Previous Article in Journal
Research Progress on Biomarkers and Their Detection Methods for Benzene-Induced Toxicity: A Review
Previous Article in Special Issue
Resistive-Based Nanostructured CeO2 Gas Sensors: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Manganese Nanoparticles for Heavy Metal Detection vs. Noble and Base Metal Nanoparticles; Prospects, Limitations, and Applications in Electroanalysis

Analytical Chemistry Laboratory, School of Chemistry, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
*
Author to whom correspondence should be addressed.
Chemosensors 2025, 13(8), 313; https://doi.org/10.3390/chemosensors13080313 (registering DOI)
Submission received: 17 July 2025 / Revised: 14 August 2025 / Accepted: 15 August 2025 / Published: 17 August 2025

Abstract

This review examines the emerging role of manganese-based nanoparticles (Mn-NPs) in detecting heavy metal pollutants in environmental matrices. Heavy metals such as cadmium, lead, zinc, and copper pose serious environmental and health concerns due to their tendency to persist in ecosystems and accumulate in living organisms. As a result, there is a growing need for reliable methods to detect and remove these pollutants. Manganese nanoparticles offer unique advantages that scientists could consider as replacing other metal nanoparticles, which may be more expensive or more toxic. The physicochemical properties of Mn-NPs—including their multiple oxidation states, magnetic susceptibility, catalytic capabilities, and semiconductor conductivity—enable the development of multi-modal sensing platforms with exceptional sensitivity and selectivity. While Mn-NPs exhibit inherently low electrical conductivity, strategies such as transition metal doping and the formation of composites with conductive materials have successfully addressed this limitation. Compared to noble metal nanoparticles (Au, Ag, Pd) and other base metal nanoparticles (Bi, Fe3O4), Mn-NPs demonstrate competitive performance without the drawbacks of high cost, complex synthesis, poor distribution control, or significant aggregation. Preliminary studies retrieved from the Scopus database highlight promising applications of manganese-based nanomaterials in electrochemical sensing of heavy metals, with recent developments showing detection limits in the sub-ppb range. Future research directions should focus on addressing challenges related to scalability, cost-effectiveness, and integration with existing water treatment infrastructure to accelerate the transition from laboratory findings to practical environmental applications.

1. Introduction

In recent years, nanoparticles (NPs) have become one of the key materials in the design of modern technologies such as sensors, drug delivery, water treatment, batteries, fuel cells, and others. Heavy metal contamination significantly threatens the environment and public health, making effective remediation strategies necessary. Metals such as cadmium, lead and copper are considered important environmental pollutants of concern due to their non-biodegradable nature, environmental persistence, and bioaccumulative potential, which can lead to long-term ecotoxicological effects at all trophic levels as they tend to accumulate in the food chain, even at very low concentrations, causing multiple health problems [1]. As for zinc, it shares some of these characteristics and is widely recognized for its biocompatibility, resulting in its extensive use in biological and biomedical applications, particularly in the form of ZnO nanoparticles [2] and nanostructures [3,4].
These metal pollutants exist as natural components within environmental matrices. However, anthropogenic activities have significantly enhanced their prevalence and concentration gradients in ecosystems. Such anthropogenic sources include industrial production processes, intensive agricultural methodologies, fossil fuel combustion technologies, agrochemical applications, vehicle emissions, and wastewater discharge mechanisms. These xenobiotics present significant ecotoxicological concerns due to their environmental persistence, non-biodegradability, and tendency to bioaccumulate and bio magnify in trophic hierarchies, thereby potentially causing chronic systemic toxicity to biological organisms after prolonged periods of exposure [5].
On the other hand, nanotechnology has emerged as a prominent research field in recent decades, attracting significant scientific interest due to its revolutionary impact in multiple fields [6,7]. Nanomaterials and nanoparticles (NPs) are fundamental components of nanotechnology, distinguished by their remarkable flexibility, adaptable architecture, and superior performance. According to the literature, nanomaterials, especially those with a size below 100 nm, although still in an early stage of research, show significant prospects for the detection, measurement, and removal of heavy metal pollutants from various environmental samples thanks to their unique physicochemical characteristics combined with the extensive surface-to-volume ratio that gives them unique properties [8].
More specifically, Mn-NPs possess unique characteristics, such as high magnetic susceptibility, remarkable catalytic activity, excellent biocompatibility, and strong adsorption capacities [9], compared to other metal nanoparticles that have been used, such as AuNPs, BiNPs, etc. In comparison, manganese nanoparticles (Mn-NPs) exhibit increased potential as adsorbents due to their significantly larger surface area, higher surface-to-volume ratio, as in the case of Ramat et al., who synthesized a three-dimensional MnO2 nanofibrous network without the use of any template by a hydrothermal method, with a large surface area [10], and a higher number of defect sites. These defect sites refer to structural imperfections in the nanoparticles, which may include surface atoms with unsaturated bonding [11]. Such sites are chemically more reactive than atoms in a perfect crystal lattice, thus enhancing the ability of nanoparticles to adsorb pollutants.
However, several critical aspects must be considered to ensure the successful implementation of nanomaterial-based remediation technologies. Recent studies by various scientists suggest that manganese nanoparticles can act as effective nano adsorbents for heavy metal ions, e.g., in aqueous environments, thanks to their high surface-to-volume ratio [12] and abundance of active defect sites [13]. Furthermore, their incorporation into electrochemical and optical sensors has shown promising results for the sensitive and selective detection of trace amounts of heavy metals such as lead and cadmium, MnO2@RGO Nanocomposite-Based Electrochemical Sensor [14].
Despite these encouraging developments, few works have been published on heavy metal detection. This review aims to fill this gap by providing a systematic overview of the current progress in the synthesis, functionalization, and application of manganese-based nanomaterials in environmental monitoring.
In addition, scalability, cost-effectiveness, and integration with other remediation techniques are key parameters for the practical application and widespread adoption of nanomaterial-based solutions. Furthermore, ongoing research efforts should focus on addressing key challenges such as the stability, longevity, and multifunctionality of nanomaterials to improve their performance and reliability in real-world conditions. Collaboration between key stakeholders in academia, industry, government agencies, and communities is essential to drive innovation, enhance technology transfer, and accelerate the conversion of research findings into actionable solutions for heavy metal detection using advanced nanomaterials such as Mn-NPs.

2. Chemical and Physical Properties of Manganese

Manganese (Mn), one of the most abundant elements in the Earth’s crust, ranked twelfth overall and third among transition metals after iron and titanium [15], has already established a prominent position among modern nanomaterials (NMs) [16], although its exploitation is still at an early stage. Beyond its geochemical abundance, manganese is an essential micronutrient, playing a vital role in growth, reproduction, and disease prevention in aquatic and terrestrial organisms [17].
Taking advantage of this physical significance, Mn-NPs exhibit unique physicochemical characteristics that significantly differentiate them from their bulk counterparts, such as improved magnetic susceptibility, superior catalytic functionality, excellent biocompatibility, and remarkable adsorption efficiency. These advantageous properties mainly stem from their nanoscale dimensions, which confer an increased surface-to-volume ratio, increased concentration of active defect sites, and optimized surface-to-volume atom distribution. Among the various manganese nanomaterials, manganese oxide (MnxOy) nanostructures represent a particularly versatile class that exhibits remarkable structural diversity [18]. These Mn-NPs are increasingly recognized as promising materials for sustainable nanotechnology due to their exceptional physical and chemical characteristics [15]. Their exceptional utility stems largely from the rich structural and synthetic diversity of manganese oxides, such as MnO, Mn5O8, Mn2O3, MnO2, and Mn3O4, which distinguishes them among three-dimensional transition metal oxides [19].
The multifunctional nature of these nanomaterials is further enhanced by their distinctive optical properties, characterized by specific fluorescence signatures and unique UV-visible absorption profiles [20]. Furthermore, they exhibit strong catalytic capabilities in facilitating various redox processes [21,22], along with excellent adsorption capabilities [23,24,25]. As a result of these combined characteristics, Mn-NPs exhibit remarkable versatility due to their physicochemical properties [26]. Given their exceptional properties—including excellent redox capabilities, high specific surface area, and abundant surface atoms—in combination with their diverse crystalline phases and morphologies, Mn-NPs continue to attract increasing interest due to their unique combination of properties [9,27]. Manganese-based nanomaterials—especially MnO2 and related oxides—combine rich redox chemistry, versatile coordination behavior, intrinsic magnetic and catalytic activities, and semiconductor conductivity, all at low cost and environmental benignity.

2.1. Redox Properties

Manganese exhibits multiple oxidation states ranging from −3 to +7, with the +2 and +4 states being the most thermodynamically stable. Oxidation states of +3 and above readily participate in electron transfer reactions [28,29]. This variability allows manganese to participate in a wide array of redox reactions, making it highly valuable for the development of chemical sensors. The higher oxidation states of the manganese act as strong oxidizing agents, enabling selective interactions with various heavy metal ions. These redox transitions occur at distinct electrochemical potentials, producing unique electrochemical signatures that can be exploited for the selective detection of metal ions [30]. Moreover, manganese oxides possess additional unique features, such as diverse crystal structures, surface nano-architectures, and high surface area, which further enhance their effectiveness in sensor applications through improved analyte interaction and signal transduction [29].

2.2. Magnetic Properties

Manganese-based nanomaterials often exhibit paramagnetic [31] or ferromagnetic behavior, which can be used for magnetic separation, preconcentration of analytes, or the development of magnetic resonance-based detection methods. The inherent magnetic susceptibility of manganese compounds, especially at the nanoscale, facilitates their integration into magnetically assisted detection platforms. This property allows for the efficient separation of target analytes from complex environmental samples, significantly reducing the effects of matrix interference and lowering the detection limits [18].

2.3. Catalytic Activity

Manganese oxides exhibit remarkable catalytic activity [32] and prove to be extremely useful for various oxidation processes, such as the degradation of organic pollutants. Their surface properties and their ability to facilitate electron exchange mechanisms make them effective for such processes.
Manganese-based catalysts have attracted the interest of scientists as economical substitutes for precious metals such as gold, given their flexible functionality and desirable material characteristics. The effectiveness of manganese oxides in catalytic processes depends largely on physical factors such as particle size and geometry, which directly affect their performance in reactions where structure plays a critical role [33].

2.4. Electrical Conductivity

Electrical conductivity is a critical parameter influencing the performance of electrochemical systems, as it governs charge transfer efficiency and the diffusion kinetics of electrolyte ions across the electrode–electrolyte interface during redox reactions. Although manganese oxides are widely recognized for their favorable redox behavior, their inherently low electrical conductivity [34], typically in the range of 10−3 to 10−4 S m−1, presents a significant limitation [35]. To address this limitation, recent studies have focused on strategies such as doping MnOx with transition metals, including Cu, Ni, Co, and Fe, which serve to enhance electron transport by improving the material’s intrinsic conductivity [36]. Additionally, the semiconducting nature of MnOx nanostructures provides distinct characteristics. While they do not match the conductivity of noble metals, these materials exhibit suitable band gap energies and charge carrier mobilities. Upon exposure to chemical species, MnOx-based materials undergo electronic band structure modifications that lead to measurable variations in conductivity, impedance, or potential [37].

2.5. Toxicity and Environmental Considerations of Manganese-Based Nanomaterials

Although manganese is an essential trace element for many biological systems and has been widely utilized in environmental remediation applications due to its high catalytic activity, chronic exposure to elevated levels can result in neurotoxic effects, a condition known as “manganism”, characterized by Parkinsonian-like symptoms in humans [38]. Furthermore, the use of manganese-based nanomaterials has raised concerns regarding their environmental safety. Nevertheless, at the nanoscale, manganese may exhibit altered bioavailability and toxicity profiles. Studies have shown that certain Mn-based nanoparticles can induce oxidative stress and neurotoxicity, as well as negatively affect aquatic organisms [39]. Moreover, the potential release of Mn2+ ions into the environment is an additional risk factor, directly related to the stability [40] of the nanoparticles under real-world conditions. Therefore, a comprehensive evaluation of manganese nanomaterials is imperative, including dissolution behavior, toxicological assessments, and surface modification strategies to mitigate adverse effects [24].

3. Overview of Previous Studies

3.1. Mn-Oxides Nanostructures and Morphology Effects on Heavy Metal Detection

MnO2 nanostructures have emerged as one of the most widely used materials for heavy metal detection, as they combine significant advantages such as cost-effectiveness, excellent adsorption capacities, unique nanomorphological structures, biocompatibility, and catalytic properties [41]. Different morphologies of MnO2, such as nanoparticles, nanocups, and nanotubes, exhibit distinct interaction characteristics when used for heavy metal detection. In particular, the mutual contribution between Cd(II), Pb(II) and Zn(II) [42] ions, and even Cu(II), directly depends on the geometry and surface structure of the nanomaterial. The sensor geometry determines the interactions between different metal ions during detection, and three different morphologies show some regularity in their analytical performance, with MnO2 nanocups outperforming MnO2 nanotubes and MnO2 nanoparticles. This suggests that careful control of nanomaterial morphology is critical for optimizing analytical performance and reducing interference in multi-metal systems [43]. MnFe2O4 has been applied in sensing due to its high electrochemical activity, outstanding adsorption capability, good water dispersibility, and excellent stability [44]. The superior electrochemical performance of MnFe2O4 compared to single-component oxides stems from complementary contributions of the constituent metals. Manganese provides the primary redox-active sites and catalytic activity, while iron enhances structural stability, contributes magnetic properties, and offers additional adsorption sites. The observed performance enhancement results from synergistic interactions between Mn and Fe components that modify their individual electronic and structural properties, creating emergent characteristics that exceed simple additive effects.
Manganese-based nanomaterials—especially MnO2 and related oxides—enable multi-modal sensing platforms (electrochemical, colorimetric, magnetic separation) for heavy metals in water, soil, and sediments. A hallmark study by Siyamthanda et al. demonstrated a MnO2@RGO electrochemical sensor achieving simultaneous Zn2+/Cd2+/Cu2+ detection down to 0.002–0.015 µg L−1 (S/N = 3) in surface water [14]. The excellent performance of the MnO2@RGO composite material stems from the complementary roles of the components. MnO2, on the one hand, provides electroactive sites and catalytic properties for the detection of heavy metals through multiple oxidation states and affinity for metal ions, while on the other hand, RGO contributes high surface area, conductivity and structural support. The remarkable sensitivity of the composite material demonstrates synergistic enhancement of the detection capabilities of MnO2 beyond the additive effects.
Further to this, Prashanth et al. developed an electroactive Mns/graphitic carbon electrode material with which the simultaneous determination of Pb(II), Cd(II) was successfully achieved. Its advantages and originality include environmental friendliness, long-term storage capability, and operational stability for a period of several months without significant loss of electrocatalytic activity, making it suitable for in situ applications [45]. In Mn–carbon composites, the performance improvement appears to stem from complementary properties. The carbon matrix provides structural integrity, electron transport, and stability, while the Mn species provide electroactive sites and catalytic activity for metal ion sensing. This functional synergy explains the superior functional stability compared to pure Mn-based electrodes.
For biosensor applications, MnO2-modified carbon electrodes are simple, inexpensive, and non-toxic to prepare, while remaining stable against interfering species like persulfate and oxygen, unlike other modifiers such as Prussian Blue or ferrocene-polyaniline. They function effectively in neutral and slightly alkaline conditions, making them compatible with oxidase-catalyzed reactions and superior in sensitivity to platinum-modified electrodes. This has led to commercial adoption by Roche Diagnostics for glucose, lactate, urea, and creatinine biosensors [46].
An important advantage of manganese oxides lies in their natural ability to dissociate interfering compounds in complex sample matrices, which improves the accuracy of sensors and potentially reduces measurement errors from unwanted chemical interactions. So far, a brief review of the literature indicates a multitude of applications related, for example, to the biomedical sector [24,47,48], while their application in the detection of heavy metals has now begun due to their magnetic properties, low toxicity, and catalytic properties.
As a result of these combined characteristics, Mn-NPs exhibit remarkable versatility in a wide range of applications, including molecular sieves, solar cells, batteries, catalysts, magnetic materials, optoelectronics, drug delivery systems, imaging contrast agents, magnetic storage devices, water treatment and purification technologies [26].
Given their exceptional properties—including excellent redox capabilities, high specific surface area, and abundant surface atoms—in combination with their diverse crystalline phases and morphologies, Mn-NPs continue to attract increasing interest for applications in energy storage [27] (batteries, supercapacitors), energy conversion, and environmental catalysis [9].

3.2. Manganese Nanoparticles for Electrochemical Detection and Removal of Heavy Metals

Mn-NPs demonstrate significant potential across various applications due to their unique structural and electrochemical properties. Preliminary evaluations have shown promising results when using these nanoparticles as electrochemical modifiers for the voltammetric detection of heavy metal ions, particularly lead (Pb2+) and cadmium (Cd2+). This approach represents an important advancement in environmental monitoring technologies.
Although not yet fully exploited, the literature highlights several notable examples demonstrating the application of MnCo2O4 and related Mn-NPs in environmental monitoring. For example, iron-manganese oxide nanomaterials have been used to remove heavy metals from aqueous media, with their capacity and efficiency being influenced by synthesis methods and environmental conditions [30]. Fe-Mn composites combine Fe’s magnetic separability with Mn’s metal ion affinity, resulting in enhanced removal efficiency through complementary properties. A similar approach was reported by Zhang et al. [23], who used manganese dioxide (MnO2) nanoparticles supported on biochar (BC@MnO2) to adsorb Cd(II), Cu(II), Zn(II), and Pb(II) ions from water samples. In BC@MnO2, biochar provides porosity and structural support while MnO2 contributes active sites and metal ion binding. The superior performance versus individual components demonstrates synergistic effects where biochar enhances accessibility while MnO2 provides selective interactions. Hybrid nanosorbents, such as HMN-MnO2 nanoparticles, have also shown promising results in the removal of heavy metals from wastewater and the passivation of metal contaminants in soil [49].
In 2017, Junjie et al. successfully used a kind of monodispersed ZnS: Mn/SiO2 nanoparticles containing a ZnS: Mn core inside them for the detection of copper in seawater. The concentration of Cu2+ ions in water samples ranged from 8.16 × 10−8 to 4.96 × 10−4 mol/L, thus providing a highly sensitive fluorescence sensor for the detection of Cu2+ in seawater [50]. While Mn2O3 nanoparticles assembled into hierarchical hollow spheres (H-Mn2O3) were used as nanocarriers for the immobilization of glycoxidase (GOx), horseradish peroxidase (HRP), and sequence signal (S3), forming the S3/H-Mn2O3/HRP/GOx complex. The hierarchical structure of the hollow spheres provides a large specific surface area and efficient loading of biomolecules, while the GOx and HRP enzymes allow the catalytic deposition of polyaniline (PANI) with excellent electrochemical activity, acting as a signal-enhancing label for Pb2+ detection [51].
Zhou et al. synthesized and evaluated L-cysteine functionalized mesoporous MnFe2O4 nanospheres (MnFe2O4@Cys) for electrochemical detection of Pb(II), Hg(II), Cu(II), and Cd(II) via SWASV. Characterization confirmed the successful functionalization, and electrochemical tests revealed enhanced activity due to the presence of L-cysteine. The sensor showed high sensitivity, especially for Pb(II), and enabled both single and simultaneous detection of multiple ions with remarkable selectivity, stability, and reproducibility, making it a powerful platform for environmental monitoring [52]. MnFe2O4@Cys performance enhancement demonstrates complementary roles: the MnFe2O4 core provides electrochemical activity and magnetic properties, while L-cysteine functionalization adds specific binding sites and selectivity. This shows that surface functionalization is crucial for optimizing the fundamental sensing capabilities of the Mn-Fe composite.
A MnFe2O4/graphene oxide (GO) nanocomposite was also used for sensitive detection of Pb(II) by applying anodic stripping square wave voltammetry (SWASV). The successful incorporation of MnFe2O4 with GO improved the material surface area, conductivity, and electrochemical activity. The resulting sensor exhibited high sensitivity, low detection limit, and excellent selectivity even in the presence of Zn(II), along with good stability, reproducibility, and practical application for the detection of heavy metal ions [53]. In the MnFe2O4/GO system, GO provides conductivity and surface area, Fe contributes magnetic properties and stability, while Mn supplies electroactive sites and catalytic activity. The superior performance versus binary combinations demonstrates that optimal sensing requires synergistic integration of conductivity enhancement, magnetic separation, and electrochemical activity. A related advancement was reported by Siyamthanda et al., who developed a modified glassy carbon electrode incorporating MnO2 supported on reduced graphene oxide (MnO2@RGO) for the simultaneous detection of Cu(II), Cd(II), and Zn(II) in surface waters using Differential Pulse Anodic Stripping Voltammetry (DPASV). The sensor demonstrated excellent electrocatalytic activity and stability, which were attributed to its high surface area and low charge transfer resistance. Optimization using response surface methodology significantly enhanced its performance, enabling direct analysis of water samples without the need for pretreatment.
Furthermore, Mn2O3 nanoparticles were synthesized and used to modify a carbon paste electrode, creating a highly sensitive sensor for the detection of Cd(II) and Pb(II) ions. The sensor demonstrated excellent detection limits and was successfully applied to real wastewater samples, achieving recovery rates of 80–95%, further confirming its potential for environmental applications [54]. Mn2O3-modified carbon paste electrodes show performance enhancement through complementary roles: carbon paste provides the conductive matrix and support, while Mn2O3 introduces electroactive sites and catalytic properties for selective heavy metal detection. Transition metal oxide nanoparticles, such as MnO2 (and Fe2O3) are also widely used in energy storage devices, such as lithium batteries and supercapacitors, due to their high capacity and large surface area, and offer performance and safety benefits in defense applications [55]. Amany et al. in their published work incorporated MnO2 nanoparticles, aiming to enhance the electrochemical properties of the 5-BHAHS@NC they had already synthesized. This low-cost detection platform, 5-BHAHS@NC/MnO2 CPE, was successfully used for the rapid and simultaneous determination of cadmium and lead ions in fish samples, tap water, and wastewater [56].
In another study, a simple colorimetric method uses octahedral Mn3O4 nanoparticles as nanozymes with oxidase-mimicking activity. Oligonucleotides inhibit the catalytic activity of Mn3O4 NPs through noncompetitive inhibition. When heavy metals like Hg(II) and Cd(II) bind to oligonucleotides, they change the oligonucleotide conformation, restoring the nanozyme activity and producing visible color changes. The method requires no hydrogen peroxide, making the color response stable and intuitive. This simple system contains only unmodified octahedral Mn3O4 NPs and oligonucleotides, offering a cost-effective approach for environmental and agricultural monitoring that can be extended to other nanozymes and detection targets [57]. On the other hand, Mei et al. reported that materials such as FeOOH and FMBO prepared by the oxidation-co-precipitation method gave satisfactory results in the removal of Cd(II), Ni(II), and Pb(II) in single and multiple systems [58].
Overall, these findings highlight the significant potential of manganese-based nanomaterials as advanced electrochemical sensing platforms for routine environmental monitoring of heavy metal pollutants. The comparative analysis demonstrates that Mn-based materials provide an optimal balance of performance, cost-effectiveness, and environmental compatibility. While pure Mn systems offer excellent sensitivity and low cost, the incorporation of carbon materials significantly enhances performance through synergistic effects, whereas Fe incorporation primarily improves stability at the cost of sensitivity. This positions Mn-based composites as highly competitive alternatives to precious metal-based sensors for practical environmental applications. When combined with voltammetric techniques, Mn-NPs have demonstrated high efficiency in the detection and quantification of heavy metal ions, particularly Cd(II) and Pb(II), in various aqueous matrices, including distilled water, tap water, and river samples. Their characteristic porous, flower-like morphology enhances the interaction of analytes, making these materials particularly suitable for environmental monitoring applications [59].

4. Comparative Analysis of Mn-NPs with Other Metallic Nanoparticles

Recent advances in nanotechnology have led to the widespread replacement of conventional chemical reagents with nanomaterials for the detection of heavy metal ions. Materials such as metal nanoparticles (MNPs) and various metal-based nanostructures offer remarkable physicochemical properties, including a high surface area, enhanced surface energy, and superior electron transfer capabilities. These characteristics significantly contribute to improved sensitivity and selectivity in electrochemical sensing platforms. Researchers have exploited both noble and base metal nanoparticles for heavy metal sensing, taking advantage of their high surface-to-volume ratios, tunable surface chemistry, and catalytic/optical properties.

4.1. Noble-Metal Nanoparticles

Noble metal nanoparticles (NMNPs), including gold (AuNPs), silver (AgNPs), and palladium (PdNPs), hold significant promise for next-generation point-of-need diagnostic technologies and environmental sensing applications due to their unique physicochemical, optical, thermal, and electronic properties. These properties—such as size, shape, elemental composition, and internal structure—play a critical role in determining their performance as signal reporters and their interactions with target molecules in analyte samples. Their high surface-to-volume ratios, stability, and biocompatibility further enable functionalization with a variety of ligands for the selective detection of clinically relevant biomarkers and heavy metal ions. Therefore, achieving large-scale production of NMNPs with controlled and consistent size and shape is essential for the development of high-performance biosensing platforms [60].
Due to their distinctive physicochemical properties, gold nanoparticles (AuNPs) have emerged as significant components in nanotechnology-based sensing applications, particularly for heavy metal detection. Their stability, biocompatibility, and exceptional optical and electronic characteristics, combined with high surface-to-volume ratios, enable functionalization with diverse ligands for selective analyte detection [61].
In environmental monitoring, AuNPs excel at detecting heavy metal ions, including Pb2+ and Cd2+ [62,63]. When integrated with complementary materials, they form composite systems with enhanced performance for specific metal targets.
The morphology and dimensions of AuNPs significantly influence sensing efficiency. While spherical particles (4–298 nm) predominate in electrochemical applications, various shapes can be synthesized through controlled reaction conditions, as Dutta et al. demonstrated in their work with gold nano-stars [64].
A notable electrode configuration combining AuNPs with L-cysteine and reduced graphene oxide exhibited exceptional performance for cadmium detection across diverse water samples. This same sensor achieved remarkably high sensitivity for Pb(II) detection, illustrating the synergistic benefits of AuNP-nanomaterial combinations in electrochemical sensing applications [63].
Despite their advantages, gold nanoparticle (AuNP) sensors face significant limitations. Their performance varies considerably with size and shape, affecting optical properties and particle distribution, leading to inconsistent responses [65]. Toxicological concerns persist due to limited information on long-term effects and their ability to circulate without immune rejection [66]. AuNPs can accumulate in ecosystems and enter food chains [67], with toxicity influenced by physicochemical properties [68,69]. Their large surface area-to-volume ratio increases reactivity and cellular penetration [70], potentially enhancing toxicity on a mass dosage basis [46]. Furthermore, economic factors restrict widespread adoption, such as gold’s inherent expense, sophisticated synthesis requirements, and specialized equipment make large-scale production challenging, particularly for environmental monitoring in resource-limited settings.
Silver nanoparticles (AgNPs), a distinct subclass of nanomaterials, have garnered considerable interest as electrode modifiers in the electrochemical detection of heavy metal ions such as Cd(II), Pb(II), and Cu(II) [71,72]. This interest is primarily attributed to their favorable properties, including excellent electrical conductivity, large surface area, and facile synthesis [73]. Numerous studies have demonstrated that integrating nanomaterials with electrochemical methods significantly enhances the speed and efficiency of heavy metal ion detection. While electrodeposition of AgNPs onto the electrode surface offers significant advantages, including the absence of unwanted by-products, elimination of chemical oxidants and reductants, and reduced synthesis time, it also presents notable limitations. These disadvantages include poor control over nanoparticle distribution, time-intensive processing requirements, and the tendency for nanoparticles to aggregate on the electrode surface, potentially compromising the desired functional properties of the modified electrode [74].
Furthermore, palladium nanoparticles (PdNPs) have been extensively investigated as sensing platforms for heavy metal detection. The predominant methodology involves synthesizing porous activated carbon (PAC) followed by decoration with PdNPs through a one-step thermal reduction process, with minor variations in experimental parameters. Literature reports indicate that spherical Pd nanoparticles (20–30 nm diameter) have been successfully employed for both individual and simultaneous detection of Cd(II), Pb(II), and Cu(II) via square wave anodic stripping voltammetry (SWASV) [75,76]. Although this approach demonstrated efficacy in real environmental sample analyses, a significant limitation was observed: detection limits during simultaneous multi-element measurements were substantially elevated compared to single-metal analyses, thereby compromising analytical performance in complex environmental matrices.

4.2. Base-Metal Nanoparticles

On the other hand, several research groups have investigated bismuth nanoparticles (BiNPs) for the effective detection of heavy metals such as cadmium and lead ions [77,78]. Through the application of anodic square wave stripping voltammetry and modified bismuth-nano powder electrodes, the detection of zinc, cadmium, and lead ions can be readily achieved [79]. Despite their advantages, bismuth nanoparticles (BiNPs) present certain limitations. A significant challenge is the interference from elevated copper(II) ion concentrations, which can compromise detection accuracy [62,80]. Additionally, the synthesis of particles with uniform size distribution remains a considerable technical challenge.
Iron nanoparticles (Fe3O4) have emerged as prominent candidates among various metal oxides due to their excellent affinity for heavy metal ions, remarkable chemical stability, biocompatibility, superparamagnetism, low toxicity, simplicity of synthesis, high surface-to-volume ratio, and cost-effectiveness [81]. Daniel et al. reported that zero-valent iron nanoparticles have demonstrated significant antibacterial properties, for example against B. subtilis, Pseudomonas fluorescens, and Gram-negative E. coli [82].
These advantageous properties have allowed Fe3O4-based nanoparticles to facilitate the simultaneous detection of multiple heavy metal species such as Zn(II), Pb(II) [83], and Cd(II) from environmental waters and soil matrices. Iron oxide nanoparticles have also been reported as effective adsorbents for highly toxic heavy metals, such as Pb, Zn, and Cd, as well as Hg and Ni, which exhibit harmful effects on plants, animals, and humans even at low concentrations [84,85,86].
Consequently, Fe3O4-derived nanomaterials have been extensively developed as electrochemical sensing platforms, particularly when forming composites with functional materials, including reduced graphene oxide (RGO) [87,88] and EDTA-functionalized polymers [89]. It is worth noting, however, that Fe3O4 nanoparticles inherently tend to aggregate, significantly reducing their electrochemical performance capabilities. This limitation explains the relatively limited literature on the applications of unmodified iron oxide nanoparticles in electrochemical sensing systems [90,91].
Based on the comparative analysis in Table 1, manganese oxide nanoparticles (MnOx-NPs) represent a balanced solution for heavy metal detection compared to other types of nanoparticles. The stable surface area range of 30–200 m2/g ensures reproducible electrochemical behavior across different synthesis methods. Key advantages include the multiple oxidation states (−3 to +7) that allow for flexible surface chemistry and the inherent catalytic activity that enhances signal amplification.
Furthermore, although MnOx nanoparticles do not reach the extremely low detection limits of gold nanoparticles (AuNPs, e.g., 0.0064 ng/L for Cu2+), they still offer satisfactory sensitivity. While the sensitivity may be lower than others, MnOx nanoparticles provide a consistent surface area (30–200 m2/g) and hold a significant advantage in terms of low cost and overall practicality. In conclusion, MnOx nanoparticles combine low synthesis cost, adequate sensitivity, as well as magnetic separation capability—a feature shared only by Fe3O4 nanoparticles. Overall, all of this makes them a practical and cost-effective option for multi-metal detection.

5. Conclusions

Mn-NPs have emerged as a promising class of materials for the detection of heavy metals in environmental matrices, offering an exciting alternative to more extensively studied nanomaterials. Their unique combination of redox flexibility, magnetic properties, catalytic activity, and environmental compatibility provides significant advantages in sensing applications. The redox chemistry of manganese, spanning oxidation states from −3 to +7, enables dynamic interactions with target analytes, thereby enhancing both the sensitivity and specificity of detection.
Compared to noble metal nanoparticles such as gold, silver, and palladium, as well as base metal alternatives like bismuth and iron oxide (Fe3O4), Mn-NPs demonstrate competitive performance at lower cost and with simpler synthesis processes. Specifically, MnO2@RGO-NPs achieve detection limits of 15.1 ng/L for Cd2+, 9.3 ng/L for Cu2+, and 13.7 ng/L for Zn2+, while maintaining excellent stability and a surface area of 148.7 m2/g. Moreover, they offer improved control over particle distribution and reduced aggregation. Although their inherently low electrical conductivity poses a challenge, this limitation can be effectively addressed through strategies such as doping with transition metals and forming composites with conductive materials.
Electrochemical techniques have shown that Mn-NPs can achieve highly sensitive detection of toxic heavy metals—including Cd(II), Pb(II), Cu(II), and Zn(II)—with detection limits reaching the sub-ppb range. The stable surface area range of 30–200 m2/g for various Mn-NPs ensures reproducible electrochemical behavior across different synthesis methods. These characteristics make them particularly attractive for environmental monitoring, where high sensitivity, selectivity, and cost-effectiveness are essential.
Despite these advantages, challenges remain regarding the long-term stability, standardization, and field deployment of Mn-NP-based sensors. Ongoing research should focus on enhancing sensor robustness, developing scalable fabrication techniques, and integrating these systems into existing monitoring infrastructure. Furthermore, advanced surface functionalization strategies may facilitate the development of multifunctional nanocomposites capable of both detecting and removing heavy metal pollutants.
To fully realize the potential of Mn-NPs in addressing global heavy metal contamination, collaborative efforts among academia, industry, and regulatory agencies are crucial. Additionally, considerations such as scalability, cost-effectiveness, and integration with complementary remediation technologies are key to the practical implementation and widespread adoption of Mn-NP-based environmental solutions.

Author Contributions

Conceptualization, V.K., S.G.; methodology, V.K., S.G.; software, S.G., V.K.; validation, V.K., S.G.; formal analysis, V.K.; investigation, V.K., S.G.; resources, V.K., S.G.; data curation, V.K., S.G.; writing—original draft preparation, V.K.; writing—review & editing, S.G., V.K.; visualization, S.G., V.K.; supervision, S.G.; Project administration, V.K., S.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bian, L.; Li, Y.-J.; Li, J.; Nie, J.-N.; Dong, F.-Q.; Song, M.-X.; Wang, L.-S.; Dong, H.-L.; Li, H.-L.; Nie, X.-Q.; et al. Photovoltage response of (XZn)Fe2O4-BiFeO3 (X = Mg, Mn or Ni) interfaces for highly selective Cr3+, Cd2+, Co2+ and Pb2+ ions detection. J. Hazard. Mater. 2017, 336, 174–187. [Google Scholar] [CrossRef]
  2. Hoseinpour, V.; Souri, M.; Ghaemi, N.; Shakeri, A. Optimization of green synthesis of ZnO nanoparticles by Dittrichia graveolens (L.) aqueous extract. Health Biotechnol. Biopharma 2017, 1, 39–49. [Google Scholar]
  3. Xu, M.; Hadi, P.; Chen, G.; McKay, G. Removal of cadmium ions from wastewater using innovative electronic waste-derived material. J. Hazard. Mater. 2014, 273, 118–123. [Google Scholar] [CrossRef]
  4. Yu, S.; Pang, H.; Huang, S.; Tang, H.; Wang, S.; Qiu, M.; Chen, Z.; Yang, H.; Song, G.; Fu, D. Recent advances in metal-organic framework membranes for water treatment: A review. Sci. Total Environ. 2021, 800, 149662. [Google Scholar] [CrossRef]
  5. Aragay, G.; Pons, J.; Merkoçi, A. Recent trends in macro-, micro-, and nanomaterial-based tools and strategies for heavy-metal detection. Chem. Rev. 2011, 111, 3433–3458. [Google Scholar] [CrossRef] [PubMed]
  6. Hoseinpour, V.; Ghaemi, N. Novel ZnO–MnO2–Cu2O triple nanocomposite: Facial synthesis, characterization, antibacterial activity and visible light photocatalytic performance for dyes degradation—A comparative study. Mater. Res. Express 2018, 5, 085012. [Google Scholar] [CrossRef]
  7. Ahmed, S.; Annu; Chaudhry, S.A.; Ikram, S. A review on biogenic synthesis of ZnO nanoparticles using plant extracts and microbes: A prospect towards green chemistry. J. Photochem. Photobiol. B Biol. 2017, 166, 272–284. [Google Scholar] [CrossRef]
  8. Servos, M.R. Nanotechnology for Water Treatment and Purification; Springer International Publishing: Cham, Switzerland, 2014. [Google Scholar]
  9. Yadav, P.; Bhaduri, A.; Thakur, A. Manganese Oxide Nanoparticles: An Insight into Structure, Synthesis and Applications. ChemBioEng Rev. 2023, 10, 363–610. [Google Scholar] [CrossRef]
  10. Rahmat, M.; Rehman, A.; Rahmat, S.; Bhatti, H.N.; Iqbal, M.; Khan, W.S.; Bajwa, S.Z.; Rahmat, R.; Nazir, A. Highly efficient removal of crystal violet dye from water by MnO2 based nanofibrous mesh/photocatalytic process. J. Mater. Res. Technol. 2019, 8, 5149–5159. [Google Scholar] [CrossRef]
  11. Musil, M.; Choi, B.; Tsutsumi, A. Morphology and Electrochemical Properties of α-, β-, γ-, and δ-MnO2 Synthesized by Redox Method. J. Electrochem. Soc. 2015, 162, A2058–A2065. [Google Scholar] [CrossRef]
  12. Prashanth, S.A.; Ashoka, S.; Pandurangappa, M. Fabrication of new Calix[4]arene functionalized Mn3O4 nanoparticles based modified glassy carbon electrode as fast responding sensor towards Pb2+ and Cd2+ ions. Anal. Methods 2019, 11, 813–820. [Google Scholar]
  13. Kong, J.; Coolahana, K.; Mugweru, A. Manganese based magnetic nanoparticles for heavy metal detection and environmental remediation. Anal. Methods 2013, 5, 5128–5133. [Google Scholar] [CrossRef]
  14. Siyamthanda, H.M.; Munonde, T.S.; Nomngongo, P.N. MnO2@Reduced Graphene Oxide Nanocomposite-Based Electrochemical Sensor for the Simultaneous Determination of Trace Cd(II), Zn(II) and Cu(II) in Water Samples. Membranes 2021, 11, 517. [Google Scholar]
  15. Hoseinpour, V.; Ghaemi, N. Green synthesis of manganese nanoparticles: Applications and future perspective–A review. J. Photochem. Photobiol. B Biol. 2018, 189, 234–243. [Google Scholar] [CrossRef]
  16. Panda, B.; Lenka, A.; Dixit, P.K.; Dash, S.K. Biosynthesis, Biofunctionalization, and Bioapplications of Manganese Nanomaterials: An Overview. In Biomaterials-Based Sensors; Springer: Singapore, 2023; pp. 73–100. [Google Scholar]
  17. Asaikkutti, A.; Bhavan, P.S.; Vimala, K.; Karthik, M.; Cheruparambath, P. Dietary supplementation of green synthesized manganese-oxide nanoparticles and its effect on growth performance, muscle composition and digestive enzyme activities of the giant freshwater prawn Macrobrachium rosenbergii. J. Trace Elem. Med. Biol. 2016, 35, 7–17. [Google Scholar] [CrossRef] [PubMed]
  18. Ding, B.; Zheng, P.; Ma, P.; Lin, J. Manganese oxide nanomaterials: Synthesis, properties, and theranostic applications. Adv. Mater. 2020, 32, e1905823. [Google Scholar] [CrossRef]
  19. Prasad, A.S. Green synthesis of nanocrystalline manganese (II, III) oxide. Mater. Sci. Semicond. Process. 2017, 71, 342–347. [Google Scholar] [CrossRef]
  20. Tirukoti, M.; Shiddappa, L.B.; Vadiraj, K.T. Manganese oxide nanoparticles synthesis route, characterization and optical properties. Mater. Today Proc. 2023, 75, 72–76. [Google Scholar]
  21. Stobbe, E.; de Boer, B.; Geus, J. The reduction and oxidation behaviour of manganese oxides. Catal. Today 1999, 47, 161–167. [Google Scholar] [CrossRef]
  22. Mao, L.; Sotomura, T.; Nakatsu, K.; Koshiba, N.; Zhang, D.; Ohsaka, T. Electrochemical characterization of catalytic activities of manganese oxides to oxygen reduction in alkaline aqueous solution. J. Electrochem. Soc. 2002, 149, A504–A507. [Google Scholar] [CrossRef]
  23. Zhang, H.; Xu, F.; Xue, J.; Chen, S.; Wang, J.; Yang, Y. Enhanced removal of heavy metal ions from aqueous solution using manganese dioxide-loaded biochar: Behavior and mechanism. Sci. Rep. 2020, 10, 6067. [Google Scholar] [CrossRef]
  24. Haque, S.; Tripathy, S.; Patra, C.R. Manganese-based advanced nanoparticles for biomedical applications: Future opportunity and challenges. Nanoscale 2021, 13, 16405–16426. [Google Scholar] [CrossRef] [PubMed]
  25. Ozkan-Ariksoysal, D.; Pantelidou, E.; Dendrinou-Samara, C.; Girousi, S. Nanoparticle-Based DNA Biosensor: Synthesis of Novel Manganese Nanoparticles Applied in the Development of a Sensitive Electrochemical Double-Stranded/Single-Stranded DNA Biosensor. Micromachines 2025, 16, 232. [Google Scholar] [CrossRef]
  26. Dippong, T.; Levei, E.A.; Cadar, O. Recent Advances in Synthesis and Applications of MFe2O4 (M = Co, Cu, Mn, Ni, Zn) Nanoparticles. Nanomaterials 2021, 11, 1560. [Google Scholar] [CrossRef] [PubMed]
  27. Nazish, P.; Sajid, A.A.; Mohd, Z.A.; Mohammad, O.A. Manganese oxide as an effective electrode material for energy storage: A review. Environ. Chem. Lett. 2021, 20, 283–309. [Google Scholar] [CrossRef]
  28. Sobanska, Z.; Roszak, J.; Kowalczyk, K.; Stepnik, M. Applications and Biological Activity of Nanoparticles of Manganese and Manganese Oxides in In Vitro and In Vivo Models. Nanomaterials 2021, 11, 1084. [Google Scholar] [CrossRef]
  29. Carney, J.R.; Dillon, B.R.; Thomas, S.P. Recent Advances of Manganese Catalysis for Organic Synthesis. Eur. J. Org. Chem. 2016, 2016, 3912–3929. [Google Scholar] [CrossRef]
  30. Li, M.; Kuang, S.; Kang, Y.; Ma, H.; Dong, J.; Guo, Z. Recent advances in application of iron-manganese oxide nanomaterials for removal of heavy metals in the aquatic environment. Sci. Total Environ. 2022, 819, 153157. [Google Scholar] [CrossRef]
  31. Kim, E.-J.; Lee, C.-S.; Chang, Y.-Y.; Chang, Y.-S. Hierarchically Structured Manganese Oxide-Coated Magnetic Nanocomposites for the Efficient Removal of Heavy Metal Ions from Aqueous Systems. ACS Appl. Mater. Interfaces 2013, 5, 9628–9634. [Google Scholar] [CrossRef]
  32. Li, L.; Zhao, L.; Ma, J.; Tian, Y. Preparation of graphene oxide/chitosan complex and its adsorption properties for heavy metal ions. Green Process. Synth. 2020, 9, 294–303. [Google Scholar] [CrossRef]
  33. Dey, S.; Kumar, V.V.P. The performance of highly active manganese oxide catalysts for ambient conditions carbon monoxide oxidation. Curr. Res. Green Sustain. Chem. 2020, 3, 100012. [Google Scholar] [CrossRef]
  34. Yu, L.; Li, Y.; Ruan, Y. Fe–Mn Oxides Based Multifunctional Adsorptive/Electrosensing Nanoplatforms: Dynamic Site Rearrangement for Metal Ion Selectivity. Environ. Sci. Technol. 2021, 55, 3967–3975. [Google Scholar] [CrossRef]
  35. Hu, Y.; Wu, Y.; Wang, J. Manganese-Oxide-Based Electrode Materials for Energy Storage Applications: How Close Are We to the Theoretical Capacitance? Adv. Mater. 2018, 30, 1802569. [Google Scholar] [CrossRef]
  36. Huang, M.; Li, F.; Dong, F.; Zhang, Y.X.; Zhang, L.L. MnO2-based nanostructures for high-performance supercapacitors. J. Mater. Chem. A 2015, 3, 21380. [Google Scholar] [CrossRef]
  37. Sonika, D.; Gupta, A.; Manita, K.; Biplab, B.; Ganesh, L.; Niranjan, P. Manganese dioxide nanoparticles: Synthesis, application and challenges. Bull. Mater. Sci. 2020, 43, 277. [Google Scholar] [CrossRef]
  38. Chen, P.; Bornhorst, J.; Aschner, M. Manganese metabolism in humans. Front. Biosci. 2018, 23, 1655–1679. [Google Scholar] [CrossRef] [PubMed]
  39. Neeraj, K.; Supriya, T.T.; Kotha, S.R. Multi biomarker approach to assess manganese and manganese nanoparticles toxicity in Pangasianodon hypophthalmus. Sci. Rep. 2023, 13, 8505. [Google Scholar] [CrossRef] [PubMed]
  40. Xue, S.; Jing, P.; Xu, W. Hemin on graphene nanosheets functionalized with flower-like MnO2 and hollow AuPd for the electrochemical sensing lead ion based on the specific DNAzyme. Biosens. Bioelectron. 2016, 86, 958–965. [Google Scholar] [CrossRef]
  41. Bhuvaneswari, K.; Radha, S.; Sreeja, B.S.; Kumar, P.S. Development of in-situ electrochemical heavy metal ion sensor using integrated 1D/0D/1D hybrid by MWCNT and CQDs supported MnO2 nanomaterial. Environ. Res. 2023, 225, 115570. [Google Scholar] [CrossRef]
  42. Zhang, Q.X.; Wen, H.; Peng, D.; Fu, Q.; Huang, X.J. Interesting interference evidences of electrochemical detection of Zn(II), Cd(II) and Pb(II) on three different morphologies of MnO2 nanocrystals. J. Electroanal. Chem. 2015, 739, 89–96. [Google Scholar] [CrossRef]
  43. Zhang, Q.X.; Peng, D.; Huang, X.J. Effect of morphology of α-MnO2 nanocrystals on electrochemical detection of toxic metal ions. Electrochem. Commun. 2013, 34, 270–273. [Google Scholar] [CrossRef]
  44. Zhang, Y.; Yang, X.; Zhang, Y.; Li, W.; Hao, X.; Liu, T.; Liang, X.; Liu, F.; Liu, F.; Yan, X.; et al. High-Performance Electrochemical Sensor Based on Mn1-xZnxFe2O4 Nanoparticle/Nafion-Modified Glassy Carbon Electrode for Pb2+ Detection. J. Electrochem. Soc. 2019, 166, B341–B348. [Google Scholar] [CrossRef]
  45. Prashanth, S.A.; Vijaya, K.G.; Craig, E.B.; Ashoka, S. One-pot synthesis of Mn3O4/graphitic carbon nanoparticles for simultaneous nanomolar detection of Pb(II), Cd(II) and Hg(II). J. Mater. Sci. 2017, 53, 4961–4973. [Google Scholar]
  46. Negussie, W.B.; Kotzian, P.; Schachl, K.; Alemu, H.; Turkušic, E.; Copra, A.; Moderegger, H.; Švancara, I.; Vytras, K.; Kalcher, K. (Bio)sensors based on manganese dioxide-modified carbon substrates: Retrospections, further improvements and applications. Talanta 2004, 64, 1151–1159. [Google Scholar]
  47. Hsu, B.Y.W.; Kirby, G.; Tan, A.; Seifalian, A.M.; Li, X.; Wang, J. Relaxivity and toxicological properties of Manganese oxide nanoparticles for MRI applications. RSC Adv. 2016, 6, 45462. [Google Scholar] [CrossRef] [PubMed]
  48. Hyon, B.N.; Jung, H.L.; Kwangjin, A.; Yong, I.P.; Mihyun, P.; In, S.L.; Do-Hyun, N.; Sung, T.K.; Seung-Hoon, K.; Sang-Wook, K.; et al. Development of a T1 Contrast Agent for Magnetic Resonance Imaging Using MnO Nanoparticles. Angew. Chem. Int. Ed. 2007, 46, 5397–5401. [Google Scholar]
  49. Zhang, H.; Zhang, Y.; Pan, Y.; Wang, F.; Sun, Y.; Wang, S.; Wang, Z.; Wu, A.; Zhang, Y. Efficient removal of heavy metal ions from wastewater and fixation of heavy metals in soil by manganese dioxide nanosorbents with tailored hollow mesoporous structure. Chem. Eng. J. 2023, 459, 141583. [Google Scholar] [CrossRef]
  50. Qin, J.; Dong, B.; Gao, R.; Su, G.; Han, J.; Li, X.; Liu, W.; Wang, W.; Cao, L. Water-soluble silica-coated ZnS:Mn nanoparticles as fluorescent sensor for ultratrace copper(II) ions in seawater. Anal. Methods 2017, 9, 322–328. [Google Scholar] [CrossRef]
  51. Ge, S.; Wu, K.; Zhang, Y.; Yan, M.; Yu, J. Paper-based biosensor relying on flower-like reduced graphene guided enzymatically deposition of polyaniline for Pb2+ detection. Biosens. Bioelectron. 2016, 80, 215–221. [Google Scholar] [CrossRef] [PubMed]
  52. Zhou, S.-F.; Wang, J.-J.; Gan, L.; Han, X.-J.; Fan, H.-L.; Mei, L.-Y.; Huang, J.; Liu, Y.-Q. Individual and simultaneous electrochemical detection toward heavy metal ions based on L-cysteine modified mesoporous MnFe2O4 nanocrystal clusters. J. Alloys Compd. 2017, 721, 492–500. [Google Scholar] [CrossRef]
  53. Zhou, S.-F.; Han, X.-J.; Fan, H.-L.; Huang, J.; Liu, Y.-Q. Enhanced electrochemical performance for sensing Pb(II) based on graphene oxide incorporated mesoporous MnFe2O4 nanocomposites. J. Alloys Compd. 2018, 747, 447–454. [Google Scholar] [CrossRef]
  54. Idrissi, G.E.; Achache, M.; El Haddaoui, H.; El-Haddar, S.; Draoui, K.; Bouchta, D.; Choukairi, M. Carbon paste electrode modified with Mn2O3 nanoparticles for simultaneous detection of heavy metals Cd(II) and Pb(II) in wastewater sample. J. Water Process Eng. 2025, 75, 107928. [Google Scholar] [CrossRef]
  55. Altammar, K.A. A review on nanoparticles: Characteristics, synthesis, applications, and challenges. Front. Microbiol. 2023, 14, 1155622. [Google Scholar] [CrossRef]
  56. Amany, E.S.; Doaa, G.G.; Omar, R.; Eslam, M.E.-N.; Ahmed, S.; Rabeay, Y.A.H. Highly selective detection of heavy metal ions in food and water using a 5-BHAHS@NC/MnO2-based electrochemical sensor. Biochem. Eng. J. 2025, 216, 109660. [Google Scholar]
  57. Wang, J.; Wang, J.; Zhou, P.; Tao, H.; Wang, X.; Wu, Y. Oligonucleotide-induced regulation of the oxidase-mimicking activity of octahedral Mn3O4 nanoparticles for colorimetric detection of heavy metals. Microchim. Acta 2020, 187, 99. [Google Scholar] [CrossRef] [PubMed]
  58. Li, M.; Kang, Y.; Ma, H.; Dong, J.; Wang, Y.; Kuang, S. Efficient removal of heavy metals from aqueous solutions using Mn-doped FeOOH: Performance and mechanisms. Environ. Res. 2023, 231 Pt 1, 116161. [Google Scholar] [CrossRef]
  59. Antunović, V.; Ilić, M.; Baošić, R.; Jelić, D.; Lolić, A. Synthesis of MnCo2O4 nanoparticles as modifiers for simultaneous determination of Pb(II) and Cd(II). PLoS ONE 2019, 14, e0210904. [Google Scholar] [CrossRef]
  60. Nilghaz, A.; Mousavi, S.M.; Tian, J.; Cao, R.; Guijt, R.M.; Wang, X. Noble-Metal Nanoparticle-Based Colorimetric Diagnostic Assays for Point-of-Need Applications. ACS Appl. Nano Mater. 2021, 4, 12808–12824. [Google Scholar] [CrossRef]
  61. Adeniji, T.M.; Stine, K.J. Nanostructure Modified Electrodes for Electrochemical Detection of Contaminants of Emerging Concern. Coatings 2023, 13, 381. [Google Scholar] [CrossRef]
  62. Sawan, S.; Maalouf, R.; Errachid, A.; Jaffrezic-Renault, N. Metal and metal oxide nanoparticles in the voltammetric detection of heavy metals ions: A review. Trends Anal. Chem. 2020, 131, 116014. [Google Scholar] [CrossRef]
  63. Jelić, D.; Zeljković, S.; Škundrić, B.; Mentus, S. Thermogravimetric study of the reduction of CuO–WO3 oxide mixtures in the entire range of molar ratios. J. Therm. Anal. Calorim. 2018, 132, 77–90. [Google Scholar] [CrossRef]
  64. Lu, D.; Sullivan, C.; Brack, E.M.; Drew, C.P.; Kurup, P. Simultaneous voltammetric detection of cadmium(II), arsenic(III), and selenium(IV) using gold nanostar–modified screen-printed carbon electrodes and modified Britton-Robinson buffer. Anal. Bioanal. Chem. 2020, 412, 4113–4125. [Google Scholar] [CrossRef] [PubMed]
  65. Kusuma, S.A.F.; Harmonis, J.A.; Pratiwi, R.; Hasanah, A.N. Gold Nanoparticle-Based Colorimetric Sensors: Properties and Application in Detection of Heavy Metals and Biological Molecules. Sensors 2023, 23, 8172. [Google Scholar] [CrossRef] [PubMed]
  66. Yah, C.S. The toxicity of Gold Nanoparticles in relation to their physiochemical properties. Biomed. Res. 2013, 24, 400–413. [Google Scholar]
  67. Renault, S.; Baudrimont, M.; Mesmer-Dudons, N.; Gonzalez, P.; Mornet, S.; Brisson, A. Impacts of gold nanoparticle exposure on two freshwater species: A phytoplanktonic alga (Scenedesmus subspicatus) and a benthic bivalve (Corbicula fluminea). Gold Bull. 2008, 41, 116–126. [Google Scholar] [CrossRef]
  68. Jennings, T.; Strouse, G. Past, present, and future of gold nanoparticles. Adv. Exp. Med. Biol. 2007, 620, 34. [Google Scholar]
  69. Sani, A.; Cao, C.; Cui, D. Toxicity of gold nanoparticles (AuNPs): A review. Biochem. Biophys. Rep. 2021, 26, 100991. [Google Scholar] [CrossRef]
  70. Van Doren, E.A.F.; Temmerman, P.R.H.D.; Francisco, A.D.; Mast, J. Determination of the volume specific surface area by using transmission electron tomography for characterization and definition of nanomaterials. J. Nanobiotechnol. 2011, 9, 17. [Google Scholar] [CrossRef]
  71. Hassan, K.M.; Elhaddad, G.M.; AbdelAzzem, M. Voltammetric determination of cadmium(II), lead(II) and copper(II) with a glassy carbon electrode modified with silver nanoparticles deposited on poly(1,8-diaminonaphthalene). Microchim. Acta 2019, 186, 440. [Google Scholar] [CrossRef]
  72. Shirsat, M.D.; Hianik, T. Electrochemical Detection of Heavy Metal Ions Based on Nanocomposite Materials. J. Compos. Sci. 2023, 7, 473. [Google Scholar] [CrossRef]
  73. Zhang, K.; Zhang, N.; Zhang, L.; Xu, J.; Wang, H.; Wang, C.; Geng, T. Amperometric sensing of hydrogen peroxide using a glassy carbon electrode modified with silver nanoparticles on poly(alizarin yellow R). Microchim. Acta 2011, 173, 135–141. [Google Scholar] [CrossRef]
  74. Abbas, A.; Amin, H.M.A. Silver nanoparticles modified electrodes for electroanalysis: An updated review and a perspective. Microchem. J. 2022, 175, 107166. [Google Scholar] [CrossRef]
  75. Zhang, T.; Jin, H.; Fang, Y.; Guan, J.; Ma, S.; Pan, Y.; Zhang, M.; Zhu, H.; Liu, X.; Du, M. Detection of trace Cd2+, Pb2+ and Cu2+ ions via porous activated carbon supported palladium nanoparticles modified electrodes using SWASV. Mater. Chem. Phys. 2019, 225, 433–442. [Google Scholar] [CrossRef]
  76. Alaqarbeh, M.; Adil, S.F.; Ghrear, T.; Khan, M.; Bouachrine, M.; Al-Warthan, A. Recent Progress in the Application of Palladium Nanoparticles: A Review. Catalysts 2023, 13, 1343. [Google Scholar] [CrossRef]
  77. Cadevall, M.; Ros, J.; Merkoc, A. Bismuth nanoparticles integration into heavy metal electrochemical stripping sensor. Electrophoresis 2015, 36, 1872–1879. [Google Scholar] [CrossRef] [PubMed]
  78. 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. Environ. Sci. Pollut. Res. 2021, 28, 58994–59002. [Google Scholar] [CrossRef]
  79. Lee, G.J.; Kim, C.K.; Lee, M.K.; Rhee, C.K. Simultaneous voltammetric determination of Zn, Cd and Pb at bismuth nanopowder electrodes with various particle size distributions. Electroanalysis 2010, 22, 530–535. [Google Scholar] [CrossRef]
  80. He, Y.; Ma, L.; Zhou, L.; 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]
  81. Kulpa-Koterwa, A.; Ossowski, T.; Niedziałkowski, P. Functionalized Fe3O4 Nanoparticles as Glassy Carbon Electrode Modifiers for Heavy Metal Ions Detection—A Mini Review. Materials 2021, 14, 7725. [Google Scholar] [CrossRef]
  82. Daniel, S.; Malathi, S.; Balasubramanian, S.; Sivakumar, M.; Sironmani, T.A. Multifunctional silver, copper and zero valent iron metallic nanoparticles for wastewater treatment. In Application of Nanotechnology in Water; John Wiley & Sons: Hoboken, NJ, USA, 2014; pp. 435–457. [Google Scholar]
  83. Jjagwea, J.; Olupota, P.Q.; Kulabako, R.; Carrara, S. Electrochemical sensors modified with iron oxide nanoparticles/nanocomposites for voltammetric detection of Pb (II) in water: A review. Heliyon 2024, 10, e29743. [Google Scholar] [CrossRef]
  84. Cheng, Z.; Tan, A.L.K.; Tao, Y.; Shan, D.; Ting, K.E.; Yin, X.J. Synthesis and characterization of iron oxide nanoparticles and applications in the removal of heavy metals from industrial wastewater. Int. J. Photoenergy 2012, 2012, 608298. [Google Scholar] [CrossRef]
  85. Yadav, V.K.; Ali, D.; Khan, S.H.; Gnanamoorthy, G.; Choudhary, N.; Yadav, K.K.; Thai, V.N.; Hussain, S.A.; Manhrdas, S. Synthesis and characterization of amorphous iron oxide nanoparticles by the sonochemical method and their application for the remediation of heavy metals from wastewater. Nanomaterials 2020, 10, 1551. [Google Scholar] [CrossRef]
  86. Jasaswini, T.; Akanshya, M.; Mayank, P.; Rakesh, R.T.; Sasmita, C.; Prangya, R.R.; Muhammad, K.S. Advances in Nanoparticles and Nanocomposites for Water and Wastewater Treatment: A Review. Water 2024, 16, 1481. [Google Scholar] [CrossRef]
  87. Xiong, S.; Yang, B.; Cai, D.; Qiu, G.; Wu, Z. Individual and simultaneous stripping voltammetric and mutual interference analysis of Cd2+, Pb2+ and Hg2+ with reduced graphene oxide-Fe3O4 nanocomposites. Electrochim. Acta 2015, 185, 52–61. [Google Scholar] [CrossRef]
  88. Sun, Y.-F.; Chen, W.-K.; Li, W.-J.; Jiang, T.-J.; Liu, J.-H.; Liu, Z.-G. Selective detection toward Cd2+ using Fe3O4/rGO nanoparticle modified glassy carbon electrode. J. Electroanal. Chem. 2014, 714, 97–102. [Google Scholar] [CrossRef]
  89. Ghasemi, E.; Heydari, A.; Sillanpää, M. Superparamagnetic Fe3O4@EDTA nanoparticles as an efficient adsorbent for simultaneous removal of Ag(I), Hg(II), Mn(II), Zn(II), Pb(II) and Cd(II) from water and soil environmental samples. Microchem. J. 2017, 131, 51–56. [Google Scholar] [CrossRef]
  90. Deshmukh, S.; Kandasamy, G.; Upadhyay, R.K.; Bhattacharya, G.; Banerjee, D.; Maity, D.; Deshusses, M.A.; Roy, S.S. Terephthalic acid capped iron oxide nanoparticles for sensitive electrochemical detection of heavy metal ions in water. J. Electroanal. Chem. 2017, 788, 91–98. [Google Scholar] [CrossRef]
  91. Dai, P.; Yang, Z. Sensor for lead (II) ion based on a glassy carbon electrode modified with double-stranded DNA and ferric oxide nanoparticles. Microchim. Acta 2012, 176, 109–115. [Google Scholar] [CrossRef]
  92. Mustafa, G.; Aslı, E.A.; Samet, Y.; Abdullah, A.; Kariper, İ.A.; İlknur, Ü. Evaluation of nanomanganese decorated typha tassel carbonaceous electrode: Preparation, characterization, and simultaneous determination of Cd2+ and Pb2+. Chem. Pap. 2019, 73, 2869–2878. [Google Scholar]
Table 1. Comparative summary of nanoparticles for heavy metal detection: advantages and limitations.
Table 1. Comparative summary of nanoparticles for heavy metal detection: advantages and limitations.
Nanoparticle TypeSurface AreaDetection LimitAdvantagesDisadvantagesReferences
MnO2@RGO-NPs148.7 m2/gCd2+: 15.1 ng/L
Cu2+: 9.3 ng/L
Zn2+: 13.7 ng/L
Multiple oxidation states (−3 to +7)
Magnetic separation
Catalytic activity
Low cost
Low electrical conductivity
Requires doping for performance
[14,24,29,31,33,36,92]
MnO2-CT-NPs-Cd2+: 50 ng/L
Pb2+: 10 ng/L
Mn3O4-NPs30–70 m2/g-
MnO2-NPs20–200 m2/g-
Fe3O4/MnO2-NPs118 m2/g-
AuNPs-Cd2+: 99.95 ng/L
Pb2+: 49.73 ng/L
Cu2+: 0.0064 ng/L
Stability, biocompatibility
High optical/electronic response
High surface-to-volume ratio
Toxicity concerns
High cost
[61,63,65]
AgNPs@p-1,8-DAN/GC-Cd2+: 19 ng/L
Pb2+: 30 ng/L
Cu2+: 6 ng/L
High conductivity
Simple synthesis
Aggregation issues
Time-intensive processing
[38,72,74]
Pd@PAC-NPs881.79 m2/gCd2+: 56.205 ng/L
Pb2+: 103.60 ng/L
Cu2+: 31.775 ng/L
Effective for multi-ion detection (Cd2+, Pb2+, Cu2+)
Compatible with SWASV
Reduced sensitivity in complex matrices[61,76]
BiNPs13.47 m2/gCd2+: 90 ng/L
Pb2+: 170 ng/L
Strong metal affinity
Compatible with SWASV
Cu2+ interference
Size distribution issues
[65,66,77]
BiNPs@Ti3C2Tx-Cd2+: 1.40 ng/L
Pb2+: 2.24 ng/L
Fe3O4@MPC-NPs75.7–238.4 m2/gPb2+: 2507.12–7086.24 ng/LStrong heavy metal affinity
Chemical stability
Biocompatible, superparamagnetic
Low-cost, simple synthesis
Multi-metal detection capability
Aggregation tendency
Reduced electrochemical performance
Limited applications when unmodified
[67,81,87]
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

Keramari, V.; Girousi, S. Manganese Nanoparticles for Heavy Metal Detection vs. Noble and Base Metal Nanoparticles; Prospects, Limitations, and Applications in Electroanalysis. Chemosensors 2025, 13, 313. https://doi.org/10.3390/chemosensors13080313

AMA Style

Keramari V, Girousi S. Manganese Nanoparticles for Heavy Metal Detection vs. Noble and Base Metal Nanoparticles; Prospects, Limitations, and Applications in Electroanalysis. Chemosensors. 2025; 13(8):313. https://doi.org/10.3390/chemosensors13080313

Chicago/Turabian Style

Keramari, Vasiliki, and Stella Girousi. 2025. "Manganese Nanoparticles for Heavy Metal Detection vs. Noble and Base Metal Nanoparticles; Prospects, Limitations, and Applications in Electroanalysis" Chemosensors 13, no. 8: 313. https://doi.org/10.3390/chemosensors13080313

APA Style

Keramari, V., & Girousi, S. (2025). Manganese Nanoparticles for Heavy Metal Detection vs. Noble and Base Metal Nanoparticles; Prospects, Limitations, and Applications in Electroanalysis. Chemosensors, 13(8), 313. https://doi.org/10.3390/chemosensors13080313

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

Article Metrics

Back to TopTop