Next Article in Journal
Electron Scattering from NO2: Cross Sections in the Energy Range of 1–1000 eV
Previous Article in Journal
Toxicity of High-Density Polyethylene Nanoparticles in Combination with Silver Nanoparticles to Caco-2 and HT29MTX Cells Growing in 2D or 3D Culture
Previous Article in Special Issue
Atypical Analysis of a Graphite-Based Anode Prepared Using Aqueous Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Analytical Determination of Heavy Metals in Water Using Carbon-Based Materials

Department of Chemistry, Faculty of Natural Sciences and Geography, Abai Kazakh National Pedagogical University, 13, Dostyk Ave., Almaty 050010, Kazakhstan
*
Authors to whom correspondence should be addressed.
First author: Zhazira Mukatayeva.
Molecules 2026, 31(1), 5; https://doi.org/10.3390/molecules31010005
Submission received: 26 November 2025 / Revised: 11 December 2025 / Accepted: 12 December 2025 / Published: 19 December 2025

Abstract

This review presents a critical and comparative analysis of carbon-based electrochemical sensing platforms for the determination of heavy metal ions in water, with emphasis on Pb2+, Cd2+, and Hg2+. The growing discharge of industrial and mining effluents has led to persistent contamination of aquatic environments by toxic metals, creating an urgent need for sensitive, rapid, and field-deployable analytical technologies. Carbon-based nanomaterials, including graphene, carbon nanotubes (CNTs), and MXene, have emerged as key functional components in modern electrochemical sensors due to their high electrical conductivity, large surface area, and tunable surface chemistry. Based on reported studies, typical detection limits for Pb2+ and Cd2+ using differential pulse voltammetry (DPV) on glassy carbon and thin-film electrodes are in the range of 0.4–1.2 µg/L. For integrated thin-film sensing systems, limits of detection of 0.8–1.2 µg/L are commonly achieved. MXene-based platforms further enhance sensitivity and enable Hg2+ detection with linear response ranges typically between 1 and 5 µg/L, accompanied by clear electrochemical or optical signals. Beyond conventional electrochemical detection, this review specifically highlights self-sustaining visual sensors based on MXene integrated with enzyme-driven bioelectrochemical systems, such as glucose oxidase (GOD) and Prussian blue (PB) assembled on ITO substrates. These systems convert chemical energy into measurable colorimetric signals without external power sources, enabling direct visual identification of Hg2+ ions. Under optimized conditions (e.g., 5 mg/mL GOD and 5 mM glucose), stable and distinguishable color responses are achieved for rapid on-site monitoring. Overall, this review not only summarizes current performance benchmarks of carbon-based sensors but also identifies key challenges, including long-term stability, selectivity under multi-ion interference, and large-scale device integration, while outlining future directions toward portable multisensor water-quality monitoring systems.

Graphical Abstract

1. Introduction

Heavy metal contamination of water resources has become a critical global environmental problem due to the persistence, bioaccumulation, and high toxicity of these pollutants [1,2,3,4,5,6,7]. Industrial discharge, mining activities, domestic wastewater, and agricultural runoff continuously introduce toxic metal ions into aquatic environments, leading to long-term ecological instability and serious risks to human health [5,6]. Even at trace concentrations, heavy metals can induce oxidative stress, DNA damage, metabolic disorders, and irreversible organ dysfunction [7,8,9,10,11,12]. In this review, the term “heavy metals” specifically refers to Pb, Cd, and Hg, which are among the most hazardous and environmentally relevant toxic elements.
Recent toxicological studies further demonstrate that exposure to heavy metals induces a wide range of molecular, genetic, and epigenetic disturbances in humans and animals. Chromium(VI), arsenic, cadmium, and mercury have been shown to trigger DNA damage, chromosomal instability, oxidative stress, and disrupted immune responses, contributing to increased carcinogenic risk and metabolic disorders [13,14,15,16,17,18,19,20]. Long-term exposure even at low concentrations may alter antioxidant defense systems, impair renal and hepatic function, and interfere with neurodevelopment, particularly in vulnerable populations such as children and adolescents [21,22,23,24]. Moreover, chromium and arsenic compounds exhibit complex mechanisms of toxicity involving DNA adduct formation, aberrant methylation patterns, and protein damage, further exacerbating their hazardous impact on human health [25,26,27,28]. Together, these findings highlight the urgent need for sensitive, accurate, and rapid detection strategies to monitor heavy metal contamination in environmental water sources.
Reliable detection of heavy metals in water is essential for environmental monitoring, pollution control, and public health protection. Numerous analytical techniques have been developed for this purpose, including spectroscopic methods such as atomic spectroscopy and X-ray fluorescence (XRF) [29,30,31,32,33,34,35,36], chromatographic techniques such as high-performance liquid chromatography (HPLC) and ion chromatography (IC) [37,38], mass spectrometry-based methods such as inductively coupled plasma mass spectrometry (ICP-MS) [39], and optical methods including colorimetry, ultraviolet–visible spectroscopy (UV–Vis), and chemiluminescence (CL) [40,41,42]. Although these laboratory-based techniques offer excellent sensitivity and selectivity, they require expensive instrumentation, well-controlled laboratory environments, trained personnel, and complex sample pretreatment. As a result, they are poorly suited for rapid, low-cost, real-time, and on-site monitoring of heavy metals in field conditions.
Currently, the following methods are used to detect heavy metal ions in water resources (as shown in Figure 1):
Electrochemical detection techniques have emerged as highly attractive alternatives due to their inherent advantages, including high sensitivity, fast response, low cost, simple instrumentation, and excellent compatibility with portable and miniaturized devices [43]. Common electrochemical approaches for heavy metal detection include ion-selective electrodes (ISEs) [44,45,46,47], anodic stripping voltammetry (ASV) [48,49], polarographic analysis [50], and potentiometric stripping analysis (PSA) [51]. Despite their strong analytical performance, practical electrochemical sensing still faces several challenges, such as electrode surface fouling, limited reproducibility, matrix interference in complex water samples, long-term electrode stability, and the demand for low-power, field-deployable operation.
To overcome these limitations, carbon-based materials have been extensively explored as advanced electrode modifiers and sensing platforms. Materials such as graphene, multi-walled carbon nanotubes, and MXene exhibit outstanding electrical conductivity, large specific surface area, chemical stability, and tunable surface functionalization. These properties significantly enhance electron-transfer kinetics, adsorption capacity toward metal ions, and overall sensor sensitivity. Consequently, carbon-based electrochemical sensors offer a powerful strategy for achieving sensitive, stable, and portable detection of heavy metal ions in water. Furthermore, the rapid development of self-powered and visual sensing platforms based on carbon materials, particularly MXene-assisted bioelectrochemical systems, has opened new opportunities for in situ and on-site environmental monitoring without the need for external power sources. These systems integrate electrochemical energy conversion with optical signal output, enabling direct, rapid, and user-friendly detection of toxic metal ions.
Table 1 provides a summary of LOD and MDL values for Pb2+, Cd2+ and Hg2+ obtained using the developed electrochemical sensing platforms. LOD values were calculated using the standard equation LOD = 3σ/S, while MDL values were determined following the EPA statistical protocol. The obtained limits are significantly below WHO and EPA regulatory thresholds, confirming that the sensors are applicable for early-stage detection of heavy metal ions in water.
Table 2 provides a comparison of the analytical performance of the developed sensors with selected literature reports. The table summarizes electrode types, detection techniques, target metal ions, and corresponding LOD values. The results demonstrate that the bare GCE provides competitive sensitivity even without surface modification, while the MXene-based self-powered platform enables Hg2+ detection within environmentally relevant concentration ranges.
Table 3 presents key electrochemical performance parameters of the developed sensors, including linear range, sensitivity, repeatability (RSD%), and operational stability. The low RSD values (≤7%) indicate good repeatability, while stable sensor responses confirm suitability for portable and field-applicable monitoring of Pb2+, Cd2+ and Hg2+ ions.
In this review, we systematically summarize recent advances in carbon-based electrochemical sensors for the determination of Pb2+, Cd2+, and Hg2+ in water. Key detection mechanisms, analytical performance, and practical limitations are critically discussed. Special attention is given to portable, self-powered, and visual sensing systems, highlighting their potential for future field-applicable water monitoring technologies.

2. Development of Electrochemical Systems Based on Carbon Materials for the Detection of Heavy Metals

2.1. Conventional Carbon Nanomaterials

Carbon nanomaterials are widely used in the development of electrochemical sensors for the detection of heavy metals due to their excellent electrical conductivity, large specific surface area, and strong adsorption ability. Carbon nanotubes (CNTs) are generally classified into single-walled carbon nanotubes (SWCNTs) and multi-walled carbon nanotubes (MWCNTs). Compared to sp3 hybridization, the sp2 hybridization of carbon atoms provides CNTs with high tensile strength and outstanding electrical properties. The π-bonding formed by the overlap of p-electron orbitals enables fast electron transfer, while the large surface area enhances catalytic activity. These features make CNTs highly suitable for electrochemical detection of toxic heavy metals [52].
From a sensing mechanism perspective, the detection of heavy metal ions on carbon-based electrodes is governed by surface adsorption, electrostatic interaction, and redox reactions at the electrode–solution interface. Functional groups on carbon nanomaterials facilitate the accumulation of Pb2+, Cd2+, and Hg2+ ions during the preconcentration step. Subsequent anodic stripping produces distinct current peaks, enabling sensitive quantitative analysis.
Preconcentration of analytes on the electrode surface remains a key strategy for achieving low detection limits in portable electrochemical systems. EDTA-functionalized polymer–carbon nanocomposites, MOF-integrated carbon electrodes, and magnetic sorbent-assisted platforms have been shown to effectively capture Pb2+ and Cd2+ ions through coordination and adsorption interactions, resulting in improved stripping efficiency, lower detection limits, and enhanced repeatability [53,54,55,56,57].
Matlou et al. [58] reported that SWCNTs modified with polyaminobenzene sulfonic acid were successfully applied for the detection of Hg2+ in water using SWASV. The modified electrode showed improved sensitivity and a low LOD of 0.06 μM.
Wang et al. [59] reported that functionalized MWCNTs were employed as electrode modifiers for Pb2+ detection using DPASV. A good linear relationship between peak current and metal ion concentration was observed, with a LOD of 0.01 μM. These studies confirm that CNT-based electrochemical sensors exhibit high sensitivity and low detection limits for heavy metal ions.
Figure 2 illustrates the preparation process of the modified electrode and the mechanism of electroanalysis.
Graphene is a two-dimensional carbon material with exceptional electrical conductivity and mechanical strength, making it an ideal material for electrochemical sensor design. Owing to its large surface area, graphene facilitates high loading of active sites and rapid electron transfer. Therefore, graphene and its derivatives such as graphene oxide (GO), reduced graphene oxide (rGO), and graphene quantum dots (GQDs) are widely used in electrochemical sensing systems [60].
Guan Huanan et al. [61] developed a graphene-modified glassy carbon electrode for the detection of Hg2+ and Cu2+, achieving high sensitivity and low detection limits. Chen Xi et al. [62] synthesized nitrogen-doped graphene for heavy metal detection and reported a LOD of 2.0 nM for Cd2+ and 0.1 nM for Pb2+. Priya et al. [63] applied a porous rGO/AuNP composite for simultaneous detection of Pb2+ and Cd2+, achieving detection limits of 2 ng/L and 3 ng/L, respectively.
Xu He et al. [64] synthesized graphene quantum dots and developed a modified GCE, with LODs of 16.1 nM for Pb2+ and 35.7 nM for Cd2+. These results demonstrate the high electrochemical activity of graphene-based materials in trace heavy metal detection (Figure 2). Furthermore, the interaction mechanism between Cd2+/Pb2+ ions and the PrGO/AuNPs/Sal-Cys-modified electrode is illustrated in Figure 3, which provides a clearer understanding of the adsorption and electrochemical stripping behavior on graphene-based hybrid surfaces.
C60 is an important carbon isomer composed entirely of sp2-hybridized carbon atoms. C60 tends to accept electrons and, under certain conditions, donate them, exhibiting rich electrochemical properties [65]. Like other carbon nanomaterials, C60 is widely used as an electrode material. Han et al. [66] developed an ASV-based method for the detection of heavy metal ions using C60. They utilized C60 as an electrode modifier and prepared a novel, highly sensitive modified electrode (C60-chitosan/GCE) for electrochemical detection of heavy metal ions based on C60 and chitosan. This method features a low detection limit, good repeatability, and excellent performance for simultaneous detection of copper, mercury, lead, and cadmium.

2.2. Carbon Materials Derived from MOFs

New metal–organic frameworks (MOFs), consisting of metal ions and multifunctional organic ligands, possess several advantages such as high BET-specific surface area, rich metal–organic content, large pore volume, and the ability to tune their structure and composition. These features make them promising self-sacrificing templates and precursors for the synthesis of various carbon nanomaterials [67]. Compared to other carbon-based catalysts, carbon nanomaterials derived from MOFs can serve as direct catalysts or catalyst supports for many important reactions due to their tunable morphology, high layered porosity, and ease of functionalization with other dopant atoms or metals/metal oxides [68,69]. Tan [70] and co-authors combined an amino-functionalized zirconium-based MOF (NH2-UiO-66) with a zeolitic imidazolate framework based on zinc (ZIF-8) to develop a core–shell hybrid material (NH2-UiO-66@ZIF-8, NU66@Z8). The NU66@Z8, combined with carboxylated multi-walled carbon nanotubes (CMWCNTs), was deposited on a GCE to fabricate an electrochemical sensor for the detection of lead and copper ions. Under optimized conditions, the detection limit for lead using this sensor was 1 nM.

2.3. Biomass-Based Carbon Materials

Compared with conventional fossil-derived carbon materials, biomass-derived carbon materials exhibit abundant surface functional groups, hierarchical porosity, and low production cost [71]. Such materials are typically prepared through carbonization and activation of agricultural waste and plant residues.
The use of biomass materials not only helps to effectively address the complex production process of traditional carbon materials but also reduces waste, which holds significant practical importance for energy conservation, emission reduction, and environmental protection. Biomaterials are mainly converted into porous carbon materials through a combination of carbonization and chemical activation. Impurities in the biomolecular framework can be removed from the biomass through high-temperature carbonization in an inert gas atmosphere, resulting in the formation of a porous carbon skeleton [72,73]. Due to the wide variety of biomass sources, diverse biomaterials such as persimmon branches [74], tea waste [75], corn cobs [76], cork oak [77], eggplants [78], palm shells [79], etc., can be used to prepare biomass-based porous carbon materials. Figure 4 shows a carbon-based biomaterial synthesized from tea waste using a simple method that combines carbonization and activation [80]. Zhang et al. [81] reported a biomass-derived porous carbon prepared from grapefruit peel, which showed excellent electrochemical performance for the detection of Pb2+, Cd2+, and Cu2+ using SWASV. The sensor exhibited high sensitivity, good repeatability, and low detection limits. These results indicate that biomass-based carbon materials are promising sustainable alternatives for electrochemical heavy metal detection.

2.4. Electrochemical Sensing Mechanisms of Carbon-Based Materials

The electrochemical detection of heavy metal ions on carbon-based electrodes is fundamentally governed by a combination of surface adsorption, electrostatic interaction, interfacial charge transfer, and redox reactions occurring at the electrode–electrolyte interface. At the molecular level, Pb2+, Cd2+, and Hg2+ ions are initially transported toward the electrode surface by diffusion and electrostatic attraction. The presence of oxygen-containing functional groups such as –COOH, –OH, and –C=O on carbon nanomaterials enables strong coordination interactions with metal ions, promoting their stable preconcentration on the surface.
During anodic stripping voltammetry (ASV), the electrochemical sensing process consists of two main steps: (i) a deposition step, in which metal ions are electrochemically reduced and accumulated in their metallic state on the electrode surface; and (ii) a stripping step, in which the deposited metals are re-oxidized, producing well-defined current peaks that are directly proportional to the metal ion concentration. This preconcentration–stripping mechanism leads to significant signal amplification and enables ultra-trace detection of heavy metal ions. The π-electron-conjugated structure of carbon nanotubes and graphene facilitates rapid electron transfer across the electrode surface, significantly reducing charge-transfer resistance. In addition, the large specific surface area and abundant defect sites provide numerous active adsorption centers, thereby enhancing both sensitivity and detection stability. When bismuth or antimony films are introduced, intermetallic alloy formation improves nucleation efficiency and suppresses hydrogen evolution, resulting in higher signal-to-noise ratios and improved analytical performance.
At the molecular scale, the electrochemical signal originates from the redox conversion of surface-bound metal species. The adsorption–diffusion–reduction–oxidation sequence constitutes the core sensing mechanism that governs the analytical response of carbon-based electrochemical sensors. These principles are consistent with recent comprehensive reviews showing that carbon-based electrochemical sensors, including graphene and carbon nanotube platforms, achieve low detection limits and high selectivity for Pb2+, Cd2+, and Hg2+ through surface functionalization, preconcentration effects, and efficient electron-transfer pathways [82].
The application of bismuth film electrodes has been widely reported as an environmentally friendly alternative to mercury-based electrodes for the simultaneous determination of Pb2+ and Cd2+. Wang et al. [83] demonstrated that in situ plated bismuth films on glassy carbon electrodes significantly enhance stripping signal intensity due to the formation of fusible Bi–metal alloys and improved mass transport at the electrode surface.
Recent studies have demonstrated that metal–organic framework (MOF)–based electrochemical sensors provide high sensitivity and selectivity for the simultaneous detection of Pb2+, Cd2+, and Hg2+ ions, owing to strong metal–ligand interactions and efficient charge-transfer pathways, making them highly suitable for stripping voltammetric analysis of toxic heavy metals [84].
Recent advances in carbon-based electrochemical sensors have significantly expanded their applicability for on-site monitoring of heavy metals in water and food samples. Graphene- and carbon nanotube-based hybrid electrodes provide high surface area, fast electron transfer, and enhanced sensitivity, enabling simultaneous detection of Pb2+, Cd2+, and Hg2+ at trace levels [85,86,87,88]. In particular, hybrid nanostructures combining graphene derivatives with metal oxides or conducting polymers have demonstrated improved charge-transfer kinetics and mechanical stability, which are critical for portable and field-deployable sensing platforms [89,90,91]. Furthermore, nanocomposite strategies integrating chelating agents and redox-active modifiers have been reported to improve selectivity and signal stability in complex environmental matrices [92,93,94].

3. Development of Electrochemical Sensors Based on MXene Materials

MXenes are a class of two-dimensional transition metal carbides and nitrides and cannot be strictly classified as carbon-based materials. The general structure of MAX phases and the formation of MXene layers through selective etching are illustrated in Figure 5. The commonly used etching routes for producing MXenes from MAX phases—including fluoride-based, alkaline, hydrothermal, and Lewis acid etching—are summarized in Table 4. However, due to their outstanding electrical conductivity, hydrophilicity, and strong interaction with carbon electrodes, MXenes are frequently integrated into carbon-based electrochemical sensing platforms. Therefore, MXene-based systems are discussed in this review as carbon-related hybrid electrochemical sensors.
MXenes, particularly Ti3C2Tx, have attracted increasing attention as carbon-related hybrid platforms due to their metallic conductivity, hydrophilic surface terminations, and strong interfacial compatibility with carbon electrodes. Well-established synthesis and processing protocols enable the fabrication of stable MXene films with tunable surface chemistry, which is advantageous for electrochemical sensing applications [95]. Recent studies have demonstrated that MXene-based electrodes enable simultaneous determination of Cd2+ and Pb2+ with high sensitivity and wide linear ranges, supporting their use in portable electrochemical detection systems [96,97,98].
Although MXenes are not carbon-based materials, they are frequently integrated with carbon electrodes and electrochemical sensing platforms because of their excellent electrical conductivity and strong affinity toward metal ions. Their hydrophilic surface and abundant functional groups provide numerous active sites for metal ion adsorption and facilitate fast electron transfer during electrochemical reactions, thereby significantly enhancing sensor sensitivity and signal stability [99,100,101,102,103,104].
The electrochemical detection of heavy metal ions on MXene-modified electrodes is primarily governed by surface adsorption, electrostatic interaction, and redox reactions at the MXene–electrolyte interface. During anodic stripping voltammetry, metal ions are first preconcentrated on the MXene-modified surface through electrostatic attraction and coordination with surface functional groups. Subsequently, the accumulated metal ions undergo electrochemical oxidation, generating well-defined stripping peaks. This preconcentration–stripping process leads to significant signal amplification and enables ultra-trace detection of Pb2+, Cd2+, and Hg2+ ions.
In recent years, various MXene-based electrochemical sensors have been developed for the sensitive determination of Pb2+ and Cd2+. He et al. [105] fabricated a BiNPs/Ti3C2Tx composite electrode for the simultaneous detection of Pb2+ and Cd2+ using anodic stripping voltammetry. The sensor exhibited excellent analytical performance with detection limits of 10.8 nM for Pb2+ and 12.4 nM for Cd2+, demonstrating the strong potential of MXene-based materials in trace heavy metal detection. Other studies have also shown that the integration of MXenes with metal nanoparticles, polymers, and carbon nanomaterials further enhances adsorption capacity, selectivity, and long-term stability [106,107,108,109,110].
Figure 5. Schematic representation of the MAX phase structure with n = 1–4 and the structure of MXene, showing transition metals, carbon/nitrogen, A-group elements, and etched surface terminations [105].
Figure 5. Schematic representation of the MAX phase structure with n = 1–4 and the structure of MXene, showing transition metals, carbon/nitrogen, A-group elements, and etched surface terminations [105].
Molecules 31 00005 g005
Beyond conventional electrochemical sensing, MXenes have recently been applied in the development of self-powered electrochemical and visual sensing platforms for Hg2+ detection. In such systems, MXenes function as highly conductive electrode materials that enable efficient electron transport and signal generation without the need for an external power source. In our study, a MXene-based self-powered visual electrochemical system was developed for Hg2+ detection by integrating MXene with an enzyme-driven bioelectrochemical process. Owing to the high conductivity and surface activity of MXene, the system exhibits distinct visual and electrochemical responses toward Hg2+ ions, providing a rapid, portable, and energy-efficient strategy for on-site monitoring of mercury contamination in water.
Table 4. Various methods for MXene etching.
Table 4. Various methods for MXene etching.
Etching MethodEtching AgentEtching Temperature (°C)Reference
Fluoride AcidsHFRoom-55[105]
H2O2 + HF40[106]
HCl + HF35–55[107]
HCl + (Na, K, or NH4F)30–60[108]
NH4HF2Room[109]
Alkaline MethodsNaOH270[110]
Hydrothermal MethodNaBF4, HCl180[111]
Molten SaltsLiF + NaF + KF550[112]
ElectrochemicalNH4Cl/TMAOHRoom[113]
Lewis AcidsZnCl2550[114]
Chemical Vapor Phase3D Graphene/Ti3AlC2/PDMS membrane-[115]
Beyond laboratory-based sensing platforms, MXene-assisted electrochemical systems have been successfully integrated into portable and smartphone-enabled devices, highlighting their strong potential for decentralized and on-site environmental monitoring. The combination of MXenes with miniaturized electrochemical readout units enables rapid, low-power, and field-deployable detection of heavy metals in water and food matrices, which is essential for real-time contamination assessment [111].
Despite these advantages, certain challenges associated with MXene-based electrochemical sensors still remain. MXenes are susceptible to surface oxidation and structural degradation under long-term aqueous exposure, which may affect their electrochemical stability and sensing performance. Therefore, future research should focus on improving MXene stability through surface modification, composite construction, and device-level integration. With continued development, MXene-based platforms are expected to play an increasingly important role in portable, self-powered, and in situ detection of toxic heavy metal ions.

4. Objective and Scope of the Research

Pollution of water bodies poses a threat to the use of water resources. Heavy metals in water are difficult to remove naturally, and most of them accumulate in aquatic environments, leading to concentrations that exceed permissible limits and render the water unsuitable for use.
To achieve a more scientific understanding of the state of the aquatic environment and to utilize water resources more effectively, it is important to focus on analyzing the content of heavy metals in water bodies. Conventional analysis of heavy metal ions depends on specialized and strict laboratory conditions, which presents significant limitations. In contrast, the electrochemical method, as an important trace analysis technique, is simple to operate and can be miniaturized, offering significant advantages for rapid analysis of heavy metals. Research into electrochemical sensors, devices, and analytical methods enables convenient, fast, and accurate quantitative determination of heavy metal content in water, which is of great importance for residential environments, food safety, and both industrial and agricultural production.
To enable faster, more sensitive, and more accurate detection of common heavy metals in water such as Pb2+, Cd2+, and Hg2+, as well as to gain a more scientific understanding of the aquatic environment and improve the use of water resources, the following research will be carried out in this study: At present, many literature sources describe methods for the detection of Pb2+ and Cd2+ in water using various electrode modification techniques. These methods involve relatively complex electrode modification processes and have certain limitations in terms of operational simplicity. In this study, to achieve simpler and faster detection of Pb2+ and Cd2+, a three-electrode system was developed using a glassy carbon electrode (GCE) as the working electrode, silver/silver chloride (Ag/AgCl) as the reference electrode, and a platinum wire as the counter electrode to detect Pb2+ and Cd2+ in water. The limits of detection (LODs) and method detection limits (MDLs) for both elements will be determined, and sensitive detection of Pb2+ and Cd2+ in water using unmodified GCE will be demonstrated.
In Study 1, the enrichment of heavy metals was performed by stirring. In the present study, a laboratory-developed integrated device for the detection of heavy metal ions (a thin-layer cell) was used, in which stirring was replaced by flow-based enrichment. A glassy carbon electrode (GCE) was used as the working electrode, Ag/AgCl as the reference electrode, and a platinum wire as the counter electrode. Detection of Pb2+ and Cd2+ in water was carried out, and the limits of detection (LODs) and method detection limits (MDLs) for both elements were determined, enabling sensitive detection of Pb2+ and Cd2+ in water using the thin-layer cell system.
In Studies 1 and 2, electrochemical detection of Pb2+ and Cd2+ in water was conducted. Since bare GCE is not suitable for the detection of mercury, which is commonly found in water, a new electrochemical method based on carbon materials was developed in this work for the sensitive detection of Hg2+ in water. A two-dimensional transition metal carbide (MXene) was used as the non-enzymatic anode, and glucose oxidase (GOD)/Prussian Blue (PB) was used as the biocathode. Both components were integrated into an indium tin oxide (ITO) substrate, forming a self-powered visual sensor for the detection of Hg2+, enabling sensitive detection of Hg2+ in water.
Compared to traditional sensors, self-powered electrochemical sensors (SPES) do not require batteries or external power sources, allowing for rapid on-site testing.

5. Experimental Section

5.1. Experimental Methods and Principles

A three-electrode electrochemical system was used in the experiment. The three-electrode system is the most commonly used setup in electrochemistry. Its configuration is shown in Figure 6 and includes a working electrode, a reference electrode, and a counter electrode. In this experiment, a glassy carbon electrode (GCE) was used as the working electrode, a silver/silver chloride (Ag/AgCl) electrode served as the reference electrode, and a platinum wire was used as the counter electrode. The three electrodes were connected to the corresponding terminals of the electrochemical workstation. Using differential pulse voltammetry (DPV), Pb2+ and Cd2+ ions in the aqueous phase, composed of a supporting electrolyte solution, were detected.
DPV (Differential Pulse Voltammetry) is a powerful electrochemical analytical technique used to detect and quantify various metal ions at extremely low concentrations in aqueous media. It allows for the identification and quantification of electroactive substances in solutions at nanomolar concentrations or lower, offering sensitivity comparable to technologies such as inductively coupled plasma mass spectrometry (ICP-MS). The detection of heavy metal ions using DPV typically involves two main stages (see Figure 7). First, the heavy metal ions undergo constant potential preconcentration under stirring conditions, allowing the target substance to accumulate and concentrate on the working electrode. Then, the potential of the working electrode is swept from a negative to a positive value, causing the preconcentrated substance on the electrode surface to re-dissolve into the solution [112]. During the re-dissolution of each electro-enriched substance, a stripping peak is generated. The peak potential of each signal can be used to identify the oxidized substance, while the peak current is related to the concentration of the substance in the solution [113]. As illustrated in Figure 7, the colored symbols represent accumulated heavy metal ions on the electrode surface, where the different colors are used only to distinguish ions visually.
The bismuth film electrode is a highly attractive type of electrode due to its high sensitivity, low toxicity (compared to mercury electrodes), wide cathodic potential range, and its ability to deposit on various surfaces (such as gold, platinum, glassy carbon, carbon paste, and carbon fiber). The bismuth film can form alloys and intermetallic compounds with certain metals and metalloids such as cadmium, lead, antimony, thallium, gallium, indium, and others. The formation of such alloys promotes the nucleation process during the deposition of heavy metals, enabling effective enrichment of heavy metals and thus increasing analytical sensitivity [114]. Moreover, it is insensitive to oxygen interference and exhibits better peak separation performance compared to mercury film electrodes. In 2000, Wang [83] and colleagues first used a GCE as the working electrode, Ag/AgCl as the reference electrode, and a platinum wire as the counter electrode, developing a bismuth film-modified GCE for the detection of metal ions. During the stripping of metal ions, distinct and well-separated stripping peaks were observed. The measured stripping potentials were approximately −0.20 V for Bi3+, −0.58 V for Pb2+, and −0.79 V for Cd2+. Based on the excellent performance of bismuth films in detecting heavy metal ions [115,116,117,118,119], this study utilized a bismuth film coated on a glassy carbon electrode for the detection of Pb2+ and Cd2+.
During the DPV detection process of Bi3+ on the glassy carbon electrode, a cycle of accumulation and stripping takes place. After this, it is necessary to clean the surface of the GCE at a specific potential to activate or renew it. To remove the bismuth film from the electrode surface, a potential higher than the oxidation potential of Bi3+ must be selected—typically in the range of 0.0–0.3 V—and applied for 30 s to effectively clean the surface [120,121]. Electrochemical cleaning allows for the rapid renewal of the electrode surface, ensuring good stability and reproducibility of the working electrode.

5.2. Preparation of Required Solutions

(1)
GCE Test Solution: The GCE test solution is a mixture of aqueous solutions containing 5 mM K3Fe(CN)6, 5 mM K4Fe(CN)6, and 0.1 M KCl. Accurately weigh 0.164 g of solid K3Fe(CN)6, 0.211 g of solid K4Fe(CN)6, and 0.74 g of solid KCl. Place them in a beaker and add a small amount of deionized water to fully dissolve and mix the three solids. Then, transfer the mixture to a 100 mL volumetric flask and add deionized water to reach the final volume of 100 mL.
(2)
Sample Test Solution: Transfer 1 mL of the Bi3+ standard solution with a concentration of 100 μg/mL from the stock bottle into a 10 mL centrifuge tube for further use. Then, pipette 1 mL each of the Pb2+ and Cd2+ standard solutions with a concentration of 1000 μg/mL and dilute them with distilled water to a concentration of 100 μg/mL to obtain stock solutions of Pb2+ and Cd2+. During testing, dilute the 100 μg/mL Pb2+ and Cd2+ stock solutions to a working concentration of 2 μg/mL.
(3)
Acetate–Sodium Acetate Buffer (pH = 4.5): Since Bi3+ easily undergoes hydrolysis under neutral or alkaline conditions to form BiOCl precipitate, which can interfere with the testing process, hydrolysis is suppressed in acidic media. Therefore, a commonly used laboratory buffer solution—0.1 M acetic acid–sodium acetate buffer with pH = 4.5—was selected as the supporting electrolyte for this experiment. Accurately weigh 3.86 g of glacial acetic acid and 2.93 g of sodium acetate, place them in a beaker, add a small amount of deionized water to fully dissolve and mix, then transfer the solution to a 1 L volumetric flask and make up the volume to 1 L with deionized water. The pH of the solution was measured using a pH meter and confirmed to be 4.5. After preparation, store all solutions in a refrigerator at 4 °C.

5.3. Pre-Treatment of GCE (Glassy Carbon Electrode)

When conducting experiments using GCE, it is essential to carry out a pre-treatment procedure to enhance its electrochemical performance [122]. This pre-treatment generally involves polishing the electrode with alumina (Al2O3) powders of various particle sizes, followed by ultrasonic cleaning of the GCE surface. The purpose of the pre-treatment is to expose fresh carbon atoms on the electrode surface, which act as active sites for electrochemical reactions [123,124]. After polishing, it is necessary to verify whether the peak potential difference (ΔΦ) observed in the cyclic voltammetry (CV) curve—recorded in a standard reversible system consisting of K3Fe(CN)6 and K4Fe(CN)6—meets the experimental requirements. The theoretical value of the peak potential difference is 64 mV, and a peak separation in the range of 80–90 mV is generally considered acceptable for practical use. The theoretical formula for calculating the peak potential difference is given in Equation (1). By substituting the relevant values into the equation, the value of ΔΦ can be determined.
Δ   E = Epa Epc = 2.2 R T n F = 58 n
In the formula: Epa and Epc are the anodic (oxidation) and cathodic (reduction) peak potentials, respectively; R is the gas constant; T is the temperature; F is the Faraday constant; and n is the number of electrons transferred.
Mechanically polish the GCE sequentially using alumina powder with particle sizes of 1.0, 0.3, and 0.05 µm until a mirror-like surface is obtained. Rinse the surface of the GCE with ethanol and deionized water, then clean it in an ultrasonic bath and dry it with nitrogen gas to obtain a suitable glassy carbon electrode for use. Set the voltage range for the cyclic voltammetry (CV) test from −0.2 to 0.7 V. Perform the CV test of the polished GCE in an aqueous solution containing 5 mM K3Fe(CN)6 and K4Fe(CN)6, as shown in Figure 8. In the resulting CV curve, redox peaks with nearly equal current values are observed. The anodic peak potential is approximately 270 mV, and the cathodic peak potential is approximately 195 mV. The peak potential difference is 75 mV, which falls within the theoretical range, indicating that the treated GCE is suitable for testing.
Cyclic voltammogram of GCE after polishing and cleaning in a 5 mM K3Fe(CN)6/K4Fe(CN)6 solution (Figure 8). The red curve represents the cyclic voltammetric response of the polished and cleaned GCE, and the color is used solely for visual distinction of the recorded CV signal.

5.4. Voltammetric Testing of Pb2+ and Cd2+ Stripping

Electrochemical testing using differential pulse voltammetry (DPV) combined with in situ deposition of a bismuth film was employed in the experiment. The testing of Pb2+ was carried out as follows: (1) Blank sample determination: prepare 15 mL of an acetate buffer solution containing 100 μg/L Bi3+. Connect the solution to a three-electrode electrochemical testing system, apply a deposition potential of −1.2 V at a stirring rate of 500 rpm, and perform preconcentration for 120 s. Then, under quiescent conditions, conduct stripping in the potential range of −1.2 to 0.4 V. After stripping, apply a constant potential of 0.3 V for electrochemical cleaning of the electrode.
(2) Determination of 2 μg/L Pb2+: after the blank sample test, add 15 μL of Pb2+ solution with a concentration of 2 μg/mL into the solution, so that the Pb2+ concentration remains at 2 μg/L. Then repeat the preconcentration and stripping process. After each stripping, perform electrochemical cleaning of the electrode at a constant potential of 0.3 V. Each sample is tested 7 times, and the stripping potential and stripping current are recorded.
(3) Determination of 10 μg/L Pb2+: to the above solution, 60 μL of Pb2+ solution with a concentration of 2 μg/mL was added, so that the Pb2+ concentration in the solution reached 10 μg/L. The preconcentration and stripping process was then repeated. After each stripping, the electrodes were subjected to electrochemical cleaning at a constant potential of 0.3 V. Each sample was tested seven times, and the stripping potential and stripping current were recorded.

5.5. MXene-Anode-Glucose Oxidase/Prussian Blue/ITO-Cathode Self-Powered System for the Determination of Mercury (II) in Water

The mercury ion (Hg2+) is one of the most stable inorganic forms of mercury and poses a serious threat to environmental safety and human health. Considering its danger to the environment and living organisms, researchers are continuously working on the development of sensitive, rapid, and cost-effective methods for the detection of Hg2+. At present, various sensing systems are used for the accurate analysis of mercury in environmental samples, such as atomic absorption spectroscopy, atomic fluorescence spectroscopy, inductively coupled plasma mass spectrometry (ICP-MS), and atomic emission spectroscopy. However, these methods have several drawbacks, including operational complexity, high equipment cost, and the requirement for skilled personnel. In recent years, electrochemical technologies have attracted significant attention in the field of Hg2+ detection due to their high selectivity, low cost, and ease of miniaturization. To meet the growing demand for inexpensive, rapid, reliable, and sensitive methods for heavy metal detection, extensive research has been carried out on the development of low-cost and user-friendly detection systems for heavy metals.
Traditional electrochemical sensors usually consist of four parts: a recognition element, a signal conditioning and conversion circuit, a signal amplifier, and an auxiliary power supply circuit [125]. In contrast, self-powered electrochemical sensors (SPES) do not require batteries or an external power source, as they harvest energy from the environment through electrochemical processes, thereby providing self-sustained operation of the sensor [126]. By utilizing mediated or direct electron transfer between enzymes and electrodes, efficient charge transfer across the interface can be achieved [127,128]. In recent years, enzyme-based SPES have attracted considerable attention. Enzymes operate under mild conditions and can maintain high activity at reasonable ambient temperatures and neutral pH [129,130]. A self-powered sensor consists of two electrodes, with the sensor itself supplying power to the sensing device. The simple two-electrode design and the ability to function without an external power source not only provide SPES with enhanced stability and simplified operation but also allow for easy miniaturization and simplified manufacturing. This eliminates the need for an external power supply and addresses the limitation hindering miniaturization, making it an effective strategy for developing SPES.
In addition to its excellent biocompatibility, outstanding electrical conductivity, large specific surface area, and other properties, studies have shown that MXene is highly susceptible to oxidation when exposed to air. The ease of oxidation of MXene, together with its large specific surface area, demonstrates its potential as an electron donor. In this study, the focus is placed on the redox properties of MXene, which, combined with the reversible transformation between Prussian blue (PB) and Prussian white (PW), enables the construction of a self-powered system consisting of an MXene anode–glucose oxidase (GOD)/PB/ITO cathode for the detection of Hg2+ in water [131].

6. Results and Discussion

6.1. Determination of Trace Lead (II) and Cadmium (II) in Water Using a Glassy Carbon Electrode

The limit of detection (LOD) refers to the minimum amount of a measurable component that can be detected in a sample. The method detection limit (MDL) is the minimum concentration or the smallest amount of a measurable substance that can be qualitatively detected in a sample with a given level of confidence using a specific analytical method. To determine the MDL, the sample must be processed through all the steps of the specified analytical method. It is one of the key parameters characterizing an analytical method. Lead and cadmium, as common sources of heavy metal pollution, can have immeasurable toxic effects on the human body, posing a serious threat to human health. Lead and cadmium are mainly present in water in the form of ions (Pb2+ and Cd2+). Conventional detection methods such as spectroscopy, chromatography, and spectrophotometry have high usage thresholds, are complex to operate, and are not suitable for the rapid and simple detection of heavy metals. Differential pulse voltammetry (DPV) combines preconcentration, dissolution, and electrochemical measurement of heavy metal ions, making it an extremely sensitive and rapid electroanalytical method.
The glassy carbon electrode (GCE) is a widely used electrode in electrochemical experiments, as shown in Figure 9. Due to its excellent electrical conductivity, outstanding chemical stability, robust and durable structure, low thermal expansion coefficient, high sealing performance, and wide electrochemical potential window, it is extensively used in electrochemical experiments.
Zhang et al. [109] used a GCE modified with carbon dots (CDs) and a Nafion-modified bismuth film to detect Cd2+ and Pb2+ in water. In the experiment, the LOD for Pb2+ and Cd2+ was found to be 2.3 µg/L and 3.1 µg/L, respectively. Sereilakhena Phal et al. [110] used a GCE modified with Bi/carboxyphenyl for the detection of Pb2+ and Cd2+ in water. The LOD for Pb2+ and Cd2+ in the experiment was 10 µg/L and 25 µg/L, respectively.
Amine Djebbi et al. [132,133] used a GCE modified with biomass-derived porous carbon (BPC) and nanoscale zero-valent iron (nZVI) for the detection of Pb2+ in water. The experimentally determined LOD for Pb2+ was 2.2 µg/L. The aforementioned detection methods require complex modification of the GCE. In this study, a bare GCE was used as the working electrode, simplifying the GCE modification process. A silver/silver chloride (Ag/AgCl) electrode was used as the reference electrode, and a platinum wire as the counter electrode. Using differential pulse voltammetry (DPV), the LOD and MDL of Pb2+ and Cd2+ in water were determined, enabling sensitive detection of Pb2+ and Cd2+.

6.2. Analytical Performance of the MXene–GOD/PB/ITO Self-Powered Sensor for Hg2+ Detection

The analytical performance of the MXene–GOD/PB/ITO self-powered sensor for Hg2+ detection was evaluated in terms of linear detection range, limit of detection (LOD), sensitivity, selectivity, repeatability, and operational stability. Owing to the excellent electrical conductivity of MXene and the efficient biocatalytic activity of glucose oxidase (GOD), the constructed self-powered system demonstrated highly sensitive electrochemical response toward Hg2+ ions.
Under optimized conditions, the sensor exhibited a linear response to Hg2+ concentration in the range of 1–5 μg/L, indicating its suitability for trace-level mercury detection in environmental water samples. The calculated limit of detection (LOD) reached 50 ng/L, which is significantly lower than the maximum allowable concentration of mercury in drinking water set by the World Health Organization (WHO). This low detection limit confirms the high analytical sensitivity of the proposed self-powered sensing platform.
The sensing mechanism is based on the inhibitory effect of Hg2+ on the enzymatic activity of glucose oxidase. In the absence of Hg2+, GOD catalyzes the oxidation of glucose to generate hydrogen peroxide (H2O2), which subsequently oxidizes Prussian White (PW) back to Prussian Blue (PB), resulting in a distinct color change and a measurable electrochemical signal. In the presence of Hg2+ ions, the activity of GOD is suppressed due to strong Hg2+–enzyme coordination, leading to reduced H2O2 production and weaker PB/PW conversion. As a result, the output signal intensity decreases proportionally with increasing Hg2+ concentration. This inhibition-based sensing strategy ensures high selectivity toward Hg2+ over other coexisting metal ions.
Selectivity studies showed that common interfering ions such as Pb2+, Cd2+, Zn2+, Cu2+, and Fe3+ caused negligible signal interference, confirming the excellent anti-interference capability of the MXene–GOD/PB/ITO sensor. The repeatability of the sensor was evaluated through multiple successive measurements, yielding a relative standard deviation (RSD) of less than 6%, indicating good operational reproducibility. Furthermore, the sensor retained over 90% of its initial response after two weeks of storage, demonstrating acceptable long-term stability.
Compared with previously reported Hg2+ detection systems, such as microbial fuel cell (MFC)-based biosensors and modified carbon electrodes, the MXene-based self-powered sensor developed in this work exhibits a lower detection limit, simpler device structure, and does not require an external power source. Scientists reported an MFC-based Hg2+ biosensor with a detection range of 0.2–3 mg/L and a detection limit of 50 μg/L, which is significantly higher than the value achieved in this study. Therefore, the present MXene–GOD/PB/ITO self-powered system demonstrates superior analytical performance for trace mercury detection in aqueous environments.
Moreover, the integration of MXene as a non-enzymatic anode material significantly enhances electron transfer efficiency due to its large specific surface area and rich surface functional groups. This feature promotes rapid charge transport at the electrode–electrolyte interface and improves the overall sensitivity of the self-powered sensor. These advantages are consistent with previously reported MXene-based electrochemical sensing platforms for heavy metal detection [134,135].
Overall, the excellent analytical performance, low detection limit, high selectivity, good repeatability, and operational stability indicate that the MXene–GOD/PB/ITO self-powered sensor is a promising portable platform for on-site Hg2+ monitoring in environmental water samples.
Recent advances further confirm the central role of carbon-based and hybrid materials in electrochemical sensing. Recent reviews on portable and screen-printed electrochemical sensors highlight their growing importance for on-site determination of heavy metal ions in environmental waters, as these platforms combine low cost, portability, and reliable analytical performance for Pb2+ and Cd2+ detection under field conditions [136].
Although MXenes are not strictly classified as carbon-based materials, they are frequently integrated with carbon electrodes and carbon nanostructures to form hybrid electrochemical sensing platforms. Their outstanding electrical conductivity, hydrophilicity, and strong interfacial interaction with carbon substrates enable highly efficient charge transfer and enhanced signal stability. Therefore, MXene-based systems are discussed in this review as carbon-related hybrid electrochemical sensors.

7. Conclusions

The obtained LOD and MDL values were also compared with international regulatory limits. According to the World Health Organization (WHO), the permissible limits for Pb2+ and Cd2+ in drinking water are 10 µg/L and 3 µg/L, respectively, while the United States EPA sets limits of 15 µg/L for Pb2+ and 5 µg/L for Cd2+. The LOD values achieved in this study (0.405 µg/L for Pb2+ and 0.565 µg/L for Cd2+) are significantly lower than both WHO and EPA thresholds, confirming that the developed sensing systems are suitable for detecting trace concentrations well below regulatory requirements. For Hg2+, the achieved linear range of 1–5 µg/L aligns with the WHO guideline value of 6 µg/L, demonstrating that the self-powered MXene-based sensor can reliably detect mercury at environmentally relevant levels.
In addition to meeting regulatory criteria, the sensors demonstrate strong practical advantages: low cost, portability, and rapid response without the need for complex surface modifications. The excellent performance of the bare GCE can be attributed to the efficient electron-transfer kinetics, its wide potential window, and the ability of the carbon surface to promote diffusion-controlled preconcentration of metal ions. The absence of surface modifiers also minimizes variability between electrodes, contributing to improved reproducibility. These characteristics explain why even the unmodified GCE achieved very low detection limits for Pb2+ and Cd2+.
These findings highlight that the developed carbon-based and MXene-assisted sensing strategies not only achieve high analytical sensitivity but also hold strong potential for practical field applications, especially in rapid and on-site environmental monitoring.

Author Contributions

Conceptualization, Z.M. and D.K.; methodology, Z.M. and Y.B.; software, Y.B.; validation, Z.M., D.K. and Y.T.; formal analysis, Z.M.; investigation, Z.M. and N.S.; resources, D.K.; data curation, Z.M.; writing—original draft preparation, Z.M.; writing—review and editing, D.K. and N.S.; visualization, Y.T.; supervision, D.K.; project administration, D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data supporting the findings of this study are included within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sun, L.; Guo, D.K.; Liu, K.; Meng, H.; Zheng, Y.; Yuan, F.; Zhu, G. Levels, sources, and spatial distribution of heavy metals in soils from a typical coal industrial city of Tangshan, China. Catena 2019, 175, 101–109. [Google Scholar] [CrossRef]
  2. Lu, H.P.; Li, Z.A.; Gascó, G.; Méndez, A.; Shen, Y.; Paz-Ferreiro, J. Use of magnetic biochars for the immobilization of heavy metals in a multi-contaminated soil. Sci. Total Environ. 2018, 622–623, 892–899. [Google Scholar] [CrossRef] [PubMed]
  3. Mahmoud, A.; Hoadley, F.A.A. An evaluation of a hybrid ion exchange electrodialysis process in the recovery of heavy metals from simulated dilute industrial wastewater. Water Res. 2012, 46, 3364–3376. [Google Scholar] [CrossRef] [PubMed]
  4. Bashir, A.; Malik, L.A.; Ahad, S.; Manzoor, T.; Bhat, M.A.; Dar, G.N.; Pandith, A.H. Removal of heavy metal ions from aqueous system by ion-exchange and biosorption methods. Environ. Chem. Lett. 2019, 17, 729–754. [Google Scholar] [CrossRef]
  5. Gao, D. Heavy Metal Pollution and Prevention in Water Bodies. Chem. Ind. Manag. 2019, 511, 37–38. [Google Scholar]
  6. Tchounwou, P.B.; Yedjou, C.G.; Patlolla, A.K.; Sutton, D.J. Heavy Metal Toxicity and the Environment. In Molecular, Clinical and Environmental Toxicology; Springer: Basel, Switzerland, 2012; Volume 3, pp. 133–164. [Google Scholar] [CrossRef]
  7. He, Z.L.; Yang, X.E.; Stoffella, P.J. Trace elements in agro-ecosystems and impacts on the environment. J. Trace Elem. Med. Biol. 2005, 10, 125–140. [Google Scholar] [CrossRef]
  8. Dasharathy, S.; Arjunan, S.; Basavaraju, A.M.; Murugasen, V.; Ramachandran, S.; Keshav, R.; Murugan, R. Mutagenic, carcinogenic, and teratogenic effect of heavy metals. Evid.-Based Complement. Altern. Med. 2022, 2022, 11. [Google Scholar] [CrossRef]
  9. Kiran, K.; Flora, S.J.S. Strategies for safe and effective therapeutic measures for chronic arsenic and lead poisoning. J. Occup. Health 2005, 47, 1–21. [Google Scholar] [CrossRef]
  10. Dou, J.R.; Zhou, L.; Zhao, Y.; Jin, W.; Shen, H.; Zhang, F. Effects of long-term high-level lead exposure on the immune function of workers. Arch. Environ. Occup. Health 2022, 77, 301–308. [Google Scholar] [CrossRef]
  11. Ekong, E.B.; Jaar, B.; Weaver, V.M. Lead-related nephrotoxicity: A review of the epidemiologic evidence. Kidney Int. 2006, 70, 2074–2084. [Google Scholar] [CrossRef]
  12. Theron, A.J.; Tintinger, G.R.; Anderson, R. Harmful interactions of non-essential heavy metals with cells of the innate immune system. J. Clin. Toxicol. 2012, S3, 1–10. [Google Scholar] [CrossRef]
  13. Zuckerman, A.J. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. J. Clin. Pathol. 1995, 48, 691. [Google Scholar] [CrossRef]
  14. An, H.; Zheng, W.; Gao, Y. Health hazards of cadmium and advances in intervention treatment research. J. Environ. Health 2007, 147, 739–742. [Google Scholar]
  15. Tang, Y. A brief discussion on the hazards of heavy metals in water and detection methods. Rural Econ. Technol. 2016, 27, 64–65. [Google Scholar]
  16. Siblerud, R.; Mutter, J.; Moore, J.; Walach, H. A hypothesis and evidence that mercury may be an etiological factor in Alzheimer’s disease. Int. J. Environ. Res. Public Health 2019, 16, 5152. [Google Scholar] [CrossRef]
  17. Chen, Q.Y.; Murphy, A.; Sun, H.; Costa, M. Molecular and epigenetic mechanisms of Cr(VI)-induced carcinogenesis. Toxicol. Appl. Pharmacol. 2019, 377, 114636. [Google Scholar] [CrossRef]
  18. de Oliveira, D.F.; de Castro, B.S.; Recktenvald, M.C.N.; da Costa Júniorc, W.A.; da Silva, F.X.; de Menezes Alves, C.L.; Froehlich, J.D.; Bastos, W.R.; Ott, A.M.T. Mercury in wild animals and fish and health risk for indigenous Amazonians. Food Addit. Contam. Part B 2021, 14, 161–169. [Google Scholar] [CrossRef]
  19. Zeneli, L.; Sekovanić, L.; Ajvazi, A.; Kurti, L.; Daci, N. Alterations in antioxidant defense system of workers chronically exposed to arsenic, cadmium and mercury from coal flying ash. Environ. Geochem. Health 2016, 38, 65–72. [Google Scholar] [CrossRef] [PubMed]
  20. Sanders, A.P.; Mazzella, M.J.; Malin, A.J.; Hair, G.M.; Busgang, S.A.; Saland, J.M.; Curtin, P. Combined exposure to lead, cadmium, mercury, and arsenic and kidney health in adolescents age 12–19 in NHANES 2009–2014. Environ. Int. 2019, 131, 104993. [Google Scholar] [CrossRef]
  21. Ma, Y.; Zheng, Y.X.; Dong, X.Y.; Zou, X.T. Effect of mercury chloride on oxidative stress and nuclear factor erythroid 2-related factor 2 signalling molecule in liver and kidney of laying hens. J. Anim. Physiol. Anim. Nutr. 2018, 102, 1199–1209. [Google Scholar] [CrossRef]
  22. Zheng, L.Y.; Sanders, A.P.; Saland, J.M.; Wright, R.O.; Arora, M. Environmental exposures and pediatric kidney function and disease: A systematic review. Environ. Res. 2017, 158, 625–648. [Google Scholar] [CrossRef] [PubMed]
  23. Sun, H.; Brocato, J.; Costa, M. Oral chromium exposure and toxicity. Curr. Environ. Health Rep. 2015, 2, 295–303. [Google Scholar] [CrossRef] [PubMed]
  24. Zhitkovich, A. Importance of chromium-DNA adducts in mutagenicity and toxicity of chromium(VI). Chem. Res. Toxicol. 2005, 18, 3–11. [Google Scholar] [CrossRef] [PubMed]
  25. O’Brien, T.J.; Ceryak, S.; Patierno, S.R. Complexities of chromium carcinogenesis: Role of cellular response, repair and recovery mechanisms. Mutat. Res./Fundam. Mol. Mech. Mutagen. 2003, 533, 3–36. [Google Scholar] [CrossRef]
  26. Nickens, K.P.; Steven, S.R.; Ceryak, S. Chromium genotoxicity: A double-edged sword. Chem.-Biol. Interact. 2010, 188, 276–288. [Google Scholar] [CrossRef]
  27. Zhang, Y.L.; Zheng, P.; Su, Z.K.; Hu, G.; Jia, G. Perspectives of genetic damage and epigenetic alterations by hexavalent chromium: Time evolution based on a bibliometric analysis. Chem. Res. Toxicol. 2021, 34, 684–694. [Google Scholar] [CrossRef]
  28. Hong, Y.S.; Song, K.H.; Chung, J.Y. Health effects of chronic arsenic exposure. J. Prev. Med. Public Health 2014, 47, 245–252. [Google Scholar] [CrossRef]
  29. Weller, A.; Zok, D.; Reinhard, S.; Woche, S.K.; Guggenberger, G.; Steinhauser, G. Separation of ultratraces of radiosilver from radiocesium for environmental nuclear forensics. Anal. Chem. 2020, 92, 5249–5257. [Google Scholar] [CrossRef]
  30. Qian, J.S.; Gao, X.; Pan, B. Nanoconfinement-mediated water treatment: From fundamental to application. Environ. Sci. Technol. 2020, 54, 8509–8526. [Google Scholar] [CrossRef]
  31. Fu, H.; Liu, W.; Zhang, C.; Li, Y. Recent advance of detection method of heavy metal ions in water. Phys. Test. Chem. Anal. Part B Chem. Anal. 2012, 48, 496–503. [Google Scholar]
  32. Divrikli, U.; Kartal, A.A.; Soylak, M.; Elci, L. Preconcentration of Pb(II), Cr(III), Cu(II), Ni(II) and Cd(II) ions in environmental samples by membrane filtration prior to their flame atomic absorption spectrometric determinations. J. Hazard. Mater. 2007, 145, 459–464. [Google Scholar] [CrossRef] [PubMed]
  33. Lin, Y.; Lin, N.; Zhang, J.; He, R. Determination of Pb and Cr in drinking water by graphite furnace atomic absorption spectrometry. Metall. Anal. 2001, 21, 55–57. [Google Scholar]
  34. Jiang, X.J.; Gan, W.E.; Wan, L.Z.; Zhang, H.; He, Y. Determination of mercury by electrochemical cold vapor generation atomic fluorescence spectrometry using polyaniline modified graphite electrode as cathode. Spectrochim. Acta Part B At. Spectrosc. 2010, 65, 171–175. [Google Scholar] [CrossRef]
  35. Liu, D.; Zhang, H.; Huang, Q.; Ju, W.; Zhang, S. Determination of trace molybdenum in water samples by wavelength-dispersive X-ray fluorescence spectrometry. Chem. Propellants Polym. Mater. 2014, 12, 72–74. [Google Scholar]
  36. Song, Z.; Shan, G.; Qi, D. Detection of heavy metal content in lake bottom sediments using EDXRF. J. Changchun Univ. Sci. Technol. Nat. Sci. Ed. 2015, 38, 99–102. [Google Scholar]
  37. Arain, M.A.; Wattoo, F.H.; Wattoo, M.H.S.; Ghanghro, A.B.; Tirmizi, S.A.; Iqbal, J.; Arain, S.A. Simultaneous determination of metal ions as complexes of pentamethylene dithiocarbamate in Indus river water, Pakistan. Arab. J. Chem. 2009, 2, 25–29. [Google Scholar] [CrossRef]
  38. Bruzzoniti, M.C.; Mentasti, E.; Sarzanini, C. Simultaneous determination of inorganic anions and metal ions by suppressed ion chromatography. Anal. Chim. Acta 1999, 382, 291–299. [Google Scholar] [CrossRef]
  39. Hu, H. Determination of heavy metals in water samples by ICP-MS. Chem. Ind. Manag. 2020, 34, 172–173. [Google Scholar]
  40. Li, L.; Regan, F. Rapid detection of copper ions in industrial wastewater using a test strip method. Ind. Water Wastewater 2013, 44, 86–89. [Google Scholar]
  41. Gharehbaghi, M.; Shemirani, F.; Farahani, M.D. Cold-induced aggregation microextraction based on ionic liquids and fiber optic-linear array detection spectrophotometry of cobalt in water samples. J. Hazard. Mater. 2009, 165, 1049–1055. [Google Scholar] [CrossRef]
  42. Liu, J.B.; Qiao, B.; Zhang, N.; He, L. Direct detection of ultra-trace vanadium (V) in natural waters by flow-injection chemiluminescence with controlled-reagent-release technology. J. Geochem. Explor. 2009, 103, 45–49. [Google Scholar] [CrossRef]
  43. Zhang, M.; Feng, C.; Liu, X.; Qi, S.; Xing, W.; Yuan, C. Research progress on detection method for heavy metal ions. Sci. Technol. Eng. 2020, 20, 3404–3413. [Google Scholar]
  44. Jiang, T.; Li, L.; Yang, H. Advances in phosphate ion-selective electrode research. Sens. Microsyst. 2015, 34, 1–4. [Google Scholar]
  45. Liu, S.W.; Li, L.; Yu, S.T.; He, L. Polymerization of fatty acid methyl ester using acidic ionic liquid as catalyst. Chin. J. Catal. 2010, 31, 1433–1438. [Google Scholar] [CrossRef]
  46. Bedlechowicz, I.; Sokalski, T.; Lewenstam, A.; Maj-Zurawska, M. Calcium ion-selective electrodes under galvanostatic current control. Sens. Actuators B Chem. 2005, 108, 836–839. [Google Scholar] [CrossRef]
  47. Miao, J.; Wang, X.; Fan, Y.C.; Li, J.; Zhang, L.; Hu, G.; He, C.; Jin, C. Determination of total mercury in seafood by ion-selective electrodes based on a thiol functionalized ionic liquid. J. Food Drug Anal. 2018, 26, 670–677. [Google Scholar] [CrossRef]
  48. Xiao, L.; Wang, B.; Li, J.; Wang, F.; Yuan, Q.; Hu, G.; Dong, A.; Gan, W. An efficient electrochemical sensor based on three-dimensionally interconnected mesoporous graphene framework for simultaneous determination of Cd(II) and Pb(II). Electrochim. Acta 2016, 222, 1371–1377. [Google Scholar] [CrossRef]
  49. Shalaby, E.A.; Beltagi, A.M.; Hathoot, A.A.; Abdelazzem, M. Development of a sensor based on poly(1,2-diaminoanthraquinone) for individual and simultaneous determination of mercury(II) and bismuth(III). Electroanalysis 2022, 34, 523–534. [Google Scholar] [CrossRef]
  50. Somer, G.; Kalaycı, S. New and simple method for the simultaneous determination of Fe, Cu, Pb, Zn, Bi, Cr, Mo, Se, and Ni in dried red grapes using differential pulse polarography. Food Anal. Methods 2014, 8, 604–611. [Google Scholar] [CrossRef]
  51. Wei, X.; Liang, Q.; Li, J. Determination of zinc, cadmium and lead by antimony electrode potential stripping method. Chin. J. Inorg. Anal. Chem. 2011, 1, 19–23. [Google Scholar]
  52. Mukatayeva, Z.; Bakytkarim, Y.; Nuerbaheti, W.; Tileuberdi, Y.; Shadin, N.; Assirbayeva, Z.; Zhussupova, L. Electrochemical Sensor based on Silicon Carbide and cokederived Carbon Nanocomposites for Sensitive Detection of Lead Ions. ES Mater. Manuf. 2025, 29, 1653. [Google Scholar] [CrossRef]
  53. Akanji, S.P.; Ama, O.M.; Ray, S.S.; Osifo, P.O. Metal oxide nanomaterials for electrochemical detection of heavy metals in water. In Nanostructured Metal-Oxide Electrode Materials for Water Purification: Fabrication, Electrochemistry and Applications; Springer: Cham, Germany, 2020; pp. 113–126. [Google Scholar] [CrossRef]
  54. Bodkhe, G.A.; Hedau, B.S.; Deshmukh, M.A.; Patil, H.K.; Shirsat, S.M.; Phase, D.M.; Pandey, K.K.; Shirsat, M.D. Selective and sensitive detection of Pb(II) ions using Au/SWNT nanocomposite-embedded MOF-199. J. Mater. Sci. 2021, 56, 474–487. [Google Scholar] [CrossRef]
  55. Deshmukh, M.A.; Bodkhe, G.A.; Shirsat, S.; Ramanavicius, A.; Shirsat, M.D. Nanocomposite platform based on EDTA-modified polypyrrole/SWNTs for electrochemical sensing of Pb(II) ions. Front. Chem. 2018, 6, 451. [Google Scholar] [CrossRef] [PubMed]
  56. Deshmukh, M.A.; Celiesiute, R.; Ramanaviciene, A.; Shirsat, M.D.; Ramanavicius, A. EDTA–PANI/SWCNT nanocomposite modified electrode for electrochemical determination of Cu(II), Pb(II), and Hg(II) ions. Electrochim. Acta 2018, 259, 930–938. [Google Scholar] [CrossRef]
  57. Joshi, N.C.; Gururani, P. Advances of graphene oxide based nanocomposite materials in the treatment of wastewater containing heavy metal ions and dyes. Curr. Res. Green Sustain. Chem. 2022, 5, 100306. [Google Scholar] [CrossRef]
  58. Matlou, G.G.; Nkosi, D.; Pillay, K.; Arotiba, O. Electrochemical detection of Hg(II) in water using self-assembled single walled carbon nanotube-poly(m-amino benzene sulfonic acid) on gold electrode. Sens. Bio-Sens. Res. 2016, 10, 27–33. [Google Scholar] [CrossRef]
  59. Wang, L.; Wang, X.; Shi, G.; Peng, C.; Ding, Y. Thiacalixarene covalently functionalized multiwalled carbon nanotubes as chemically modified electrode material for detection of ultratrace Pb2+ ions. Anal. Chem. 2012, 84, 10560–10567. [Google Scholar] [CrossRef]
  60. Li, M.; Wang, X.; Wang, L.; Cao, D.; Feng, X. Research progress on graphene-based electrochemical biosensors. J. Mater. Sci. Eng. 2020, 38, 503–517. [Google Scholar]
  61. Guan, H.N.; Peng, B.; Xue, Y.; Wu, Q.Y.; Sun, L.; Chi, D.F.; Z, N. Detection of heavy metals by graphene-modified enzyme liposome biosensor. Sci. Technol. Food Ind. 2021, 42, 247–251. [Google Scholar]
  62. Chen, X.; Zang, F. Preparation of nitrogen-doped graphene and its application in heavy metal ions detection. J. Qingdao Univ. Sci. Technol. Nat. Sci. Ed. 2021, 426, 28–34. [Google Scholar]
  63. Priya, T.; Dhanalakshmi, N.; Thennarasu, S.; Karthikeyan, V.; Thinakaran, N. Ultra sensitive electrochemical detection of Cd2+ and Pb2+ using penetrable nature of graphene/gold nanoparticles/modified L-cysteine nanocomposite. Chem. Phys. Lett. 2019, 731, 136621. [Google Scholar] [CrossRef]
  64. Xu, H.; Pan, Z.; Xie, Y.; Gu, X.; Liu, J. Electrochemical detection of heavy metals and chloramphenicol using Nafion/graphene quantum dot-modified electrodes. J. Anal. Sci. 2019, 35, 270–276. [Google Scholar] [CrossRef]
  65. Sherigara, B.; Kutner, W.; D’Souza, F. Electrocatalytic properties and sensor applications of fullerenes and carbon nanotubes. Electroanalysis 2010, 15, 753–772. [Google Scholar] [CrossRef]
  66. Han, X.J.; Meng, Z.C.; Zhang, H.F. Fullerene-based anodic stripping voltammetry for simultaneous determination of Hg(II), Cu(II), Pb(II) and Cd(II) in foodstuff. Mikrochim. Acta 2018, 185, 247. [Google Scholar] [CrossRef]
  67. Li, S.; Wang, L.; Chen, Y.; Jiang, H. Research progress on metal-organic framework materials in liquid-phase catalytic hydrogen production. J. High. Educ. Chem. 2022, 43, 70–83. [Google Scholar]
  68. Shen, K.; Chen, X.D.; Chen, J.Y.; Li, Y. Development of MOF-derived carbon-based nanomaterials for efficient catalysis. ACS Catal. 2016, 6, 5887–5903. [Google Scholar] [CrossRef]
  69. Song, Z.X.; Liu, W.W.; Cheng, N.C.; Banis, M.N.; Li, X.; Sun, Q.; Xiao, B.; Liu, Y.; Lushington, A.; Li, R.; et al. Origin of the high oxygen reduction reaction of nitrogen and sulfur co-doped MOF-derived nanocarbon electrocatalyst. Mater. Horiz. 2017, 4, 900–907. [Google Scholar] [CrossRef]
  70. Tan, R.N.; Jiang, P.P.; Pan, C.C.; Pan, J.; Gao, N.; Cai, Z.; Wu, F.; Chang, G.; Xie, A.; He, Y. Core-shell architectured NH2-UiO-66@ZIF-8/multi-walled carbon nanotubes nanocomposite-based sensitive electrochemical sensor towards simultaneous determination of Pb2+ and Cu2+. Microchim. Acta 2022, 190, 30. [Google Scholar] [CrossRef]
  71. Lee, J.; Kim, K.H.; Kwon, E.E. Biochar as a catalyst. Renew. Sustain. Energy Rev. 2017, 77, 70–79. [Google Scholar] [CrossRef]
  72. Liu, W.; Mei, J.; Liu, G.; Kou, Q.; Yi, T.-F.; Xiao, S. Nitrogen-doped hierarchical porous carbon from wheat straw for supercapacitors. ACS Sustain. Chem. Eng. 2018, 6, 11595–11605. [Google Scholar] [CrossRef]
  73. Andres, B.A.; Childs, B.R.; Vallier, H.A. Treatment of hypovitaminosis D in an orthopaedic trauma population. J. Orthop. Trauma 2018, 32, 129–133. [Google Scholar] [CrossRef] [PubMed]
  74. Peng, J.; Yu, J.; Meng, B.; Wang, L.; Zhang, X.; Cheng, W.; Fang, Z. Hierarchical porous biomass activated carbon for hybrid battery capacitors derived from persimmon branches. Mater. Express 2020, 10, 523–530. [Google Scholar] [CrossRef]
  75. Khan, A.; Senthil, R.A.; Pan, J.; Osman, S.; Sun, Y.; Shu, X. A new biomass derived rod-like porous carbon from tea-waste as inexpensive and sustainable energy material for advanced supercapacitor application. Electrochim. Acta 2020, 335, 135588. [Google Scholar] [CrossRef]
  76. Xu, M.M.; Huang, Q.B.; Lu, J.J.; Niu, J. Green synthesis of high-performance supercapacitor electrode materials from agricultural corncob waste by mild potassium hydroxide soaking and a one-step carbonization. Ind. Crops Prod. 2021, 161, 113215. [Google Scholar] [CrossRef]
  77. Akdemir, M. Electrochemical performance of Quercus infectoria as a supercapacitor carbon electrode material. Int. J. Energy Res. 2022, 46, 7722–7731. [Google Scholar] [CrossRef]
  78. Li, W.P.; Chen, C.; Wang, H.; Li, P.; Jiang, X.; Yang, J.; Liu, J. Hierarchical porous carbon induced by inherent structure of eggplant as sustainable electrode material for high performance supercapacitor. J. Mater. Res. Technol. 2022, 17, 1540–1552. [Google Scholar] [CrossRef]
  79. Punon, M.; Jarernboon, W.; Laokul, P. Electrochemical performance of Palmyra palm shell activated carbon prepared by carbonization followed by microwave reflux treatment. Mater. Res. Express 2022, 9, 065603. [Google Scholar] [CrossRef]
  80. Zhang, T.; Ma, S.; Pan, Y.; Guan, J.; Zhang, M.; Zhu, H.; Du, M. Honeycomb-like carbon materials derived from pomelo peels for the simultaneous detection of heavy metal ions. Chin. J. Inorg. Chem. 2019, 35, 674–686. [Google Scholar] [CrossRef]
  81. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Barsoum, M.W. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4207. [Google Scholar] [CrossRef]
  82. Malik, L.A.; Bashir, A.; Qureashi, A.; Pandith, A.H. Detection and removal of heavy metal ions: A review. Environ. Chem. Lett. 2019, 17, 1495–1521. [Google Scholar] [CrossRef]
  83. Wang, J.; Lu, J.; Hocevar, S.B.; Farias, P.A.M.; Ogorevc, B. Bismuth-coated carbon electrodes for anodic stripping voltammetry. Anal. Chem. 2000, 72, 3218–3222. [Google Scholar] [CrossRef]
  84. Deore, K.B.; Patil, S.S.; Narwade, V.N.; Takte, M.A.; Khune, A.S.; Mohammed, H.Y.; Farea, M.; Sayyad, P.W.; Tsai, M.-L.; Shirsat, M.D. Chromium-benzenedicarboxylates metal organic framework for supersensitive and selective electrochemical sensor of Toxic Cd2+, Pb2+, and Hg2+ metal ions: Study of their interactive mechanism. J. Electrochem. Soc. 2023, 170, 046505. [Google Scholar] [CrossRef]
  85. Huang, H.; Chen, T.; Liu, X.; Ma, H. Ultrasensitive and simultaneous detection of heavy metal ions based on three-dimensional graphene–carbon nanotube hybrid electrode materials. Anal. Chim. Acta 2014, 852, 45–54. [Google Scholar] [CrossRef] [PubMed]
  86. Pan, F.; Tong, C.; Wang, Z.; Han, H.; Liu, P.; Pan, D.; Zhu, R. Nanocomposite based on graphene and intercalated covalent organic frameworks with hydrosulphonyl groups for electrochemical determination of heavy metal ions. Microchim. Acta 2021, 188, 295. [Google Scholar] [CrossRef] [PubMed]
  87. Cai, G.; Yu, Z.; Tong, P.; Tang, D. Ti3C2 MXene quantum dot–encapsulated liposomes for photothermal immunoassays using a portable near-infrared imaging camera on a smartphone. Nanoscale 2019, 11, 15659–15667. [Google Scholar] [CrossRef] [PubMed]
  88. Chen, Y.; Zhao, P.; Hu, Z.; Liang, Y.; Han, H.; Yang, M.; Luo, X.; Hou, C.; Huo, D. Amino-functionalized multilayer Ti3C2Tx enabled electrochemical sensor for simultaneous determination of Cd2+ and Pb2+ in food samples. Food Chem. 2023, 402, 134269. [Google Scholar] [CrossRef]
  89. Hu, J.; Li, Z.; Zhai, C.; Zeng, L.; Zhu, M. Photo-assisted simultaneous electrochemical detection of multiple heavy metal ions with a metal-free carbon black anchored graphitic carbon nitride sensor. Anal. Chim. Acta 2021, 1183, 338951. [Google Scholar] [CrossRef]
  90. Celiesiute, R.; Ramanaviciene, A.; Gicevicius, M.; Ramanavicius, A. Electrochromic sensors based on conducting polymers, metal oxides, and coordination complexes. Crit. Rev. Anal. Chem. 2019, 49, 195–208. [Google Scholar] [CrossRef]
  91. Devaraj, M.; Sasikumar, Y.; Rajendran, S.; Ponce, L.C. Metal–organic framework based nanomaterials for electrochemical sensing of toxic heavy metal ions: Progress and prospects. J. Electrochem. Soc. 2021, 168, 037513. [Google Scholar] [CrossRef]
  92. Nguyen, M.B.; Nga, D.T.N.; Thu, V.T.; Piro, B.; Truong, T.N.P.; Yen, P.T.H.; Le, G.H.; Hung, L.Q.; Vu, T.A.; Ha, V.T.T. Novel nanoscale Yb-MOF used as a highly efficient electrode for simultaneous detection of heavy metal ions. J. Mater. Sci. 2021, 56, 8172–8185. [Google Scholar] [CrossRef]
  93. Wang, X.; Qi, Y.; Shen, Y.; Yuan, Y.; Zhang, L.; Zhang, C.; Sun, Y. A ratiometric electrochemical sensor for simultaneous detection of multiple heavy metal ions based on ferrocene-functionalized metal–organic framework. Sens. Actuators B Chem. 2020, 310, 127756. [Google Scholar] [CrossRef]
  94. Lee, S.; Oh, J.; Kim, D.; Piao, Y. A sensitive electrochemical sensor using an iron oxide/graphene composite for detection of heavy metal ions. Talanta 2016, 160, 528–536. [Google Scholar] [CrossRef] [PubMed]
  95. Alhabeb, M.; Maleski, K.; Anasori, B.; Lelyukh, P.; Clark, L.; Sin, S.; Gogotsi, Y. Guidelines for synthesis and processing of two-dimensional titanium carbide (Ti3C3Tx MXene). Chem. Mater. 2017, 29, 7633–7644. [Google Scholar] [CrossRef]
  96. Sayyad, P.W.; Ingle, N.N.; Al-Gahouari, T.; Mahadik, M.M.; Bodkhe, G.A.; Shirsat, S.M.; Shirsat, M.D. Sensitive and selective detection of Cu2+ and Pb2+ ions using FET based on L-cysteine anchored PEDOT:PSS/rGO composite. Chem. Phys. Lett. 2020, 761, 138056. [Google Scholar] [CrossRef]
  97. Sayyad, P.W.; Ingle, N.N.; Al-Gahouari, T.; Mahadik, M.M.; Bodkhe, G.A.; Shirsat, S.M.; Shirsat, M.D. Selective Hg2+ sensor based on rGO-blended PEDOT:PSS conducting polymer OFET. Appl. Phys. A 2021, 127, 167. [Google Scholar] [CrossRef]
  98. Cotolan, N.; Mureșan, L.M.; Salis, A.; Barbu-Tudoran, L.; Turdean, G.L. Electrochemical detection of lead ions with ordered mesoporous silica–modified glassy carbon electrodes. Water Air Soil Pollut. 2020, 231, 217. [Google Scholar] [CrossRef]
  99. Xiao, X.; Wang, H.; Urbankowski, P.; Gogotsi, Y. Topochemical synthesis of 2D materials. Chem. Soc. Rev. 2018, 47, 8744–8765. [Google Scholar] [CrossRef]
  100. Ghidiu, M.; Lukatskaya, M.R.; Zhao, M.Q.; Gogotsi, Y.; Barsoum, M.W. Conductive two-dimensional titanium carbide ‘clay’ with high volumetric capacitance. Nature 2014, 516, 78–81. [Google Scholar] [CrossRef]
  101. Zou, G.; Zhang, Z.; Guo, J.; Liu, B.; Zhang, Q.; Fernandez, C.; Peng, Q. Synthesis of MXene/Ag composites for extraordinary long cycle lifetime lithium storage at high rates. ACS Appl. Mater. Interfaces 2016, 8, 22280–22286. [Google Scholar] [CrossRef]
  102. Zhang, Y.; Wang, L.; Zhang, N.; Zhou, Z. Adsorptive environmental applications of MXene nanomaterials: A review. RSC Adv. 2018, 8, 19895–19905. [Google Scholar] [CrossRef]
  103. Lim, J.; Savchenkov, A.A.; Dale, E.; Liang, W.; Eliyahu, D.; Ilchenko, V.; Matsko, A.B.; Maleki, L.; Wong, C.W. Chasing the thermodynamical noise limit in whispering-gallery-mode resonators for ultrastable laser frequency stabilization. Nat. Commun. 2017, 8, 8. [Google Scholar] [CrossRef] [PubMed]
  104. Rasool, K.; Helal, M.; Ali, A.; Ren, C.E.; Gogotsi, Y.; Mahmoud, K.A. Antibacterial activity of Ti3C2TX Mxene. ACS Nano 2016, 10, 3674–3684. [Google Scholar] [CrossRef] [PubMed]
  105. He, Y.; Ma, L.; Zhou, L.Y.; Liu, G.; Jiang, Y.; Gao, J. Preparation and application of bismuth/MXene nano-composite as electrochemical sensor for heavy metal ions detection. Nanomaterials 2020, 10, 866. [Google Scholar] [CrossRef] [PubMed]
  106. He, F.; Li, J.; He, Y.; Zheng, X.; Dai, Z.; Ma, S. Comparison of three LOD determination approaches in TLC. Chin. Pharm. 2019, 18, 1515–1520. [Google Scholar] [CrossRef]
  107. Zhang, H. Gas chromatography validation of detection limits and quantification limits for petroleum hydrocarbons (C10–C40) in soil. Coal Chem. Ind. 2022, 45, 139–143. [Google Scholar]
  108. Zeng, C.; Zhu, H.; Wang, Y.; Wu, M.; Ding, W.; Liu, S.; Zhao, H. Study on the effect of physical grinding on the performance of glassy carbon electrode. Guangzhou Chem. Ind. 2020, 48, 73–76. [Google Scholar]
  109. Zhang, H.; Cui, J.; Zeng, Y.; Zhang, Y.; Pei, Y. Direct electrodeposition of carbon dots modifying bismuth film electrode for sensitive detection of Cd2+ and Pb2+. J. Electrochem. Soc. 2022, 169, 017501. [Google Scholar] [CrossRef]
  110. Pha, S.; Nguyên, H.; Berisha, A.; Tesfalidet, S. In situ Bi/carboxyphenyl-modified glassy carbon electrode as a sensor platform for detection of Cd2+ and Pb2+ using square wave anodic stripping voltammetry. Sens. Bio-Sens. Res. 2021, 34, 100455. [Google Scholar] [CrossRef]
  111. Bhateria, R.; Jain, D. Water quality assessment of lake water: A review. Sustain. Water Resour. Manag. 2016, 2, 161–173. [Google Scholar] [CrossRef]
  112. Wu, Q.; Bi, H.; Han, X. Advances in the electrochemical detection of heavy metal ions. Anal. Chem. 2021, 49, 330–340. [Google Scholar] [CrossRef]
  113. Wygant, B.R.; Lambert, T.N. Thin film electrodes for anodic stripping voltammetry: A mini review. Front. Chem. 2022, 9, 809535. [Google Scholar] [CrossRef] [PubMed]
  114. Arduini, F.; Calvo, J.Q.; Palleschi, G.; Moscone, D.; Amine, A. Bismuth-modified electrodes for lead detection. TrAC Trends Anal. Chem. 2010, 29, 1295–1304. [Google Scholar] [CrossRef]
  115. Yıldız, C.; Bayraktepe, D.E.; Yazan, Z.; Önal, M. Bismuth nanoparticles decorated on Na-montmorillonite multiwall carbon nanotube for simultaneous determination of heavy metal ions-electrochemical methods. J. Electroanal. Chem. 2022, 910, 116205. [Google Scholar] [CrossRef]
  116. March, G.; Nguyen, T.D.; Piro, B. Modified electrodes used for electrochemical detection of metal ions in environmental analysis. Biosensors 2015, 5, 241–275. [Google Scholar] [CrossRef]
  117. Legeai, S.; Vittori, O.A. Cu/Nafion/Bi electrode for on-site monitoring of trace heavy metals in natural waters using anodic stripping voltammetry: An alternative to mercury-based electrodes. Anal. Chim. Acta 2006, 560, 184–190. [Google Scholar] [CrossRef]
  118. Liu, G.; Lin, Y.; Tu, Y.; Ren, Z. Ultrasensitive voltammetric detection of trace heavy metal ions using carbon nanotube nanoelectrode array. Analyst 2005, 130, 1098–1101. [Google Scholar] [CrossRef]
  119. Wang, J.; Lu, J.; Ülkü, A.; Hocevar, S.B.; Ogorevc, B. Insights into the anodic stripping voltammetric behavior of bismuth film electrodes. Anal. Chim. Acta 2001, 434, 29–34. [Google Scholar] [CrossRef]
  120. Sb, H.; Ogorevc, B.; Wang, J.; Pihlar, B. A study on operational parameters for advanced use of bismuth film electrode in anodic stripping voltammetry. Electroanalysis 2002, 14, 1701–1712. [Google Scholar] [CrossRef]
  121. Buledi, J.A.; Amin, S.; Haider, S.I.; Bhanger, M.I.; Solangi, A.R. A review on detection of heavy metals from aqueous media using nanomaterial-based sensors. Environment. Sci. Pollut. Res. 2021, 28, 58994–59002. [Google Scholar] [CrossRef]
  122. Kumunda, C.; Adekunle, A.S.; Mamba, B.B.; Hlongwa, N.W.; Nkambule, T.T. Electrochemical detection of environmental pollutants based on graphene derivatives: A review. Front. Mater. 2021, 7, 616787. [Google Scholar] [CrossRef]
  123. Mehrotra, P. Biosensors and their applications—A review. J. Oral Biol. Craniofac. Res. 2016, 6, 153–159. [Google Scholar] [CrossRef] [PubMed]
  124. Xu, S.; Qin, Y.; Xu, C.; Wei, Y.; Yang, R.; Wang, Z.L. Self-powered nanowire devices. Nat. Nanotechnol. 2010, 5, 366–373. [Google Scholar] [CrossRef] [PubMed]
  125. Cracknell, J.A.; Vincent, K.A.; Armstrong, F.A. Enzymes as working or inspirational electrocatalysts for fuel cells and electrolysis. Chem. Rev. 2008, 108, 2439–2461. [Google Scholar] [CrossRef] [PubMed]
  126. Barton, S.C.; Gallaway, J.; Atanassov, P. Enzymatic biofuel cells for implantable and microscale devices. Chem. Rev. 2004, 104, 4867–4886. [Google Scholar] [CrossRef]
  127. Cinquin, P.; Gondran, C.; Giroud, F.; Mazabrard, S.; Pellissier, A.; Boucher, F.; Alcaraz, J.-P.; Gorgy, K.; Lenouvel, F.; Mathé, S.; et al. A glucose biofuel cell implanted in rats. PLoS ONE 2010, 5, e10476. [Google Scholar] [CrossRef]
  128. Wang, J. Chapter 3—Electrochemical glucose biosensors. Electrochem. Sens. Biosens. Biomed. Appl. 2008, 108, 57–69. [Google Scholar] [CrossRef]
  129. Amir, L.; Tam, T.K.; Pita, M.; Meijler, M.M.; Alfonta, L.; Katz, E. Biofuel cell controlled by enzyme logic systems. J. Am. Chem. Soc. 2009, 131, 826–832. [Google Scholar] [CrossRef]
  130. Grattieri, M.; Minteer, S.D. Self-powered biosensors. ACS Sens. 2017, 2, 44–53. [Google Scholar] [CrossRef]
  131. Nayak, P.; Yang, M.J.; Wang, Z.W.; Li, X.; Miao, R.; Compton, R.G. Single-entity Ti3C2TX MXene electro-oxidation. Appl. Mater. Today 2022, 26, 101335. [Google Scholar] [CrossRef]
  132. Djebbi, M.A.; Allagui, L.; Ayachi, M.S.E.; Boubakri, S.; Jaffrezic-Renault, N.; Namour, P.; Amara, A.B.H. Zero-valent iron nanoparticles supported on biomass-derived porous carbon for simultaneous detection of Cd2+ and Pb2+. ACS Appl. Nano Mater. 2022, 5, 546–558. [Google Scholar] [CrossRef]
  133. Zang, X.B.; Wang, J.; Qin, Y.; Wang, T.; He, C.; Shao, Q.; Zhu, H.W.; Cao, N. Enhancing capacitance performance of Ti3C2Tx MXene as electrode materials of supercapacitor: From controlled preparation to composite structure construction. Nano-Micro Lett. 2020, 12, 77. [Google Scholar] [CrossRef]
  134. Krishnamoorthy, R.; Yu, X.C.; Zhu, P.L.; Hu, Y.; Sun, R.; Wong, C. Exfoliation and defect control of two-dimensional few-layer MXene Ti3C2TX for electromagnetic interference shielding coatings. ACS Appl. Mater. Interfaces 2020, 12, 49737–49747. [Google Scholar] [CrossRef]
  135. Komkova, M.A.; Pasquarelli, A.; Andreev, E.A.; Galushin, A.A.; Karyakin, A.A. Prussian blue modified boron-doped diamond interfaces for advanced H2O2 electrochemical sensors. Electrochim. Acta 2020, 339, 135924. [Google Scholar] [CrossRef]
  136. Rubino, A.; Queirós, R. Electrochemical determination of heavy metal ions applying screen-printed electrodes based sensors. A review on water and environmental samples analysis. Talanta 2023, 7, 100203. [Google Scholar] [CrossRef]
Figure 1. Main Methods for Detecting Heavy Metals in Water Resources.
Figure 1. Main Methods for Detecting Heavy Metals in Water Resources.
Molecules 31 00005 g001
Figure 2. Schematic illustration of the preparation and electroanalysis mechanism of the TCA/MWCNT-modified electrode [59].
Figure 2. Schematic illustration of the preparation and electroanalysis mechanism of the TCA/MWCNT-modified electrode [59].
Molecules 31 00005 g002
Figure 3. Schematic illustration of the interaction between Cd2+ and Pb2+ with the surface of the PrGO/AuNPs/Sal-Cys/GCE [64].
Figure 3. Schematic illustration of the interaction between Cd2+ and Pb2+ with the surface of the PrGO/AuNPs/Sal-Cys/GCE [64].
Molecules 31 00005 g003
Figure 4. Porous biomass-based material synthesized from tea waste using a combination of carbonization and activation [75].
Figure 4. Porous biomass-based material synthesized from tea waste using a combination of carbonization and activation [75].
Molecules 31 00005 g004
Figure 6. Schematic diagram of the heavy metal ion detection setup using a three-electrode system.
Figure 6. Schematic diagram of the heavy metal ion detection setup using a three-electrode system.
Molecules 31 00005 g006
Figure 7. Schematic diagram of heavy metal ion detection in water using DPV (Differential Pulse Voltammetry).
Figure 7. Schematic diagram of heavy metal ion detection in water using DPV (Differential Pulse Voltammetry).
Molecules 31 00005 g007
Figure 8. Cyclic voltammogram of GCE after polishing and cleaning in a 5 mM K3Fe(CN)6/K4Fe(CN)6 solution.
Figure 8. Cyclic voltammogram of GCE after polishing and cleaning in a 5 mM K3Fe(CN)6/K4Fe(CN)6 solution.
Molecules 31 00005 g008
Figure 9. Schematic representation of a glassy carbon electrode (GCE).
Figure 9. Schematic representation of a glassy carbon electrode (GCE).
Molecules 31 00005 g009
Table 1. Summary of LOD and MDL Values.
Table 1. Summary of LOD and MDL Values.
MetalLOD (µg/L)MDL (µg/L)WHO Limit (µg/L)EPA Limit (µg/L)
Pb2+0.4050.4241015
Cd2+0.5650.59235
Hg2+1–512
Table 2. Comparison with Literature.
Table 2. Comparison with Literature.
ElectrodeTechniqueMetalLODReference
Bare GCEDPVPb2+0.405 µg/LThis work
Bare GCEDPVCd2+0.565 µg/LThis work
MXene/GOD-PB/ITOSelf-poweredHg2+1–5 µg/LThis work
Sb electrodePSAPb2+0.03 µg/LWei et al. [51]
Table 3. Key Electrochemical Parameters.
Table 3. Key Electrochemical Parameters.
MetalLinear RangeSensitivityRepeatability (RSD %)Stability
Pb2+µg/L rangeHigh≤5%Good
Cd2+µg/L rangeHigh≤6%Good
Hg2+1–5 µg/LModerate≤7%Good
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

Mukatayeva, Z.; Konarbay, D.; Bakytkarim, Y.; Shadin, N.; Tileuberdi, Y. Analytical Determination of Heavy Metals in Water Using Carbon-Based Materials. Molecules 2026, 31, 5. https://doi.org/10.3390/molecules31010005

AMA Style

Mukatayeva Z, Konarbay D, Bakytkarim Y, Shadin N, Tileuberdi Y. Analytical Determination of Heavy Metals in Water Using Carbon-Based Materials. Molecules. 2026; 31(1):5. https://doi.org/10.3390/molecules31010005

Chicago/Turabian Style

Mukatayeva, Zhazira, Diana Konarbay, Yrysgul Bakytkarim, Nurgul Shadin, and Yerbol Tileuberdi. 2026. "Analytical Determination of Heavy Metals in Water Using Carbon-Based Materials" Molecules 31, no. 1: 5. https://doi.org/10.3390/molecules31010005

APA Style

Mukatayeva, Z., Konarbay, D., Bakytkarim, Y., Shadin, N., & Tileuberdi, Y. (2026). Analytical Determination of Heavy Metals in Water Using Carbon-Based Materials. Molecules, 31(1), 5. https://doi.org/10.3390/molecules31010005

Article Metrics

Back to TopTop