Next Article in Journal
A Dynamic Range Preservation Readout Integrated Circuit for Multi-Gas Sensor Array Applications
Previous Article in Journal
Electrochemical Sensors, Biosensors, and Optical Sensors for the Detection of Opioids and Their Analogs: Pharmaceutical, Clinical, and Forensic Applications
Previous Article in Special Issue
Potentiometric Phosphate Ion Sensor Based on Electrochemically Modified All-Solid-State Copper Electrode for Phosphate Ions’ Detection in Real Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Aptasensors for the Detection of Environmental Contaminants of High Concern in Water Bodies: A Systematic Review

by
Eduardo Canek Reynoso
1,
Patrick Severin Sfragano
2,
Mario González-Perea
3,
Ilaria Palchetti
2 and
Eduardo Torres
4,*
1
Centro de Investigación en Salud Poblacional, Instituto Nacional de Salud Pública, Cuernavaca 62100, Mexico
2
Department of Chemistry “Ugo Schiff”, University of Florence, 50019 Sesto Fiorentino, Italy
3
Facultad de Ciencias Químicas, Benemérita Universidad Autónoma de Puebla, Puebla 72570, Mexico
4
Centro de Química, Instituto de Ciencias, Benemérita Universidad Autónoma de Puebla, Puebla 72570, Mexico
*
Author to whom correspondence should be addressed.
Chemosensors 2024, 12(4), 59; https://doi.org/10.3390/chemosensors12040059
Submission received: 8 February 2024 / Revised: 29 March 2024 / Accepted: 6 April 2024 / Published: 9 April 2024
(This article belongs to the Special Issue Chemical Sensors and Analytical Methods for Environmental Monitoring)

Abstract

:
With the advancement of technology and increasing industrial activity, anthropogenic contaminants are currently detected where there is no record of their presence or insufficient information about their toxicological impact. Consequently, there are not sufficiently robust local or global regulations, the ecotoxicological and human health risks are critical, and they may not be routinely monitored despite being ubiquitous. The interest in studying environmental contaminants, including micropollutants and emerging contaminants, in complex environmental water samples has grown in the last decade. Due to the concentrations in which they are typically found in the environment and the rapid global dispersion, the detection procedures for these substances must be capable of measuring very low concentrations. Many efforts have been made to improve remediation procedures or develop novel analytical methods for their determination. Although there are several robust and reliable standard analytical techniques for their monitoring, pollutant contamination requires simple and inexpensive methods for massive, in situ monitoring campaigns. In this regard, biosensors have emerged as devices with high selectivity, sensitivity, easy operation, and short analysis times. Aptasensors are biosensors based on a nucleic acid recognition element (aptamer). Due to their synthetic nature, stability, and easy production, aptamers are frequently employed to develop bioassays. This work presents a systematic review of the trends in using aptasensors for detecting environmental contaminants present in environmental water samples, as well as the estimation of the potential technological contribution these devices might give to environmental monitoring.

1. Global Problem of Contaminants in Water

The growing requirements of consumers for new products for everyday use have generated the appearance of many contaminants in the environment, mainly in aquifer systems. These substances, known as emerging contaminants (ECs), are found in trace concentrations in water bodies. ECs are currently of high environmental interest because they are recalcitrant compounds, ubiquitous throughout the ecosystems and, in many cases, have unknown toxicological effects [1,2].
The sources of environmental contaminants are mainly due to industrial practices or anthropogenic activities [2,3]. As shown in Figure 1, several contaminants are transferred by different routes to water systems. Wastewater treatment plants (WWTPs) constitute one of the main factors for the spread of contaminants. When contaminant degradation or removal processes are not appropriate, these substances are released into water bodies in high amounts due to an enrichment and accumulation effect [3,4,5,6]. After release from WWTPs to water systems, contaminants can bioaccumulate in aquatic species [7,8,9] and be transported to different water bodies (groundwater, river water, lakes) where risk exists to reach water for human consumption (tap water, bottled water) [2,10,11,12,13].
Regarding the risks for the ecosystems and human health due to the presence of environmental contaminants in the hydric system, the World Health Organization (WHO) and the governments of different regions have established minimum concentration levels of some of these substances in water [14,15,16,17,18]. The concentrations allowed in water bodies, depending on the type of compound, are, generally, in the order of pico- to micro-molar in heavy metals [18,19,20], pesticides (insecticides, fungicides, herbicides) [21,22,23], pharmaceutical substances (anti-inflammatories, analgesics, antibiotics, hormones, diuretics, anxiolytics) [24,25,26], plasticizers and micro- and nano-plastics [27,28].
Traditional methods for their monitoring are based on chromatography coupled to mass spectrometry (GC-MS and LC-MS) and direct detection methods based on luminescence (absorption and fluorescence). GC-MS and LC-MS are frequently based on complex analytical procedures, requiring well-trained personnel and laborious sample treatments with associated high costs and low throughput. These techniques are undoubtedly suitable as a confirmatory method. By contrast, the development of screening analytical methods to be performed directly in the field is mandatory for monitoring programs. Screening methods allow the use of confirmatory methods only on positive findings, thus decreasing the number of samples to be analyzed and reducing the cost of the whole analysis. In this context, methods based on biosensors [29] are of particular interest. Indeed, any screening method must meet the Affordable, Sensitive, Specific, User-friendly, Rapid and Robust, Equipment-free, and Deliverable to End-users (ASSURED) criteria established by the WHO [18,30]. Recently, aptasensors have emerged as interesting tools in the detection of environmental contaminants due to their high selectivity in binding to the target and their synthetic nature, compared to other biorecognition elements, such as antibodies or enzymes [31,32,33].
Aptamers are single-stranded oligonucleotides with high-affinity binding to a given target analyte, produced by in vitro selection. They fold into complex three-dimensional structures upon association with their target, providing multiple molecular interactions of different types (electrostatic, hydrogen bonding, etc.) that govern selective recognition. Aptamers are produced by chemical synthesis, avoiding biological raw materials, or living animals. Aptamers are selected by a universal approach that does not depend on a particular analyte with the possibility to use toxins as well as molecules that do not elicit a good immune response [34,35,36].
Moreover, aptamers have other interesting features such as high thermal stability and the opportunities for further chemical modifications that provide the immobilization of aptamers onto solid support with mild alteration of selectivity. Due to these features, aptamers are frequently used in biosensor assembly and are frequently combined with a wide variety of nanomaterials, improving the sensing performance [35,37,38,39,40].
Considering the above-mentioned advantages, this work aims to systematically review the use of aptasensors for detecting environmental contaminants in water bodies to evaluate the current trend, sensitivity projections, assembly materials, and the ability to reuse the devices. Although there has been a noticeable increase in the number of review papers related to aptasensors for detecting contaminants in environmental water samples in the last five years (see Figure 2), the vast majority are focused on describing only the original research on the development and functionality of aptamers, in a specific type of transducer (optical or electrochemical) [38,41,42,43,44,45,46,47,48], the effectiveness of certain materials (metallic nanoparticles, metal–organic framework, quantum dots, single- and multi-walled carbon nanotubes) [49,50,51,52], or type of contaminants (heavy metals, pharmaceutical compounds, pesticides, endocrine disruptors) [53,54,55,56,57,58,59,60,61,62]. Nevertheless, a systematic analysis that includes these aspects, emphasizing the analysis of water samples of environmental origin or that had undergone purification treatment, has not yet been published.

2. Materials and Methods

This review followed the preferred reporting items for the systematic review and meta-analysis statement (PRISMA). The literature search focused on reports published in peer-reviewed journals indexed in Web of Science and Scopus databases and available in English.
The search for studies concerning aptasensors was performed by evaluating keywords, titles, and abstracts. Studies that did not assess the applicability of aptasensors in detecting contaminants in environmental water (river, lake, groundwater, wastewater) were excluded. Only articles published from the past five years were considered (2019–2023).
Focus questions were prepared based on the problem, intervention, comparison, and outcome (PICO) method: (P) what water contaminants are usually detected using aptasensors? (I) What configurations are frequently applied for aptasensors elaboration for contaminant detection? (C) Are aptasensors sensitive enough for environmental contaminant detection at ultra-low concentrations in complex environmental water samples? (O) What aptasensor exhibits the higher performance stability for analyzing real water samples?
The search was performed using the following components for every database:
Search component 1 (SC1), including the key terms: Aptasensor OR Aptamer AND Water; Search component 2 (SC2), including the key terms: Aptasensor OR Aptamer AND water OR wastewater OR groundwater OR river OR Sewage OR Environment.
The following data were extracted and captured in an Excel spreadsheet featuring the following information: article title, year, analyte, analyte classification, transducer type, detection principle, sensitivity, water sample type, test on real samples, aptasensor selectivity/specificity, reproducibility/repeatability, stability, and reusability. Table 1 shows the concept and description of the data searched.
This review considered only original research articles in English directly related to detecting or quantifying water environmental contaminants using aptamers as recognizing elements. The exclusion criteria for full-text articles were articles written in a language different from English, aptasensors applied for the detection of other chemicals not considered environmental contaminants, abstract-only papers as proceeding papers, and conference, editorial, and author response theses and books. Finally, aptasensor applications in model water samples (distilled, buffer, or synthetic environmental water samples) or uncomplex water samples, such as those from drinking or tap water, were excluded as they were not useful to answer the research questions.

3. Results and Discussion

A total of 250 studies were found in the Scopus database and 286 in the Web of Science database, totaling 536 from 2019 until 2023. After duplicate exclusions, 151 studies that met the inclusion criteria remained. In addition, 42 reviews on aptasensors for contaminants application were identified in the same search period; of these, 13 reviews related to the detection of water contaminants were selected to identify some other original publications not collected in the initial search. The rest of the reviews were not selected because they focused on the development of aptasensors for a particular contaminant or a family of them (arsenic, bisphenol, pesticides, mycotoxins, metals, viruses, microorganisms, antibiotics) in another type of compartment (food, biological fluids, or plants); some other reviews focused on the detection of biomolecules in biological fluids, for the diagnosis of diseases (Alzheimer, cancer), molecular monitoring of metabolites or drugs in the body; and other reviews focused on the transduction system (electrochemical, optical). None of the 42 reviews were of the systematic type but rather were of the narrative type. In the eligibility step, the rest of the 13 reviews were excluded and no additional papers evaluated were chosen. The consultation of the full texts of the articles to determine the inclusion and exclusion criteria led to the exclusion of 36 articles, mainly because the articles excluded in this step applied the aptasensor in simple water samples such as buffer or distilled water and those applied in food samples, although the abstract mentioned that it could be applied in environmental samples or environmental monitoring. In total, 73 studies that fully met the inclusion criteria were subsequently analyzed to answer the research questions. The PRISMA flow diagram template used in this systematic review is shown in Figure 3. Table S1 of the Supporting Information details the 73 studies with the concepts developed in Table 1.

3.1. Production of Aptamers as Recognizing Elements

The aptamer selection process, called Systematic Ligand Evolution by Exponential Enrichment (SELEX), is a technique developed almost simultaneously by Tuerk and Gold [63], and Ellington and Szostak [64] in the 1990s. The SELEX technique consists of the following steps: selection, partitioning, and amplification (Figure 4). To carry out the selection, it is necessary to synthesize a library of approximately 1013–1018 random oligonucleotide sequences. Each oligonucleotide contains random bases (20–50 NTs) flanked by two conserved primer binding sites, which are used for PCR amplification. In the selection step, the oligonucleotide library is incubated with target molecules, which are immobilized on solid phase supports; after incubation, the unbound sequences are separated using different methods. The target-bound sequences are amplified by PCR (DNA SELEX) or reverse transcription PCR (RNA SELEX); the products are used for the next selection round, performing the same sequence and target molecule interaction process. After several rounds of selection, the enriched oligonucleotides are sequenced and evaluated for their binding capabilities [63,64,65,66]. However, there are some deficiencies to be overcome; one of them is stringency, as SELEX assumes that the most enriched sequences are the most specific binders, which is not always the case, and high-affinity binders can sometimes be overlooked due to insufficient stringency in selection conditions. The process can enrich for non-specific binders that bind to the matrix used rather than the target molecule, leading to false positive results; in addition, minimal mistakes in the initial library result in a biased library; another restriction is the limited scale. Moreover, SELEX can be limited in its ability to identify specific binders for targets with low binding specificity or for targets with highly structured regions [37,67,68,69,70,71]. Over time, different modified SELEX procedures have been developed and continue to be developed to improve the efficiency, specificity, or speed of the selection process, including variations in the separation stage. Some examples are shown in Table S2.

3.2. Overview of Data Collection

As can be seen in Figure 5 and Table S1, the 73 papers reported the detection of 30 aqueous contaminants: 5 metals, 6 pesticides, 4 industrial chemicals, 2 toxins, and 13 pharmaceutical compounds. In particular, the study of antibiotics stands out due to the promotion of bacterial resistance, an issue of global concern. Different environmental water matrices were analyzed: municipal and industrial wastewaters, lakes, rivers, ponds, and canal water. However, most of the assays were performed in spiked water samples, allowing for the study of the effects of natural interferents, but the concentrations used in these assays are usually much higher than the environmental ones. Most of the papers studied the selectivity of aptasensors using chemically similar analytes, added to the assay individually or in a mixture of interferents, in significantly higher concentrations. However, very few addressed the study of the reusability of the aptasensor, a parameter of major importance in the environmental area, where a massive number of readouts are performed, which implies a high cost for an adequate determination of contamination in time and space.

3.3. Analysis of Categorizations

3.3.1. Environmental Contaminants

The systematic analysis of the literature on applying aptasensors in environmental water over the past five years shows that pharmaceuticals (PhCs) are the most studied, with 26 papers and 13 compounds reported (Table S1). In recent years, there has been a boom in the literature on monitoring PhCs as emerging contaminants [72,73]. Antibiotics are undoubtedly the most studied PhCs, and there are several reasons for this. First, they are the most widely used drugs worldwide for human, animal, and plant health; second, because of the above, they are expected to be discharged into the environmental compartments by various household, hospital, and industrial discharges. Thirdly, they are the cause of bacterial resistance, an issue of great concern worldwide, where the global action plans against antimicrobial resistance promote the monitoring of both resistant microorganisms and the antimicrobials commonly used as a reference [74]. In a recent work [2], it was found that, among the 53 compounds reported, those with the highest calculated relevance were PhCs, with antibiotics having the highest proportion.
In second place was the detection of metals, with 21 works for detecting copper, mercury, lead, cadmium, and arsenic. Metals are pollutants with a broader concentration range, from mg/L to ng/L, because they are produced by intensive industrial activity from several sectors; in addition, they can be released into the environment from natural sources. They have been studied for many years, and their toxic effects and environmental impact are known. Hence, the search for accurate, sensitive, and stable methods or devices will promote the development of research and innovation in the field. Indeed, aptasensors have been widely applied to detect metals, as reported in recent literature reviews [32,53].
The other family of contaminants analyzed using aptasensors is pesticides. These compounds are widely discharged into aqueous compartments upon spraying over large areas, so they are prone to being partially deposited outside the target sites and carried into streams or reservoirs by runoff or filtration. Their presence as contaminants is documented, as is their effects on human and environmental health [75,76,77]. Although applied in large quantities, some undergo chemical and biological transformation reactions, leading to their detection at µg/L or ng/L concentrations. The most studied pesticides are diazinon, chlorpyrifos, atrazine, acetamiprid, quinclorac, and malathion, as reported in 13 papers.
Fewer papers have been devoted to the determination of industrial chemicals (8 papers, 4 chemicals) and toxins (5 papers, 2 toxins), as reported in Table S1. Studies performed within complex samples and in the presence of many interferents indicate that aptamers are robust molecules capable of maintaining their affinity towards the analyte of interest in environmental conditions.

3.3.2. Sensor Design and Sensitivity

Regarding the most used sensor configurations (Figure 6), 77% of the cases (56 articles) used the aptamer without further modification or a simple modification, such as functionalization with an amino group or thiol group to anchor at the surface of the transducer. On the transducer side, three types were documented. Electrochemical, optical, and opto-electrochemical techniques were used within these transducers. With the term opto-electrochemical, we include both electrochemiluminescent and photoelectrochemical systems.
To enhance aptasensor performance, nanomaterials have become a prevalent strategy. Their well-documented properties enable them to exert significant influence on several critical aspects, including controlling assembly density, regulating the accumulation of the aptamers, optimizing the orientation of these elements for target interaction, and facilitating the rate of electron transfer at the sensor interface [78]. In this review, 60% of the papers used composites of two or more materials of metallic or carbonaceous type, or their combination. Among those, noble metal nanoparticles, graphene and graphene oxide, carbon nanotubes, quantum dots, and metal–organic frameworks are the most widely employed. Particularly, 18 papers used electrochemical transducers modified with nanocomposites, while 18 papers used opto-electrochemical transducers with a surface composed of two or more materials, and only 9 were optical with nanocomposites. A total of 28 papers used a “simple” interface with only one or no nanomaterial, most using the optical transducer (colorimetric analysis). Figure 6a shows the number of papers classified based on the transducing principle. Figure 6b–d describe a simplified subdivision of the works in line with the heterostructure used, i.e., (i) simple: bare surface or one nanomaterial; (ii) composite: heterostructures of two nanomaterials; (iii) complex: more complex heterostructures. Moreover, a further classification according to sensitivity is also reported.
Regarding sensitivity, in the selected works, the reported LODs displayed a wide interval of values, from mg/L to ag/L. A total of 82% of LODs (60 papers) fall in the value here defined as high to ultrahigh sensitivity (LOD < 1 µg/L), 14% showed a medium sensitivity (0.1 mg/L > LOD > 1µg/L), and only three papers showed a low LOD (LOD > 0.1 mg/L) (Figure 6). It is well known that aptamers present dissociation constants like those shown by antibodies, with values of approximately 10−9 M, which helps to achieve high sensitivity. Regarding the type of transducer on the sensitivity parameter, 100% of the papers using opto-electrochemical transducers showed high and ultra-high sensitivity, followed by 90% of the electrochemical and 50% of the optical transducers (Figure 6).
The data presented in Table 2 and Table S1 show that the high sensitivity of the aptasensors can be attributed to the coupling of specific aptamers with nanomaterials.
As an example of an electrochemical transducer, Zhao et al. [98] recently reported an aptasensor for bisphenol A (BPA) detection (Figure 7). Their design employed a glassy carbon electrode (GCE) modified with a composite material consisting of multi-walled carbon nanotubes (MWCNTs), amino-functionalized magnetite (NH2-Fe3O4), and gold nanoparticles (Au NPs). The combination of these elements resulted in a limit of detection (LOD) of 0.08 aM (18 ag/L) for BPA, along with a linear detection range spanning from 10−19 M to 10−14 M. The authors reported an advantage in using a dual-signal amplification effect based on (i) the synergic properties of the composite material and (ii) the conductivity of the SWCNTs. The specific aptamer also facilitated a “signal-on” sensing scheme, contributing to a simple and efficient protocol.
Besides nanomaterials, an interesting signal enhancement mechanism that can be introduced when dealing with aptamers relies on nucleic acid amplification techniques. In an optical aptasensor developed by Tian et al. [88], the detection of Hg2+ was accomplished through a dual recycling amplification strategy (Figure 8). The method utilizes a functional aptamer (DNA 1) harboring Hg2+ recognition sites and amplification regions. Additionally, hairpin DNA conjugated to magnetic nanoparticles and a SERS probe comprised of capture DNA and labeling DNA (Rox-DNA) immobilized on gold nanoparticles (AuNPs) were employed. Upon Hg2+ addition, the first amplification cycle generates numerous trigger DNA strands. These unfold the hairpin DNA so that a second amplification cycle can start through the capture of SERS probes by the unfolded hairpin DNA. This complex formation triggers the release of more trigger DNA strands through the sequential action of polymerase and endonuclease enzymes. These released strands initiate the formation of additional complexes with hairpin DNA and SERS probes. Magnetic separation of these complexes effectively eliminates background noise caused by excess SERS probes. In this way, a limit of detection (LOD) of 0.11 fM (22 fg/L) was obtained, with a linear detection range of 0.2–125 fM.
As seen in the first quartiles of Figure 9, the aptasensor configurations can reach levels lower than 1 ng/L, which is sufficiently sensitive for detecting contaminants in real environmental samples. Nine works reported LODs below the ag/L level (Table S1). These ultralow values could be found in environmental water a certain time after the discharge event when the original concentrations had decreased significantly due to spontaneous transformations (biodegradation, adsorption, chemical oxidation, etc.). The highest LOD values reported were in the order of mg/L for metals (2–10 mg/L) (Table S1). These values seem unrealistic in environmental samples unless they can be found at sites of high industrial activity. At this point, it can be concluded that aptasensors could have enough sensitivity to detect contaminants in real environmental water samples. However, further testing on real, unfortified, contaminated samples is needed to determine their feasibility more accurately.
The limit of detection (LOD) analysis demonstrates the exceptional sensitivity of aptasensors for contaminant detection in real environmental water samples, reaching concentrations quite below 1 µg/L (Figure 9). Notably, the median LOD achieved with optical transducer systems was 0.045 μg/L. This sensitivity significantly improved for opto-electrochemical transducers, reaching a median LOD of 0.00056 μg/L. The most impressive performance was observed with electrochemical detection, achieving an ultra-low median LOD of 0.000064 μg/L. These remarkable sensitivities suggest the aptasensors’ suitability for real-world environmental monitoring applications where contaminants are often present at trace levels.

3.3.3. Accuracy and Precision

In terms of accuracy and precision, aptasensors showed real application potential when tested in different environmental water matrices, from surface water (river or lake) to effluents from municipal, hospital, and industrial treatment plants. Although in 92% of the cases, the tests were performed with spiked samples, in the remaining 8%, the aptasensors were applied in a real analyte concentration (Table S1). For the spiked samples, the accuracy was measured as recovery (comparison of the added to the measured concentration). In the cases where the natural concentration was measured, the accuracy was obtained by comparing the measured amount with the concentration measured with a reference standard technique, such as high-performance liquid chromatography coupled to mass spectrometry (HPLC-MS) [102,103], inductively coupled plasma (ICP) [104,105], atomic absorption spectroscopy [106] or enzyme-linked immunosorbent assay (ELISA) [107]). Both accuracy and precision were interesting, with recovery values in the 90–110% range, and precision, as measured by the coefficient of variation of the replicates made, was less than 10%.
As an example, Wang et al. [108] reported a detection method for microcystin-LR in spiked water samples from Jinshan Lake using a laser-induced graphene-based electrochemical aptasensor. The accuracy and precision were then compared with the HPLC-MS/MS technique, observing a correlation between the two methodologies. A good agreement between the results achieved with a photoelectrochemical aptasensor and HPLC were also found by Zhang et al. [109] for atrazine detection in real environmental water samples.

3.3.4. Selectivity

In 71 of the 73 papers analyzed, the selectivity of the method or device was investigated. Selectivity was measured in different ways, for example, by quantifying the response in the presence of a similar chemical compound but in the absence of the analyte of interest; it can also be performed in the presence of the analyte and adding one interferent at the same time, or it can be performed in the presence of the analyte and a mixture of interferents. Many papers (50 out of 71) reported the addition of an interferent with a similar chemical structure to the pollutant tested; the remaining 21 works did the same but also studied the mixture of compounds. Usually, the interferents were added in concentrations up to 100 times higher than the analyte under study. Table 3 shows some examples of studies that examined the selectivity using structurally similar interferents, both in individual and mixed forms.
As reported in Table 3, Salandari-Jolge et al. [87] observed no effects (<5%) on the current response of the aptasensor for the detection of Hg2+ even in presence of a 100-fold higher concentration of other metal ions. Working with another aptasensor configuration, Xu et al. [100] developed a fluorescent aptasensor based on graphene oxide for tetracycline detection. The selectivity towards tetracycline was tested in the presence of other chemically related compounds, such as doxycycline, chlortetracycline, minocycline, demeclocycline, lymecycline, methacycline, sarecycline, and their mixture. Significant signals were only obtained when in the presence of tetracycline, except for a low response with oxytetracycline due to the dramatic homologue structures between the two molecules.
Lin et al. (2023) [89] developed a rapid lateral flow assay based on aptamers for the simultaneous detection of ampicillin (AMP) and kanamycin (KAN) utilizing G-quadruplex fragments as an internal standard to achieve high assay selectivity (Figure 10). The latter was demonstrated by testing it with various non-target antibiotics (streptomycin sulfate, oxytetracycline hydrochloride, etc.). In the presence of these antibiotics, there was no significant signal change on the AMP and KAM lines, confirming that the aptamers specifically bind only to the target antibiotics. The method proposed successfully detected antibiotics in water samples from diverse sources, including hospital wastewater, chicken farm wastewater, tap water, and aquaculture water.
The reported studies exhibited excellent selectivity in most cases, with interference values below 10%. Notably, the use of environmental water samples did not significantly impact analyte quantification, suggesting a tolerance for potential natural interferents. However, it is crucial to acknowledge that these findings may not translate directly to real-world scenarios. Environmental water matrices are inherently complex, often harboring mixtures of pollutants. Even minimal cross-reactivity with common ions, organic compounds, or a confluence of changing physicochemical and environmental conditions can lead to inaccurate readings or false positives. Therefore, while the current research suggests adequate selectivity, further optimization in this area remains paramount for reliable on-site water quality monitoring applications. Developing aptasensors with high and unequivocal target selectivity is essential to ensure the accuracy and robustness of these analytical tools in complex environmental matrices.

3.3.5. Stability

An important performance criterion in environmental monitoring is the stability of the method or device. In 43 papers, stability studies were reported (Table S1). The stability reported was mainly storage stability, which is the ability of the device or method to give the same response after being stored in suitable laboratory conditions, such as 4 °C, neutral pH, etc. The results varied in testing time; the most prolonged time reported was 30 days, where the same response was maintained as on day 1 [84,104,111,112]. Some papers discussed other types of stability such as signal stability; in the case of opto-electrochemical aptasensors, the device’s response to different cycles of light stimuli in the same test was registered. Reports showed that the photocurrent response almost remained stable after 20 cycles [79,94], 7 cycles [113], 12 cycles [83,85] and 50 scanning cycles [84] under light off-on irradiation. An additional type of stability is given by considering the reusability or regeneration of the device. Liu et al. [84] reported an electrochemical aptasensor for Hg2+ detection consisting of a gold electrode modified with Cu@carbon nanoneedles, in situ constructed through a controllable pyrolysis process of melamine and CuCl2. The principle of detection involved an exonuclease-III-assisted cycling amplification strategy. To regenerate the aptasensor, the modified electrode was immersed in distilled water for 10 min at 80 °C to dissociate the DNA-based signal-reporting structure, rinsed with buffer and subjected to another reading cycle. The stability was measured through 30 cycles and the signal response remained almost unchanged.
In the work of Yildirim-Tirgil et al. [112], a biosensor-containing probe-DNA immobilized on functionalized SWCNTs for oxytetracycline (OTC) detection was developed. The protocol involved an initial incubation step where OTC was exposed to its specific aptamer for 3 min. Following the completion of this binding phase, the incubated mixture was injected across the surface of a gold chip. This facilitated the interaction of any remaining free aptamers with their complementary immobilized DNA counterparts on the semiconducting surface. The sensing surface could be regenerated for upwards of 20 cycles while maintaining minimal signal loss (less than 15%). This regeneration process was achieved through a simple washing step with a 0.5% SDS solution for 5 min, followed by a rinsing step with a buffer solution at pH 7.2. The reusability of this sensor was attributed to the precise control of assembly parameters during the aptasensor fabrication process, which ultimately led to the creation of a stable system.
In a separate study, Song et al. [96] developed electrochemical aptasensors using bimetallic AgMo heteronanostructures to detect bisphenol A (BPA). To demonstrate the reusability of the sensor, after each BPA measurement, the aptasensor was rinsed with 1 mM NaOH at room temperature for 5 min. This step disrupted the bond between the aptamer (recognition molecule) and BPA. The sensor was then rinsed with a large volume of phosphate buffer solution, allowing the aptamer to return to its original shape. Finally, the electrode was dipped into a fresh BPA solution for the next detection cycle. Notably, the sensor’s response signal showed minimal variation even after seven cycles.
While aptasensors have shown promising stability in controlled environments, their suitability for commercial applications, particularly in pollution monitoring, requires further investigation. Environmental monitoring demands frequent and geographically dispersed measurements, making reusability crucial for cost-effective implementation. In-depth studies incorporating cost–benefit analyses are necessary to determine the optimal number of reuse cycles that balance sensor performance with economic viability. This will help establish whether aptasensors can offer a cost-competitive and sustainable solution for environmental monitoring.

3.3.6. Scalability

All the inherent advantages of aptasensors position them favorably for the design of user-friendly and portable devices for in-field analyses. Indeed, (micro)fluidic approaches have gained wide interest as a tool in the automation of sample collection and handling, allowing for on-line and continuous measurements [114,115]. New smartphone-assisted platforms or other compact analyzers based on colorimetric and electrochemical readouts have been introduced in recent years [116,117]. C. Xu and co-workers [118], for instance, replaced the common microplate reader used for optical readouts with the camera of a smartphone for the detection of acetamiprid. J. Wei et al. [80] developed a sunlight-driven self-powered portable system based on a digital multimeter and aptamers for the on-site detection of microcystin-arginine-arginine. Unfortunately, only very few works have exploited their advantages and scalability for real-time decentralized monitoring of contaminants; thus, the traditional laboratory-based procedure remains the most effective approach.

4. Conclusions

The detection of environmental contaminants in water bodies is becoming increasingly imperative. Thus, detection methodologies capable of rapidly detecting such contaminants for screening purposes have been studied and developed in the last decade. Among those, aptasensors, namely biosensors that rely on aptamers as biorecognition elements, have received great attention. This systematic review provides an in-depth analysis of the trends in the use of aptasensors for the detection of contaminants in environmental water samples. All scientific papers on this topic, published from 2019 to 2023, were identified, screened, and evaluated according to precise inclusion criteria. A total of 73 studies passed the phase of eligibility and were further analyzed and categorized based on the environmental contaminant(s) examined, transduction system employed, sensitivity, accuracy and precision, selectivity, and stability achieved. A total of 29 aqueous contaminants were investigated, including pesticides, metals, industrial chemicals, toxins, and pharmaceutical compounds, with the latter being the most studied.
Most of the works focused on optical transducers; nevertheless, electrochemical transducers showed lower limits of detection overall. Aptasensors have demonstrated recoveries of 90–110% in the detection of environmental contaminants in spiked and real environmental water samples. Stability—considered either as the reusability of the device or the ability to provide comparable outputs upon long storage times—was evaluated by fewer works among those under analysis. Although promising, the results cannot ensure proper stability for commercial applications over long and continuous monitoring periods.
On balance, aptasensors have proven to be a valuable and cost-effective tool in the detection of environmental contaminants of high concern in water bodies. Indeed, thanks to their versatility and ease of synthesis, which overcome ethical problems linked to the more widely used antibodies, aptamers might encounter increasing demand in the future over their protein counterparts. Nonetheless, considering that most of the works herein reported analyzed the analytes in buffered solutions or spiked samples, further investigations focused on real contaminated samples over long-term monitoring are required to better assess their strengths and limits.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemosensors12040059/s1, Table S1: Studies included through systematic review. Table S2: SELEX variants and purposes employing different technologies. References [119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167] are cited in the supplementary materials.

Author Contributions

Conceptualization, E.C.R., I.P. and E.T.; methodology, E.C.R. and P.S.S.; formal analysis, E.T. and I.P.; investigation, E.C.R. and M.G.-P.; writing—original draft preparation, E.C.R. and P.S.S.; writing—review and editing, E.T. and I.P. All authors have read and agreed to the published version of the manuscript.

Funding

Eduardo Canek Reynoso was funded by the Consejo de Ciencia y Tecnología del Estado de Puebla (CONCYTEP, Puebla, Mexico) through the grant Estímulos a la Investigación para Doctoras y Doctores 2023 (SNII-01/2023) and by Consejo Nacional de Humanidades, Ciencia y Tecnología (CONAHCYT) (Grant Number 754592), and received financial support from the Instituto de Ciencias and Posgrado en Ciencias Ambientales of Benemérita Universidad Autónoma de Puebla for an international internship. Patrick Severin Sfragano and Ilaria Palchetti were funded by the European Climate, Infrastructure and Environment Executive Agency, HORIZON-MISS-2022-OCEAN-01, project iMERMAID, grant number 101112824.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shehu, Z.; Nyakairu, G.W.A.; Tebandeke, E.; Odume, O.N. Overview of African Water Resources Contamination by Contaminants of Emerging Concern. Sci. Total Environ. 2022, 852, 158303. [Google Scholar] [CrossRef] [PubMed]
  2. Yang, Y.; Zhang, X.; Jiang, J.; Han, J.; Li, W.; Li, X.; Yee Leung, K.M.; Snyder, S.A.; Alvarez, P.J.J. Which Micropollutants in Water Environments Deserve More Attention Globally? Environ. Sci. Technol. 2022, 56, 13–29. [Google Scholar] [CrossRef]
  3. Warner, W.; Licha, T.; Nödler, K. Qualitative and Quantitative Use of Micropollutants as Source and Process Indicators. A Review. Sci. Total Environ. 2019, 686, 75–89. [Google Scholar] [CrossRef]
  4. Kumar, N.M.; Sudha, M.C.; Damodharam, T.; Varjani, S. Chapter 3—Micro-Pollutants in Surface Water: Impacts on the Aquatic Environment and Treatment Technologies. In Current Developments in Biotechnology and Bioengineering; Varjani, S., Pandey, A., Tyagi, R.D., Ngo, H.H., Larroche, C., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 41–62. ISBN 978-0-12-819594-9. [Google Scholar]
  5. Belete, B.; Desye, B.; Ambelu, A.; Yenew, C. Micropollutant Removal Efficiency of Advanced Wastewater Treatment Plants: A Systematic Review. Environ. Health Insights 2023, 17, 11786302231195158. [Google Scholar] [CrossRef]
  6. Barcellos, D.d.S.; Procopiuck, M.; Bollmann, H.A. Management of Pharmaceutical Micropollutants Discharged in Urban Waters: 30 years of Systematic Review Looking at Opportunities for Developing Countries. Sci. Total Environ. 2022, 809, 151128. [Google Scholar] [CrossRef] [PubMed]
  7. Goutte, A.; Alliot, F.; Budzinski, H.; Simonnet-Laprade, C.; Santos, R.; Lachaux, V.; Maciejewski, K.; Le Menach, K.; Labadie, P. Trophic Transfer of Micropollutants and Their Metabolites in an Urban Riverine Food Web. Environ. Sci. Technol. 2020, 54, 8043–8050. [Google Scholar] [CrossRef]
  8. Desiante, W.L.; Minas, N.S.; Fenner, K. Micropollutant Biotransformation and Bioaccumulation in Natural Stream Biofilms. Water Res. 2021, 193, 116846. [Google Scholar] [CrossRef] [PubMed]
  9. Fonseca, V.F.; Duarte, I.A.; Duarte, B.; Freitas, A.; Pouca, A.S.V.; Barbosa, J.; Gillanders, B.M.; Reis-Santos, P. Environmental Risk Assessment and Bioaccumulation of Pharmaceuticals in a Large Urbanized Estuary. Sci. Total Environ. 2021, 783, 147021. [Google Scholar] [CrossRef]
  10. Sackaria, M.; Elango, L. Organic Micropollutants in Groundwater of India—A Review. Water Environ. Res. 2020, 92, 504–523. [Google Scholar] [CrossRef]
  11. Kandie, F.J.; Krauss, M.; Beckers, L.-M.; Massei, R.; Fillinger, U.; Becker, J.; Liess, M.; Torto, B.; Brack, W. Occurrence and Risk Assessment of Organic Micropollutants in Freshwater Systems within the Lake Victoria South Basin, Kenya. Sci. Total Environ. 2020, 714, 136748. [Google Scholar] [CrossRef]
  12. Borrull, J.; Colom, A.; Fabregas, J.; Borrull, F.; Pocurull, E. Presence, Behaviour and Removal of Selected Organic Micropollutants through Drinking Water Treatment. Chemosphere 2021, 276, 130023. [Google Scholar] [CrossRef]
  13. Cai, Y.; Tian, T.; Huang, Y.; Yao, H.; Qi, X.; Fan, J.; Kuang, Y.; Chen, J.; Li, X.; Kadokami, K. Occurrence and Health Risks of Organic Micropollutants in Tap Water in Dalian. Chem. Res. Toxicol. 2023, 36, 1938–1946. [Google Scholar] [CrossRef]
  14. McGinley, J.; Healy, M.G.; Ryan, P.C.; Harmon O’Driscoll, J.; Mellander, P.-E.; Morrison, L.; Siggins, A. Impact of Historical Legacy Pesticides on Achieving Legislative Goals in Europe. Sci. Total Environ. 2023, 873, 162312. [Google Scholar] [CrossRef]
  15. Morin-Crini, N.; Lichtfouse, E.; Liu, G.; Balaram, V.; Ribeiro, A.R.L.; Lu, Z.; Stock, F.; Carmona, E.; Teixeira, M.R.; Picos-Corrales, L.A.; et al. Worldwide Cases of Water Pollution by Emerging Contaminants: A Review. Environ. Chem. Lett. 2022, 20, 2311–2338. [Google Scholar] [CrossRef]
  16. Mutzner, L.; Furrer, V.; Castebrunet, H.; Dittmer, U.; Fuchs, S.; Gernjak, W.; Gromaire, M.-C.; Matzinger, A.; Mikkelsen, P.S.; Selbig, W.R.; et al. A Decade of Monitoring Micropollutants in Urban Wet-Weather Flows: What Did We Learn? Water Res. 2022, 223, 118968. [Google Scholar] [CrossRef]
  17. Khan, S.; Naushad, M.; Govarthanan, M.; Iqbal, J.; Alfadul, S.M. Emerging Contaminants of High Concern for the Environment: Current Trends and Future Research. Environ. Res. 2022, 207, 112609. [Google Scholar] [CrossRef]
  18. Venkateswara Raju, C.; Hwan Cho, C.; Mohana Rani, G.; Manju, V.; Umapathi, R.; Suk Huh, Y.; Pil Park, J. Emerging Insights into the Use of Carbon-Based Nanomaterials for the Electrochemical Detection of Heavy Metal Ions. Coord. Chem. Rev. 2023, 476, 214920. [Google Scholar] [CrossRef]
  19. Rasheed, T.; Shafi, S.; Sher, F. Smart Nano-Architectures as Potential Sensing Tools for Detecting Heavy Metal Ions in Aqueous Matrices. Trends Environ. Anal. Chem. 2022, 36, e00179. [Google Scholar] [CrossRef]
  20. Escandar, G.M.; Olivieri, A.C. A Critical Review on the Development of Optical Sensors for the Determination of Heavy Metals in Water Samples. The Case of Mercury(II) Ion. ACS Omega 2022, 7, 39574–39585. [Google Scholar] [CrossRef]
  21. Nangare, S.N.; Patil, S.R.; Patil, A.G.; Khan, Z.G.; Deshmukh, P.K.; Tade, R.S.; Mahajan, M.R.; Bari, S.B.; Patil, P.O. Structural Design of Nanosize-Metal–Organic Framework-Based Sensors for Detection of Organophosphorus Pesticides in Food and Water Samples: Current Challenges and Future Prospects. J. Nanostructure Chem. 2022, 12, 729–764. [Google Scholar] [CrossRef]
  22. Mukherjee, S.; Ghosh, K.; Bhattacharyya, S.; Behera, B.K.; Singh, O.K.; Pal, S. A Review on Recent Trends in Advancement of Bio-Sensory Techniques Toward Pesticide Detection. Food Anal. Methods 2022, 15, 3416–3434. [Google Scholar] [CrossRef]
  23. Reynoso, E.; Torres, E.; Bettazzi, F.; Palchetti, I. Trends and Perspectives in Immunosensors for Determination of Currently-Used Pesticides: The Case of Glyphosate, Organophosphates, and Neonicotinoids. Biosensors 2019, 9, 20. [Google Scholar] [CrossRef]
  24. Papagiannaki, D.; Belay, M.H.; Gonçalves, N.P.F.; Robotti, E.; Bianco-Prevot, A.; Binetti, R.; Calza, P. From Monitoring to Treatment, How to Improve Water Quality: The Pharmaceuticals Case. Chem. Eng. J. Adv. 2022, 10, 100245. [Google Scholar] [CrossRef]
  25. Bustos Bustos, E.; Sandoval-González, A.; Martínez-Sánchez, C. Detection and Treatment of Persistent Pollutants in Water: General Review of Pharmaceutical Products. ChemElectroChem 2022, 9, e202200188. [Google Scholar] [CrossRef]
  26. Yu, X.; Yu, F.; Li, Z.; Zhan, J. Occurrence, Distribution, and Ecological Risk Assessment of Pharmaceuticals and Personal Care Products in the Surface Water of the Middle and Lower Reaches of the Yellow River (Henan Section). J. Hazard. Mater. 2023, 443, 130369. [Google Scholar] [CrossRef]
  27. Gugliandolo, E.; Licata, P.; Crupi, R.; Albergamo, A.; Jebara, A.; Lo Turco, V.; Potortì, A.G.; Mansour, H.B.; Cuzzocrea, S.; Di Bella, G. Plasticizers as Microplastics Tracers in Tunisian Marine Environment. Front. Mar. Sci. 2020, 7, 89398. [Google Scholar] [CrossRef]
  28. Koelmans, A.A.; Mohamed Nor, N.H.; Hermsen, E.; Kooi, M.; Mintenig, S.M.; De France, J. Microplastics in Freshwaters and Drinking Water: Critical Review and Assessment of Data Quality. Water Res. 2019, 155, 410–422. [Google Scholar] [CrossRef]
  29. Kanan, S.; Moyet, M.; Obeideen, K.; El-Sayed, Y.; Mohamed, A.A. Occurrence, Analysis and Removal of Pesticides, Hormones, Pharmaceuticals, and Other Contaminants in Soil and Water Streams for the Past Two Decades: A Review. Res. Chem. Intermed. 2022, 48, 3633–3683. [Google Scholar] [CrossRef]
  30. Otoo, J.A.; Schlappi, T.S. REASSURED Multiplex Diagnostics: A Critical Review and Forecast. Biosensors 2022, 12, 124. [Google Scholar] [CrossRef]
  31. Zhao, Y.; Yavari, K.; Liu, J. Critical Evaluation of Aptamer Binding for Biosensor Designs. TrAC Trends Anal. Chem. 2022, 146, 116480. [Google Scholar] [CrossRef]
  32. McConnell, E.M.; Nguyen, J.; Li, Y. Aptamer-Based Biosensors for Environmental Monitoring. Front. Chem. 2020, 8, 434. [Google Scholar] [CrossRef] [PubMed]
  33. Palchetti, I.; Mascini, M. Nucleic Acid Biosensors for Environmental Pollution Monitoring. Analyst 2008, 133, 846–854. [Google Scholar] [CrossRef] [PubMed]
  34. Mascini, M.; Palchetti, I.; Tombelli, S. Nucleic Acid and Peptide Aptamers: Fundamentals and Bioanalytical Aspects. Angew. Chem. Int. Ed. 2012, 51, 1316–1332. [Google Scholar] [CrossRef] [PubMed]
  35. Palchetti, I.; Mascini, M. Electrochemical Nanomaterial-Based Nucleic Acid Aptasensors. Anal. Bioanal. Chem. 2012, 402, 3103–3114. [Google Scholar] [CrossRef] [PubMed]
  36. Labuda, J.; Brett, A.M.O.; Evtugyn, G.; Fojta, M.; Mascini, M.; Ozsoz, M.; Palchetti, I.; Paleček, E.; Wang, J. Electrochemical Nucleic Acid-Based Biosensors: Concepts, Terms, and Methodology (IUPAC Technical Report). Pure Appl. Chem. 2010, 82, 1161–1187. [Google Scholar] [CrossRef]
  37. Kudłak, B.; Wieczerzak, M. Aptamer Based Tools for Environmental and Therapeutic Monitoring: A Review of Developments, Applications, Future Perspectives. Crit. Rev. Environ. Sci. Technol. 2020, 50, 816–867. [Google Scholar] [CrossRef]
  38. Rapini, R.; Marrazza, G. Electrochemical Aptasensors for Contaminants Detection in Food and Environment: Recent Advances. Bioelectrochemistry 2017, 118, 47–61. [Google Scholar] [CrossRef] [PubMed]
  39. Geleta, G.S.; Zhao, Z.; Wang, Z. Electrochemical Biosensors for Detecting Microbial Toxins by Graphene-Based Nanocomposites. J. Anal. Test. 2018, 2, 20–25. [Google Scholar] [CrossRef]
  40. Kaur, H.; Shorie, M. Nanomaterial Based Aptasensors for Clinical and Environmental Diagnostic Applications. Nanoscale Adv. 2019, 1, 2123–2138. [Google Scholar] [CrossRef]
  41. Geleta, G.S. A Colorimetric Aptasensor Based on Two Dimensional (2D) Nanomaterial and Gold Nanoparticles for Detection of Toxic Heavy Metal Ions: A Review. Food Chem. Adv. 2023, 2, 100184. [Google Scholar] [CrossRef]
  42. Ding, R.; Li, Z.; Xiong, Y.; Wu, W.; Yang, Q.; Hou, X. Electrochemical (Bio)Sensors for the Detection of Organophosphorus Pesticides Based on Nanomaterial-Modified Electrodes: A Review. Crit. Rev. Anal. Chem. 2023, 53, 1766–1791. [Google Scholar] [CrossRef] [PubMed]
  43. Hayat, A.; Marty, J.L. Aptamer Based Electrochemical Sensors for Emerging Environmental Pollutants. Front. Chem. 2014, 2, 41. [Google Scholar] [CrossRef] [PubMed]
  44. Sun, Y.; Lu, J. Chemiluminescence-Based Aptasensors for Various Target Analytes. Luminescence 2018, 33, 1298–1305. [Google Scholar] [CrossRef] [PubMed]
  45. Li, F.; Yu, Z.; Han, X.; Lai, R.Y. Electrochemical Aptamer-Based Sensors for Food and Water Analysis: A Review. Anal. Chim. Acta 2019, 1051, 1–23. [Google Scholar] [CrossRef] [PubMed]
  46. Musarurwa, H.; Tawanda Tavengwa, N. Extraction and Electrochemical Sensing of Pesticides in Food and Environmental Samples by Use of Polydopamine-Based Materials. Chemosphere 2021, 266, 129222. [Google Scholar] [CrossRef] [PubMed]
  47. Kurup, C.P.; Mohd-Naim, N.F.; Ahmed, M.U. Recent Trends in Nanomaterial-Based Signal Amplification in Electrochemical Aptasensors. Crit. Rev. Biotechnol. 2022, 42, 794–812. [Google Scholar] [CrossRef] [PubMed]
  48. Farid, S.; Ghosh, S.; Dutta, M.; Stroscio, M.A. Aptamer-Based Optical and Electrochemical Sensors: A Review. Chemosensors 2023, 11, 569. [Google Scholar] [CrossRef]
  49. Hashem, A.; Hossain, M.A.M.; Marlinda, A.R.; Mamun, M.A.; Simarani, K.; Johan, M.R. Nanomaterials Based Electrochemical Nucleic Acid Biosensors for Environmental Monitoring: A Review. Appl. Surf. Sci. Adv. 2021, 4, 100064. [Google Scholar] [CrossRef]
  50. Zhang, N.; Li, J.; Liu, B.; Zhang, D.; Zhang, C.; Guo, Y.; Chu, X.; Wang, W.; Wang, H.; Yan, X.; et al. Signal Enhancing Strategies in Aptasensors for the Detection of Small Molecular Contaminants by Nanomaterials and Nucleic Acid Amplification. Talanta 2022, 236, 122866. [Google Scholar] [CrossRef]
  51. Azzouz, A.; Kumar, V.; Hejji, L.; Kim, K.-H. Advancements in Nanomaterial-Based Aptasensors for the Detection of Emerging Organic Pollutants in Environmental and Biological Samples. Biotechnol. Adv. 2023, 66, 108156. [Google Scholar] [CrossRef]
  52. Rahimizadeh, K.; Zahra, Q.; Chen, S.; Le, B.T.; Ullah, I.; Veedu, R.N. Nanoparticles-Assisted Aptamer Biosensing for the Detection of Environmental Pathogens. Environ. Res. 2023, 238, 117123. [Google Scholar] [CrossRef] [PubMed]
  53. Guo, W.; Zhang, C.; Ma, T.; Liu, X.; Chen, Z.; Li, S.; Deng, Y. Advances in Aptamer Screening and Aptasensors’ Detection of Heavy Metal Ions. J. Nanobiotechnol. 2021, 19, 166. [Google Scholar] [CrossRef] [PubMed]
  54. Dolati, S.; Ramezani, M.; Abnous, K.; Taghdisi, S.M. Recent Nucleic Acid Based Biosensors for Pb2+ Detection. Sens. Actuators B Chem. 2017, 246, 864–878. [Google Scholar] [CrossRef]
  55. Khoshbin, Z.; Housaindokht, M.R.; Verdian, A.; Bozorgmehr, M.R. Simultaneous Detection and Determination of Mercury (II) and Lead (II) Ions through the Achievement of Novel Functional Nucleic Acid-Based Biosensors. Biosens. Bioelectron. 2018, 116, 130–147. [Google Scholar] [CrossRef] [PubMed]
  56. Zhao, L.; Huang, Y.; Dong, Y.; Han, X.; Wang, S.; Liang, X. Aptamers and Aptasensors for Highly Specific Recognition and Sensitive Detection of Marine Biotoxins: Recent Advances and Perspectives. Toxins 2018, 10, 427. [Google Scholar] [CrossRef]
  57. Kim, S.H.; Thoa, T.T.T.; Gu, M.B. Aptasensors for Environmental Monitoring of Contaminants in Water and Soil. Curr. Opin. Environ. Sci. Health 2019, 10, 9–21. [Google Scholar] [CrossRef]
  58. Yue, F.; Li, F.; Kong, Q.; Guo, Y.; Sun, X. Recent Advances in Aptamer-Based Sensors for Aminoglycoside Antibiotics Detection and Their Applications. Sci. Total Environ. 2021, 762, 143129. [Google Scholar] [CrossRef]
  59. Li, T.; Wang, J.; Zhu, L.; Li, C.; Chang, Q.; Xu, W. Advanced Screening and Tailoring Strategies of Pesticide Aptamer for Constructing Biosensor. Crit. Rev. Food Sci. Nutr. 2023, 63, 10974–10994. [Google Scholar] [CrossRef]
  60. Khosropour, H.; Kalambate, P.K.; Kalambate, R.P.; Permpoka, K.; Zhou, X.; Chen, G.Y.; Laiwattanapaisal, W. A Comprehensive Review on Electrochemical and Optical Aptasensors for Organophosphorus Pesticides. Mikrochim. Acta 2022, 189, 362. [Google Scholar] [CrossRef]
  61. Qin, N.; Liu, J.; Li, F.; Liu, J. Recent Advances in Aptasensors for Rapid Pesticide Residues Detection. Crit. Rev. Anal. Chem. 2023, 53, 1–22. [Google Scholar] [CrossRef]
  62. D’Adamo, I.; Gastaldi, M.; Giannini, M.; Nizami, A.-S. Environmental Implications and Levelized Cost Analysis of E-Fuel Production under Photovoltaic Energy, Direct Air Capture, and Hydrogen. Environ. Res. 2024, 246, 118163. [Google Scholar] [CrossRef]
  63. Tuerk, C.; Gold, L. Systematic Evolution of Ligands by Exponential Enrichment: RNA Ligands to Bacteriophage T4 DNA Polymerase. Science (1979) 1990, 249, 505–510. [Google Scholar] [CrossRef]
  64. Ellington, A.D.; Szostak, J.W. In Vitro Selection of RNA Molecules That Bind Specific Ligands. Nature 1990, 346, 818–822. [Google Scholar] [CrossRef]
  65. ZOU, X.-M.; ZHOU, J.-W.; SONG, S.-H.; CHEN, G.-H. Screening of Oligonucleotide Aptamers and Application in Detection of Pesticide and Veterinary Drug Residues. Chin. J. Anal. Chem. 2019, 47, 488–499. [Google Scholar] [CrossRef]
  66. Robertson, D.L.; Joyce, G.F. Selection in Vitro of an RNA Enzyme That Specifically Cleaves Single-Stranded DNA. Nature 1990, 344, 467–468. [Google Scholar] [CrossRef] [PubMed]
  67. Zhuo, Z.; Yu, Y.; Wang, M.; Li, J.; Zhang, Z.; Liu, J.; Wu, X.; Lu, A.; Zhang, G.; Zhang, B. Recent Advances in SELEX Technology and Aptamer Applications in Biomedicine. Int. J. Mol. Sci. 2017, 18, 2142. [Google Scholar] [CrossRef] [PubMed]
  68. Prante, M.; Segal, E.; Scheper, T.; Bahnemann, J.; Walter, J. Aptasensors for Point-of-Care Detection of Small Molecules. Biosensors 2020, 10, 108. [Google Scholar] [CrossRef]
  69. Kohlberger, M.; Gadermaier, G. SELEX: Critical Factors and Optimization Strategies for Successful Aptamer Selection. Biotechnol. Appl. Biochem. 2022, 69, 1771–1792. [Google Scholar] [CrossRef]
  70. Kadam, U.S.; Hong, J.C. Advances in Aptameric Biosensors Designed to Detect Toxic Contaminants from Food, Water, Human Fluids, and the Environment. Trends Environ. Anal. Chem. 2022, 36, e00184. [Google Scholar] [CrossRef]
  71. Lyu, C.; Khan, I.M.; Wang, Z. Capture-SELEX for Aptamer Selection: A Short Review. Talanta 2021, 229, 122274. [Google Scholar] [CrossRef]
  72. Luo, Y.; Guo, W.; Ngo, H.H.; Nghiem, L.D.; Hai, F.I.; Zhang, J.; Liang, S.; Wang, X.C. A Review on the Occurrence of Micropollutants in the Aquatic Environment and Their Fate and Removal during Wastewater Treatment. Sci. Total Environ. 2014, 473–474, 619–641. [Google Scholar] [CrossRef] [PubMed]
  73. Bhatt, P.; Bhandari, G.; Bilal, M. Occurrence, Toxicity Impacts and Mitigation of Emerging Micropollutants in the Aquatic Environments: Recent Tendencies and Perspectives. J. Environ. Chem. Eng. 2022, 10, 107598. [Google Scholar] [CrossRef]
  74. Reynoso, E.C.; Laschi, S.; Palchetti, I.; Torres, E. Advances in Antimicrobial Resistance Monitoring Using Sensors and Biosensors: A Review. Chemosensors 2021, 9, 232. [Google Scholar] [CrossRef]
  75. Mali, H.; Shah, C.; Raghunandan, B.H.; Prajapati, A.S.; Patel, D.H.; Trivedi, U.; Subramanian, R.B. Organophosphate Pesticides an Emerging Environmental Contaminant: Pollution, Toxicity, Bioremediation Progress, and Remaining Challenges. J. Environ. Sci. 2023, 127, 234–250. [Google Scholar] [CrossRef] [PubMed]
  76. Singh, S.; Rawat, M.; Malyan, S.K.; Singh, R.; Tyagi, V.K.; Singh, K.; Kashyap, S.; Kumar, S.; Sharma, M.; Panday, B.K.; et al. Global Distribution of Pesticides in Freshwater Resources and Their Remediation Approaches. Environ. Res. 2023, 225, 115605. [Google Scholar] [CrossRef] [PubMed]
  77. Bettazzi, F.; Romero Natale, A.; Torres, E.; Palchetti, I. Glyphosate Determination by Coupling an Immuno-Magnetic Assay with Electrochemical Sensors. Sensors 2018, 18, 2965. [Google Scholar] [CrossRef] [PubMed]
  78. Berti, F.; Lozzi, L.; Palchetti, I.; Santucci, S.; Marrazza, G. Aligned Carbon Nanotube Thin Films for DNA Electrochemical Sensing. Electrochim. Acta 2009, 54, 5035–5041. [Google Scholar] [CrossRef]
  79. Qi, Z.; Yan, P.; Qian, J.; Zhu, L.; Li, H.; Xu, L. A Photoelectrochemical Aptasensor Based on CoN/g-C3N4 Donor-Acceptor Configuration for Sensitive Detection of Atrazine. Sens. Actuators B Chem. 2023, 387, 133792. [Google Scholar] [CrossRef]
  80. Wei, J.; Ding, J.; Hu, Q.; Tian, X.; Bai, M.; Qian, J.; Wang, K. Internal Reference Self-Powered Aptasensor for on-Site Detection of MC-RR Used Sunlight as Light Source and CoMoS4 Hollow Nanospheres as Photocathode. Anal. Chim. Acta 2023, 1262, 341239. [Google Scholar] [CrossRef]
  81. Zhang, Z.; Karimi-Maleh, H. In Situ Synthesis of Label-Free Electrochemical Aptasensor-Based Sandwich-like AuNPs/PPy/Ti3C2Tx for Ultrasensitive Detection of Lead Ions as Hazardous Pollutants in Environmental Fluids. Chemosphere 2023, 324, 138302. [Google Scholar] [CrossRef]
  82. Li, L.; Chen, B.; Liu, X.; Jiang, P.; Luo, L.; Li, X.; You, T. ‘On-off-on’ Electrochemiluminescent Aptasensor for Hg2+ Based on Dual Signal Amplification Enabled by a Self-Enhanced Luminophore and Resonance Energy Transfer. J. Electroanal. Chem. 2022, 907, 116063. [Google Scholar] [CrossRef]
  83. Chen, Y.; Xu, L.; Yang, M.; Jia, Y.; Yan, Y.; Qian, J.; Chen, F.; Li, H. Design of 2D/2D CoAl LDH/g-C3N4 Heterojunction-Driven Signal Amplification: Fabrication and Assay for Photoelectrochemical Aptasensor of Ofloxacin. Sens. Actuators B Chem. 2022, 353, 131187. [Google Scholar] [CrossRef]
  84. Liu, T.; Lin, B.; Yuan, X.; Chu, Z.; Jin, W. In Situ Fabrication of Urchin-like Cu@carbon Nanoneedles Based Aptasensor for Ultrasensitive Recognition of Trace Mercury Ion. Biosens. Bioelectron. 2022, 206, 114147. [Google Scholar] [CrossRef]
  85. Chen, Y.; Xu, L.; Dong, J.; Yan, P.; Chen, F.; Qian, J.; Li, H. An Enhanced Photoelectrochemical Ofloxacin Aptasensor Using NiFe Layered Double Hydroxide/Graphitic Carbon Nitride Heterojunction. Electrochim. Acta 2021, 368, 137595. [Google Scholar] [CrossRef]
  86. Hamami, M.; Bouaziz, M.; Raouafi, N.; Bendounan, A.; Korri-Youssoufi, H. MoS2/PPy Nanocomposite as a Transducer for Electrochemical Aptasensor of Ampicillin in River Water. Biosensors 2021, 11, 311. [Google Scholar] [CrossRef]
  87. Salandari-Jolge, N.; Ensafi, A.A.; Rezaei, B. Ultra-Sensitive Electrochemical Aptasensor Based on Zeolitic Imidazolate Framework-8 Derived Ag/Au Core-Shell Nanoparticles for Mercury Detection in Water Samples. Sens. Actuators B Chem. 2021, 331, 129426. [Google Scholar] [CrossRef]
  88. Tian, C.; Zhao, L.; Zhu, J.; Zhang, S. Ultrasensitive Detection of Trace Hg2+ by SERS Aptasensor Based on Dual Recycling Amplification in Water Environment. J. Hazard. Mater. 2021, 416, 126251. [Google Scholar] [CrossRef]
  89. Lin, J.; Shi, A.; Zheng, Z.; Huang, L.; Wang, Y.; Lin, H.; Lin, X. Simultaneous Quantification of Ampicillin and Kanamycin in Water Samples Based on Lateral Flow Aptasensor Strip with an Internal Line. Molecules 2021, 26, 3806. [Google Scholar] [CrossRef]
  90. Qi, H.; Huang, X.; Wu, J.; Zhang, J.; Wang, F.; Qu, H.; Zheng, L. A Disposable Aptasensor Based on a Gold-Plated Coplanar Electrode Array for on-Site and Real-Time Determination of Cu2+. Anal. Chim. Acta 2021, 1183, 338991. [Google Scholar] [CrossRef]
  91. Song, Y.; Xu, M.; Liu, X.; Li, Z.; Wang, C.; Jia, Q.; Zhang, Z.; Du, M. A Label-Free Enrofloxacin Electrochemical Aptasensor Constructed by a Semiconducting CoNi-Based Metal–Organic Framework (MOF). Electrochim. Acta 2021, 368, 137609. [Google Scholar] [CrossRef]
  92. Yuan, R.; Ding, L.; You, F.; Wen, Z.; Liu, Q.; Wang, K. B, N Co-Doped Graphene Synergistic Catalyzed ZnO Quantum Dots with Amplified Cathodic Electrochemiluminescence for Fabricating Microcystin-LR Aptasensor. Sens. Actuators B Chem. 2021, 349, 130795. [Google Scholar] [CrossRef]
  93. Yuan, R.; Wen, Z.; You, F.; Jiang, D.; Wang, K. Catalysis-Induced Performance Enhancement of an Electrochemical Microcystin-LR Aptasensor Based on Cobalt-Based Oxide on a B, N Co-Doped Graphene Hydrogel. Analyst 2021, 146, 2574–2580. [Google Scholar] [CrossRef]
  94. Li, J.; Shan, X.; Jiang, D.; Chen, Z. An Ultrasensitive Electrochemiluminescence Aptasensor for the Detection of Diethylstilbestrol Based on the Enhancing Mechanism of the Metal–Organic Framework NH2-MIL-125(Ti) in a 3,4,9,10-Perylenetetracarboxylic Acid/K2S2O8 System. Analyst 2020, 145, 3306–3312. [Google Scholar] [CrossRef]
  95. Mohammadi, A.; Heydari-Bafrooei, E.; Foroughi, M.M.; Mohammadi, M. Electrochemical Aptasensor for Ultrasensitive Detection of PCB77 Using Thionine-Functionalized MoS2-RGO Nanohybrid. Microchem. J. 2020, 155, 104747. [Google Scholar] [CrossRef]
  96. Song, Y.; Xu, M.; Li, Z.; He, L.; Hu, M.; He, L.; Zhang, Z.; Du, M. Ultrasensitive Detection of Bisphenol A under Diverse Environments with an Electrochemical Aptasensor Based on Multicomponent AgMo Heteronanostructure. Sens. Actuators B Chem. 2020, 321, 128527. [Google Scholar] [CrossRef]
  97. Xu, L.; Duan, W.; Chen, F.; Zhang, J.; Li, H. A Photoelectrochemical Aptasensor for the Determination of Bisphenol A Based on the Cu (I) Modified Graphitic Carbon Nitride. J. Hazard. Mater. 2020, 400, 123162. [Google Scholar] [CrossRef]
  98. Zhao, Z.; Zheng, J.; Nguyen, E.P.; Tao, D.; Cheng, J.; Pan, H.; Zhang, L.; Jaffrezic-Renault, N.; Guo, Z. A Novel SWCNT-Amplified “Signal-on” Electrochemical Aptasensor for the Determination of Trace Level of Bisphenol A in Human Serum and Lake Water. Microchim. Acta 2020, 187, 500. [Google Scholar] [CrossRef]
  99. Fan, L.; Zhang, C.; Shi, H.; Zhao, G. Design of a Simple and Novel Photoelectrochemical Aptasensor for Detection of 3,3′,4,4′-Tetrachlorobiphenyl. Biosens. Bioelectron. 2019, 124–125, 8–14. [Google Scholar] [CrossRef]
  100. Liu, X.; Hu, M.; Wang, M.; Song, Y.; Zhou, N.; He, L.; Zhang, Z. Novel Nanoarchitecture of Co-MOF-on-TPN-COF Hybrid: Ultralowly Sensitive Bioplatform of Electrochemical Aptasensor toward Ampicillin. Biosens. Bioelectron. 2019, 123, 59–68. [Google Scholar] [CrossRef]
  101. Zhou, N.; Ma, Y.; Hu, B.; He, L.; Wang, S.; Zhang, Z.; Lu, S. Construction of Ce-MOF@COF Hybrid Nanostructure: Label-Free Aptasensor for the Ultrasensitive Detection of Oxytetracycline Residues in Aqueous Solution Environments. Biosens. Bioelectron. 2019, 127, 92–100. [Google Scholar] [CrossRef]
  102. Song, J.; Huang, M.; Jiang, N.; Zheng, S.; Mu, T.; Meng, L.; Liu, Y.; Liu, J.; Chen, G. Ultrasensitive Detection of Amoxicillin by TiO2-g-C3N4@AuNPs Impedimetric Aptasensor: Fabrication, Optimization, and Mechanism. J. Hazard. Mater. 2020, 391, 122024. [Google Scholar] [CrossRef]
  103. Kim, W.H.; Lee, J.U.; Jeon, M.J.; Park, K.H.; Sim, S.J. Three-Dimensional Hierarchical Plasmonic Nano-Architecture Based Label-Free Surface-Enhanced Raman Spectroscopy Detection of Urinary Exosomal MiRNA for Clinical Diagnosis of Prostate Cancer. Biosens. Bioelectron. 2022, 205, 114116. [Google Scholar] [CrossRef]
  104. Li, S.; Ma, X.; Pang, C.; Tian, H.; Xu, Z.; Yang, Y.; Lv, D.; Ge, H. Fluorometric Aptasensor for Cadmium(II) by Using an Aptamer-Imprinted Polymer as the Recognition Element. Microchim. Acta 2019, 186, 823. [Google Scholar] [CrossRef]
  105. Memon, A.G.; Xing, Y.; Zhou, X.; Wang, R.; Liu, L.; Zeng, S.; He, M.; Ma, M. Ultrasensitive Colorimetric Aptasensor for Hg2+ Detection Using Exo-III Assisted Target Recycling Amplification and Unmodified AuNPs as Indicators. J. Hazard. Mater. 2020, 384, 120948. [Google Scholar] [CrossRef]
  106. Rabai, S.; Benounis, M.; Catanante, G.; Baraket, A.; Errachid, A.; Jaffrezic Renault, N.; Marty, J.-L.; Rhouati, A. Development of a Label-Free Electrochemical Aptasensor Based on Diazonium Electrodeposition: Application to Cadmium Detection in Water. Anal. Biochem. 2021, 612, 113956. [Google Scholar] [CrossRef]
  107. Wei, T.; Zhang, Y.; Wang, H.; Li, H.; Fang, T.; Wang, Z.; Dai, Z. Self-Assembled Electroactive MOF–Magnetic Dispersible Aptasensor Enables Ultrasensitive Microcystin-LR Detection in Eutrophic Water. Chem. Eng. J. 2023, 466, 142809. [Google Scholar] [CrossRef]
  108. Wang, Y.; Li, Y.; Liu, C.; Dong, N.; Liu, D.; You, T. Laser Induced Graphene Electrochemical Aptasensor Based on Tetrahedral DNA for Ultrasensitive On-Site Detection of Microcystin-LR. Biosens. Bioelectron. 2023, 239, 115610. [Google Scholar] [CrossRef]
  109. Zhang, Z.; Ding, X.; Lu, G.; Du, B.; Liu, M. A Highly Sensitive and Selective Photoelectrochemical Aptasensor for Atrazine Based on Au NPs/3DOM TiO2 Photonic Crystal Electrode. J. Hazard. Mater. 2023, 451, 131132. [Google Scholar] [CrossRef]
  110. Xu, J.; Qing, T.; Jiang, Z.; Zhang, P.; Feng, B. Graphene Oxide-Regulated Low-Background Aptasensor for the “Turn on” Detection of Tetracycline. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2021, 260, 119898. [Google Scholar] [CrossRef]
  111. Mat Zaid, M.H.; Abdullah, J.; Rozi, N.; Mohamad Rozlan, A.A.; Abu Hanifah, S. A Sensitive Impedimetric Aptasensor Based on Carbon Nanodots Modified Electrode for Detection of 17ß-Estradiol. Nanomaterials 2020, 10, 1346. [Google Scholar] [CrossRef]
  112. Yildirim-Tirgil, N.; Lee, J.; Cho, H.; Lee, H.; Somu, S.; Busnaina, A.; Gu, A.Z. A SWCNT Based Aptasensor System for Antibiotic Oxytetracycline Detection in Water Samples. Anal. Methods 2019, 11, 2692–2699. [Google Scholar] [CrossRef]
  113. Chen, Y.; Yan, P.; Lu, G.; Chen, J.; Chen, F.; Xu, L. A Highly Selective Photoelectrochemical Chloramphenicol Aptasensor Based on AgBr/BiOBr Heterojunction. Inorg. Chem. Commun. 2021, 123, 108333. [Google Scholar] [CrossRef]
  114. Aryal, P.; Hefner, C.; Martinez, B.; Henry, C.S. Microfluidics in Environmental Analysis: Advancements, Challenges, and Future Prospects for Rapid and Efficient Monitoring. Lab Chip 2024, 24, 1175–1206. [Google Scholar] [CrossRef]
  115. Sfragano, P.S.; Reynoso, E.C.; Rojas-Ruíz, N.E.; Laschi, S.; Rossi, G.; Buchinger, M.; Torres, E.; Palchetti, I. A Microfluidic Card-Based Electrochemical Assay for the Detection of Sulfonamide Resistance Genes. Talanta 2024, 271, 125718. [Google Scholar] [CrossRef]
  116. Aslan, Y.; Atabay, M.; Chowdhury, H.K.; Göktürk, I.; Saylan, Y.; Inci, F. Aptamer-Based Point-of-Care Devices: Emerging Technologies and Integration of Computational Methods. Biosensors 2023, 13, 569. [Google Scholar] [CrossRef]
  117. Lan, Y.; He, B.; Tan, C.S.; Ming, D. Applications of Smartphone-Based Aptasensor for Diverse Targets Detection. Biosensors 2022, 12, 477. [Google Scholar] [CrossRef]
  118. Xu, C.; Lin, M.; Song, C.; Chen, D.; Bian, C. A Gold Nanoparticle-Based Visual Aptasensor for Rapid Detection of Acetamiprid Residues in Agricultural Products Using a Smartphone. RSC Adv. 2022, 12, 5540–5545. [Google Scholar] [CrossRef]
  119. Deiminiat, B.; Rounaghi, G.H. Fabrication of a Novel Photoelectrochemical Aptasensor Using Gold Nanoparticle-Sensitized TiO2 Film for Quantitative Determination of Diazinon in Solutions. Electrocatalysis 2023, 14, 484–498. [Google Scholar] [CrossRef]
  120. Fan, P.; Qian, X.; Li, Q.; Jiang, P.; Wu, Q.; Huang, G.; Zhang, Z.; Li, L. A Novel Label-Free Dual-Mode Aptasensor Based on the Mutual Regulation of Silver Nanoclusters and MoSe2 Nanosheets for Reliable Detection of Ampicillin. Anal. Chim. Acta 2023, 1251, 340997. [Google Scholar] [CrossRef]
  121. Gaviria-Arroyave, M.I.; Arango, J.P.; Barrientos Urdinola, K.; Cano, J.B.; Peñuela Mesa, G.A. Fluorescent Nanostructured Carbon Dot-Aptasensor for Chlorpyrifos Detection: Elucidating the Interaction Mechanism for an Environmentally Hazardous Pollutant. Anal. Chim. Acta 2023, 1278, 341711. [Google Scholar] [CrossRef]
  122. Irfan, M.; Murtaza, G.; Fu, S.; Chen, A.; Qu, F.; Su, X. Molecular Simulation-Guided Aptasensor Design of Robust and Sensitive Lateral Flow Strip for Cadmium Ion Detection. Analyst 2023, 148, 1961–1969. [Google Scholar] [CrossRef]
  123. Kushwah, M.; Yadav, R.; Berlina, A.N.; Gaur, K.; Gaur, M.S. Development of an Ultrasensitive RGO/AuNPs/SsDNA-Based Electrochemical Aptasensor for Detection of Pb2+. J. Solid. State Electrochem. 2023, 27, 559–574. [Google Scholar] [CrossRef]
  124. Macagno, J.; Gerlero, G.S.; Satuf, M.L.; Berli, C.L.A. Field-Deployable Aptasensor with Automated Analysis of Stain Patterns for the Detection of Chlorpyrifos in Water. Talanta 2023, 252, 123782. [Google Scholar] [CrossRef]
  125. Mou, Y.; Zhang, Y.; Lin, X.; Chen, M.; Xia, Y.; Zhu, S.; Wei, C.; Luo, X. Construction of a Novel Fluorescent DNA Aptasensor for the Fast-Response and Sensitive Detection of Copper Ions in Industrial Sewage. Anal. Methods 2023, 15, 3466–3475. [Google Scholar] [CrossRef]
  126. Muhammad, I.; Murtaza, G.; Zhao, Y.; Rizvi, A.S.; Fu, S.; Su, X.; Qu, F. Exploration of the Interaction of Cadmium and Aptamer by Molecular Simulation and Development of Sensitive Capillary Zone Electrophoresis-Based Aptasensor. J. Chem. Inf. Model. 2023, 63, 2783–2793. [Google Scholar] [CrossRef]
  127. Su, C.; Dong, C.; Jiang, D.; Shan, X.; Chen, Z. Construction of Electrochemiluminescence Aptasensor for Acetamiprid Detection Using Flower-Liked SnO2 Nanocrystals Encapsulated Ag3PO4 Composite as Luminophore. Microchem. J. 2023, 187, 108374. [Google Scholar] [CrossRef]
  128. Suo, Z.; Liang, R.; Liu, R.; Wei, M.; He, B.; Jiang, L.; Sun, X.; Jin, H. A Convenient Paper-Based Fluorescent Aptasensor for High-Throughput Detection of Pb2+ in Multiple Real Samples (Water-Soil-Food). Anal. Chim. Acta 2023, 1239, 340714. [Google Scholar] [CrossRef]
  129. Wang, H.; Chen, Z.; Zhu, C.; Du, H.; Mao, J.; Qin, H.; She, Y.; Yan, M. An Interference-Free SERS-Based Aptasensor for Chlorpyrifos Detection. Anal. Chim. Acta 2023, 1268, 341398. [Google Scholar] [CrossRef]
  130. Yu, H.; Wang, C.; Xiong, X.; Dai, B.; Wang, Y.; Feng, Z.; Luo, H.; Zhu, J.; Shen, G.; Deng, Y.; et al. Development of Fe-N-C Single-Atom Nanozymes Assisted Aptasensor for the Detection of Acetamiprid in Water Samples. Microchem. J. 2023, 193, 109174. [Google Scholar] [CrossRef]
  131. Yuan, M.; Yang, Y.; Chau, N.T.Q.; Zhang, Q.; Wu, X.; Chen, J.; Wu, Z.; Zhong, H.; Li, Y.; Xu, F. A Novel Fluorescent Aptasensor for Arsenic(III) Detection Based on a Triple-Helix Molecular Switch. Molecules 2023, 28, 2341. [Google Scholar] [CrossRef]
  132. Zhang, Y.; Zhu, Y.; Zeng, T.; Qiao, L.; Zhang, M.; Song, K.; Yin, N.; Tao, Y.; Zhao, Y.; Zhang, C.; et al. Self-Powered Photoelectrochemical Aptasensor Based on Hollow Tubular g-C3N4/Bi/BiVO4 for Tobramycin Detection. Anal. Chim. Acta 2023, 1250, 340951. [Google Scholar] [CrossRef]
  133. Zhang, Z.; Karimi-Maleh, H. Label-Free Electrochemical Aptasensor Based on Gold Nanoparticles/Titanium Carbide MXene for Lead Detection with Its Reduction Peak as Index Signal. Adv. Compos. Hybrid. Mater. 2023, 6, 68. [Google Scholar] [CrossRef]
  134. Zhu, Y.; Li, X.; Wu, M.; Shi, M.; Tian, Q.; Fu, L.; Tsai, H.-S.; Xie, W.-F.; Lai, G.; Wang, G.; et al. A Novel Electrochemical Aptasensor Based on Eco-Friendly Synthesized Titanium Dioxide Nanosheets and Polyethyleneimine Grafted Reduced Graphene Oxide for Ultrasensitive and Selective Detection of Ciprofloxacin. Anal. Chim. Acta 2023, 1275, 341607. [Google Scholar] [CrossRef]
  135. Sun, Z.; Zhao, L.; Li, C.; Jiang, Y.; Wang, F. Direct Z-Scheme In2S3/Bi2S3 Heterojunction-Based Photoelectrochemical Aptasensor for Detecting Oxytetracycline in Water. J. Environ. Chem. Eng. 2022, 10, 107404. [Google Scholar] [CrossRef]
  136. Wei, Q.; Zhang, P.; Pu, H.; Sun, D.-W. A Fluorescence Aptasensor Based on Carbon Quantum Dots and Magnetic Fe3O4 Nanoparticles for Highly Sensitive Detection of 17β-Estradiol. Food Chem. 2022, 373, 131591. [Google Scholar] [CrossRef]
  137. Yang, L.; Ye, X.; Li, X.; Huang, Z.; Chen, F.; Yang, W.; Wang, Z. Colorimetric Aptasensor for Sensitive Detection of Quinclorac Based on Exonuclease III-Assisted Cyclic Release of Phosphorodiamidate Morpholino Oligomer Mimic Enzyme Strategy. Anal. Chim. Acta 2022, 1207, 339815. [Google Scholar] [CrossRef]
  138. Chen, Y.; Wang, Z.; Liu, S.; Zhao, G. A Highly Sensitive and Group-Targeting Aptasensor for Total Phthalate Determination in the Environment. J. Hazard. Mater. 2021, 412, 125174. [Google Scholar] [CrossRef]
  139. Chen, Y.; Zhu, Q.; Zhou, X.; Wang, R.; Yang, Z. Reusable, Facile, and Rapid Aptasensor Capable of Online Determination of Trace Mercury. Environ. Int. 2021, 146, 106181. [Google Scholar] [CrossRef]
  140. Fan, L.; Liang, G.; Yan, W.; Guo, Y.; Bi, Y.; Dong, C. A Highly Sensitive Photoelectrochemical Aptasensor Based on BiVO4 Nanoparticles-TiO2 Nanotubes for Detection of PCB72. Talanta 2021, 233, 122551. [Google Scholar] [CrossRef]
  141. Fan, L.; Zhang, C.; Liang, G.; Yan, W.; Guo, Y.; Bi, Y.; Dong, C. Highly Sensitive Photoelectrochemical Aptasensor Based on MoS2 Quantum Dots/TiO2 Nanotubes for Detection of Atrazine. Sens. Actuators B Chem. 2021, 334, 129652. [Google Scholar] [CrossRef]
  142. Ding, J.; Zhang, D.; Liu, Y.; Zhan, X.; Lu, Y.; Zhou, P.; Zhang, D. An Electrochemical Aptasensor for Pb2+ Detection Based on Metal–Organic-Framework-Derived Hybrid Carbon. Biosensors 2020, 11, 1. [Google Scholar] [CrossRef]
  143. Li, L.; Chen, B.; Luo, L.; Liu, X.; Bi, X.; You, T. Sensitive and Selective Detection of Hg2+ in Tap and Canal Water via Self-Enhanced ECL Aptasensor Based on NH2–Ru@SiO2-NGQDs. Talanta 2021, 222, 121579. [Google Scholar] [CrossRef]
  144. Nguyen, D.K.; Jang, C.-H. A Cationic Surfactant-Decorated Liquid Crystal-Based Aptasensor for Label-Free Detection of Malathion Pesticides in Environmental Samples. Biosensors 2021, 11, 92. [Google Scholar] [CrossRef]
  145. Su, C.; Song, Q.; Jiang, D.; Dong, C.; Shan, X.; Chen, Z. An Electrochemiluminescence Aptasensor for Diethylstilbestrol Assay Based on Resonance Energy Transfer between Ag3PO4-Cu-MOF(Ii) and Silver Nanoparticles. Analyst 2021, 146, 4254–4260. [Google Scholar] [CrossRef]
  146. Wang, Y.; Yan, X.; Kou, Q.; Sun, Q.; Wang, Y.; Wu, P.; Yang, L.; Tang, J.; Le, T. An Ultrasensitive Label-Free Fluorescent Aptasensor Platform for Detection of Sulfamethazine. Int. J. Nanomed. 2021, 16, 2751–2759. [Google Scholar] [CrossRef]
  147. Zou, L.; Li, X.; Lai, Y. Colorimetric Aptasensor for Sensitive Detection of Kanamycin Based on Target-Triggered Catalytic Hairpin Assembly Amplification and DNA-Gold Nanoparticle Probes. Microchem. J. 2021, 162, 105858. [Google Scholar] [CrossRef]
  148. Chen, X.-X.; Lin, Z.-Z.; Hong, C.-Y.; Yao, Q.-H.; Huang, Z.-Y. A Dichromatic Label-Free Aptasensor for Sulfadimethoxine Detection in Fish and Water Based on AuNPs Color and Fluorescent Dyeing of Double-Stranded DNA with SYBR Green I. Food Chem. 2020, 309, 125712. [Google Scholar] [CrossRef]
  149. Chen, Y.; Wang, Y.; Yan, P.; Ouyang, Q.; Dong, J.; Qian, J.; Chen, J.; Xu, L.; Li, H. Co3O4 Nanoparticles/Graphitic Carbon Nitride Heterojunction for Photoelectrochemical Aptasensor of Oxytetracycline. Anal. Chim. Acta 2020, 1125, 299–307. [Google Scholar] [CrossRef]
  150. Khoshbin, Z.; Housaindokht, M.R.; Verdian, A. A Low-Cost Paper-Based Aptasensor for Simultaneous Trace-Level Monitoring of Mercury (II) and Silver (I) Ions. Anal. Biochem. 2020, 597, 113689. [Google Scholar] [CrossRef]
  151. Blidar, A.; Feier, B.; Tertis, M.; Galatus, R.; Cristea, C. Electrochemical Surface Plasmon Resonance (EC-SPR) Aptasensor for Ampicillin Detection. Anal. Bioanal. Chem. 2019, 411, 1053–1065. [Google Scholar] [CrossRef]
  152. Chen, X.-X.; Lin, Z.-Z.; Hong, C.-Y.; Zhong, H.-P.; Yao, Q.-H.; Huang, Z.-Y. Label-Free Fluorescence-Based Aptasensor for the Detection of Sulfadimethoxine in Water and Fish. Appl. Spectrosc. 2019, 73, 294–303. [Google Scholar] [CrossRef]
  153. Fan, L.; Wang, G.; Liang, W.; Yan, W.; Guo, Y.; Shuang, S.; Dong, C.; Bi, Y. Label-Free and Highly Selective Electrochemical Aptasensor for Detection of PCBs Based on Nickel Hexacyanoferrate Nanoparticles/Reduced Graphene Oxides Hybrids. Biosens. Bioelectron. 2019, 145, 111728. [Google Scholar] [CrossRef]
  154. Chen, Z.; Li, P.; Cheng, X.; Yang, W.; Wu, Y.; Chen, Q.; Fu, F. Multicolor Aptasensor Based on DNA-Induced Au–Ag Nanorods for Simultaneous and Visual Detection of Inorganic and Organic Mercury. ACS Omega 2019, 4, 15112–15119. [Google Scholar] [CrossRef]
  155. Saberi, Z.; Rezaei, B.; Ensafi, A.A. Fluorometric Label-Free Aptasensor for Detection of the Pesticide Acetamiprid by Using Cationic Carbon Dots Prepared with Cetrimonium Bromide. Microchim. Acta 2019, 186, 273. [Google Scholar] [CrossRef]
  156. Yang, R.; Liu, J.; Song, D.; Zhu, A.; Xu, W.; Wang, H.; Long, F. Reusable Chemiluminescent Fiber Optic Aptasensor for the Determination of 17β-Estradiol in Water Samples. Microchim. Acta 2019, 186, 726. [Google Scholar] [CrossRef]
  157. Ellington, A.D.; Szostak, J.W. Selection in Vitro of Single-Stranded DNA Molecules That Fold into Specific Ligand-Binding Structures. Nature 1992, 355, 850–852. [Google Scholar] [CrossRef]
  158. Jenison, R.D.; Gill, S.C.; Pardi, A.; Polisky, B. High-Resolution Molecular Discrimination by RNA. Science (1979) 1994, 263, 1425–1429. [Google Scholar] [CrossRef]
  159. Mendonsa, S.D.; Bowser, M.T. In Vitro Evolution of Functional DNA Using Capillary Electrophoresis. J. Am. Chem. Soc. 2004, 126, 20–21. [Google Scholar] [CrossRef]
  160. Mendonsa, S.D.; Bowser, M.T. In Vitro Selection of High-Affinity DNA Ligands for Human IgE Using Capillary Electrophoresis. Anal. Chem. 2004, 76, 5387–5392. [Google Scholar] [CrossRef]
  161. Ashley, J.; Ji, K.; Li, S.F.Y. Selection of Bovine Catalase Aptamers Using Non-SELEX. Electrophoresis 2012, 33, 2783–2789. [Google Scholar] [CrossRef]
  162. Jing, M.; Bowser, M.T. Isolation of DNA Aptamers Using Micro Free Flow Electrophoresis. Lab Chip 2011, 11, 3703–3709. [Google Scholar] [CrossRef]
  163. Park, S.; Ahn, J.-Y.; Jo, M.; Lee, D.; Lis, J.T.; Craighead, H.G.; Kim, S. Selection and Elution of Aptamers Using Nanoporous Sol-Gel Arrays with Integrated Microheaters. Lab Chip 2009, 9, 1206–1212. [Google Scholar] [CrossRef]
  164. Daniels, D.A.; Chen, H.; Hicke, B.J.; Swiderek, K.M.; Gold, L. A Tenascin-C Aptamer Identified by Tumor Cell SELEX: Systematic Evolution of Ligands by Exponential Enrichment. Proc. Natl. Acad. Sci. USA 2003, 100, 15416–15421. [Google Scholar] [CrossRef]
  165. Mi, J.; Liu, Y.; Rabbani, Z.N.; Yang, Z.; Urban, J.H.; Sullenger, B.A.; Clary, B.M. In Vivo Selection of Tumor-Targeting RNA Motifs. Nat. Chem. Biol. 2010, 6, 22–24. [Google Scholar] [CrossRef]
  166. Nguyen Quang, N.; Perret, G.; Ducongé, F. Applications of High-Throughput Sequencing for In Vitro Selection and Characterization of Aptamers. Pharmaceuticals 2016, 9, 76. [Google Scholar] [CrossRef]
  167. Stoltenburg, R.; Reinemann, C.; Strehlitz, B. FluMag-SELEX as an Advantageous Method for DNA Aptamer Selection. Anal. Bioanal. Chem. 2005, 383, 83–91. [Google Scholar] [CrossRef]
Figure 1. (1) Environmental contaminants from a synthetic origin are produced by different kinds of industrial activities for several applications. (2) Industrial, urban, and rural activities release contaminants in waste landfills and liquid discharge lines (3 and 4). These liquid waste dumps must be treated (5) before being released into natural water bodies (7). However, in various regions of the world, they are disposed of directly without prior treatment (6). Even when treated, in most cases it is not possible to eliminate or degrade contaminants. Thus, they will be then released into natural waters and eventually used for (8) industrial, urban, and rural activities.
Figure 1. (1) Environmental contaminants from a synthetic origin are produced by different kinds of industrial activities for several applications. (2) Industrial, urban, and rural activities release contaminants in waste landfills and liquid discharge lines (3 and 4). These liquid waste dumps must be treated (5) before being released into natural water bodies (7). However, in various regions of the world, they are disposed of directly without prior treatment (6). Even when treated, in most cases it is not possible to eliminate or degrade contaminants. Thus, they will be then released into natural waters and eventually used for (8) industrial, urban, and rural activities.
Chemosensors 12 00059 g001
Figure 2. History of the number of review papers on the detection of contaminants in environmental water samples using aptasensors. Data extracted from Scopus on 20 March 2024.
Figure 2. History of the number of review papers on the detection of contaminants in environmental water samples using aptasensors. Data extracted from Scopus on 20 March 2024.
Chemosensors 12 00059 g002
Figure 3. Exclusion and inclusion results of the systematic review in 2019–2023.
Figure 3. Exclusion and inclusion results of the systematic review in 2019–2023.
Chemosensors 12 00059 g003
Figure 4. General scheme of SELEX and its main applications.
Figure 4. General scheme of SELEX and its main applications.
Chemosensors 12 00059 g004
Figure 5. Applications of aptasensor in the environmental field from 2018 to 2023. The term opto-electrochemical includes photoelectrochemical and electroluminescent sensors. The numbers indicate the number of total papers. Created with flourish.studio.
Figure 5. Applications of aptasensor in the environmental field from 2018 to 2023. The term opto-electrochemical includes photoelectrochemical and electroluminescent sensors. The numbers indicate the number of total papers. Created with flourish.studio.
Chemosensors 12 00059 g005
Figure 6. Configuration and sensitivity of aptasensor for contaminant detection in environmental water from 2015 to 2023. Upper part: total aptasensor according to transductor used (a). Type of interface and sensitivity for optical (b), electrochemical (c), and opto-electrochemical (d) transductor. Created with flourish.studio.
Figure 6. Configuration and sensitivity of aptasensor for contaminant detection in environmental water from 2015 to 2023. Upper part: total aptasensor according to transductor used (a). Type of interface and sensitivity for optical (b), electrochemical (c), and opto-electrochemical (d) transductor. Created with flourish.studio.
Chemosensors 12 00059 g006
Figure 7. (A) Schematic diagram of the fabrication of a “signal-on” sensing model. (B) Differential pulse voltammetry responses at different concentrations of BPA (10−19 M, 10−18 M, 10−17 M, 10−16 M, 10−15 M, 10−14 M) in lake water. (C) BPA calibration curve in lake water. Reprinted from reference [98] with permission from Springer Nature.
Figure 7. (A) Schematic diagram of the fabrication of a “signal-on” sensing model. (B) Differential pulse voltammetry responses at different concentrations of BPA (10−19 M, 10−18 M, 10−17 M, 10−16 M, 10−15 M, 10−14 M) in lake water. (C) BPA calibration curve in lake water. Reprinted from reference [98] with permission from Springer Nature.
Chemosensors 12 00059 g007
Figure 8. (a) Schematic diagram of SERS aptasensor based on dual recycling amplification for trace Hg2+ detection; (b) SERS spectra under different concentrations of Hg2+; (c) linear relationship between Raman intensity and different mercury ion concentrations. Reprinted from reference [88] with permission from Elsevier.
Figure 8. (a) Schematic diagram of SERS aptasensor based on dual recycling amplification for trace Hg2+ detection; (b) SERS spectra under different concentrations of Hg2+; (c) linear relationship between Raman intensity and different mercury ion concentrations. Reprinted from reference [88] with permission from Elsevier.
Chemosensors 12 00059 g008
Figure 9. Limits of detection for the different configurations of aptasensors. The red line represents 1 µg/L. The points are the total papers according to the LOD of the aptasensor.
Figure 9. Limits of detection for the different configurations of aptasensors. The red line represents 1 µg/L. The points are the total papers according to the LOD of the aptasensor.
Chemosensors 12 00059 g009
Figure 10. (a) Schematic of lateral flow assay for simulation detection of kanamycin and ampicillin. (b) Fluorescent images for selectivity analysis and (c) its corresponding intensities. The concentrations of KAM and AMP on strip 6 were 30 ng/L, and the concentrations of KAM and AMP on strip 7 were 80 ng/L, respectively. The concentrations of other interfering antibiotics (1, 2, 3, 4, and 5) were 80 ng/L. Con. and Int. represent the control line and internal line, respectively. Reprinted from ref. [89] under the terms and conditions of the Creative Commons Attribution license (http://creativecommons.org/licenses/by/4.0/) accessed on 10 March 2024.
Figure 10. (a) Schematic of lateral flow assay for simulation detection of kanamycin and ampicillin. (b) Fluorescent images for selectivity analysis and (c) its corresponding intensities. The concentrations of KAM and AMP on strip 6 were 30 ng/L, and the concentrations of KAM and AMP on strip 7 were 80 ng/L, respectively. The concentrations of other interfering antibiotics (1, 2, 3, 4, and 5) were 80 ng/L. Con. and Int. represent the control line and internal line, respectively. Reprinted from ref. [89] under the terms and conditions of the Creative Commons Attribution license (http://creativecommons.org/licenses/by/4.0/) accessed on 10 March 2024.
Chemosensors 12 00059 g010
Table 1. Description of the primary data sought in the systematic review.
Table 1. Description of the primary data sought in the systematic review.
ConceptDescription
AnalyteName of environmental contaminant.
Analyte classificationThe environmental contaminants were classified according to their chemical family: metals, pesticides, toxins, industrial chemicals, and pharmaceutical compounds.
Transducer typeElectrochemical, optical, photoelectrochemical/electroluminescence (opto-electrochemical) transducers.
SensitivitySensitivity according to the LOD 1 using the following ranges as a basis: low (LOD > 0.1 mg/L), medium (0.1 mg/L > LOD ≥ 1 μg/L), high (1 μg/L > LOD > 0.1 ng/L), and ultra-high (LOD < 0.1 ng/L).
Water sample typeWater used from a complex matrix: river water, lake water, wastewater, or seawater.
Test on real sampleRefers to whether the target analytes were found in real water samples, or the compounds were spiked to the water samples.
Selectivity/specificityPresence of interferents in the complex matrix and whether the assay was performed: (1) with the target in the presence of the interferents in the same sample (mixed with interferents); (2) whether the target and interferents were analyzed separately (individual); (3) if the target was analyzed with one of the interferents (individual interferents with the target); and 4) if this test was not reported (NR).
Reproducibility/repeatabilityReported RSD 2.
StabilityTo evaluate the behavior of the aptasensor over time.
ReusabilityDetermines if the same device can be used in different periods.
1 Limit of detection; 2 relative standard deviation.
Table 2. Examples of aptasensors for environmental applications.
Table 2. Examples of aptasensors for environmental applications.
Sensitivity (LOD)TransducerNanomaterialsTarget ClassificationReference
7.11 ag/LOpto-electrochemicalGraphitic carbon loaded by CoN nanoparticles (CoN/g-C3N4)Pesticides[79]
0.33 pg/LElectrochemicalCoMoS4 hollow nanospheresMycotoxins[80]
2.07 pg/LElectrochemicalNanocomposite structure of AuNPs/PPy/Ti3C2TxHeavy metals[81]
0.601 pg/LOpto-electrochemicalRu(bpy)32+-doped silica nanoparticle-nitrogen-doped graphene quantum dots (Ru@SiO2-NGQDs)Heavy metals[82]
3.01 fg/LOpto-electrochemical(CoAl LDH/g-C3N4) two-dimensional/two-dimensional structurePharmaceutical
compounds
[83]
0.742 pg/LElectrochemicalCu@carbon nanoneedles (Cu@CNNs)Heavy metals[84]
12.64 pg/LOpto-electrochemicalNiFe layered double hydroxide (NiFe LDH)/graphitic carbon nitride (g-CN) heterojunctionPharmaceutical
compounds
[85]
97.8 pg/LElectrochemicalNanostructure composed of MoS2 nanosheets and conductive polypyrrole nanoparticles (PPyNPs)Pharmaceutical
compounds
[86]
3.6 fg/LElectrochemicalZeolitic Imidazolate Framework-8 (ZIF-8)-derived Ag@Au core–shell nanoparticles (Ag@Au/ZIF-8)Heavy metals[87]
22 fg/LOpticMagnetic beads and gold nanoparticles (AunNPs)Heavy metals[88]
15 pg/LOpticNot applicablePharmaceutical
compounds
[89]
0.189 pg/LElectrochemicalGold-plated coplanar electrode arrayHeavy metals[90]
0.22 pg/LElectrochemicalCoNi-based metal–organic framework (MOF), CoxNi3-x,(HITP)2,Pharmaceutical
Compounds
[91]
29.86 pg/LElectrochemicalZnO quantum dots decorated B, N co-doped graphene (BNG/ZnO)Mycotoxins[92]
29.86 pg/LElectrochemical3D cobalt-based oxide modified boron and nitrogen co-doped graphene hydrogel (3D BNG/Co)Mycotoxins[93]
75 fg/LOpto-electrochemicalMetal–organic framework NH2-MIL-125(Ti)Pharmaceutical
compounds
[94]
80 fg/LElectrochemicalThionine (Thi)-functionalized MoS2-rGO nanocompositeIndustrial chemicals[95]
0.20 pg/LElectrochemicalNanohybrid of Ag, Ag2O, Ag2S, and ultra-thin MoS2 nanosheet (Ag/Ag2O/Ag2S/MoS2(600))Industrial chemicals[96]
15 pg/LOpto-electrochemicalCu(I) modified carbon nitride (Cu/g-C3N4)Industrial chemicals[97]
18 ag/LElectrochemicalMulti-walled carbon nanotubes (MWCNT), amino-functionalized magnetite, and gold nanoparticles (NH2-Fe3O4/Au NPs)Industrial chemicals[98]
99.86 pg/LOpto-electrochemicalN-doped TiO2 nanotubes (N-doped TiO2 NTs)Industrial chemicals[99]
0.22 pg/LElectrochemicalCo-based metal–organic frameworks (Co-MOF) and terephthalonitrile-based covalent organic framework (TPN-COF) (Co-MOF@TPN-COF)Pharmaceutical
compounds
[100]
17.4 pg/LElectrochemicalNanohybrids of Covalent organic framework (COF) and Ce-based metal organic framework (Ce-MOF) (Ce-MOF@COF hybrid nanostructure)Pharmaceutical
compounds
[101]
Table 3. Selectivity and interference assessment of some aptasensors: criteria (ultrahigh) and (individual and mixed with interferents) selectivity.
Table 3. Selectivity and interference assessment of some aptasensors: criteria (ultrahigh) and (individual and mixed with interferents) selectivity.
AptasensorContaminantSelectivity% InterferenceReference
Sandwich-like AuNPs/PPy/Ti3C2TxPb2+Individual and mixed interferentsThe aptasensor was used to test the response towards eleven other ions; excluding Pb2+ and Mix, all other ions caused negligible response changes[81]
Aptamer linked with AuNPs and
Ru@SiO2-NGQD
Hg2+Individual and mixed with interferentsTen different interfering ions. The response caused by individual interfering ions or their mixtures was nearly negligible[82]
Urchin-like Cu@carbon nanoneedles modified electrodeHg2+Individual and mixed with interferentsEach of eight interferents with the concentration of 1 μM
produced a negligible signal
response compared to that
generated by 1 nM Hg2+
[84]
ZIF-8-derived Ag@Au core–shell nanoparticles (Ag@Au/ZIF-8)Hg2+Individual interferents with the targetThe presence of a 100-fold higher concentration of eight metal ions produced negligible effect on the current response of aptasensor [87]
CoNi-based metal–organic framework (MOF), CoxNi3−x,(HITP)2EnrofloxacinIndividual and mixed with interferentsNo significant response was observed for each individual interferent (thirteen antibiotics, small biomolecules, and harmful ions). In addition, the response with a mix is comparable to that of a pure enrofloxacin solution (104.4%)[91]
Metal–organic frameworks NH2-MIL-125(Ti)DiethylstilbestrolIndividual and mixed with interferentsThere was no significant difference between the response of the sensor with diethylstilbestrol and the response of three different interferents or their mixture [94]
Nanohybrid of Ag, Ag2O, Ag2S, and ultra-thin MoS2 nanosheetBisphenol AIndividual and mixed with interferentsThe response with nine interferents showed fluctuations of approximately 2.3–4.8%. When all the interferences were mixed with bisphenol A, the obtained value was 106.28% compared to that of pure bisphenol[96]
Multi-walled carbon nanotubes (MWCNT), amino-functionalized magnetite, and gold nanoparticles (NH2-Fe3O4/AuNPs)Bisphenol AIndividual interferents with the targetIn the presence of four interferents and bisphenol A, the response is close to that of the bisphenol A alone, with <3% difference in value [98]
Nanohybrids of Covalent organic framework (COF) and Ce-based metal organic framework (Ce-MOF) (Ce-MOF@COF hybrid nanostructure)OxytetracyclineIndividual and mixed with interferentsNegligible response variation in the presence of eleven interferents
(some ions, biomolecules, and antibiotics), except for oxytetracycline
[101]
Graphene oxide (GO)TetracyclineIndividual and mixed with interferentsThe system responded only to tetracycline, whereas other analogs (eight antibiotics) did not produce significant signal changes[110]
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

Reynoso, E.C.; Sfragano, P.S.; González-Perea, M.; Palchetti, I.; Torres, E. Aptasensors for the Detection of Environmental Contaminants of High Concern in Water Bodies: A Systematic Review. Chemosensors 2024, 12, 59. https://doi.org/10.3390/chemosensors12040059

AMA Style

Reynoso EC, Sfragano PS, González-Perea M, Palchetti I, Torres E. Aptasensors for the Detection of Environmental Contaminants of High Concern in Water Bodies: A Systematic Review. Chemosensors. 2024; 12(4):59. https://doi.org/10.3390/chemosensors12040059

Chicago/Turabian Style

Reynoso, Eduardo Canek, Patrick Severin Sfragano, Mario González-Perea, Ilaria Palchetti, and Eduardo Torres. 2024. "Aptasensors for the Detection of Environmental Contaminants of High Concern in Water Bodies: A Systematic Review" Chemosensors 12, no. 4: 59. https://doi.org/10.3390/chemosensors12040059

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