Next Article in Journal
Detection of Road Risk Sources Based on Multi-Scale Lightweight Networks
Previous Article in Journal
A Lightweight Pathological Gait Recognition Approach Based on a New Gait Template in Side-View and Improved Attention Mechanism
Previous Article in Special Issue
Microscale Sensor Arrays for the Detection of Dopamine Using PEDOT:PSS Organic Electrochemical Transistors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electrochemical Sensors for Antibiotic Detection: A Focused Review with a Brief Overview of Commercial Technologies

Sensor Engineering Department, Faculty of Science and Engineering, Maastricht University, P.O. Box 616, 6200 MD Maastricht, The Netherlands
*
Author to whom correspondence should be addressed.
Sensors 2024, 24(17), 5576; https://doi.org/10.3390/s24175576
Submission received: 2 August 2024 / Revised: 21 August 2024 / Accepted: 26 August 2024 / Published: 28 August 2024

Abstract

:
Antimicrobial resistance (AMR) poses a significant threat to global health, powered by pathogens that become increasingly proficient at withstanding antibiotic treatments. This review introduces the factors contributing to antimicrobial resistance (AMR), highlighting the presence of antibiotics in different environmental and biological matrices as a significant contributor to the resistance. It emphasizes the urgent need for robust and effective detection methods to identify these substances and mitigate their impact on AMR. Traditional techniques, such as liquid chromatography-mass spectrometry (LC-MS) and immunoassays, are discussed alongside their limitations. The review underscores the emerging role of biosensors as promising alternatives for antibiotic detection, with a particular focus on electrochemical biosensors. Therefore, the manuscript extensively explores the principles and various types of electrochemical biosensors, elucidating their advantages, including high sensitivity, rapid response, and potential for point-of-care applications. Moreover, the manuscript investigates recent advances in materials used to fabricate electrochemical platforms for antibiotic detection, such as aptamers and molecularly imprinted polymers, highlighting their role in enhancing sensor performance and selectivity. This review culminates with an evaluation and summary of commercially available and spin-off sensors for antibiotic detection, emphasizing their versatility and portability. By explaining the landscape, role, and future outlook of electrochemical biosensors in antibiotic detection, this review provides insights into the ongoing efforts to combat the escalating threat of AMR effectively.

1. Introduction

Pathogens are dangerous classes of microorganisms, including bacteria, viruses, fungi, or parasites, that can cause different types of diseases in their host organisms [1]. These microbes have the ability to infect humans, animals, and plants, succeeding in multiplying within the tissues of the host, thus potentially leading to infections and various health problems. While prevention methods such as proper hand hygiene and vaccination are still the most cost-effective and sustainable approaches to stopping the spread of bacterial infections, the discovery of antibiotics was also crucial in controlling and treating them [2]. However, pathogens are generating increasingly more issues for the environment, healthcare, and food safety due to their ability to develop adaptivity to diverse conditions, including resistance to specific drugs used to eradicate them. Antimicrobial resistance (AMR) occurs when microorganisms, including bacteria, evolve mechanisms to evade the effects of drugs designed to inhibit their growth or kill them. This resistance can arise not only from the evolution of new resistant strains but also from the expansion of naturally resistant populations or through intrinsic resistance mechanisms [3]. In addition, it is important to note that some antibiotics act by inhibiting bacterial growth (bacteriostatic) rather than directly killing bacteria (bactericidal) [4]. AMR, or antimicrobial resistance, refers to the ability of various microbes—including bacteria, viruses, fungi, and parasites—to withstand the effects of a broad range of antimicrobial agents. According to the WHO, antibiotics are a specific subset of antimicrobial agents that target bacteria. Thus, antibiotic resistance (AR or ABR) specifically pertains to bacterial resistance to these drugs. It is important to note that in many scientific publications, the terms AMR and AR are often used interchangeably when discussing bacterial resistance [5]. The problem of antimicrobial and antibiotic resistance is becoming increasingly overwhelming. According to the Centers for Disease Control and Prevention (CDC), the US alone experiences over 2.8 million AMR-related infections each year, leading to more than 35,000 deaths. These infections include both community-acquired and nosocomial infections, which differ in their origins and settings. Community-acquired infections occur outside of healthcare settings, while nosocomial infections are acquired during hospital stays. Both types are significantly impacted by AMR, highlighting the widespread nature of the problem [6,7]. This acquired resistance creates infections that are challenging or even impossible to treat, which further increases the spreading of diseases, severe illnesses, and death. Antimicrobial-resistant pathogens cause health complications such as pneumonia and liver abscesses, as well as several infections ranging from skin and soft tissue infections to urinary tract infections [8]. The six main nosocomial antimicrobial-resistant pathogens that are primarily to blame for these health complications have been coined under the acronym ESKAPE and they include Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter species [9].
The development of AMR occurs through two mechanisms. Firstly, gene inheritance occurs via mobile genetic elements through horizontal gene transfer (HGT), allowing antibiotic resistance to be transferred among different species. There are three types of HGT: transformation, conjugation, and transduction [10]. Transformation occurs when bacteria take up DNA from their environment, including exogenous DNA that encodes antibiotic-resistant genes, which may have been released by dead bacteria. This DNA is then incorporated into the host bacterium with the help of specific competence factors. Conjugation is the process of direct gene transfer between bacteria via a specialized conjugation tube or pilus. Finally, transduction is the action of bacteriophages transferring genes from one cell to another, including AMR genes to bacterial cells. After gene transfer, pathogens continue to adapt through various mechanisms, including genetic mutations and the acquisition of new resistance genes. In the case of antibiotic resistance, the overuse of antibiotics employs natural selection pressure, which promotes the proliferation of resistant strains [11]. In addition, random genetic mutations can take place, which result in variations in translated proteins, or increases/decreases in gene expression, that allow for antibiotic resistance [12]. When antibiotics are administered to humans or animals, they kill or prevent the growth of the target bacteria. Inevitably, only bacteria that did not develop resistance are killed, leaving behind those with natural or acquired resistance. This process is a form of natural selection at the microbial level and those with antibiotic-resistant genes can pass on these advantageous traits to their offspring [13]. Over time, this can lead to the proliferation of antibiotic-resistant bacterial strains, as their resistance genes are inherited by subsequent generations. However, it is important to note that resistance genes may impose a fitness cost on bacteria, with the consequence that without antibiotic selection pressure, the frequency of these resistant strains may decrease [14]. The main mechanism of AR may be separated into four principal strategies employed by bacteria: (a) limiting the uptake of a drug, namely reducing the permeability of the cell membranes to prevent the drug from entering the cell; (b) modifying a drug target, a genetic ability that bacteria can use to modify the molecules or the receptors targeted by the antibiotics, (c) inactivating a drug, which usually happens through the production of specific enzymes that can chemically modify or degrade the antibiotic, and lastly, (d) active drug efflux, namely the ability of bacteria to actively transport the drug outside the cell before it reaches the target [15]. Notably, Gram-negative bacteria can use all four processes, whereas Gram-positive bacteria do not have the capacity for certain drug efflux mechanisms or the ability to limit the uptake of the drug due to lacking a lipopolysaccharide outer membrane, making Gram-negative bacteria generally more difficult to fight [16].
Numerous causes may be attributed to the development of AMR and AR, including antibiotic misuse, inaccurate prescribing, excessive use in agriculture and livestock, inefficient wastewater treatment, and lack of novel antibiotics [17]. The excessive use of antibiotics in agriculture results in the transfer of antibiotic-resistant bacteria to humans through products such as cereals, meat, eggs, milk, vegetables, fruit, and even water [18,19]. In addition, the redundant use of antibiotics in livestock leads to detrimental effects on the environment. For example, up to 90% of antibiotics administered to livestock are excreted through urine and feces and are then spread to local groundwater [20]. Antibiotic misuse in humans is another prevalent cause of AR, with antibiotics being administered too indiscriminately [21]. The importance of the detection of antibiotics and AMR microorganisms has been highlighted by the “One Health” approach adopted by the WHO, the World Organization for Animal Health (WOAH—formerly OIE), the United Nations Food and Agriculture Organization (FAO), and the United Nations Environment Programme (UNEP). The “One Health“ approach is built on the concept that human health is intricately connected to those of other animals and the environment that they inhabit [22].
Besides environmental applications, antibiotic detection is important in clinical development [23,24]. In fact, a key step during the development of novel antibiotics is translational Pharmacokinetic/Pharmacodynamic (PK/PD) studies, which require accurate antibiotic concentration measurements to predict the efficacy of antibiotics in animal models and humans [25,26]. During patient treatment, antibiotic detection plays a crucial role in Therapeutic Drug Monitoring (TDM) to allow for personalized medicine [27]. In 2019, the WHO published a list of drugs that require TDM, which contained all major classes of antibiotics [28]. Previously, this concern was addressed through pharmaceutical development, with new antibiotics being synthesized to overcome the issue of resistance. In 2021, an annual WHO analysis identified 27 new antibiotics in clinical development targeting priority pathogens [28]. Furthermore, between 2017 and 2022, 12 novel antibiotics were approved, with 10 of these belonging to antibiotic classes that are particularly susceptible to resistance [29]. Due to this continuous drive towards the development of new antibiotics, the development of fast, accessible, and ready-to-use antibiotic detection methods has acquired paramount importance for both industry and academia, leading to the evolution of technologies and novel publications presenting a wide variety of alternatives to fight the increasing number of problems generated by AR and AMR [30]. In addition, many countries have adopted strict maximum residue limits for antibiotics in water and food to limit the effect of residual antibiotics, resulting in the need for sensitive detection techniques for antibiotics [31].
Currently, the gold-standard analytical methods for quantification of antibiotics in biological, food, and environmental samples are liquid chromatography (LC), typically paired with mass spectrometry (MS), gas chromatography (GC), and capillary electrophoresis (CE) [32]. The most prevailing LC technique is high-performance liquid chromatography (HPLC), which is commonly employed due to its sensitivity as well as specificity for identifying antibiotics such as β-lactam compounds [33], namely one of the most prescribed antibiotic classes that include penicillins, carbapenems, and monobactams [34]. β-lactams’ mechanism of action includes interruption of cell wall formation due to covalent binding to penicillin-based proteins (PBPs), enzymes involved in the vital cross-linking of peptidoglycan in both Gram-negative and Gram-positive bacteria [35]. The LoDs for the detection of β-lactams in serum and interstitial fluids were found to be 0.003 mg/L [36].
On the other hand, an alternative to the gold-standard technologies is represented by immunoassays such as enzyme-linked immunosorbent assays (ELISAs) and immunochromatographic assays (ICAs). These technologies are common examples of screening techniques that utilize antibodies for screening antibiotic residues [37]. They enable the detection of antibiotics through the linking of the analyte to a specific antibody [38]. Since these reactions are stoichiometric, it is possible to determine the concentration of a given antibiotic based on the determination of either free or bound antigens [39]. ELISAs have high sensitivity, detects multiple samples at once, and allows for the screening of many small-volume samples [40,41,42]. Immunoassays may be characterized by low limits of detection (LoDs), as well as high specificity and sensitivity [43]. An example is the indirect competitive chemiluminescence enzyme immunoassay (IC-CLEIA) developed for the detection of chloramphenicol (CAP) residues in shrimp [43]. It was found that LoDs for the detection of various antibiotics ranged from 0.02 ng/mL to 0.5 ng/mL [44,45].
The methods for antibiotic detection presented above, although currently accepted in the industry, present numerous limitations. MS techniques require steep initial investments in equipment, as well as solid infrastructure [46]. In addition, LC-MS can struggle to produce precise quantitative results, particularly at lower concentrations of the analyte, which increases the likelihood of false-negative results [46]. In contrast, lateral flow immunoassays are designed to be more straightforward and rapid, but they typically require pre-treatment to minimize interference that could cause false positives. Such pre-treatments might include sample dilution, purification, the use of blocking agents, and buffer optimization [47]. While these steps are crucial for reducing non-specific interactions and improving assay accuracy, they can also delay the time of detection. Despite these challenges, lateral flow immunoassays have proven effective in detecting different analytes in point-of-need testing, as demonstrated, for instance, for COVID-19 [48]. These limitations impact the spread of AR as the ubiquity of antibiotics throughout the environment may not be addressed quickly enough. As an emerging alternative, biosensors have the potential to address these limitations by presenting extremely high sensitivity, rapid detection, and versatility compared to the current gold-standard techniques [49].
Biosensors are devices that can measure biological or chemical reactions through the generation of signals that are proportional to the concentration of the analyte tested. Biosensors have wide-scale commercial applications, which include food safety, metabolite detection, disease detection, etc. [50]. The most common types of biosensors are electrochemical, optical, thermal, piezoelectric, and potentiometric biosensors [51,52]. Typically, a biosensor consists of a recognition element and a transducer. Recognition elements are biomolecules that have specific binding properties related to the analyte. Enzymes, antibodies, deoxyribonucleic acids (DNA), and molecularly imprinted polymers (MIPs) are all examples of commonly used biorecognition elements [53]. The generation of a signal from the interaction between an analyte and the bio-receptor is coined as biorecognition [54]. A transducer can convert a biorecognition event into a measurable signal, which may be used to deduce the presence of an analyte and its concentration [55]. Another similar category of sensing technologies is covered by biomimetic sensors, which, similarly to biosensors, generate a signal after an interaction with an analyte. In addition, biomimetic sensors’ working principle is based on the introduction of biomaterials designed to imitate the function of biological processes, i.e., lipid membranes, polymers, etc. [56].
Optical biosensors are the most common type of biosensor used commercially. Detection of analytes by optical biosensors is based on the interaction of an optical field with a biorecognition element [57]. Generally, optical biosensors may be divided into two categories: label-based and label-free [58]. Label-based sensing relies on the use of a label and thus the optical signal is usually generated by a colorimetric, fluorescent, or luminescent response [59]. An example of this type of sensing is the detection of glucose through enzymatic oxidation, which is applied in the most common commercially used biosensor in the world, the hand-held glucose meter used by diabetics to test blood sugar levels [60,61]. In contrast, label-free biosensors generate an electrical signal directly through the interaction of the analyte with the transducer [62]. An example of this is surface plasmon resonance-based optical biosensors (SPR biosensors), which are among the most common biosensors for the detection of pathogens in food and water [63].
Despite the consistent growth of optical biosensors as a solution for antibiotic detection, they still present numerous issues. One of the key challenges that optical biosensors face is cost and reusability [64]. The commercialization and miniaturization of optical biosensors are very limited due to their optical components being extremely delicate and expensive [65]. In this scenario, electrochemical biosensors were shown to be a solid alternative for the detection of antibiotics in a wide variety of samples and matrices, paving the way for the possibility of creating versatile devices that can be employed in resource-limited environments or, more generally, applied for point-of-care analysis thanks to their portability [66]. While a few review articles have already been published on electrochemical biosensors for antibiotic detection, this work focuses on novel biorecognition elements used for the manufacture of electrochemical biosensors and offers an overview of the technologies used for antibiotic detection for commercial purposes, thereby providing fresh insights and expanding the understanding of their potential in diverse and emerging contexts [67,68,69]. The research papers and reviews that are included in this manuscript were mainly published from 2020 to 2024, presenting an updated overview of the most recent works related to electrochemical biosensors.

2. General Principles of Electrochemical Biosensors

Electrochemical biosensors are analytical devices used in various fields, including food safety, environmental monitoring, and medical and industrial diagnostics. These sensors are mainly composed of four elements: (a) a recognition element, whose function is to selectively interact with the target molecule, (b) a transducer, responsible for the conversion of the signal generated after the interaction of the target with the recognition element into a measurable output, (c) an electrolyte, namely the medium in which the electrochemical reaction takes place, and (d) the electrodes, which are divided into reference, counter, and working electrodes and can be made of different materials including carbon, gold, or other conductive materials. To date, there are multiple types of electrochemical biosensors, including amperometric, potentiometric, and impedimetric biosensors. Electrochemical-based signals generate either a measurable current (amperometry), a measurable potential or charge accumulation (potentiometry), a measurable change in impedance (electrochemical impedance spectroscopy, EIS), or measurably alter the conductive properties of a medium (conductometry) between electrodes [70]. A schematic representation of the working principle of an electrochemical biosensor is provided in Figure 1 [71]. Therefore, the aim of this chapter is to elucidate the working principles of the previously mentioned electrochemical analysis techniques.

2.1. Amperometric Biosensors

Amperometry is an electroanalytical technique that involves the application of a constant reducing or oxidizing potential to an indicator working electrode and the subsequent measurement of the generated current. Typically, the magnitude of the measured current is dependent on the concentration of the reduced/oxidized species. In the case of biosensors, the reduction or oxidation of an electroactive species is directly proportional to the concentration of the analyte being investigated [72]. In this scenario, amperometric biosensors are a type of electrochemical biosensor that enables the quantification of an analyte within a sample matrix by transducing surface-level interactions caused by an electroactive species into a current signal [73]. The intrinsic simplicity of the transducer makes these biosensors applicable for low-cost portable devices [74]. An amperometric transducer is used to study the charge transfer at the interfaces of phases, typically between two electrodes separated by an electrolyte. These types of systems are called electrochemical cells and are used to describe the system in which the reaction takes place. One of the half-cells of the electrochemical compartment is heavily controlled to study the charge transfer occurring at the other half-cell, known as the working electrode. Different amperometric methods can be used in electrochemical biosensors, including cyclic voltammetry (CV), differential pulse voltammetry (DPV), or square wave voltammetry (SWV), with the final two being used in most commercial products (such as glucose tests) [73].

2.2. Potentiometric Biosensors

Potentiometry is an electrochemical technique similar to amperometry; however, to conduct potentiometric analysis, the potential difference between a reference electrode and an indicator electrode inside of an electrolyte is measured to extract information about bioreceptor–analyte interactions [75]. This technique has several advantages compared to amperometry as it requires little power and can be compact and portable. In potentiometric biosensors, at zero current, the potential changes are correlated to the changes in concentration of a certain analyte. The electromotive force (EMF), namely a measure of the energy provided by a power source per unit charge and better known as the voltage of a cell, is dependent on the concentration. This relationship may be related using the Nernst Equation (Equation (1)) [76].
E c e l l = E c e l l ° R T z F ln Q
In this equation, Ecell represents the standard electrode potential of the cell, measured in Volt, R is the universal gas constant, T is the temperature in Kelvin, z is the ion charge (moles of electrons), F is Faraday’s constant, and Q is the reaction quotient.

2.3. Coulometric Biosensors

The coulometric analysis uses either an applied current or a potential to exhaustively convert an analyte from one oxidation state to another at the working electrode. This method measures the total current passed, either directly or indirectly, and uses this to determine the number of electrons exchanged through the electrochemical cell. As a result, Faraday’s law may be used to extract the concentration of the analyte based on the number of electrons passed (Equation (2)) [77].
Q = n F N a
where n is the number of electrons per mole of the analyte, F is Faraday’s constant, and Na represents the moles of the analyte. This relationship is valid when 100% of the applied current is used to change the oxidation state of the analyte, which is typically the case as the analyte’s conversion rate when passing through the electrodes is nearly 100%, so slight losses of current are negligible [78]. An advantage of coulometric analysis relative to other techniques is that it lacks the need for calibration curves or chemical standards for the quantification of the analyte, making it an absolute method [78].

2.4. Impedimetric Biosensors

Electrochemical impedance spectroscopy (EIS) is an analytical technique where the impedance or resistance of a system after the target–recognition site interaction is measured in Ohm [79]. EIS offers numerous advantages compared to other electrochemical sensing techniques, as it can utilize small signal analysis by investigating sinusoidal signal relaxations over a wide range of frequencies, from approximately 0.1 Hz to 1 MHz [76]. EIS may be used to analyze charge transfer, mass transfer, and diffusion processes. Furthermore, EIS has the capability of examining intrinsic material properties or processes taking place that will impact the conductance, resistance, or capacitance of a cell [80]. Analyzing impedance differs from resistance, as in a cell with a direct current (DC) applied, resistance obeys Ohm’s law whereas impedance does not. This is due to resistance opposing the flow of both DC and alternating current (AC), whereas impedance solely opposes the flow of AC [81]. By varying the excitation frequency of the applied potential over a range of frequencies, it is possible to calculate the complex impedance, which is the sum of the real impedance (resistance, which measures the opposition to a flowing current) and imaginary impedance (reactance, which measures the opposition to a change in current) of the system as a function of the frequency [70]. Consequently, the analysis of the impedimetric signals generated after the detection at the interface of a selected target can provide information about the presence of a molecule in a sample and lead to quantitative data on the target molecule itself [82].

3. Types of Electrochemical Biosensors Used for Antibiotic Detection

The method of transduction used in an electrochemical biosensor is very dependent on the type of recognition element [83]. The most common biorecognition molecules used are MIPs, antibodies, nucleic acids, aptamers, and enzymes, which are used due to their specific binding capabilities and activity [70]. The aim of this chapter is to elucidate the key biorecognition molecules used in electrochemical biosensors for antibiotic detection, highlighting the positive aspects of each sensing type, such as LoDs and durability. An overview of the biorecognition elements with their main advantages and disadvantages is provided in Table 1 [53].

3.1. Aptamer-Based Electrochemical Sensors

Aptamers are specialized molecules, typically short strands of single-stranded DNA or RNA. Due to their unique three-dimensional structures, aptamers can bind to a wide variety of targets, including ions, small molecules, proteins, and even cells, with high specificity and selectivity [84]. Aptamers are synthesized in vitro through a systematic process known as SELEX (Systematic Evolution of Ligands by EXponential enrichment) [85]. This process involves the repetitive binding, separation, and amplification of oligonucleotides to selectively produce aptamers with a high affinity for specific target molecules. The SELEX process allows for the large-scale production of uniform and high-fidelity aptamers [86]. The synthetic nature of aptamers confers several advantages, including high stability, ease of modification, and the ability to withstand harsh conditions, making them ideal for use in biosensors [87].
An example of aptamer implementation in biosensor technology, particularly in electrochemical sensors, involves the formation of covalent bonds between the aptamer and the sensor’s surface [88]. This covalent attachment ensures a stable and oriented immobilization of the aptamers on the electrode surface, enhancing the sensor’s specificity and sensitivity [89]. The binding event between the aptamer and its target leads to a detectable change in the electrochemical signal, which can be quantitatively measured (Figure 2) [90].
This method capitalizes on the specific interaction between the aptamer and the antibiotic to facilitate accurate and sensitive detection, illustrating the practical application of aptamers in the development of advanced biosensing technologies [92]. An interesting recent example of the utility of aptamers in electrochemical biosensors was published by Bao et. al. They developed a sensor for kanamycin on the surface of an integrated portable plastic gold electrode, as shown in Figure 3, resulting in a sensor with good stability and a low LoD of 0.40 μmol/L, which is highly effective for the detection of kanamycin in environmental water samples [93].
The work published by Malecka-Baturo et. al. is another example illustrating that low LoDs can be achieved with aptamer-based sensors. They utilized a 5′-ferrocene-modified ssDNA aptamer to develop a sensor for tetracycline detection in cow milk samples, as depicted in Figure 4. They achieved a LoD of 0.20 nmol/L with high selectivity for tetracycline over structurally related derivatives [94].
More examples of recently developed technologies based on aptamer biosensors, with relative target molecules, detection methods used, LoDs, and type of sample analyzed, are summarized in Table 2.

3.2. Molecularly Imprinted Polymer-Based Electrochemical Sensors

Molecularly imprinted polymers (MIPs) represent a class of highly specific polymeric materials crafted for recognizing and binding selectively to target molecules. The unique nature of MIPs lies in their synthesis process, during which a target molecule, known as the template, is mixed with functional monomers. These monomers form reversible bonds with the template due to the presence of weak forces such as hydrogen bonding, ionic interactions, π–π interactions, and acid–base interactions [106]. This arrangement is stabilized by a cross-linking agent, leading to the formation of a three-dimensional polymer matrix, typically initiated by a free radical polymerization mechanism. Numerous polymerization techniques can be used such as bulk, precipitation, emulsion, and suspension polymerization, which alter specific elements of the synthesized MIP such as particle size, sensitivity, selectivity, etc. [107]. Post-polymerization, the template is carefully extracted, leaving behind a polymeric structure with cavities containing a specific size, functional groups, and shape to fit a specific template molecule (Figure 5).
The most common application of MIPs is their combination with a portable transducer element. An example is screen-printed electrodes (SPEs), where MIPs synthesized a priori are integrated into the sensor’s surface for specific analyte detection. Nonetheless, it is becoming more and more common to implement the molecularly imprinted polymer technology directly onto the surface of interest, namely on the surface of an electrode or another material that can be used to perform electrochemical measurements. These are known as surface imprinted polymers (SIPs), which are a subclass of MIPs designed to have selective recognition sites on their surface. A key feature of SIPs is that a template molecule is imprinted directly on or near the surface of a polymer matrix, which enhances the accessibility and binding efficiency of the imprinted sites to a specific analyte. Consequently, a higher surface area-to-volume ratio is achieved, leading to improved sensitivity and binding kinetics. To this end, the functionalization is carried out via electropolymerization on the electrode’s surface, resulting in a faster and more reproducible preparation of the sensor itself [109]. Moreover, the direct functionalization of the surface reduces the batch-to-batch variations that can be present in higher amounts if an extra step (i.e., polymer synthesis) must be carried out in advance. It is difficult to categorize SIP-based electrochemical biosensors as a class of their own since they are mainly used for the detection of bigger molecules or full cells. Therefore, they have been implemented in Table 3 and combined with an overview of the recent advances for MIP-based electrochemical biosensors for antibiotic detection.
A representative example MIP-based sensor is the dual recognition sensor for the detection of amoxicillin shown in Figure 6. Li et al. combined electrochemiluminescence with fluorescence to obtain a highly sensitive sensor with a LoD of 8.3 × 10−12 mol/L [110].
A second interesting MIP-based sensor was developed by Thi Vu et al. They were able to make a selective sensor for the detection of norfloxacin in pharmaceutical and aquacultural samples. Through the incorporation of Au nanoparticles into the MIP, as shown in Figure 7, they achieved a LoD of 0.15 ng/mL [111].
Table 3. Molecularly imprinted polymer-based electrochemical biosensors for antibiotic detection.
Table 3. Molecularly imprinted polymer-based electrochemical biosensors for antibiotic detection.
Sensor TypeTarget
Molecule
Method of DetectionLimit of DetectionSampleMonomerRef.
SPE combined with an MIP prepared from dual functional monomers.Erythromycin
Clarithromycin
Azithromycin
DPV1.1–1.6 nMBuffer,
Tap water samples
m-phenylenediamine[112]
Gold screen-printed electrode (Au-SPE) functionalized via electropolymerization of custom-made conjugated monomer (Th2-NDI-PIA)Streptomycin sulfateDPV0.190 pMBuffer, tap waterTh2-NDI-PIA[113]
Electropolymerized MIPs onto a screen-printed carbon electrode (SPCE)AzithromycinDPV0.08 µMWater samples4-aminobenzoic acid[114]
MIP electrodeposited onto the surface of AuNPs/rGo/single-walled carbon nanotube-modified GCEPefloxacinDPV16 nMMilko-phenylenediamine (oPD)[115]
CO2 laser-induced graphene (LIG) with AuNPs and MIPsTetracyclineDPV0.32 nM
0.85 nM
0.80 nM
Buffer,
Milk,
Meat
oPD[116]
Nanocomposite molecularly imprinted polymer (nanoMIP) using oxidised MWNCTs and ultrathin overoxidised polypyrrole MIPSulfamethoxazoleDPV0.41 μMBufferpyrrole[117]
Dual recognition MIP-coated graphene oxide loaded with CdTe quantum dots/AuNPs (GO/CdTe/AuNPs) on an indium tin oxide (ITO) electrodeAmoxicillinDPV
EIS
8.3 pMBufferα-methacrylic acid[110]
Gold nanoparticles (AuNPs) and MIP-based electrochemical sensorsNorfloxacinEIS0.15 ng/mLAquaculture water4-aminothiophenol[111]
Bifunctional dual-template molecularly imprinted polymer-modified electrodeCeftazidime
Avibactam
EIS
SWV
35 μM
0.5 μM
Serum sampleso-PD[118]
Sensor using magnetic nanoparticles (mag) and molecularly imprinted polymerTetracyclineSWV0.15 μMMilkAcrylic acid[119]
Molecularly imprinted electrochemiluminescence (ECL) sensor using amino-functional titanium carbide nanodots (TNDs) and carbon nitride nanosheets (CNNS)CiprofloxacinECL
EIS
1.2 nMFood samples including chicken, milk, and porko-PD[120]
MIP sensor using electro-polymerization with surface-deposited AuNPsSufhaguanidine
Sulfamerazine
CV
DPV
CA
0.030 µMHuman fluidsoPD[121]
Chitosan gold nanoparticle-decorated MIP (Ch-AuMIP) modified GCECiprofloxacinCV0.210 µMWaterMethacrylic acid (MAA)[122]
MIP electrochemical sensor using Fe3N-Co2N nanoarray with high electric conductivity and large surface area for MIP growthAmpicillinCV
EIS
0.365 nMMilk samplesN-N-dimethyl bisacrylamide[123]
MIP coated on graphene oxide deposited as a thin film on GCEAmoxicillinCV
DPV
0.294 nMBufferAPTES + PTES[124]
Fe-doped porous carbon (Fe-PC)-modified Au electrode covered with MIP film electropolymerized onto an electrodeLomefloxacinCV
DPV
EIS
0.2 nMWater sampleso-PD[125]
MIP-based biomimetic layer electrodeposited onto a glassy carbon electrode (GCE)AzithromycinCV
EIS
0.85 nMSpiked plasma, tears, and urine samples3-thienyl boronic acid[126]
Aggregation-based ECL sensor using ferriferous oxide@Pt NPs for signal amplificationCiprofloxacinCV
EIS
0.60 pMMeat samples4-aminothiophenol[127]
Screen-printed electrodeErythromycinCV
EIS
0.1 nMBufferMAA[128]

3.3. Antibody-Based Electrochemical Sensors

Antibody-based electrochemical biosensors, also known as electrochemical immunosensors, leverage the specific affinity of antibodies for antigens to detect and quantify target analytes in a sample [129]. The core functionality of these biosensors involves immobilizing antibodies onto an electrode surface, acting as biorecognition elements [130]. Similarly to MIP sensors, SPEs are an attractive transducer option for immunosensors due to their ability to be mass-produced, their low cost, and their applications in portable detection devices [131]. The binding of these antibodies to their corresponding antigens induces an electrochemical signal, which is measured and correlated with the antigen’s concentration in a sample. Often, labeling components such as nanomaterials are typically employed for the detection of an analyte using immunosensors [132,133]. A common technique is the use of a redox mediator which is catalytically oxidized in the presence of an enzyme substrate [134].
A recent example is the use of concanavalin A (ConA), which showed excellent binding ability with monoclonal antibodies of arsanilic acid through lectin–sugar interactions. The working principle of the sensor is shown in Figure 8. You et al. developed an Ag nanoparticle-reduced graphene oxide nanocomposite which showed a LoD of 0.011 ng/mL [135].
There are numerous types of antibodies used for electrochemical immunosensor applications, including monoclonal, polyclonal, recombinant antibodies, etc. Some examples of their application for the detection of antibiotic residues are reported in Table 4.

4. Commercially Available Sensors for Antibiotic Detection

Detecting antibiotic residues is essential for maintaining food safety, monitoring the environment, and performing clinical diagnostics. Therefore, the main focus of this chapter is to present the technologies that are commercially used to detect the presence of antibiotics in different types of samples, flowing out to a discussion on electrochemical biosensors and electrochemical devices that have been recently introduced on the market. To date, various commercially available sensors have been developed to address this need, offering different mechanisms of detection and varying degrees of sensitivity, specificity, and practicality. In this chapter, we discuss four main categories of these technologies for the detection of antibiotics, namely (a) gold-standard technologies, (b) optical-based sensors, (c) immunoassay-based sensors, and (d) emerging electrochemical-based biosensors and devices.

4.1. Gold-Standard Technologies

A gold standard technology is a technique that has been thoroughly tested and has proven to be highly reliable. Because of the need to use effective methods for the detection of antibiotics in various samples, traditional methods such as high-performance liquid chromatography (HPLC), mass spectrometry (MS), microbiological assays, capillary electrophoresis (CE), and ultraviolet–visible (UV–Vis) spectroscopy are highly regarded and used thanks to their accuracy, sensitivity, and reliability. High-performance liquid chromatography (HPLC) combined with mass spectrometry (MS) is known for its precision in antibiotic detection. Companies such as Agilent Technologies, Santa Clara, CA, USA and Thermo Fisher Scientific, Waltham, MA, USA produce advanced HPLC-MS systems. For example, Agilent’s 1290 Infinity II LC system offers excellent resolution and throughput, while Thermo Fisher’s TSQ Quantis MS provides accurate quantification of antibiotics at very low levels (parts per billion) [146,147]. On the other hand, microbiological assays are another important technology, valued for their ability to detect a wide range of antibiotics through microbial growth inhibition. Neogen Corporation, Lansing, MI, USA and Charm Sciences, Lawrence, MA, USA are key providers in this area. Neogen’s Soleris Next Generation (NG) system and Charm Sciences’ ROSA (Rapid One Step Assay) platforms are extensively used in the food and dairy industries, delivering results within hours and detecting antibiotic residues at levels of low parts per billion [148,149]. Furthermore, capillary electrophoresis (CE) is another reliable method for analyzing antibiotics. In this scenario, the company Sciex, Framingham, MA, USA introduced the PA 800 Plus Pharmaceutical Analysis System, which achieves high-resolution separation and quantification of antibiotics in various samples, with detection limits in the low nanogram per milliliter (ng mL−1) range [150]. Ultraviolet–visible (UV–Vis) spectroscopy is another conventional technique used for antibiotic detection. As an example, Shimadzu’s UV-1900i UV–Vis Spectrophotometer (Kyoto, Japan) is widely used due to its accuracy in detecting various antibiotic compounds, with detection limits typically in the microgram per liter (μg L−1) range, which is slightly higher compared to other devices, although it offers a high degree of reliability [151]. In the last decades, many companies have introduced important technologies and products to the market with the aim of detecting the presence of antibiotics in a variety of samples. Some examples with their related working parameters are summarized in Table 5.
These traditional methods, supported by products from leading companies, provide highly sensitive, specific, and reproducible results, making them essential tools for monitoring and detecting antibiotic residues across a range of samples. Nonetheless, despite their impressive capabilities, the cost and portability of these technologies can pose significant challenges, particularly in resource-limited settings where access to sophisticated equipment and funding is constrained. This financial barrier and the often bulky nature of these devices limit their widespread application and practicality in such areas. As a result, researchers and engineers are actively exploring novel materials, design methodologies, and fabrication techniques to create next-generation technologies that are not only cost-effective but also highly portable. This ensures that these critical diagnostic tools can be installed more broadly, meeting the diverse needs of various applications and geographic regions, thus bridging the gap in accessibility and effectiveness.

4.2. Optical-Based Sensors

Optical-based sensors utilize the interaction of light with a sample to detect the presence of antibiotics. These sensors often are based on principles such as fluorescence, absorbance, and surface plasmon resonance (SPR). Commercially available optical sensors are valued for their high sensitivity and specificity. In fact, the market for optical-based sensors has grown significantly, providing rapid and sensitive detection methods. Several companies lead this field with innovative products, each offering unique performance metrics and catering to specific industry needs. As an example, Charm Sciences Inc. stands out with its Charm II System, which is particularly effective in the dairy industry [158]. This system can detect a wide range of antibiotic residues in milk and dairy products within eight minutes. With a limit of detection (LoD) of 1ppb for several antibiotics, the Charm II System combines speed and sensitivity, making it a valuable tool in ensuring food safety. Furthermore, the company R-Biopharm AG, Pfungstadt, Germany, offers the RIDAQUICK SULFONAMIDES test kit, which provides qualitative analysis of sulphonamide residues in meat and other food matrices [159]. The test delivers results in ten minutes with a LoD of 10 ppb, making it a quick and reliable choice for food safety testing. Another example of commercially available products based on optical biosensors is given by the company Neogen Corporation, where the product BetaStar Advanced is designed primarily for the dairy industry, offering results in just five minutes [160]. This product is selective for the detection of beta-lactam antibiotics with a LoD of 2 ppb. Thermo Fisher Scientific’s QExactive Orbitrap, although primarily a mass spectrometer, includes optical detection capabilities [161]. This versatile instrument can detect antibiotic residues in food and environmental samples with a LoD ranging from low ppb to ppt [162]. Besides being an expensive alternative, the results are available in less than an hour, making it suitable for high-throughput testing. Similarly, IDEXX Laboratories’ SNAP Tetracycline Test (Westbrook, ME, USA) targets milk testing, detecting tetracycline residues within ten minutes with a LoD of 4 ppb [163].
Overall, the market for optical-based sensors for detecting antibiotic residues is diverse, with leading companies offering products that vary in performance metrics, including detection limits, time to results, and pricing. These sensors provide essential tools for ensuring food and environmental safety, supporting the specific needs of various industries.

4.3. Immunoassay-Based Sensors

Immunoassay-based sensors are highly valued for their exceptional specificity and sensitivity in detecting antibiotic residues in a variety of sample types, including food, water, and biological tissues. These sensors utilize the precise binding interaction between antigens and antibodies to deliver quick and dependable results. By exploiting this high specificity, immunoassay-based sensors can accurately identify and quantify antibiotic residues in complex samples, ensuring both rapid detection and reliable performance. Several companies have developed innovative immunoassay-based products, each tailored to meet the specific needs of different industries. Therefore, this paragraph explores key players in this field and their notable products, discussing performance metrics such as time required for measurement, limit of detection (LoD), and pricing. R-Biopharm AG is a prominent competitor in the immunoassay market, offering products such as the RIDASCREEN Chloramphenicol [164]. This ELISA-based sensor is designed to detect chloramphenicol residues in various food products, providing results within two hours. With a limit of detection of 0.05 ppb, the RIDASCREEN Chloramphenicol offers high sensitivity, ensuring the reliable detection of this antibiotic. Furthermore, the company Perkin Elmer, Waltham, MA, USA offers the MaxSignal Beta-lactam ELISA Test Kit, designed for detecting beta-lactam antibiotics in food samples [165]. This immunoassay-based sensor provides results within two hours and presents a LoD of 1 ppb. Another example is given by the company Enzo Life Sciences, Farmingdale, NY, USA, which provides the Amp’d ELISA Kit for the detection of aminoglycoside residues in food and animal feed [166]. This immunoassay-based sensor offers results in under two hours with a LoD of 0.1 ppb, ensuring high sensitivity and specificity, making it a valuable tool for ensuring compliance with regulatory standards. Together with the aforementioned examples, other products based on similar technology are also present on the market. Some examples of the most significant ones are reported in Table 6.
In summary, immunoassay-based sensors are essential tools in the detection of antibiotic residues, offering high specificity and sensitivity across various sample types. Some leading companies have developed robust products that satisfy the needs of different industries. These sensors vary in their performance metrics, including detection limits, time to results, and pricing, providing versatile and effective solutions for ensuring public health and safety. Nonetheless, the cost of these technologies and their portability can still be considered a drawback, especially if the application needs to be performed in resource-limited areas. Therefore, both academia and industry are continuously researching novel alternatives to ensure that the technologies can service all needs.

4.4. Electrochemical-Based Biosensors

Electrochemical-based biosensors are a prominent category among commercially available sensors for antibiotic detection. These devices are characterized by cost-effectiveness, portability, and good precision, thus offering a robust alternative to gold-standard technologies or conventional detection methods, which, despite their accuracy, can have drawbacks such as high operational costs and complex preparatory requirements. For example, sensors developed by companies such as Randox Food Diagnostics, Crumlin, UK provide rapid and reliable detection of various antibiotics, helping to maintain food quality and safety [175]. Environmental monitoring also benefits from these biosensors, as they are used to detect antibiotic contamination in water sources. Moreover, companies such as Bio-Rad Laboratories, Hercules, Ca, USA, Thermo Fisher Scientific, and Clinical Diagnostics (Bangkok, Thailand) utilize electrochemical-based sensing detectors to monitor antibiotic levels in biological samples, assisting in effective treatment management and preventing antibiotic resistance. However, there are other examples of spin-off companies and products developed for the same purpose (Table 7).
In addition, there are some examples of commercially available products for the specific detection of various types of antibiotics. The devices listed in Table 7 are not only particular tools that can be used on-site for sensing purposes but also more general apparatus that can be applied to the detection of antibiotic residues. In particular, the Dropsens Screen-Printed Electrodes are examples of electrochemical devices that can be functionalized differently based on the needs of the user and then coupled with specific appliances such as a potentiostat to be able to obtain the desired measurements (i.e., cyclic voltammetry, differential pulse voltammetry, etc.).
Overall, the introduction of electrochemical-based devices is becoming increasingly prevalent in the sensing field thanks to their versatility and ease of use. These devices offer a wide range of applications, from environmental monitoring to medical diagnostics, providing accurate and reliable measurements. Nonetheless, besides being cheaper compared to almost all the gold-standard technologies, some options can still be considered expensive, posing a barrier to widespread adoption. However, developing highly portable technologies, such as the Sensit Smart from PalmSens, Utrecht, The Netherlands, is opening the way toward more accessible and mobile sensing applications. This portability allows for on-site testing and real-time data collection, significantly enhancing the practicality and efficiency of sensing processes across various fields.

5. General Conclusion and Future Perspectives

Looking forward, the integration of electrochemical-based devices into the sensing field marks a significant technological advancement, driven by their versatility and user-friendly nature. These devices, renowned for their wide array of applications, have become valid alternatives in domains ranging from environmental monitoring to medical diagnostics. They provide accurate and reliable measurements that are crucial for making informed decisions and ensuring public safety, overcoming some of the main issues related to the gold-standard technologies already present on the market. In fact, the key advantage of electrochemical sensors lies in their ability to deliver precise data quickly and efficiently. For instance, in environmental monitoring, these devices can detect pollutants at very low concentrations, enabling timely interventions to prevent ecological damage. In medical diagnostics, electrochemical sensors are not only used for the detection of antibiotics but also to monitor glucose levels, detect biomarkers for various diseases, and even assess the efficacy of treatments in real-time. This capability is particularly beneficial for patients who require constant monitoring, as it allows for better management of their conditions and reduces the need for frequent hospital visits. When comparing different kinds of electrochemical biosensors, it is important to consider their sensitivity, cost, and practical applications. For example, amperometric sensors are known for their high sensitivity and fast response times, making them ideal for real-time monitoring. However, they may be more expensive to produce due to the need for specialized electrodes. Potentiometric sensors, on the other hand, are generally more cost-effective but might offer lower sensitivity in comparison. Conductometric sensors, which measure changes in electrical conductivity, provide a balance between cost and sensitivity but may require more complex calibration procedures. Despite their many benefits, the adoption and marketing of electrochemical-based devices is not without challenges. While these sensors are generally more affordable than many traditional gold-standard technologies, the initial cost can still be prohibitive for some users. Furthermore, the ongoing advancements in sensor technology are likely to drive down costs and improve accessibility even further. Innovations in materials science, such as the use of more affordable and sustainable materials for sensor construction, are expected to make these devices even more economical. Additionally, improvements in manufacturing processes can enhance the scalability of sensor production, leading to lower prices and wider availability.
The high cost can limit the widespread adoption of these sensors, particularly in low-resource settings where budget constraints are a significant concern. Therefore, while the cost-effectiveness of electrochemical sensors is an advantage, further efforts are needed to reduce costs and make these technologies more reliable and accessible to a broader audience.

Author Contributions

Conceptualization, M.F. and M.P.K.; methodology, G.H.; investigation, M.F. and M.P.K.; resources, M.F., M.P.K. and G.H.; writing—original draft preparation, M.P.K., M.F. and G.H.; writing—review and editing, M.F. and G.H.; supervision, J.W.L. and K.E.; project administration, K.E. and H.D.; funding acquisition, T.J.C. and B.v.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the European Regional Development Fund through the AgrEU food project, funded by the Interreg VA Flanders-The Netherlands program, CCI grant no. 2014TC16RFCB046.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Balloux, F.; van Dorp, L. Q&A: What Are Pathogens, and What Have They Done to and for Us? BMC Biol. 2017, 15, 91. [Google Scholar] [CrossRef]
  2. Antimicrobial Resistance|Antimicrobial Resistance|CDC. Available online: https://www.cdc.gov/antimicrobial-resistance/index.html (accessed on 2 August 2024).
  3. Urban-Chmiel, R.; Marek, A.; Stępień-Pyśniak, D.; Wieczorek, K.; Dec, M.; Nowaczek, A.; Osek, J. Antibiotic Resistance in Bacteria—A Review. Antibiotics 2022, 11, 1079. [Google Scholar] [CrossRef] [PubMed]
  4. Patil, S.M.; Patel, P. Bactericidal and Bacteriostatic Antibiotics. Infect. Sepsis Dev. 2021, 3, 773–774. [Google Scholar] [CrossRef]
  5. WHO EMRO|What Is the Difference between Antibiotic and Antimicrobial Resistance?|Antimicrobial Resistance|Health Topics. Available online: https://www.emro.who.int/health-topics/drug-resistance/what-is-the-difference-between-antibiotic-and-antimicrobial-resistance.html (accessed on 2 August 2024).
  6. 2019 Antibiotic Resistance Threats Report|Antimicrobial Resistance|CDC. Available online: https://www.cdc.gov/antimicrobial-resistance/data-research/threats/index.html (accessed on 2 August 2024).
  7. Murray, C.J.; Ikuta, K.S.; Sharara, F.; Swetschinski, L.; Robles Aguilar, G.; Gray, A.; Han, C.; Bisignano, C.; Rao, P.; Wool, E.; et al. Global Burden of Bacterial Antimicrobial Resistance in 2019: A Systematic Analysis. Lancet 2022, 399, 629–655. [Google Scholar] [CrossRef] [PubMed]
  8. Denissen, J.; Reyneke, B.; Waso-Reyneke, M.; Havenga, B.; Barnard, T.; Khan, S.; Khan, W. Prevalence of ESKAPE Pathogens in the Environment: Antibiotic Resistance Status, Community-Acquired Infection and Risk to Human Health. Int. J. Hyg. Environ. Health 2022, 244, 114006. [Google Scholar] [CrossRef]
  9. Santajit, S.; Indrawattana, N. Mechanisms of Antimicrobial Resistance in ESKAPE Pathogens. Biomed Res. Int. 2016, 2016, 2475067. [Google Scholar] [CrossRef]
  10. Michaelis, C.; Grohmann, E. Horizontal Gene Transfer of Antibiotic Resistance Genes in Biofilms. Antibiotics 2023, 12, 328. [Google Scholar] [CrossRef]
  11. Burmeister, A.R. Horizontal Gene Transfer. Evol. Med. Public Heal. 2015, 2015, 193. [Google Scholar] [CrossRef]
  12. Larsson, D.G.J.; Flach, C.F. Antibiotic Resistance in the Environment. Nat. Rev. Microbiol. 2021, 20, 257–269. [Google Scholar] [CrossRef]
  13. Jian, Z.; Zeng, L.; Xu, T.; Sun, S.; Yan, S.; Yang, L.; Huang, Y.; Jia, J.; Dou, T. Antibiotic Resistance Genes in Bacteria: Occurrence, Spread, and Control. J. Basic Microbiol. 2021, 61, 1049–1070. [Google Scholar] [CrossRef]
  14. Medina, E.; Pieper, D.H. Tackling Threats and Future Problems of Multidrug-Resistant Bacteria. Curr. Top. Microbiol. Immunol. 2016, 398, 3–33. [Google Scholar] [CrossRef] [PubMed]
  15. Woodford, N.; Ellington, M.J. The Emergence of Antibiotic Resistance by Mutation. Clin. Microbiol. Infect. 2007, 13, 5–18. [Google Scholar] [CrossRef]
  16. Reygaert, W.C. An Overview of the Antimicrobial Resistance Mechanisms of Bacteria. AIMS Microbiol. 2018, 4, 482. [Google Scholar] [CrossRef] [PubMed]
  17. Ventola, C.L. The Antibiotic Resistance Crisis: Part 1: Causes and Threats. Pharm. Ther. 2015, 40, 277. [Google Scholar]
  18. Hu, M.; Ben, Y.; Wong, M.H.; Zheng, C. Trace Analysis of Multiclass Antibiotics in Food Products by Liquid Chromatography-Tandem Mass Spectrometry: Method Development. J. Agric. Food Chem. 2021, 69, 1656–1666. [Google Scholar] [CrossRef]
  19. Ben, Y.; Hu, M.; Zhang, X.; Wu, S.; Wong, M.H.; Wang, M.; Andrews, C.B.; Zheng, C. Efficient Detection and Assessment of Human Exposure to Trace Antibiotic Residues in Drinking Water. Water Res. 2020, 175, 115699. [Google Scholar] [CrossRef]
  20. Bartlett, J.G.; Gilbert, D.N.; Spellberg, B. Seven Ways to Preserve the Miracle of Antibiotics. Clin. Infect. Dis. 2013, 56, 1445–1450. [Google Scholar] [CrossRef]
  21. Blaser, M.J.; Melby, M.K.; Lock, M.; Nichter, M. Accounting for Variation in and Overuse of Antibiotics among Humans. BioEssays 2021, 43, 2000163. [Google Scholar] [CrossRef]
  22. Pitt, S.J.; Gunn, A. The One Health Concept. Br. J. Biomed. Sci. 2024, 81, 12366. [Google Scholar] [CrossRef]
  23. Gorham, J.; Taccone, F.S.; Hites, M. Ensuring Target Concentrations of Antibiotics in Critically Ill Patients through Dose Adjustment. Expert Opin. Drug Metab. Toxicol. 2022, 18, 177–187. [Google Scholar] [CrossRef]
  24. Miethke, M.; Pieroni, M.; Weber, T.; Brönstrup, M.; Hammann, P.; Halby, L.; Arimondo, P.B.; Glaser, P.; Aigle, B.; Bode, H.B.; et al. Towards the Sustainable Discovery and Development of New Antibiotics. Nat. Rev. Chem. 2021, 5, 726–749. [Google Scholar] [CrossRef]
  25. Bissantz, C.; Zampaloni, C.; David-Pierson, P.; Dieppois, G.; Guenther, A.; Trauner, A.; Winther, L.; Stubbings, W. Translational PK/PD for the Development of Novel Antibiotics—A Drug Developer’s Perspective. Antibiotics 2024, 13, 72. [Google Scholar] [CrossRef] [PubMed]
  26. Craig, W.A. Pharmacokinetic/Pharmacodynamic Parameters: Rationale for Antibacterial Dosing of Mice and Men. Clin. Infect. Dis. 1998, 26, 1–12. [Google Scholar] [CrossRef] [PubMed]
  27. Garzón, V.; Bustos, R.H.; Pinacho, D.G. Personalized Medicine for Antibiotics: The Role of Nanobiosensors in Therapeutic Drug Monitoring. J. Pers. Med. 2020, 10, 147. [Google Scholar] [CrossRef] [PubMed]
  28. World Health Organization (WHO). Request for Advice from the Expert Committee on Selection and Use of Essential Medicines on Prioritization of Medicines Requiring Therapeutic Drug Monitoring; WHO: Geneva, Switzerland, 2019. [Google Scholar]
  29. Huang, J.-W.; Zhong, X.-F.; Gao, Y.-Z. New Antibiotic against Multi-Drug Resistant Bacteria. Innov. Life 2024, 2, 100057. [Google Scholar] [CrossRef]
  30. Zaman, S.B.; Hussain, M.A.; Nye, R.; Mehta, V.; Mamun, K.T.; Hossain, N. A Review on Antibiotic Resistance: Alarm Bells Are Ringing. Cureus 2017, 9, e1403. [Google Scholar] [CrossRef]
  31. Sachi, S.; Ferdous, J.; Sikder, M.H.; Azizul Karim Hussani, S.M. Antibiotic Residues in Milk: Past, Present, and Future. J. Adv. Vet. Anim. Res. 2019, 6, 315–332. [Google Scholar] [CrossRef]
  32. Pauter, K.; Szultka-Młýnska, M.; Buszewski, B. Determination and Identification of Antibiotic Drugs and Bacterial Strains in Biological Samples. Molecules 2020, 25, 2556. [Google Scholar] [CrossRef]
  33. Hagel, R.B.; Waysek, E.H.; Cort, W.M. High-Pressure Liquid Chromatography Analysis of Antibiotic Susceptibility Disks. Antimicrob. Agents Chemother. 1979, 16, 372. [Google Scholar] [CrossRef]
  34. Lin, C.C.; Wu, J.Y.; Huang, P.Y.; Sung, H.L.; Tung, Y.C.; Lai, C.C.; Wei, Y.F.; Fu, P.K. Comparing Prolonged Infusion to Intermittent Infusion Strategies for Beta-Lactam Antibiotics in Patients with Gram-Negative Bacterial Infections: A Systematic Review and Meta-Analysis. Expert Rev. Anti. Infect. Ther. 2024, 1–11. [Google Scholar] [CrossRef]
  35. Bush, K.; Bradford, P.A. β-Lactams and β-Lactamase Inhibitors: An Overview. Cold Spring Harb. Perspect. Med. 2016, 6, a025247. [Google Scholar] [CrossRef]
  36. Riezk, A.; Wilson, R.C.; Rawson, T.M.; Vasikasin, V.; Arkel, P.; Ferris, T.J.; Haigh, L.D.; Cass, A.E.G.; Holmes, A.H. A Rapid, Simple, High-Performance Liquid Chromatography Method for the Clinical Measurement of Beta-Lactam Antibiotics in Serum and Interstitial Fluid. Anal. Methods 2023, 15, 829–836. [Google Scholar] [CrossRef] [PubMed]
  37. Peng, J.; Wang, Y.; Liu, L.; Kuang, H.; Li, A.; Xu, C. Multiplex Lateral Flow Immunoassay for Five Antibiotics Detection Based on Gold Nanoparticle Aggregations. RSC Adv. 2016, 6, 7798–7805. [Google Scholar] [CrossRef]
  38. Frigoli, M.; Lowdon, J.W.; Caldara, M.; Cleij, T.J.; Diliën, H.; Eersels, K.; van Grinsven, B. Emerging Biomimetic Sensor Technologies for the Detection of Pathogenic Bacteria: A Commercial Viability Study. ACS Omega 2024, 9, 23155–23171. [Google Scholar] [CrossRef]
  39. Haiping, L.; Jiangyue, W.; Fanping, M.; Aifeng, L. Immunochromatographic Assay for the Detection of Antibiotics in Animal-Derived Foods: A Review. Food Control. 2021, 130, 108356. [Google Scholar] [CrossRef]
  40. Zhao, C.; Pan, B.; Wang, M.; Si, Y.; Taha, A.Y.; Liu, G.; Pan, T.; Sun, G. Improving the Sensitivity of Nanofibrous Membrane-Based ELISA for On-Site Antibiotics Detection. ACS Sens. 2022, 7, 1458–1466. [Google Scholar] [CrossRef] [PubMed]
  41. Mohammadzadeh, M.; Montaseri, M.; Hosseinzadeh, S.; Majlesi, M.; Berizi, E.; Zare, M.; Derakhshan, Z.; Ferrante, M.; Conti, G.O. Antibiotic Residues in Poultry Tissues in Iran: A Systematic Review and Meta-Analysis. Environ. Res. 2022, 204, 112038. [Google Scholar] [CrossRef]
  42. Aslam, S.; Zhang, Z.; Muhammad, A. A Review on Validation of Enzyme Linked Immunosorbent Assay (Elisa) Techniques for Detection and Quantification of Different Contaminant in Aquatic Environment. J. Bioresour. Manag. 2023, 10, 6. [Google Scholar]
  43. Ahmed, S.; Ning, J.; Peng, D.; Chen, T.; Ahmad, I.; Ali, A.; Lei, Z.; Abu bakr Shabbir, M.; Cheng, G.; Yuan, Z. Current Advances in Immunoassays for the Detection of Antibiotics Residues: A Review. Food Agric. Immunol. 2020, 31, 268–290. [Google Scholar] [CrossRef]
  44. Shanin, I.A.; Shaimardanov, A.R.; Thai, N.T.D.; Eremin, S.A. Determination of Fluoroquinolone Antibiotic Levofloxacin in Urine by Fluorescence Polarization Immunoassay. J. Anal. Chem. 2015, 70, 712–717. [Google Scholar] [CrossRef]
  45. Pastor-Navarro, N.; Gallego-Iglesias, E.; Maquieira, Á.; Puchades, R. Immunochemical Method for Sulfasalazine Determination in Human Plasma. Anal. Chim. Acta 2007, 583, 377–383. [Google Scholar] [CrossRef]
  46. Gaspar, V.P.; Ibrahim, S.; Zahedi, R.P.; Borchers, C.H. Utility, Promise, and Limitations of Liquid Chromatography-mass Spectrometry-based Therapeutic Drug Monitoring in Precision Medicine. J. Mass Spectrom. 2021, 56, e4788. [Google Scholar] [CrossRef] [PubMed]
  47. Chan, C.L.; Wai, H.K.F.; Wu, P.; Lai, S.W.; Chan, O.S.K.; Tun, H.M. A Universal LC-MS/MS Method for Simultaneous Detection of Antibiotic Residues in Animal and Environmental Samples. Antibiotics 2022, 11, 845. [Google Scholar] [CrossRef] [PubMed]
  48. Budd, J.; Miller, B.S.; Weckman, N.E.; Cherkaoui, D.; Huang, D.; Decruz, A.T.; Fongwen, N.; Han, G.-R.; Broto, M.; Estcourt, C.S.; et al. Lateral Flow Test Engineering and Lessons Learned from COVID-19. Nat. Rev. Bioeng. 2023, 1, 13–31. [Google Scholar] [CrossRef]
  49. Haleem, A.; Javaid, M.; Singh, R.P.; Suman, R.; Rab, S. Biosensors Applications in Medical Field: A Brief Review. Sens. Int. 2021, 2, 100100. [Google Scholar] [CrossRef]
  50. Bahadir, E.B.; Sezgintürk, M.K. Applications of Commercial Biosensors in Clinical, Food, Environmental, and Biothreat/Biowarfare Analyses. Anal. Biochem. 2015, 478, 107–120. [Google Scholar] [CrossRef] [PubMed]
  51. Mehrotra, P. Biosensors and Their Applications—A Review. J. Oral Biol. Craniofacial Res. 2016, 6, 153–159. [Google Scholar] [CrossRef]
  52. Frigoli, M.; Lowdon, J.W.; Caldara, M.; Arreguin-Campos, R.; Sewall, J.; Cleij, T.J.; Diliën, H.; Eersels, K.; van Grinsven, B. Thermal Pyocyanin Sensor Based on Molecularly Imprinted Polymers for the Indirect Detection of Pseudomonas Aeruginosa. ACS Sens. 2023, 8, 353–362. [Google Scholar] [CrossRef]
  53. Morales, M.A.; Halpern, J.M. Guide to Selecting a Biorecognition Element for Biosensors. Bioconjug. Chem. 2018, 29, 3231. [Google Scholar] [CrossRef]
  54. Fatima, T.; Bansal, S.; Husain, S.; Khanuja, M. Biosensors. In Electrochemical Sensors: From Working Electrodes to Functionalization Miniaturized Devices; Woodhead Publishing: Cambridge, UK, 2022; pp. 1–30. [Google Scholar] [CrossRef]
  55. Bhalla, N.; Jolly, P.; Formisano, N.; Estrela, P. Introduction to Biosensors. Essays Biochem. 2016, 60, 1–8. [Google Scholar] [CrossRef]
  56. Lu, L.; Hu, X.; Zhu, Z. Biomimetic Sensors and Biosensors for Qualitative and Quantitative Analyses of Five Basic Tastes. TrAC Trends Anal. Chem. 2017, 87, 58–70. [Google Scholar] [CrossRef]
  57. Singh, A.K.; Mittal, S.; Das, M.; Saharia, M. Optical biosensors: A decade in review. Alexandria. 2023, 67, 673–691. [Google Scholar] [CrossRef]
  58. Qian, W.; Zhou, J.; Chen, Y.; Liu, H.; Ding, P. Label-free electrochemical immunosensor based on staphylococcal protein a and AgNPs-rGO-Nf for sensitive detection of virginiamycin M1. Bioelectrochemistry 2023, 153, 108489. [Google Scholar] [CrossRef] [PubMed]
  59. Peltomaa, R.; Glahn-Martinez, B.; Benito-Peña, E.; Moreno-Bondi, M. COptical Biosensors for Label-Free Detection of Small Molecules. Sensors. 2018, 18, 4126. [Google Scholar] [CrossRef] [PubMed]
  60. Price, C.P. Point-of-Care Testing in Diabetes Mellitus. Clin. Chem. Lab. Med. 2003, 41, 1213–1219. [Google Scholar] [CrossRef] [PubMed]
  61. D’Costa, E.J.; Higgins, I.J.; Turner, A.P.F. Quinoprotein Glucose Dehydrogenase and Its Application in an Amperometric Glucose Sensor. Biosensors 1986, 2, 71–87. [Google Scholar] [CrossRef]
  62. Kabashin, A.V.; Kravets, V.G.; Grigorenko, A.N. Label-Free Optical Biosensing: Going beyond the Limits. Chem. Soc. Rev. 2023, 52, 6554–6585. [Google Scholar] [CrossRef]
  63. Ravindran, N.; Kumar, S.; M, Y.; S, R.; CA, M.; Thirunavookarasu, S.N.; CK, S. Recent Advances in Surface Plasmon Resonance (SPR) Biosensors for Food Analysis: A Review. Crit. Rev. Food Sci. Nutr. 2023, 63, 1055–1077. [Google Scholar] [CrossRef]
  64. Sharma, K.; Sharma, M. Optical Biosensors for Environmental Monitoring: Recent Advances and Future Perspectives in Bacterial Detection. Environ. Res. 2023, 236, 116826. [Google Scholar] [CrossRef]
  65. Herrera-Domínguez, M.; Morales-Luna, G.; Mahlknecht, J.; Cheng, Q.; Aguilar-Hernández, I.; Ornelas-Soto, N. Optical Biosensors and Their Applications for the Detection of Water Pollutants. Biosensors 2023, 13, 370. [Google Scholar] [CrossRef]
  66. Rojin, S.; Pour, S.; Calabria, D.; Emamiamin, A.; Lazzarini, E.; Pace, A.; Guardigli, M.; Zangheri, M.; Mirasoli, M. Electrochemical vs. Optical Biosensors for Point-of-Care Applications: A Critical Review. Chemosens 2023, 11, 546. [Google Scholar] [CrossRef]
  67. Wang, Q.; Xue, Q.; Chen, T.; Li, J.; Liu, Y.; Shan, X.; Liu, F.; Jia, J. Recent Advances in Electrochemical Sensors for Antibiotics and Their Applications. Chin. Chem. Lett. 2021, 32, 609–619. [Google Scholar] [CrossRef]
  68. Singh, A.K.; Bhat, A.; Dutta, K.M. Electrochemical Biosensors for the Detection of Antibiotics in Milk: Recent Trends and Future Perspectives. Biosens. 2023, 13, 867. [Google Scholar] [CrossRef] [PubMed]
  69. Luo, Y.; Sun, Y.; Wei, X.; He, Y.; Wang, H.; Cui, Z.; Ma, J.; Liu, X.; Shu, R.; Lin, H.; et al. Detection Methods for Antibiotics in Wastewater: A Review. Bioprocess Biosyst. Eng. 2024, 47, 1433–1451. [Google Scholar] [CrossRef] [PubMed]
  70. Grieshaber, D.; MacKenzie, R.; Vörös, J.; Reimhult, E. Electrochemical Biosensors—Sensor Principles and Architectures. Sensors 2008, 8, 1400. [Google Scholar] [CrossRef]
  71. Lopes, L.C.; Santos, A.; Bueno, P.R. An Outlook on Electrochemical Approaches for Molecular Diagnostics Assays and Discussions on the Limitations of Miniaturized Technologies for Point-of-Care Devices. Sens. Actuators Rep. 2022, 4, 100087. [Google Scholar] [CrossRef]
  72. Adeloju, S.B. Amperometry. Encycl. Anal. Sci. 2005, 70–79. [Google Scholar] [CrossRef]
  73. Hammond, J.L.; Formisano, N.; Estrela, P.; Carrara, S.; Tkac, J. Electrochemical Biosensors and Nanobiosensors. Essays Biochem. 2016, 60, 69. [Google Scholar] [CrossRef]
  74. Pedersen, T.; Fojan, P.; Pedersen, A.K.N.; Magnusson, N.E.; Gurevich, L. Amperometric Biosensor for Quantitative Measurement Using Sandwich Immunoassays. Biosensors 2023, 13, 519. [Google Scholar] [CrossRef]
  75. Walker, N.L.; Roshkolaeva, A.B.; Chapoval, A.I.; Dick, J.E. Recent Advances in Potentiometric Biosensing. Curr. Opin. Electrochem. 2021, 28, 100735. [Google Scholar] [CrossRef]
  76. Magar, H.S.; Hassan, R.Y.A.; Mulchandani, A. Electrochemical Impedance Spectroscopy (EIS): Principles, Construction, and Biosensing Applications. Sensors 2021, 21, 6578. [Google Scholar] [CrossRef] [PubMed]
  77. Ziyatdinova, G.; Budnikov, H.; Analysis Chemosensors, A.; Rodriguez-Mendez, L. Analytical Capabilities of Coulometric Sensor Systems in the Antioxidants Analysis. Chemosensors 2021, 9, 91. [Google Scholar] [CrossRef]
  78. Razem, M.; Ding, Y.; Morozova, K.; Mazzetto, F.; Scampicchio, M. Analysis of Phenolic Compounds in Food by Coulometric Array Detector: A Review. Sensors 2022, 22, 7498. [Google Scholar] [CrossRef]
  79. Basics of EIS: Electrochemical Research-Impedance Gamry Instruments. Available online: https://www.gamry.com/application-notes/EIS/basics-of-electrochemical-impedance-spectroscopy/ (accessed on 2 August 2024).
  80. Lazanas, A.C.; Prodromidis, M.I. Electrochemical Impedance Spectroscopy─A Tutorial. ACS Meas. Sci. Au 2023, 3, 162–193. [Google Scholar] [CrossRef]
  81. Macdonald, J.R. Impedance Spectroscopy and Its Use in Analyzing the Steady-State AC Response of Solid and Liquid Electrolytes. J. Electroanal. Chem. Interfacial Electrochem. 1987, 223, 25–50. [Google Scholar] [CrossRef]
  82. Caldara, M.; Lowdon, J.W.; van Wissen, G.; Ferrari, A.G.M.; Crapnell, R.D.; Cleij, T.J.; Diliën, H.; Banks, C.E.; Eersels, K.; van Grinsven, B. Dipstick Sensor Based on Molecularly Imprinted Polymer-Coated Screen-Printed Electrodes for the Single-Shot Detection of Glucose in Urine Samples—From Fundamental Study toward Point-of-Care Application. Adv. Mater. Interfaces 2023, 10, 2300182. [Google Scholar] [CrossRef]
  83. Cho, I.H.; Kim, D.H.; Park, S. Electrochemical Biosensors: Perspective on Functional Nanomaterials for on-Site Analysis. Biomater. Res. 2020, 24, 6. [Google Scholar] [CrossRef]
  84. Toh, S.Y.; Citartan, M.; Gopinath, S.C.B.; Tang, T.H. Aptamers as a Replacement for Antibodies in Enzyme-Linked Immunosorbent Assay. Biosens. Bioelectron. 2015, 64, 392–403. [Google Scholar] [CrossRef]
  85. Liu, Q.; Zhang, W.; Chen, S.; Zhuang, Z.; Zhang, Y.; Jiang, L.; Lin, J.S. SELEX Tool: A Novel and Convenient Gel-Based Diffusion Method for Monitoring of Aptamer-Target Binding. J. Biol. Eng. 2020, 14, 1. [Google Scholar] [CrossRef]
  86. Lee, E.S.; Kim, E.J.; Park, T.K.; Bae, D.W.; Cha, S.S.; Kim, T.W.; Kim, Y.P. Gold Nanoparticle-Assisted SELEX as a Visual Monitoring Platform for the Development of Small Molecule-Binding DNA Aptasensors. Biosens. Bioelectron. 2021, 191, 113468. [Google Scholar] [CrossRef]
  87. Chrouda, A.; Ayed, D.; Zinoubi, K.; Majdoub, H.; Jaffrezic-Renault, N. Highly Stable and Ultra-Sensitive Amperometric Aptasensor Based on Pectin Stabilized Gold Nanoparticles on Graphene Oxide Modified GCE for the Detection of Aflatoxin M1. Food Chem. Adv. 2022, 1, 100068. [Google Scholar] [CrossRef]
  88. Radi, A.E.; Abd-Ellatief, M.R. Electrochemical Aptasensors: Current Status and Future Perspectives. Diagnostics 2021, 11, 104. [Google Scholar] [CrossRef] [PubMed]
  89. Pellitero, M.A.; Arroyo-Currás, N. Study of Surface Modification Strategies to Create Glassy Carbon-Supported, Aptamer-Based Sensors for Continuous Molecular Monitoring. Anal. Bioanal. Chem. 2022, 414, 5627. [Google Scholar] [CrossRef]
  90. Sequeira-Antunes, B.; Ferreira, H.A. Nucleic Acid Aptamer-Based Biosensors: A Review. Biomedicines 2023, 11, 3201. [Google Scholar] [CrossRef] [PubMed]
  91. Liang, G.; Song, L.; Gao, Y.; Wu, K.; Guo, R.; Chen, R.; Zhen, J.; Pan, L. Aptamer Sensors for the Detection of Antibiotic Residues— A Mini-Review. Toxics 2023, 11, 513. [Google Scholar] [CrossRef]
  92. Han, C.; Li, R.; Li, H.; Liu, S.; Xu, C.; Wang, J.; Wang, Y.; Huang, J. Ultrasensitive Voltammetric Determination of Kanamycin Using a Target-Triggered Cascade Enzymatic Recycling Couple along with DNAzyme Amplification. Microchim. Acta 2017, 184, 2941–2948. [Google Scholar] [CrossRef]
  93. Bao, Y.; Cai, K.; Liu, Y.; Li, B. Aptamer-Based Antibiotic Electrochemical Detection Platform Using Portable Plastic Gold Electrode and Multichannel Chip. Chin. J. Chem. 2024, 42, 171–176. [Google Scholar] [CrossRef]
  94. Malecka-Baturo, K.; Zaganiaris, A.; Grabowska, I.; Kurzątkowska-Adaszyńska, K. Electrochemical Biosensor Designed to Distinguish Tetracyclines Derivatives by SsDNA Aptamer Labelled with Ferrocene. Int. J. Mol. Sci. 2022, 23, 13785. [Google Scholar] [CrossRef]
  95. Akbarzadeh, S.; Khajehsharifi, H.; Hajihosseini, S. Detection of Oxytetracycline Using an Electrochemical Label-Free Aptamer-Based Biosensor. Biosensors 2022, 12, 468. [Google Scholar] [CrossRef]
  96. Li, F.; Wu, Y.; Chen, D.; Guo, Y.; Wang, X.; Sun, X. Sensitive Dual-Labeled Electrochemical Aptasensor for Simultaneous Detection of Multi-Antibiotics in Milk. Int. J. Hydrog. Energy 2021, 46, 23301–23309. [Google Scholar] [CrossRef]
  97. Chen, F.L.; Song, K.H.; Xu, J.T.; Wang, K.Z.; Feng, X.Z.; Han, G.C.; Kraatz, H.B. An Electrochemical Aptamer Sensor for Rapid Quantification Sulfadoxine Based on Synergistic Signal Amplification of Indole and MWCNTs and Its Electrooxidation Mechanism. Sens. Actuators B Chem. 2024, 401, 135008. [Google Scholar] [CrossRef]
  98. Hosu, O.; Melinte, G.; Ștefan, G.; Casian, M.; Cristea, C. Towards Selective Tetracycline Recognition in Wastewater Based on Gold Nanovoids@aptamer Sensing. Electrochim. Acta 2023, 460, 142556. [Google Scholar] [CrossRef]
  99. Wei, X.; Sun, Y.; Luo, Y.; Shu, R.; Lin, H.; Xu, D. Electrochemical Biosensor Bi2O3@Au@Apta Based on Bi2O3 Nanofibers for Sensitive Detection of Ampicillin. Biochem. Eng. J. 2023, 200, 109107. [Google Scholar] [CrossRef]
  100. Zhang, R.; Zhang, J.; Qu, X.; Li, S.; Zhao, Y.; Liu, S.; Wang, Y.; Huang, J.; Yu, J. Efficient Strand Displacement Amplification via Stepwise Movement of a Bipedal DNA Walker on an Electrode Surface for Ultrasensitive Detection of Antibiotics. Analyst 2020, 145, 2975–2981. [Google Scholar] [CrossRef] [PubMed]
  101. 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]
  102. Hu, M.; Yue, F.; Dong, J.; Tao, C.; Bai, M.; Liu, M.; Zhai, S.; Chen, S.; Liu, W.; Qi, G.; et al. Screening of Broad-Spectrum Aptamer and Development of Electrochemical Aptasensor for Simultaneous Detection of Penicillin Antibiotics in Milk. Talanta 2024, 269, 125508. [Google Scholar] [CrossRef]
  103. Yao, X.; Shen, J.; Liu, Q.; Fa, H.; Yang, M.; Hou, C. A Novel Electrochemical Aptasensor for the Sensitive Detection of Kanamycin Based on UiO-66-NH2/MCA/MWCNT@rGONR Nanocomposites. Anal. Methods 2020, 12, 4967–4976. [Google Scholar] [CrossRef]
  104. Zhang, Y.; Li, B.; Wei, X.; Gu, Q.; Chen, M.; Zhang, J.; Mo, S.; Wang, J.; Xue, L.; Ding, Y.; et al. Amplified Electrochemical Antibiotic Aptasensing Based on Electrochemically Deposited AuNPs Coordinated with PEI-Functionalized Fe-Based Metal-Organic Framework. Microchim. Acta 2021, 188, 286. [Google Scholar] [CrossRef]
  105. Gao, F.; Zhao, Y.; Dai, X.; Xu, W.; Zhan, F.; Liu, Y.; Wang, Q. Aptamer Tuned Nanozyme Activity of Nickel-Metal–Organic Framework for Sensitive Electrochemical Aptasensing of Tetracycline Residue. Food Chem. 2024, 430, 137041. [Google Scholar] [CrossRef]
  106. Reville, E.K.; Sylvester, E.H.; Benware, S.J.; Negi, S.S.; Berda, E.B. Customizable Molecular Recognition: Advancements in Design, Synthesis, and Application of Molecularly Imprinted Polymers. Polym. Chem. 2022, 13, 3387–3411. [Google Scholar] [CrossRef]
  107. Veloz Martínez, I.; Ek, J.I.; Ahn, E.C.; Sustaita, A.O. Molecularly Imprinted Polymers via Reversible Addition–Fragmentation Chain-Transfer Synthesis in Sensing and Environmental Applications. RSC Adv. 2022, 12, 9186. [Google Scholar] [CrossRef]
  108. Moro, G.; Bottari, F.; Sleegers, N.; Florea, A.; Cowen, T.; Moretto, L.M.; Piletsky, S.; De Wael, K. Conductive Imprinted Polymers for the Direct Electrochemical Detection of β-Lactam Antibiotics: The Case of Cefquinome. Sens. Actuators B Chem. 2019, 297, 126786. [Google Scholar] [CrossRef]
  109. Seguro, I.; Rebelo, P.; Pacheco, J.G.; Delerue-Matos, C. Electropolymerized, Molecularly Imprinted Polymer on a Screen-Printed Electrode—A Simple, Fast, and Disposable Voltammetric Sensor for Trazodone. Sensors 2022, 22, 2819. [Google Scholar] [CrossRef]
  110. Li, S.; Ma, X.; Pang, C.; Li, H.; Liu, C.; Xu, Z.; Luo, J.; Yang, Y. Novel Molecularly Imprinted Amoxicillin Sensor Based on a Dual Recognition and Dual Detection Strategy. Anal. Chim. Acta 2020, 1127, 69–78. [Google Scholar] [CrossRef] [PubMed]
  111. Vu, O.T.; Nguyen, Q.H.; Nguy Phan, T.; Luong, T.T.; Eersels, K.; Wagner, P.; Truong, L.T.N. Highly Sensitive Molecularly Imprinted Polymer-Based Electrochemical Sensors Enhanced by Gold Nanoparticles for Norfloxacin Detection in Aquaculture Water. ACS Omega 2022, 8, 2887–2896. [Google Scholar] [CrossRef] [PubMed]
  112. Nguyen, V.B.C.; Ayankojo, A.G.; Reut, J.; Rappich, J.; Furchner, A.; Hinrichs, K.; Syritski, V. Molecularly Imprinted Co-Polymer for Class-Selective Electrochemical Detection of Macrolide Antibiotics in Aqueous Media. Sens. Actuators B Chem. 2023, 374, 132768. [Google Scholar] [CrossRef]
  113. Frigoli, M.; Caldara, M.; Royakkers, J.; Lowdon, J.W.; Cleij, T.J.; Diliën, H.; Eersels, K.; van Grinsven, B. Gold Screen-Printed Electrodes Coupled with Molecularly Imprinted Conjugated Polymers for Ultrasensitive Detection of Streptomycin in Milk. Microchem. J. 2024, 200, 110433. [Google Scholar] [CrossRef]
  114. Rebelo, P.; Pacheco, J.G.; Cordeiro, M.N.D.S.; Melo, A.; Delerue-Matos, C. Azithromycin Electrochemical Detection Using a Molecularly Imprinted Polymer Prepared on a Disposable Screen-Printed Electrode. Anal. Methods 2020, 12, 1486–1494. [Google Scholar] [CrossRef]
  115. Shi, X.; Zuo, Y.; Jia, X.; Wu, X.; Jing, N.; Wen, B.; Mi, X. A Novel Molecularly Imprinted Sensor Based on Gold Nanoparticles/Reduced Graphene Oxide/Single-Walled Carbon Nanotubes Nanocomposite for the Detection of Pefloxacin. Int. J. Electrochem. Sci. 2020, 15, 9683–9697. [Google Scholar] [CrossRef]
  116. Abera, B.D.; Ortiz-gómez, I.; Shkodra, B.; Romero, F.J.; Cantarella, G.; Petti, L.; Salinas-castillo, A.; Lugli, P.; Rivadeneyra, A. Laser-induced Graphene Electrodes Modified with a Molecularly Imprinted Polymer for Detection of Tetracycline in Milk and Meat. Sensors 2022, 22, 269. [Google Scholar] [CrossRef]
  117. Turco, A.; Corvaglia, S.; Pompa, P.P.; Malitesta, C. An Innovative and Simple All Electrochemical Approach to Functionalize Electrodes with a Carbon Nanotubes/Polypyrrole Molecularly Imprinted Nanocomposite and Its Application for Sulfamethoxazole Analysis. J. Colloid Interface Sci. 2021, 599, 676–685. [Google Scholar] [CrossRef]
  118. Wang, X.; Liu, Y.; Liu, J.; Qu, J.; Huang, J.; Tan, R.; Yu, Y.; Wu, J.; Yang, J.; Li, Y.; et al. A Bifunctional Electrochemical Sensor for Simultaneous Determination of Electroactive and Non-Electroactive Analytes: A Universal yet Very Effective Platform Serving Therapeutic Drug Monitoring. Biosens. Bioelectron. 2022, 208, 114233. [Google Scholar] [CrossRef]
  119. Zeb, S.; Wong, A.; Khan, S.; Hussain, S.; Sotomayor, M.D.P.T. Using Magnetic Nanoparticles/MIP-Based Electrochemical Sensor for Quantification of Tetracycline in Milk Samples. J. Electroanal. Chem. 2021, 900, 115713. [Google Scholar] [CrossRef]
  120. Wang, Y.; Sun, X.; Cai, L.; Wang, H.; Zhang, B.; Fang, G.; Wang, S. A “Signal on/off” Biomimetic Electrochemiluminescence Sensor Using Titanium Carbide Nanodots as Co-Reaction Accelerator for Ultra-Sensitive Detection of Ciprofloxacin. Anal. Chim. Acta 2022, 1206, 339690. [Google Scholar] [CrossRef] [PubMed]
  121. Koçak, I. ZnO and Au Nanoparticles Supported Highly Sensitive and Selective Electrochemical Sensor Based on Molecularly Imprinted Polymer for Sulfaguanidine and Sulfamerazine Detection. J. Pharm. Biomed. Anal. 2023, 234, 115518. [Google Scholar] [CrossRef] [PubMed]
  122. Surya, S.G.; Khatoon, S.; Lahcen, A.A.; Nguyen, A.T.H.; Dzantiev, B.B.; Tarannum, N.; Salama, K.N. A Chitosan Gold Nanoparticles Molecularly Imprinted Polymer Based Ciprofloxacin Sensor. RSC Adv. 2020, 10, 12823–12832. [Google Scholar] [CrossRef] [PubMed]
  123. Liu, Z.; Fan, T.; Zhang, Y.; Ren, X.; Wang, Y.; Ma, H.; Wei, Q. Electrochemical Assay of Ampicillin Using Fe3N-Co2N Nanoarray Coated with Molecularly Imprinted Polymer. Microchim. Acta 2020, 187, 442. [Google Scholar] [CrossRef]
  124. Güney, S.; Arslan, T.; Yanık, S.; Güney, O. An Electrochemical Sensing Platform Based on Graphene Oxide and Molecularly Imprinted Polymer Modified Electrode for Selective Detection of Amoxicillin. Electroanalysis 2021, 33, 46–56. [Google Scholar] [CrossRef]
  125. Li, J.; Huang, X.; Ma, J.; Wei, S.; Zhang, H. A Novel Electrochemical Sensor Based on Molecularly Imprinted Polymer with Binary Functional Monomers at Fe-Doped Porous Carbon Decorated Au Electrode for the Sensitive Detection of Lomefloxacin. Ionics 2020, 26, 4183–4192. [Google Scholar] [CrossRef]
  126. Stoian, I.A.; Iacob, B.C.; Dudaș, C.L.; Barbu-Tudoran, L.; Bogdan, D.; Marian, I.O.; Bodoki, E.; Oprean, R. Biomimetic Electrochemical Sensor for the Highly Selective Detection of Azithromycin in Biological Samples. Biosens. Bioelectron. 2020, 155, 112098. [Google Scholar] [CrossRef]
  127. Li, S.; Pang, C.; Ma, X.; Wu, Y.; Wang, M.; Xu, Z.; Luo, J. Aggregation-Induced Electrochemiluminescence and Molecularly Imprinted Polymer Based Sensor with Fe3O4@Pt Nanoparticle Amplification for Ultrasensitive Ciprofloxacin Detection. Microchem. J. 2022, 178, 107345. [Google Scholar] [CrossRef]
  128. Ayankojo, A.G.; Reut, J.; Ciocan, V.; Öpik, A.; Syritski, V. Molecularly Imprinted Polymer-Based Sensor for Electrochemical Detection of Erythromycin. Talanta 2020, 209, 120502. [Google Scholar] [CrossRef] [PubMed]
  129. Wilkirson, E.C.; Singampalli, K.L.; Li, J.; Dixit, D.D.; Jiang, X.; Gonzalez, D.H.; Lillehoj, P.B. Affinity-Based Electrochemical Sensors for Biomolecular Detection in Whole Blood. Anal. Bioanal. Chem. 2023, 415, 3983. [Google Scholar] [CrossRef]
  130. Shen, G.; Zhang, X.; Shen, Y.; Zhang, C. Immobilization of Antibodies on Aldehyde-Functionalized Polymer/Graphene Films for the Fabrication of a Label-Free Electrochemical Immunosensor. J. Electroanal. Chem. 2015, 759, 67–71. [Google Scholar] [CrossRef]
  131. Aymard, C.; Kanso, H.; Serrano, M.J.; Pagán, R.; Noguer, T.; Istamboulie, G. Development of a New Dual Electrochemical Immunosensor for a Rapid and Sensitive Detection of Enrofloxacin in Meat Samples. Food Chem. 2022, 370, 131016. [Google Scholar] [CrossRef]
  132. Police Patil, A.V.; Chuang, Y.S.; Li, C.; Wu, C.C. Recent Advances in Electrochemical Immunosensors with Nanomaterial Assistance for Signal Amplification. Biosensors 2023, 13, 125. [Google Scholar] [CrossRef] [PubMed]
  133. Iglesias-Mayor, A.; Amor-Gutiérrez, O.; Costa-García, A.; de la Escosura-Muñiz, A. Nanoparticles as Emerging Labels in Electrochemical Immunosensors. Sensors 2019, 19, 5137. [Google Scholar] [CrossRef]
  134. Pollap, A.; Kochana, J. Electrochemical Immunosensors for Antibiotic Detection. Biosensors 2019, 9, 61. [Google Scholar] [CrossRef]
  135. You, X.; Chen, Y.; Zhou, J.; Liu, H.; Liu, Y.; Qi, Y.; Liang, C.; Ding, P.; Zhu, X.; Wang, A.; et al. Concanavalin A as Carrier for Sensitive Electrochemical Immunosensor Based on AgNPs-RGO Signal Amplification. J. Electroanal. Chem. 2024, 961, 118214. [Google Scholar] [CrossRef]
  136. Kumar, R.; Lakshmi, G.B.V.S.; Dhiman, T.K.; Singh, K.; Solanki, P.R. Highly Sensitive Amoxicillin Immunosensor Based on Aqueous Vanadium Disulphide Quantum Dots. J. Electroanal. Chem. 2021, 892, 115266. [Google Scholar] [CrossRef]
  137. Kolhe, P.; Roberts, A.; Gandhi, S. Fabrication of an Ultrasensitive Electrochemical Immunosensor Coupled with Biofunctionalized Zero-Dimensional Graphene Quantum Dots for Rapid Detection of Cephalexin. Food Chem. 2023, 398, 133846. [Google Scholar] [CrossRef] [PubMed]
  138. Yadav, A.K.; Verma, D.; Lakshmi, G.B.V.S.; Eremin, S.; Solanki, P.R. Fabrication of Label-Free and Ultrasensitive Electrochemical Immunosensor Based on Molybdenum Disulfide Nanoparticles Modified Disposable ITO: An Analytical Platform for Antibiotic Detection in Food Samples. Food Chem. 2021, 363, 130245. [Google Scholar] [CrossRef] [PubMed]
  139. Cai, M.; Zhu, L.; Ding, Y.; Wang, J.; Li, J.; Du, X. Determination of Sulfamethoxazole in Foods Based on CeO2/Chitosan Nanocomposite-Modified Electrodes. Mater. Sci. Eng. C 2012, 32, 2623–2627. [Google Scholar] [CrossRef]
  140. Sanli, S. Single-Drop Electrochemical Immunosensor with 3D-Printed Magnetic Attachment for Onsite Smartphone Detection of Amoxicillin in Raw Milk. Food Chem. 2024, 437, 137823. [Google Scholar] [CrossRef]
  141. Chomthong, K.; Kunpatee, K.; Pimpitak, U.; Puthong, S.; Komolpis, K.; Wonsawat, W.; Nuanualsuwan, S.; Yakoh, A.; Khongchareonporn, N.; Ruecha, N.; et al. Label-Free Simultaneous Detection of Quinolone Antibiotic Residues Using an Origami Paper–Based Electrochemical Immunosensor. Sens. Actuators B Chem. 2024, 410, 135667. [Google Scholar] [CrossRef]
  142. Valera, E.; Muriano, A.; Pividori, I.; Sánchez-Baeza, F.; Marco, M.P. Development of a Coulombimetric Immunosensor Based on Specific Antibodies Labeled with CdS Nanoparticles for Sulfonamide Antibiotic Residues Analysis and Its Application to Honey Samples. Biosens. Bioelectron. 2013, 43, 211–217. [Google Scholar] [CrossRef]
  143. Himanshu, J.K.; Lakshmi, G.B.V.S.; Singh, A.K.; Solanki, P.R. Reduced Graphene Oxide-Gadolinium Oxide-Functionalized Paper Based Immunosensor for Electrochemical Detection of Gentamicin. Biosens. Bioelectron. X 2024, 17, 100442. [Google Scholar] [CrossRef]
  144. Chaudhary, N.; Yadav, A.K.; Verma, D.; Sharma, J.G.; Solanki, P.R. An Electrochemical Immunosensor Based on a Nanostructured Lanthanum Oxide-Substituted Reduced Graphene Oxide Interface for Ultralow Ciprofloxacin Detection in Milk Samples. Mater. Adv. 2024, 5, 1597–1613. [Google Scholar] [CrossRef]
  145. Prusty, A.K.; Bhand, S. A Capacitive Immunosensor for Tetracycline Estimation Using Antibody Modified Polytyramine-Alkanethiol Ultra-Thin Film on Gold. J. Electroanal. Chem. 2020, 863, 114055. [Google Scholar] [CrossRef]
  146. High-End UHPLC System, 1290 Infinity II LC System|Agilent. Available online: https://www.agilent.com/en/product/liquid-chromatography/hplc-systems/analytical-hplc-systems/1290-lc-system (accessed on 2 August 2024).
  147. TSQ QuantisTM Triple Quadrupole Mass Spectrometer. Available online: https://www.thermofisher.com/order/catalog/product/TSQ02-10001 (accessed on 2 August 2024).
  148. ROSA Lateral Flow Mycotoxin Strips|Charm Sciences. Available online: https://www.charm.com/products/test-and-kits/mycotoxin-tests/rosa-lateral-flow/ (accessed on 2 August 2024).
  149. The Next Generation of Rapid Microbial Testing System|Neogen. Available online: https://www.neogen.com/soleris-next-generation/ (accessed on 2 August 2024).
  150. PA 800 Plus Pharmaceutical Analysis System. Available online: https://sciex.com/products/capillary-electrophoresis/pa-800-plus-pharmaceutical-analysis-system (accessed on 2 August 2024).
  151. UV-1900i: SHIMADZU (Shimadzu Corporation). Available online: https://www.shimadzu.com/an/products/molecular-spectroscopy/uv-vis/uv-vis-nir-spectroscopy/uv-1900i/index.html (accessed on 2 August 2024).
  152. Vanquish HPLC and UHPLC Systems—NL. Available online: https://www.thermofisher.com/uk/en/home/products-and-services/promotions/industrial/vanquish-hplc-uhplc-systems.html (accessed on 2 August 2024).
  153. Sciences, C. Antibiotic Testing|Milk Antibiotic Test Kits. Available online: https://www.charm.com/products/test-and-kits/antibiotic-tests/ (accessed on 2 August 2024).
  154. Sciex Antibiotic Testing. Available online: https://sciex.com/applications/food-and-beverage-testing/residue-analysis/antibiotic-testing (accessed on 2 August 2024).
  155. Bruker Beta-Lactamase Activity Detection|Bruker. Available online: https://www.bruker.com/en/products-and-solutions/microbiology-and-diagnostics/microbial-identification/beta-lactamase-activity-detection.html (accessed on 2 August 2024).
  156. Jeol The NMR Application for Medicinal Screening|Applications Notes|JEOL Ltd. Available online: https://www.jeol.com/solutions/applications/details/NM170019.php (accessed on 2 August 2024).
  157. Waters Method Development of Ten Antibiotic Compounds Using a Systematic Screening Protocol on an ACQUITY Arc with UV and QDa Detection|Waters. Available online: https://www.waters.com/nextgen/ee/en/library/application-notes/2021/method-development-of-ten-antibiotic-compounds-using-a-systematic-screening-protocol-on-an-acquity-arc-with-uv-and-qda-detection.html (accessed on 2 August 2024).
  158. Charm II System|Charm Sciences Food Safety Solutions. Available online: https://www.charm.com/products/equipment/monitoring-systems/charm-ii-system/ (accessed on 2 August 2024).
  159. RIDASCREEN® Sulfonamide—Food & Feed Analysis. Available online: https://food.r-biopharm.com/products/ridascreen-sulfamethazin/ (accessed on 2 August 2024).
  160. Neogen Releases Easiest Dairy Antibiotic Tests|Neogen. Available online: https://www.neogen.com/en-gb/neocenter/press-releases/neogen-releases-easiest-dairy-antibiotic-tests/ (accessed on 2 August 2024).
  161. Q Exactive Orbitrap Mass Spectrometers|Thermo Fisher Scientific—NL. Available online: https://www.thermofisher.com/nl/en/home/industrial/mass-spectrometry/liquid-chromatography-mass-spectrometry-lc-ms/lc-ms-systems/orbitrap-lc-ms/q-exactive-orbitrap-mass-spectrometers.html (accessed on 2 August 2024).
  162. Chen, Q.; Pan, X.D.; Huang, B.F.; Han, J.L.; Zhou, B. Screening of Multi-Class Antibiotics in Pork Meat by LC-Orbitrap-MS with Modified QuEChERS Extraction. RSC Adv. 2019, 9, 28119–28125. [Google Scholar] [CrossRef]
  163. SNAP Beta-Lactam ST Test: Detects Beta-Lactam Residues—IDEXX US. Available online: https://www.idexx.com/en/milk/dairy-tests/snap-beta-lactam-st/ (accessed on 2 August 2024).
  164. RIDASCREEN® Chloramphenicol—Food & Feed Analysis. Available online: https://food.r-biopharm.com/products/ridascreen-chloramphenicol/ (accessed on 2 August 2024).
  165. MaxSignal® Beta-Lactam ELISA Kit|PerkinElmer. Available online: https://www.perkinelmer.com/product/maxsignal-beta-lactam-elisa-kit-food-1065-01d (accessed on 2 August 2024).
  166. AMP’D® ELISA Signal Amplification Kit—Enzo. Available online: https://www.enzo.com/product/ampd-elisa-signal-amplification-kit/ (accessed on 2 August 2024).
  167. Meizheng Accurate Antibiotics ELISA Test Kits|Meizheng. Available online: https://mzfoodtest.com/product-category/technologies/elisa-test-kits-technologies/antibiotics-elisa-test-kits/ (accessed on 2 August 2024).
  168. Meizheng Nitroimidazole ELISA Test Kit for Animal Tissues and Eggs. Available online: https://mzfoodtest.com/product/nitroimidazole-elisa-test-kit-for-animal-tissues-and-eggs/ (accessed on 2 August 2024).
  169. Diagnostics, C. ELISA Kits for Drug Residues Detection—Creative Diagnostics. Available online: https://www.creative-diagnostics.com/news-elisa-kits-for-drug-residues-detection-73.htm (accessed on 2 August 2024).
  170. Neogen Dairy Antibiotic Testing|Food Safety|Neogen. Available online: https://www.neogen.com/categories/dairy-residues/ (accessed on 2 August 2024).
  171. Gold Standard Diagnostic SENSISpec Tetracycline ELISA, 96 Tests. Available online: https://www.goldstandarddiagnostics.com/products/veterinary-drug-residues/sensispec-tetracycline-elisa.html (accessed on 2 August 2024).
  172. Abcam Kanamycin ELISA Kit (Ab287806/K4210-100)|Abcam, Netherlands. Available online: https://www.abcam.com/en-nl/products/elisa-kits/kanamycin-elisa-kit-ab287806 (accessed on 2 August 2024).
  173. MP Multi Panel Drug Tests—Test Kits for Screening Drugs in Urine & Testing Strips. Available online: https://www.mpbio.com/us/diagnostics/drugs-of-abuse (accessed on 2 August 2024).
  174. Biovision Gentamicin ELISA Kit, BioVision, Inc.|VWR. Available online: https://us.vwr.com/store/product/23608718/gentamicin-elisa-kit-biovision-inc (accessed on 2 August 2024).
  175. Randox General Biochip—Antimicrobial Array I Ultra—Randox Food. Available online: https://www.randoxfood.com/biochip/antimicrobial-array-i-ultra/ (accessed on 2 August 2024).
  176. Sciences, C. SLBL—Charm SL Beta-Lactam Test—Charm Sciences. Available online: https://www.charm.com/products/test-and-kits/antibiotic-tests/rosa-lateral-flow/slbl-charm-sl-beta-lactam-test/ (accessed on 2 August 2024).
  177. Metrohm Gentamicin Sulfate Using Pulsed Amperometric Detection in Accordance with Ph. Eur.|Metrohm. Available online: https://www.metrohm.com/en_nl/applications/application-notes/aa-p-001-100/an-p-037.html (accessed on 2 August 2024).
  178. PalmSens PalmSens4—PalmSens. Available online: https://www.palmsens.com/product/palmsens4/ (accessed on 2 August 2024).
  179. Metrohm Screen-Printed Electrodes Archivos—Metrohm. Available online: https://metrohm-dropsens.com/category/electrodes/screen-printed-electrodes/ (accessed on 2 August 2024).
  180. PalmSens Sensit Smart—PalmSens. Available online: https://www.palmsens.com/product/sensit-smart/ (accessed on 2 August 2024).
  181. ZP ZP—Commercializing a Sensor for Antibiotic Detection|Zimmer & Peacock AS. Available online: https://www.zimmerpeacock.com/2020/07/12/zp-commercializing-a-sensor-for-antibiotic-detection/ (accessed on 2 August 2024).
  182. Analytik Jena Multi EA 5100—CNSX Micro-Elemental Analyzer—Analytik Jena. Available online: https://www.analytik-jena.com/products/chemical-analysis/combustion-elemental-analysis/cnsx/multi-ea-5100-for-micro-elemental-analysis/ (accessed on 2 August 2024).
Figure 1. Schematic representation of (a) a generic biosensor device showing a sample containing the target interacting with the bioreceptor. After the recognition, the transducer is responsible for generating a signal. (b) examples of electrochemical transduction signals. Copyright Sensors and Actuators Reports. 2022 [71].
Figure 1. Schematic representation of (a) a generic biosensor device showing a sample containing the target interacting with the bioreceptor. After the recognition, the transducer is responsible for generating a signal. (b) examples of electrochemical transduction signals. Copyright Sensors and Actuators Reports. 2022 [71].
Sensors 24 05576 g001
Figure 2. Schematic illustration of the working principle of electrochemically based aptamer sensors. Copyright Toxics 2023 [91].
Figure 2. Schematic illustration of the working principle of electrochemically based aptamer sensors. Copyright Toxics 2023 [91].
Sensors 24 05576 g002
Figure 3. Portable aptamer-based sensor for kanamycin developed by Bao et al. Copyright Chinese Jour. Of Chem., 2023 [93].
Figure 3. Portable aptamer-based sensor for kanamycin developed by Bao et al. Copyright Chinese Jour. Of Chem., 2023 [93].
Sensors 24 05576 g003
Figure 4. A modified ssDNA aptamer sensor for the selective detection of tetracycline designed by Malecka-Baturo et al. Copyright Int. J. Mol. Sci. 2022 [94]. (A) Scheme of the aptasensor preparation, and (B) representation of the steps needed to use the aptasensor.
Figure 4. A modified ssDNA aptamer sensor for the selective detection of tetracycline designed by Malecka-Baturo et al. Copyright Int. J. Mol. Sci. 2022 [94]. (A) Scheme of the aptasensor preparation, and (B) representation of the steps needed to use the aptasensor.
Sensors 24 05576 g004
Figure 5. Schematic representation of the extraction and rebinding process on an antibiotic (cefquinome) after electropolymerization on the surface of an electrode and consequent electrochemical analysis. Copyright Sensors and Act. B: Chemical, 2019 [108].
Figure 5. Schematic representation of the extraction and rebinding process on an antibiotic (cefquinome) after electropolymerization on the surface of an electrode and consequent electrochemical analysis. Copyright Sensors and Act. B: Chemical, 2019 [108].
Sensors 24 05576 g005
Figure 6. The dual recognition sensor for amoxicillin. Copyright Analytica Chimica Acta, 2020 [110].
Figure 6. The dual recognition sensor for amoxicillin. Copyright Analytica Chimica Acta, 2020 [110].
Sensors 24 05576 g006
Figure 7. Overview of the norfloxacin sensor developed by Thi Vu et al. [111]. Copyright ACS Omega, 2023.
Figure 7. Overview of the norfloxacin sensor developed by Thi Vu et al. [111]. Copyright ACS Omega, 2023.
Sensors 24 05576 g007
Figure 8. Working principle of the ConA arsanilic acid developed by You et al. Copyright J. Electroanalytical Chem., 2024 [135].
Figure 8. Working principle of the ConA arsanilic acid developed by You et al. Copyright J. Electroanalytical Chem., 2024 [135].
Sensors 24 05576 g008
Table 1. Advantages and disadvantages of specific biorecognition elements for biosensors [53].
Table 1. Advantages and disadvantages of specific biorecognition elements for biosensors [53].
Biorecognition ElementAdvantagesDisadvantages
Antibody
Enzyme
Selectivity,
Reusability,
High specificity,
Versatility
Reproducibility,
Batch variations,
Processing stability,
Cost; Limited shelf-life
Nucleic Acids
Aptamers
Sensitivity,
Reproducibility
Cost,
Non-specific binding interactions,
Stability
Molecularly imprinted polymersStability,
Reusability,
Low cost
Reproducibility,
Template removal from cavities
Surface imprinted polymersSelectivity,
Robustness
Scalability,
Template availability
Table 2. Electrochemical aptasensors for antibiotic detection.
Table 2. Electrochemical aptasensors for antibiotic detection.
Sensor TypeTarget MoleculeMethod of DetectionLimit of DetectionSampleRef.
Modified nanocomposite including multi-walled carbon nanotubes (MWCNTs), gold nanoparticles (AuNPs), reduced graphene oxide (rGO), chitosan (CS), and modified nanosheets to bind with OTC-specific aptamerOxytetracyclineDPV30 pMSpiked milk samples[95]
Dual-labeled multiple aptasensor using RNA-based aptamer strands, semiconductor quantum dots (QDs), and gold nanoshells (AuNSs). Conjugation of biotinylated aptamers to SA-coated sadmium sulfide (CdS) and lead sulfide (PbS) QDs.Kanamycin
Tobramycin
DPV0.12 nM
0.49 nM
Spiked milk[96]
Aptamer (Apt) BSA/Apt/indole/MWCNTs/GCE with the synergistic amplification of multi-walled carbon nanotubes (MWCNTs) and indoleSulfadoxineDPV0.033 μMbuffer[97]
Novel 3D honeycombed goldnanovoids@aptamer nanostructured platform with ferrocene labeling on aptamer strandsTetracyclineDPV1.2 nMWastewater[98]
Aptamer sequence bonded onto bismuth oxide nanofibers paired with AuNPsAmpicillinDPV0.88 nMWater and milk[99]
Exonuclease III (Exo III)-aided target-aptamer binding recycling (ETBR) activated bipedal DNA machineKanamycinCV
DPV
7.1fMbuffer[100]
Polyethyleneimine grafted rGO and titanium dioxide nanocomposite materialCiprofloxacinDPV
EIS
0.7 nMReal water samples[101]
Aptasensor prepared using AuNPs combined with ferroferric oxide-multi walled carbon nanotube (Fe3O4-MWCNTs) complexPenicillin antibiotics (PENs)CV
DPV
EIS
0.667 nMSpiked milk samples[102]
DNA aptamer and partially complementary short chain assembled onto integrated portable plastic gold electrode (PGE) through Au-S bonds.KanamycinSWV
CV
0.40 μMBuffer[93]
Thiolated aptamer labeled with ferrocene covalently co-immobilized onto a gold electrode surface with 6-mercaptohexan-1-olTetracyclineSWV0.20 nMSpiked milk[94]
Nanocomposite comprising a functionalized MOF, a MWCNT@reduced graphene oxide nanoribbon, and a covalent organic framework (COF). cDNA strands with terminal amino groups anchored on the surface, as well as penetration into the poresKanamycinSWV
CV
13 nMFish
Meat
Milk
[103]
Amplified aptasensor using Immobilized DNA strands on AuNPs/Fe-based metal organic framework (MOF)TobramycinSWV56 pMSpiked milk samples[104]
MOF of Ni2+-2,3,6,7,10,11-hexahydroxytriphenylene (Ni-HHTP) coated on a SPE, followed by non-covalent adsorption of tetracycline aptamer (TC-Apt) through the π-stackingTetracyclineCV
SWV
1.9 pMbuffer[105]
Table 4. Antibody-based electrochemical biosensors for antibiotic detection.
Table 4. Antibody-based electrochemical biosensors for antibiotic detection.
Sensor TypeTarget MoleculeMethod of DetectionLimit of DetectionSampleRef.
A composite o0066 chitosan (CH) and thioglycolic acid capped vanadium sulfide quantum dots (TGA-VS2QDs) was constructed on glass substrate coated with ITO film to form electrodes on which monoclonal antibodies (mAb) for amoxicillin were immobilizedAmoxicillinDPV1.65 pMSpiked fish[136]
Immunosensor based on AgNPs-reduced graphene oxide (AgNPs-rGO) and staphylococcal protein A (SPA) that was targeted to immobilize mAbVirginiamycin M1DPV0.18 ng mL−1Meat[58]
A cephalexin–bovine serum albumin (CFX-BSA) conjugate was developed to create antibodies (Abs). Graphene quantum dots (GQDs) were used for signal enhancementCephalexinDPV0.53 fMSpiked animal source food[137]
Anti-quinolone Ab immobilized onto screen-printed dual carbon electrodesEnrofloxacinDPV3 ng mL−1Meat[131]
molybdenum disulfide nanoparticles (nMoS2NPs) deposited on ITO-coated glass substrate with Abs bonded through amide linkagesAmpicillinDPV0.028 µg mL−1 Milk, orange juice, and tap water[138]
Sensor based on AgNPs-rGO nanocomposite and concanavalin A (ConA) that was bound to mAbs through lectin–sugar interactionsArsanilic acidDPV0.008 ng mL−1Buffer, chicken, eggs[135]
Nanocomposite-modified glass carbon electrode (GCE) with a biospecific CeO2-chitosan (CHIT)-modified nanocomposite on which polyclonal Abs were immobilizedSulfamethoxazoleDPV1.3 nMBuffer or food[139]
Immunosensor platform based on Ab-conjugated magnetic particles on an electrode surface that uses a 3D cell to accumulate the analyte AmoxicillinSWV0.44 µMRaw milk[140]
Origami paper-based analytical device (oPAD) with multiple Ab zones for simultaneous quinolone residue detectionNorfloxacin,
Enrofloxacin
SWV2.02 ng mL−1
1.70 ng mL−1
Food[141]
A graphite composite electrode (GEC), biofunctionalized magnetic μ-particles, and electrochemical nanoprobes prepared by labeling specific antibodies with CdS nanoparticles (CdSNP).SulfapyridineSWV0.11 μg kg−1Honey[142]
rGO-gadolinium oxide nanocomposite (rGO@Gd2O3 NC) with suspended mAbs on a SPEGentamicinCV0.424 pMMilk[143]
Nanoconstructed lanthanum oxide nanoparticle-decorated reduced graphene oxide nanocomposite (nLa2O3 NP@rGO)-based platform functionalized with 3-aminopropyltriethoxysilane (APTES) and attachment on ITO-coated substrate. mAbs immobilized onto surface.CiprofloxacinCV
DPV
EIS
0.055 μg mL−1Milk[144]
Tyramine (TA) electropolymerized resulting in an ultrathin polytyramine (PTA) film on a gold-coated silicon electrode (AuE) modified with polyclonal antibodiesTetracyclineEIS0.01 μg L−1Spiked buffer[145]
Table 5. Examples of gold-standard technologies for antibiotic detection, with analysis time and sample types.
Table 5. Examples of gold-standard technologies for antibiotic detection, with analysis time and sample types.
CompanyTechnologyProductLoDTime RequiredSample TypesRef
ThermoFisher ScientificHPLC and MSVanquish UHPLC and Orbitrap MSPpb levels30–60 min per sampleFood, water, biological fluids[152]
Charm SciencesMicrobiological AssaysROSA Lateral Flow SystemLow ppb levels1–2 h per batchFood, dairy products[153]
SciexCECapillary Electrophoresis SystemLow ng/mL range20–40 min per sampleWater, biological fluids, pharmaceuticals[154]
Bruker Corporation (Billerica, MA, USA)MSMBT STAR-Carba IVDLow ppb levels10–20 min per sampleFood, water, pharmaceuticals[155]
JEOL Ltd. (Tokyo, Japan)NMRECZ500R NMR SpectrometerPpb levels1–2 h per sampleBiological fluids, complex mixtures[156]
Waters CorporationLC-MS/UV–VisXevo TQ-S micro LC-MS/MS SystemPpt levels30–60 min per sampleFood, biological fluids, environmental samples[157]
Table 6. Examples of immunoassay-based technologies for antibiotic detection, with analysis time and sample types.
Table 6. Examples of immunoassay-based technologies for antibiotic detection, with analysis time and sample types.
CompanyProductSample TypesTime RequiredLoDRef
Meizheng (Rizhoa, China)ELISA KitFood, water, biological tissues1–2 h0.05 ppb[167]
MeizhengNitroimidazole ELISA Test KitAnimal tissues and eggs1–2 h0.2 ppb[168]
Creative Diagnostics (Shirley, NY, USA)ELISA Kits for Drug Residues DetectionWater, food, Biological tissues2 h0.01–1.5 ppb[169]
Neogen CorporationVeratox for TetracyclinesDairy30 min1 ppb[170]
Gold Standard Diagnostics (Warminster, PA, USA)SENSISpec Tetracycline ELISAMeat, milk, shrimp, and honey1–2 h0.05–2 ppb[171]
Abcam (Cambridge, UK)Antibiotic Residue Detection ELISA KitTissues, Milk1.5 h0.1 ppb[172]
MP Biomedicals (Irvine, CA, USA)Quick Test Kit for AntibioticsMilk, meat, seafood1 h0.2 ppb[173]
BioVision (Zurich, Switzerland)Antibiotic Residue ELISA KitMilk, Tissues2 h0.3 ppb[174]
Table 7. Examples of electrochemical-based devices and sensors for the detection of antibiotics.
Table 7. Examples of electrochemical-based devices and sensors for the detection of antibiotics.
CompanyProductSample TypesTime RequiredLoDRef
Randox Food DiagnosticsAntibiotic ArrayMilk, meat, fish30 min0.5 ppb[175]
Charm Sciences Inc.SLBL Beta-Lactam TestDairy products~5 min1 ppb[176]
Metrohm (Bangkok, Thailand)AN-P-037Food, water1 h0.1 ppb[177]
PalmSensPalmSens4 PotentiostatFood, environmental samples~30 min0.05 ppb[178]
Dropsens (Metrohm)Screen-Printed ElectrodesFood, water~20 min0.01–10 ppb[179]
PalmSensSensit SmartFood, clinical samples~45 min0.1 ppb[180]
ZP (Zimmer and Peacock) (Skoppum, Norway)ZP AnapotEnvironmental samples30 min0.05 ppb[181]
Analytik Jena (Jena, Germany)Multi EA 5100Water, food, biological samples1 h0.2 ppb[182]
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

Frigoli, M.; Krupa, M.P.; Hooyberghs, G.; Lowdon, J.W.; Cleij, T.J.; Diliën, H.; Eersels, K.; van Grinsven, B. Electrochemical Sensors for Antibiotic Detection: A Focused Review with a Brief Overview of Commercial Technologies. Sensors 2024, 24, 5576. https://doi.org/10.3390/s24175576

AMA Style

Frigoli M, Krupa MP, Hooyberghs G, Lowdon JW, Cleij TJ, Diliën H, Eersels K, van Grinsven B. Electrochemical Sensors for Antibiotic Detection: A Focused Review with a Brief Overview of Commercial Technologies. Sensors. 2024; 24(17):5576. https://doi.org/10.3390/s24175576

Chicago/Turabian Style

Frigoli, Margaux, Mikolaj P. Krupa, Geert Hooyberghs, Joseph W. Lowdon, Thomas J. Cleij, Hanne Diliën, Kasper Eersels, and Bart van Grinsven. 2024. "Electrochemical Sensors for Antibiotic Detection: A Focused Review with a Brief Overview of Commercial Technologies" Sensors 24, no. 17: 5576. https://doi.org/10.3390/s24175576

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