/bios16010062

Chicago/Turabian Style

Imran, K., Al Amin, Gajapaneni Venkata Prasad, Y. Veera Manohara Reddy, Lestari Intan Gita, Jeyaraj Wilson, and Tae Hyun Kim. 2026. "Two-Dimensional Carbon-Based Electrochemical Sensors for Pesticide Detection: Recent Advances and Environmental Monitoring Applications" Biosensors 16, no. 1: 62. https://doi.org/10.3390/bios16010062

APA Style

Imran, K., Amin, A., Venkata Prasad, G., Veera Manohara Reddy, Y., Gita, L. I., Wilson, J., & Kim, T. H. (2026). Two-Dimensional Carbon-Based Electrochemical Sensors for Pesticide Detection: Recent Advances and Environmental Monitoring Applications. Biosensors, 16(1), 62. https://doi.org/10.3390/bios16010062

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop