Next Article in Journal
Towards Molecularly Imprinted Polypyrrole-Based Sensor for the Detection of Methylene Blue
Previous Article in Journal
Contamination- and Perturbation-Free Fluorescent Monitoring of Zn2+ in Suspensions Using Crown Ether-Functionalized Magnetic Nanoparticles
Previous Article in Special Issue
Molecularly Imprinted Plasmonic Sensors for the Determination of Environmental Water Contaminants: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carbon Electrode Modified with Molecularly Imprinted Polymers for the Development of Electrochemical Sensor: Application to Pharmacy, Food Safety, Environmental Monitoring, and Biomedical Analysis

PAM UMR A 02.102, Food and Microbiological Processes, Institut Agro, Université Bourgogne Franche-Comté, 1 Esplanade Erasme, F-21000 Dijon, France
Chemosensors 2023, 11(11), 548; https://doi.org/10.3390/chemosensors11110548
Submission received: 31 August 2023 / Revised: 21 October 2023 / Accepted: 22 October 2023 / Published: 24 October 2023
(This article belongs to the Special Issue Molecularly Imprinted Plasmonic Sensor)

Abstract

:
This review aims to elucidate recent developments in electrochemical sensors that use functionalized carbon electrodes with molecularly imprinted polymers (MIPs) for the selective detection of organic compounds in diverse fields including pharmacy, food safety, environmental monitoring of pollutants, and biomedical analysis. The main targets include explosive compounds, dyes, antioxidants, disease biomarkers, pharmaceuticals, antibiotics, allergens, pesticides, and viruses. Following a brief overview of the molecular imprinting principle, the most significant applications are explored. The selection of the functional monomer is subsequently discussed. Notably, various types of carbon electrodes are presented, with a particular emphasis on screen-printed carbon electrodes. The most commonly employed techniques for MIP deposition such as electropolymerization, drop casting, and chemical grafting are introduced and discussed. Electrochemical transduction techniques like cyclic voltammetry, differential pulse voltammetry, square wave voltammetry, and electrochemical impedance spectroscopy are presented. Lastly, the review concludes by examining potential future directions and primary limitations concerning carbon electrodes modified with MIPs.

1. Introduction

The need to accurately detect analytes in complex matrices in real time is becoming increasingly significant [1]. This requirement spans a wide range of applications, including food safety [2], quality control for drugs and food [3], monitoring environmental air and water pollution [4], biomedical analysis [5], disease diagnosis through monitoring biomarkers in blood or urine [6], combatting food fraud [7], and enhancing civil security [8]. To address these growing demands, the development of affordable, selective, and sensitive sensors presents an interesting solution. A notable challenge in sensor development involves ensuring selective responses in complex matrices such as food, urine, or blood. Electrochemical sensors offer viable solutions due to their cost-effectiveness, maturity, user-friendliness, ease of miniaturization, and the capacity to analyze small volumes [9]. However, despite these advantages, electrochemical sensors still encounter issues related to selectivity limitations and the potential for electrode fouling [10].
Combining molecularly imprinted polymers (MIPs) with electrochemistry offers a robust strategy to enhance the selectivity of electrochemical sensors [11]. MIPs are synthetic polymers designed to replicate the selectivity found in natural systems like enzyme–substrate, antigen–antibody, or enzyme–substrate interactions. These MIPs can be deposited onto electrode surfaces through various techniques, including physical methods such as spin or drop coating, as well as chemical methods like grafting or electropolymerization.
Numerous electrode materials have been functionalized with MIPs; among them, carbon-based electrodes stand out due to their cost-effectiveness and their compelling chemical and mechanical resistance properties. They are present in different forms such as glassy carbon electrode, graphene oxide, single-walled carbon nanotubes, or multi-walled carbon nanotube, offering a wide possibility of functionalization [12]. Screen-printed carbon electrodes (SPCEs) enable the miniaturization of the detection system and the analysis of small sample volumes. Furthermore, they are disposable, requiring no maintenance treatment, and carry a minimal risk of fouling [13].
This review aims to develop the recent advancements in electrochemical sensors using carbon electrodes functionalized with MIPs for the specific detection of organic compounds across various domains, including pharmaceuticals, food safety, environmental pollutant monitoring, and biomedical analysis. A special focus has been made on the development of electrochemical sensors based on MIP modified SPCE. The novelty of this review compared to other reviews are the wide applications in several fields and the focus on the development of SPCE.

2. Molecularly Imprinted Polymers

Molecularly imprinted polymers (MIPs) are synthetic polymers that mimic specific interactions occurring in nature. These interactions include those between an antigen and its antibody, an enzyme and its substrate, or a hormone and its receptor. These interactions are generally weak and come in different types: hydrogen bonding, hydrophobic, or ionic interactions [14].
The principle of MIP synthesis is presented in (Figure 1).
Synthesizing an MIP involves selecting several reagents and determining their relative quantities. All of these choices have an impact on the morphological properties and performance of the MIPs. The reagents are the template (T), the functional monomer (FM), the solvent (S), the initiator (I), and the crosslinker (C) [16].
The template is chosen based on the application field. The choice of the functional monomer is a key step; it must demonstrate chemical complementarity to the template. The solvent must solubilize all the reagents and play the role of a porogen, which is responsible for the formation of a porous structure within the polymer [16]. The conditions for eliminating the solvent (time and temperature) play a crucial role in pore formation. The initiator allows the polymerization reaction to take place; it must be present in sufficient quantity to activate all the polymerization functions [17]. The main role of the crosslinker is to create a three-dimensional network structure within the polymer matrix [18].
It is important to make sure that all the reagents are soluble in the reaction medium. The non-solubility of one of the reagents presents a serious obstacle to successful polymerization. A non-imprinted polymer (NIP) is always synthesized in parallel with the MIP and serves as a blank polymer. The NIP is synthesized in the same way as MIPs but in the absence of the template in the reaction medium.
Molecularly imprinted polymers are mostly acrylate-based. Occasionally, sol-gel imprinted polymers using ethoxysilane or methoxysilane monomers and crosslinkers are employed. In this case, the polymerization takes place in a hydroalcoholic medium. This approach yields more eco-friendly and less toxic polymers [19,20], which are also more hydrophilic, an important criterion for applications in aqueous media. However, the main limitations of silicate-based polymers are the restricted availability of commercial monomers and the low solubility of most organic molecules in the hydroalcoholic medium.

3. Main Targets and Template

A quick search into the application of MIPs in the past 10 years in the field of electrochemical sensing reveals that the primary targets are related to explosive compounds, dyes, antioxidants, disease biomarkers, pharmaceuticals, antibiotics, allergens, pesticides, and viruses. The primary areas of application include medicine, pharmacy, the environment, public safety, and food safety (Table 1).
Examples of explosives sensing include: Triacetone triperoxide (TATP) which is a widely used explosive and is one of the main components of improvised explosive devices (IEDs). Law enforcement agencies and security experts often focus on detecting and preventing the production and use of TATP [21]. RDX is another target example of explosive sensors development: RDX is widely used. It is usually used in mixtures with some other explosives including TNT, PETN, and HMX.
Examples of dyes include synthetic dyes like sunset yellow. Determining their presence in food matrices is crucial, as they pose a potential risk to human health.
Concerning the antioxidants sensing: Catechin is a well-known natural lipid antioxidant [22] and TBHQ and BHA are synthetic antioxidants. Synthetic antioxidants are commonly used in various industries, including food, biodiesel, and cosmetics. Their role in preventing the oxidation of fats and oils make them a valuable additive. However, concerns have been raised about their potential effects on health, prompting their continuous monitoring.
Examples of disease biomarkers include cytokine interleukin-1β [23], creatinine [24], and cholesterol [25]. Cytokine interleukin-1β is a biomarker of allergic rhinitis, periodontal disease, and peri-implant disease. Its monitoring in diagnostics is of crucial importance in following up the indication and progress of cancer. Creatinine concentration in urine and blood serves as an indicator of kidney function, muscular functions, and thyroid dysfunctions. Cholesterol monitoring in the blood is very important for the heart health: It is important to maintain a balance of cholesterol levels in the body. High levels of LDL cholesterol and low levels of HDL cholesterol are often linked to an increased risk of cardiovascular diseases.
Pharmaceuticals and drugs are among the primary targets in the development of electrochemical sensors. The following targets have been studied recently in the literature: Atorvastatin belongs to a class of drugs known as statins [26]. Statins are commonly prescribed to lower cholesterol levels in the blood and reduce the risk of cardiovascular diseases, including heart attacks and strokes. Paracetamol is widely used without prescription to treat a variety of symptoms including headache, fever, arthritis, colds and pain [27]. It is considered a mild analgesic (pain reliever) and antipyretic (fever reducer). The most studied drugs are antibiotics. Several electrochemical sensors for antibiotic determination were developed: Chloramphenicol is a wide spectrum antibiotic [28]. It was commonly used to treat a variety of bacterial infections in humans and animals. Metronidazole is an antibiotic and antiprotozoal medication commonly used to treat various types of bacterial and parasitic infections [29]. It belongs to a class of drugs known as nitroimidazoles.
Only one example of allergen marker was reported. It concerns Genistein, a naturally occurring isoflavone compound that is found in various plants, particularly soybeans and soy products. It was used as a marker for the detection of soy allergens in foods [30].
Pesticides were widely studied. Among them, we found Profenofos [31] and Methidathion [32] insecticides and Isoproturon herbicide [33].
The most compelling example is the detection of HIV-1, the most common and widespread type of HIV. It is responsible for the majority of HIV infections worldwide. This detection became achievable through the selective sensing of the HIV-p24 antigen, which is the capsid protein of HIV-1. It appears at an earlier stage of HIV infection than its antibodies. Its early diagnosis helps in the prevention and control of Acquired Immune Deficiency Syndrome (AIDS) [34].
Table 1. Examples of application of MIP in electrochemical sensors using carbon-based electrode.
Table 1. Examples of application of MIP in electrochemical sensors using carbon-based electrode.
AnalyteAnalyte
Family
Matrix *ElectrodeMIP Deposition DetailsDeposition MethodDetection MethodSelectivityLODReference
Cytokine interleukin-1βProteinHuman serum solutionSPCEo-PD under-layer than electropolymerization with C2RCV electropolymerizationEISOther proteins0.23 ng/L[23]
AtorvastatinDrugWaterSPCEElectropolymerization
with 4-ABA
CV electropolymerizationDPVOther statins0.56 µg/L[26]
GenisteinAllergen markerFood
products
SPCEElectropolymerization
with o-PD
CV electropolymerizationDPVIsoflavones and
flavones
100 µg/kg[30]
ParacetamolDrugPlasmaSPCECovalent attachment of nanoMIPs after deposition of APTESChemical
grafting
DPVCaffeine, procainamide, or ethyl 4-aminobenzoate7.56 mg/L[27]
IsoproturonHerbicideGroundwater sampleGCEElectropolymerization with pyrroleCV electropolymerizationSWVCarbendazim, diuron, and carbamazepine2.2
µg/L
[33]
TryptophanAA, precursor of neurotransmittersHuman
serum
MWCNTsDrop-coating of an imprinted chitosan filmAcid
polymerization
LSVAscorbic acid,
dopamine, uric acid, and tyrosine
204 ng/L[35]
RDXExplosiveWaterGCEDrop-coating of MIP/MWCNTs mixture. MIP was prepared MAA as FMDrop castingDPVTNT and HMX4.4
ng/L
[36]
ProfenofosInsecticideVegetable samplesCNTsGrafting of SiO2 and vinyl end groups on the carboxylated CNTs, then thermal polymerization (grafting) of MIPThermal polymerizationCVCarbofuran, chlopyrifos, hydroquinone, caffeine, phenol, MgSO4 and NaCl747 ng/L[31]
Sunset
Yellow
Synthetic dyeCandy, jelly powder, juice powder, and beverageMWCNTsElectropolymerization
with acrylamide
CV electropolymerizationDPVTartrazine, erythrosine, indigo carmine, glucose, sucrose, and ascorbic acid2.26 µg/L[37]
SulfanilamideAntibioticBuffer aqueous solutionGCESynthesis of MIP by precipitation polymerization in the presence of GO. Drop-coating of MIP/GO in a chitosan matrixDrop castingSWVnot studied10
mg/L
[38]
CreatinineDisease markerBuffer aqueous solutionGCEElectropolymerization with aniline and methacrylic acid as bifunctional monomers in the presence of Ni nanoparticlesCV electropolymerizationDPVTyrosine, uric acid, dopamine, creatine, and ascorbic acid22.6 ng/L[24]
HIV-p24VirusHuman
serum
MWCNTs modified GCEGrafting of HIV-P24 protein on a drop casted chitosan layer. Polymerization of the MIP using AAM as FMRT
polymerization
DPVCEA, HCG, AFP, and BSA83
µg/L
[34]
Bisphenol AEndocrine
disruptor
Mineralized water and fresh milkGCEFunctionalization of GO with APTES. Template immobilization. Grafting of EGDMA onto the APTES coated GO. MIP thermal polymerization. GO/APTES-MIP was immobilized on a GCE using chitosanDrop castingDPVEstradiol, ethinyl estradiol, and phenol685 ng/L[39]
ParacetamolDrugPharmaceutical formulationGCENanocomposite: Oxidation of MWCNTs, then functionalization with VTMS, then MIP thermal polymerization using MAA as FM. Drop coating of the nanocompositeDrop castingSWVAcetaminophen, hydroquinone, catechol, 3,4-Dihydroxy-L-phenylalanine, ascorbic acid, and uric acid166 µg/L[40]
TBHQ, BHASynthetic
antioxidant
Soybean oil, margarine, mayonnaise, and biodieselGCENanocomposite: Oxidation of MWCNTs, then functionalization with VTMS, then MIP thermal polymerization using MAA as FM. Drop coating of the nanocompositeDrop castingDPVL-ascorbic acid, epinephrine hydrochloride, butylated hydroxyanisole, catechol, dopamine hydrochloride, hydroquinone, acetaminophen, and propyl gallate90.1 µg/L for BHA and 141.3 µg/L for TBHQ[41]
ChloramphenicolAntibioticMilk and honeyCKM-3 and P-r-GO modified GCESynthesis of a MWCNTs@MIP thermal polymerization using 3-hexadecyl-1-vinylimidazoliumchloride as functional monomer. Coating of the MWCNTs@MIP on the modified GCEDrop castingDPVGlucose, ascorbic acid, uric acid, and glutamic acid32.3 ng/L[28]
GanciclovirAntiviral drugHuman
serum
MWCNTs modified GCEElectropolymerization with 2,2’-dithiodianiline as FM in the presence of Au nanoparticlesCV electropolymerizationDPVValganciclovir, aciclovir, valaciclovir, guanine, deoxyguanosine, aniline, and cysteine383 ng/L[42]
MetronidazoleAntibioticFish meat and pharmaceutical tabletsMWCNTs modified GCEElectropolymerization with dopamineCV electropolymerizationCVRonidazole, 4-nitroimidazole, 1,2-dimethylimidazole and dimetridazole49
ng/L
[43]
MethidathionInsecticideWaste waterSPCEThermal polymerization of bulk MIP using MBAA as FM. Drop casting of MIP@sol-gel/PEG on the surface of SPCEDrop castingEISMalathion, fenthion, parathion, and chlorfenvinphos5.14 µg/L[32]
MetronidazoleAntibiotic, antiprotozoalMilk and honeyMGCESynthesis of a sol-gel and magnetic MIP using APTES as functional monomer (Fe3O4@SiO2-MIP). Coating of Fe3O4@SiO2-MIP on MGCEAttachment
using magnetic force
DPSVRonidazole and dimetridazole2.74 µg/L[29]
SucroseTable sugarSugar beet juicesMWCNTs modified GCEElectropolymerization with o-phenylenediamineCV electropolymerizationDPVRaffinose, kestose, glucose, and fructose1
mg/L
[44]
CholesterolDisease markerHydro-alcoholic solutionCCEThermal polymerization MWCNT@MIP, graphite powder, and silicon alkoxidePackingLSVCholic acid and deoxycholic acid386 ng/L[25]
* spiked samples.

4. The Rational Selection of the Functional Monomer

The selection of the functional monomer is a crucial step in the synthesis of MIPs. The most commonly used functional monomer is MAA for acrylate-based MIPs. Electropolymerization is the predominant method for depositing MIPs. In this scenario, various functional monomers are employed, including 4-ABA, o-PD, C2R, pyrrole, acrylamide, or dopamine [23,26,30,33,37,43].
An original approach consists of using bifunctional and multifunctional monomers in order to enrich the diversity of recognition sites. Rao et al. used a bifunctional monomer (methacrylic acid and aniline) for the synthesis of an MIP specific for creatinine [24]. Jiang et al. used electropolymerization of o-phenylenediamine and L-lysine for the design of an MIP at the surface of GCE for the selective detection of moxifloxacin antibiotic [45]. Dechtrirat et al. used the co-electropolymerization of 3-aminophenylboronic acid and o-phenylenediamine for the selective detection of salbutamol drug based on a graphene modified SPCE [46].
Often, the selection of the functional monomer relies on a literature search or on the theoretical chemical complementarity of interactions between the template and the functional monomer.
A rational approach to choosing the FM is through molecular modeling. It enables researchers to make informed decisions about the most suitable FMs based on rational design principles, rather than relying on time-consuming experimental trials. By utilizing computational tools, researchers can quickly identify the most appropriate MIP for a given template in terms of favorable binding characteristics [47].
In the synthesis of MIPs, molecular modeling simulates interactions between various FMs and one template, facilitating the choice of the most suitable FM based on the most stable combination (FM-Template), while considering the following factors: (i) Prediction of intermolecular interactions: Molecular modeling enables the prediction of interactions between the FM and the template molecule such as hydrogen bonding, van der Waals forces, electrostatic interactions, etc. Through the analysis of these interactions, researchers can identify which FMs exhibit the strongest and most specific binding sites for the template [48,49]. (ii) Binding site geometry: Molecular modeling provides insights into the spatial arrangement of FMs around the template molecule. This information is crucial for designing MIPs with an optimal binding site geometry. This geometry should align with the shape and size of the template molecule. A well-designed and appropriately shaped binding site enhances the selectivity and affinity of the MIP for the target molecule [50,51]. (iii) Energy calculations: Computational methods are able to calculate the interaction energy between the template and a list of FMs. Lower energy values translate to stronger and more stable binding [47,52]. (iv) Screening different monomers: With the help of molecular modelling, researchers have the possibility to screen a range of potential FMs before conducting costly and time-consuming experimental synthesis [53,54]. This means that less promising candidates can be eliminated, and efforts can be concentrated on the most promising FMs.
Several authors have employed molecular modeling and computational methods to select the most suitable functional monomers. Bakas et al. calculated the binding energy between six different functional monomers and the methidathion template. They compared the binding energy values obtained through computational simulations with the experimental charge transfer resistance of the MIP sensor prepared using the corresponding monomers. The experimental results showed a strong correlation with the computational calculations. MBAA-based MIP sensor demonstrated the largest electron transfer resistance ΔR (22.15 kΩ), which is consistent with highest binding energy between MBAA and methidathion [32]. Alanazi et al. utilized computational screening to choose the most FM for designing a MIP tailored for paracetamol. They demonstrated that the itaconic acid monomer exhibited the highest binding energy, forming six hydrogen interactions and displaying the highest interaction ratio with paracetamol among the 25 FMs tested [27]. Rebelo et al. employed quantum mechanical calculations to determine the functional monomer (FM) that forms the most stable complex with Atorvastatin. Out of ten electropolymerizable monomers, 4-ABA exhibited the highest binding energy between ABA and Atorvastatin [26].

5. Carbon-Based Electrodes

Glassy carbon electrodes (GCEs) are extensively utilized in conjunction with MIPs for shaping selective sensors. This carbon material is non-graphitizing; it does not transform into crystalline graphite upon heat treatment, resulting in a non-crystalline structure of carbon. GCEs exhibit chemical inertness, high electrical conductivity, facile functionalization, and commendable mechanical robustness. Their chemical inertness ensures stability and response accuracy. Furthermore, GCEs tolerate an extensive range of potentials, maintaining stability while manifesting minimal background current. High electrical conductivity and mechanical strength facilitate prompt electron transfer during redox reactions and durability during physical manipulation. To enhance the selective analyte detection, GCEs can be readily functionalized with an array of materials like polymers, nanoparticles, or biomolecules. GCEs are compatible with a broad spectrum of electrochemical techniques, including cyclic voltammetry, chronoamperometry, and impedance spectroscopy, among numerous others.
GCE are used alone as a base for the immobilization of MIPs, but GCE is often used in combination with GO or MWCNTs. Sadriu et al. have immobilized an MIP on the surface of a GCE using CV electropolymerization of pyrrole for the electrochemical sensing of Isoproturon herbicide in groundwater samples [33]. Rao et al. have immobilized an MIP on the surface of a GCE using CV electropolymerization of aniline and methacrylic acid as bi-functional monomers in the presence of Ni nanoparticles for the electrochemical sensing of creatinine [24].
The integration of GO into electrochemical sensors presents numerous benefits, including enhanced selectivity and simplified functionalization. By applying GO onto electrode surfaces, sensitivity and selectivity are enhanced. The substantial surface area of GO contributes to improved electron transfer kinetics and a more efficient sensor response. Electrodes modified with GO can considerably enhance the sensitivity of electrochemical sensors. The extensive surface area of GO increases interactions with the analyte, thereby lowering the detection limit of the sensor. The functionalization of GO with specific functional groups or MIPs facilitates the selective detection of target analytes. The main challenge remains in the surface electrode regeneration following the removal of MIP.
Wei et al. used a GCE modified with MIP and GO to develop a novel electrochemical sensor for the selective detection of sulfanilamide [38]. Dadkhah et al. modified a GCE with GO and MIP for the electrochemical sensing of Bisphenol A. They have used an elaborated methodology to functionalize GO with APTES. The Bisphenol A was then immobilized onto amino-functionalized GO. The next step was the grafting of EGDMA onto the APTES coated GO then the thermal polymerization of the MIP [39].
MWCNTs possess the same properties as those employed in the development of carbon-based electrochemical sensors, which include excellent electrical conductivity, mechanical stability, and the potential for functionalization. Additionally, their high specific surface area contributes to enhanced sensitivity and performance [55].
MWCNTs represent a type of nanomaterial composed of cylindrical carbon atom tubes. Their structure resembles that of SWCNTs, though with multiple layers of concentrically arranged graphene sheets. Each layer is referred to as a “wall”. The quantity of walls can vary, generally ranging from a few layers to dozens. MWNTs are an allotrope of sp2-hybridized carbon, similar to fullerenes and graphite; consequently, they exhibit significant chemical stability. MWCNTs diameters can extend up to 30 nm, in contrast to the typical SWCNT range of 0.7 to 2.0 nm. Nonetheless, SWCNTs are generally 100 to 1000 times more costly than MWCNTs.
These materials have found application as platforms for immobilizing MIPs to facilitate the specific detection of particular analytes. A straightforward and speedy approach to modify an MWCNTs electrode involves the application of a thin layer of MIP through electropolymerization. Researchers such as Arvand et al. [37], Gholivand et al. [42], Yuan et al. [43], and Shekarchizadeh et al. [44] have utilized CV electropolymerization to deposit MIP onto MWCNTs electrodes, enabling the detection of substances like sunset yellow, ganciclovir, metronidazole, and sucrose, respectively. Wu et al. [35] have drop coated a MIP-chitosan film on the surface of a GCE modified with MWCNTs for the selective sensing of tryptophan in human serum. Ma et al. drop casted a chitosan layer onto MWCNTs modified GCE. Then, they grated the HIV-p24 protein on the chitosan layer before the polymerization of a MIP. The modified electrode was used for the selective determination of human immunodeficiency virus p24 (HIV-p24) [34].
Screen-printed electrodes (SPEs) are miniaturized, disposable electrodes that fulfill the requirements of in situ monitoring, offering advantages in terms of reproducibility, cost-effectiveness, sensitivity, and portability. Additionally, in alignment with green chemistry principles, they necessitate only small sample volumes for operation, thereby reducing the usage of environmentally unfriendly solvents and minimizing waste production. SPEs represent a novel generation of miniaturized electrodes, and they are increasingly utilized in the development of electrochemical sensors due to their straightforward fabrication, low production cost, and user-friendly nature [56]. A section will be dedicated to the development of SPCEs.

6. Functionalization Methods of the Carbon-Based Electrode with MIPs

Several functionalization methods were used to deposit an MIP onto the surface of the C-based electrodes. The most widely used are electropolymerization and drop casting, and to a lesser extent chemical grafting. Other functionalization techniques that have rarely been used such as packing and magnetic force attachment will be developed.

6.1. Electropolymerization

Electropolymerization is a relatively simple process that can be conducted under mild conditions. This simplicity contributes to the ease of manufacturing sensors and their scalability. The materials employed for electropolymerization and MIP synthesis often demonstrate cost-effectiveness, rendering the production of MIP-coated electrodes economically feasible [57]. Through electropolymerization, a controlled and uniform MIP coating is achieved on the electrode surface using cyclic voltammetry [46]. The number of cycles directly correlates with the thickness of the deposited MIP layer. This uniformity guarantees reproducible results across various sensors by minimizing variations in sensor response. Electropolymerization establishes a stable and adherent MIP layer on the electrode surface, significantly reducing the possibility of detachment or degradation [58]. This layer’s durability ensures the sensor’s longevity, enabling multiple uses without a notable loss in performance.
It is the most commonly used method for the functionalization of electrodes with MIPs (Table 1). Electropolymerization has been applied directly to SPCEs, GCEs, or MWCNTs. One of the simplest examples is the electropolymerization of 4-ABA onto the surface of an SPCE for the selective sensing of atorvastatin in water samples [26].

6.2. Drop Casting

The drop casting method is relatively simple and does not necessitate complex equipment or specialized expertise. It involves placing a droplet of the MIP suspension onto the electrode surface, making it accessible to non-expert researchers. This technique is versatile and can be applied to a wide range of electrode materials and shapes. By adapting the concentration of the MIP solution and the number of casting steps, researchers can achieve a controlled and adjustable thickness of the MIP layer [59]. Drop casting is a relatively quick process, requiring less time compared to other coating methods. Gentle solvent evaporation during drop casting should be employed, as it can aid in preserving the structure and integrity of the MIP.
The major advantage of drop casting remains the possibility of synthesizing the MIP using conventional methods outside of the electrode, in the form of a monolith or powder, and then depositing the MIP on the electrode surface in a second step. This provides greater freedom in choosing the MIP formulation and the type of polymerization.
On the other hand, drop casting is suitable for small-scale production or research purposes, allowing the creation of a manageable number of sensors but not adapted to large-scale production. However, drop casting does not allow precise control over the orientation and density of binding sites on the MIP layer. While drop casting offers some control over the thickness of the MIP layer, achieving a reproducible thickness across multiple sensors can be challenging. The solvent evaporation step during drop casting can sometimes result in irregular drying, leading to cracks or defects in the MIP layer [59]. The choice of solvent for drop casting must be compatible with both the MIP and the electrode material. Care should be taken to select a solvent that does not damage the electrode or the MIP structure. If the electrode has a non-flat geometry, achieving a uniform layer of MIP through drop casting may be more difficult. PEG and chitosan are often used as immobilization matrices, but they could potentially reduce accessibility to the MIP interaction sites or introduce interferences [60].
In the examples of MIP deposition using drop casting, MIP was deposited alone or in mixture with other polymers or materials such as chitosan [39], MWCNTS [28], GO [38], or PEG [32].
In the following example (Figure 2), dos Santos Moretti et al. [40] prepared a nanocomposite by oxidation of MWCNTs, then they functionalized the oxidized material with VTMS before the polymerization of MIP using MAA as FM and TRIM as crosslinker. The prepared nanocomposite was drop casted on the surface of a GCE in the form of a suspension dispersed in DMF followed by addition of Nafion.

6.3. Chemical Grafting

Functionalizing electrodes through chemical grafting involves modifying the electrode’s surface by covalently attaching MIPs. Prior to functionalization, the electrode surface needs to be cleaned and adequately prepared to ensure the effective attachment of the functional groups of the MIPs. In many cases, the electrode surface requires activation to facilitate the covalent attachment of these functional groups. Activation methods may include chemical treatments, plasma treatments, or electrochemical approaches. The main advantage of this method is the strong attachment of MIP on the electrode, but this method is often tedious and time consuming.
In the following example (Figure 3), Kan et al. [61] functionalized a MWCNT with a MIP. Firstly, the surface of the MWCNT was activated using HNO3. Then, MWCNTs were treated with sulfoxide chloride (SOCl2) and allyl alcohol in order to transform the surface of the MWCNTs into polymerizable vinyl group (MWCNTs-CH=CH2). The final step was a selective polymerization of a MIP using dopamine as a template, TRIM as a crosslinker, and AIBN as an initiator. The obtained functionalized material was drop casted at the surface of glassy carbon electrode for the electrochemical sensing of dopamine.

6.4. Other Functionalization Methods

In this paragraph, we will discuss other infrequently utilized functionalization methods.
Tong et al. [25] employed packing of a mixture of MIP, MWCNTs, graphite powder, and silicon alkoxide to functionalize a ceramic carbon electrode (CCE). The resulting functionalized electrode was utilized for electrochemical cholesterol sensing. A notable advantage of this packing method is the flexibility it offers in synthesizing the MIP externally to the electrode. However, a significant drawback to consider is the potential risk of MIP fragments or monomers leaching into the electrolyte or the environment, leading to system contamination and potential performance effects.
In an innovative approach, Chen et al. [29] used magnetic forces to functionalize the surface of a magnetic glassy carbon electrode with a magnetic MIP. This functionalized electrode demonstrated efficacy in selectively sensing the metronidazole antibiotic in milk and honey samples.

6.5. Comparison of Functionalization Methods

Table 2 summarizes the main advantages and drawbacks of the functionalization methods.

7. Electrochemical Detection Methods

Several electrochemical detection methods were used for the sensing of organic compounds: CV, DPV, SWV, LSV, DPSV, and EIS.
DPV remains the most used transduction method. DPV enhances the sensitivity of electrochemical measurements by applying a minor voltage pulse at each potential step [62]. This leads to an improved signal-to-noise ratio and better detection limits when compared to other voltammetric techniques. DPV reduces the influence of capacitive currents, which are often met in cyclic voltammetry and can complicate data interpretation [63]. The distinctive measurement format of DPV facilitates the differentiation between faradaic and capacitive currents, yielding clearer and more precise quantification [64]. DPV displays reduced susceptibility to interference from non-faradaic processes. It is a suitable technique for complex sample matrices.
The lowest limit of detection obtained with DPV was 4.4 ng/L. It was obtained after drop casting a MIP/MWCNTs mixture onto a GCE for the detection of RDX in water [36]. This value is the second lowest, following the one determined by EIS during the electrochemical detection of the cytokine interleukin-1β protein after the electropolymerization of an MIP on the surface of an SPCE. In this case, the limit of detection was as low as 0.23 ng/L [23].
Electrochemical impedance spectroscopy (EIS) offers several distinct advantages when applied to electrochemical sensors. EIS is a quantitative technique which allows the determination of analyte concentrations based on impedance changes [65]. Moreover, EIS shows high sensitivity to alterations in the surface properties of electrodes, and it is highly effective at detecting small variations in analyte concentration or surface conditions [66]. This sensitivity is crucial for sensor applications. EIS does not necessitate the detected species to be electroactive, it involves the use of a redox probe in solution [67]. This simplifies the detection process and reduces costs. EIS is a non-invasive technique, it preserves sample integrity [68]. It is a particularly advantageous technique for applications where sample integrity must be maintained. Furthermore, EIS excels in low-frequency analysis, which is highly important for studying biological interactions, redox processes, and reactions involving slow kinetics [69,70].

8. Screen-Printed Carbon Electrode

Screen-printed carbon electrodes (SPCEs) are increasingly being used in electrochemical sensing and analysis. These electrodes are produced using a screen-printing process, which deposits a layer of carbon ink onto a substrate, typically ceramic, plastic, or paper. The screen-printing process employed in their manufacture is relatively straightforward and cost-effective, facilitating mass production while maintaining consistent quality [71]. SPCEs are designed for disposable use, making them well-suited for one-time applications or field usage. Their compact dimensions and lightweight construction contribute to their portability. Generally, the carbon surface of SPCEs exhibits a low tendency to accumulate fouling, ensuring stable and reproducible measurements over extended periods [72].
Moreover, SPCEs have a broad electrochemical potential window, which allows the investigation of various redox processes without experiencing electrode degradation [73]. The limited working area of SPCEs allows for the use of reduced sample volumes, which is highly important when working with scarce or expensive samples [74]. SPCEs are compatible with a range of electrochemical techniques, such as CV, DPV, SWV, chronoamperometry and EIS.
Additionally, SPCEs can be easily functionalized through the incorporation of different MIP materials. Table 3 provides a summary of examples illustrating the functionalization of SPCEs using MIPs.
Most of the applications concern drugs, antibiotics, and disease biomarkers. Applications also include pesticides, food toxins, and allergens. The most developed field of application is biomedical analysis, but fields such as food safety, environmental monitoring, and civil safety are growing rapidly.
The easiest, fastest, and most cost-effective method for SPCE functionalization is electropolymerization. This is why it was used in the majority of examples found in the literature. Drop coating is the second most used method for the MIP deposition onto the surface of SPCE.
Figure 4 presents an example of the functionalization of SPCE with MIP using electropolymerization of a monomer combination: 3-aminophenylboronic acid and o-phenylenediamine, in the presence of the L-hydroxyproline template. The developed electrochemical sensor was employed for the selective determination of L-hydroxyproline, a biomarker for bone disease, in human serum samples. This was achieved using EIS as the transduction method [88].
The most used electrochemical methods for the detection of analytes are DPV and SWV. EIS is often used when the analyte is not electroactive. The lowest limit of detection recorded was in the range of pM for Tau protein and insulin [87,92].

9. Conclusions and Future Directions

In conclusion, electrochemical sensors using carbon electrodes functionalized with MIPs have found applications in various fields. Among these, biomedical analysis has seen the most significant development, including the analysis of drugs, antibiotics, and disease biomarkers. Emerging applications involve the monitoring of environmental and food contaminants. Analytes vary widely, from small molecules like sucrose to 62 kDa proteins such as tau protein. One remarkable application demonstrated the selective detection of HIV viruses using an MWCNTs modified GCE and functionalized with an AAM-based MIP.
Miniature electrodes such as SPCEs are rapidly gaining attractiveness, attributed to their affordability, user-friendly nature, and disposability, which avoids the requirement for frequent maintenance seen in conventional electrodes. The inaugural application of SPCE functionalization with an MIP dates back to 1999; however, it was not until 2010 that the first real-time detection emerged. In the period covering 2010 to 2018, only six applications were documented in literature. It was not until 2019 that the utilization of MIP-modified SPCEs saw significant development, as evidenced by the publication of 4 to 5 articles annually.
Future directions lie in the development of MIP-based electrochemical sensors for the detection of contaminants for food safety and environment monitoring at low cost. The commercialization of sensitive and selective MIP-based electrochemical sensors remains a challenging objective, despite various successful attempts at the laboratory scale. The primary obstacles include relatively high costs associated with extended development periods and the capability to simultaneously detect multiple targets. Researchers often omit the timeframe required for developing new sensors and may not share unsuccessful results, which can be both time-consuming and discouraging. While MIPs have proven effective, they have, thus far, only been applied to the electrochemical detection of individual targets and cannot be utilized for multi-target detection, such as the simultaneous determination of multiple pesticide residues or food and environmental pollutants. One of the main challenges that persists is the creation of MIP-based multi-target sensors.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations

4-ABA4-aminobenzoic acid
AAAmino Acid
AAMAcrylamide
AFPAlpha fetal protein
APTES3-Aminopropyltriethoxysilane
BHAButylated hydroxyanisole
BSABovine serum albumin
C2RChromotrope 2R
CAChronoamperometry
CCECeramic carbon electrode
CEACarcinoembryonic antigen
CKM-3Mesoporous carbon
CNTsCarbon nanotube
CVCyclic voltammetry
DMFN,N-Dimethylformamide
DPSVDifferential pulse stripping voltammetry
EGDMAEthylene glycol dimethacrylate
EISElectrochemical impedance spectroscopy
FcMMAFerrocenyl-methylmethacrylate
FMFunctional monomer
GCEGlassy carbon electrode
GOGraphene oxide
HCGHuman chorionic gonadotropin
HIVHuman immune deficiency virus
HMX1,3,5,7-tetranitro-1,3,5,7-tetrazocane
LSVLinear sweep voltammetry
MAAMethacrylic acid
MBAAN-methylenebisacrylamide
MDMA3,4-methylenedioxymethamphetamine
MGCEMagnetic glassy carbon electrode
MWCNTMulti-walled carbon nanotube
o-PDO-phenylenediamine
PEGPoly ethylene glycol
PETNPentaerythritol tetranitrate
P-r-GO3-dimensional porous graphene
RDXTrinitroperhydro-1,3,5-triazine
RTRoom temperature
SPCEScreen printed carbon electrode
SWCNTSingle-walled carbon nanotube
SWVSquare wave voltammetry
TBHQTert-butylhydroquinone
TNT2,4,6-trinitrotoluene
TRIMTrimethylolpropane trimethacrylate
VTMSVinyltrimethoxysilane

References

  1. Nekrasov, N.; Jaric, S.; Kireev, D.; Emelianov, A.V.; Orlov, A.V.; Gadjanski, I.; Nikitin, P.I.; Akinwande, D.; Bobrinetskiy, I. Real-Time Detection of Ochratoxin A in Wine through Insight of Aptamer Conformation in Conjunction with Graphene Field-Effect Transistor. Biosens. Bioelectron. 2022, 200, 113890. [Google Scholar] [CrossRef]
  2. Yang, T.; Huang, H.; Zhu, F.; Lin, Q.; Zhang, L.; Liu, J. Recent Progresses in Nanobiosensing for Food Safety Analysis. Sensors 2016, 16, 1118. [Google Scholar] [CrossRef] [PubMed]
  3. Adumitrăchioaie, A.; Tertiș, M.; Cernat, A.; Săndulescu, R.; Cristea, C. Electrochemical Methods Based on Molecularly Imprinted Polymers for Drug Detection. A Review. Int. J. Electrochem. Sci. 2018, 13, 2556–2576. [Google Scholar] [CrossRef]
  4. Rebollar-Pérez, G.; Campos-Terán, J.; Ornelas-Soto, N.; Méndez-Albores, A.; Torres, E. Biosensors Based on Oxidative Enzymes for Detection of Environmental Pollutants. Biocatalysis 2016, 1, 118–129. [Google Scholar] [CrossRef]
  5. Staden, R.-I.S. Perspective—Challenges in Biomedical Analysis: From Classical Sensors to Stochastic Sensors. ECS Sens. Plus 2022, 1, 011603. [Google Scholar] [CrossRef]
  6. Leva-Bueno, J.; Peyman, S.A.; Millner, P.A. A Review on Impedimetric Immunosensors for Pathogen and Biomarker Detection. Med. Microbiol. Immunol. 2020, 209, 343–362. [Google Scholar] [CrossRef] [PubMed]
  7. Kharajinezhadian, R.; Javad Chaichi, M.; Nazari, O.; Mansour Lakouraj, M.; Hasantabar, V. Fraud Monitoring Using a New Disposable Photoluminescence Sensor in Milk. Microchem. J. 2023, 189, 108437. [Google Scholar] [CrossRef]
  8. Apak, R.; Üzer, A.; Sağlam, Ş.; Arman, A. Selective Electrochemical Detection of Explosives with Nanomaterial Based Electrodes. Electroanalysis 2023, 35, e202200175. [Google Scholar] [CrossRef]
  9. Karimi-Maleh, H.; Karimi, F.; Alizadeh, M.; Sanati, A.L. Electrochemical Sensors, a Bright Future in the Fabrication of Portable Kits in Analytical Systems. Chem. Rec. 2020, 20, 682–692. [Google Scholar] [CrossRef]
  10. Adarakatti, P.S.; Kempahanumakkagari, S.K. Modified Electrodes for Sensing. In Electrochemistry; Banks, C., McIntosh, S., Eds.; Royal Society of Chemistry: Cambridge, UK, 2018; Volume 15, pp. 58–95. ISBN 978-1-78801-373-4. [Google Scholar]
  11. Ayerdurai, V.; Cieplak, M.; Kutner, W. Molecularly Imprinted Polymer-Based Electrochemical Sensors for Food Contaminants Determination. TrAC Trends Anal. Chem. 2023, 158, 116830. [Google Scholar] [CrossRef]
  12. Karimi-Maleh, H.; Beitollahi, H.; Senthil Kumar, P.; Tajik, S.; Mohammadzadeh Jahani, P.; Karimi, F.; Karaman, C.; Vasseghian, Y.; Baghayeri, M.; Rouhi, J.; et al. Recent Advances in Carbon Nanomaterials-Based Electrochemical Sensors for Food Azo Dyes Detection. Food Chem. Toxicol. 2022, 164, 112961. [Google Scholar] [CrossRef] [PubMed]
  13. Kumar Singh, A.; Agrahari, S.; Kumar Gautam, R.; Tiwari, I. Fabrication of a Novel Screen-Printed Carbon Electrode Based Disposable Sensor for Sensitive Determination of an Endocrine Disruptor BPSIP in Environmental and Biological Matrices. Microchem. J. 2023, 193, 109031. [Google Scholar] [CrossRef]
  14. Nicholls, I.A.; Andersson, H.S.; Charlton, C.; Henschel, H.; Karlsson, B.C.G.; Karlsson, J.G.; O’Mahony, J.; Rosengren, A.M.; Rosengren, K.J.; Wikman, S. Theoretical and Computational Strategies for Rational Molecularly Imprinted Polymer Design. Biosens. Bioelectron. 2009, 25, 543–552. [Google Scholar] [CrossRef]
  15. Haupt, K. Molecularly Imprinted Polymers in Analytical Chemistry. Analyst 2001, 126, 747–756. [Google Scholar] [CrossRef] [PubMed]
  16. Cormack, P.A.G.; Elorza, A.Z. Molecularly Imprinted Polymers: Synthesis and Characterisation. J. Chromatogr. B Anal. Technol. Biomed Life Sci. 2004, 804, 173–182. [Google Scholar] [CrossRef] [PubMed]
  17. Mijangos, I.; Navarro-Villoslada, F.; Guerreiro, A.; Piletska, E.; Chianella, I.; Karim, K.; Turner, A.; Piletsky, S. Influence of Initiator and Different Polymerisation Conditions on Performance of Molecularly Imprinted Polymers. Biosens. Bioelectron. 2006, 22, 381–387. [Google Scholar] [CrossRef] [PubMed]
  18. Yan, H.; Row, K.H. Characteristic and Synthetic Approach of Molecularly Imprinted Polymer. Int. J. Mol. Sci. 2006, 7, 155–178. [Google Scholar] [CrossRef]
  19. Séverin, I.; Lionti, K.; Dahbi, L.; Loriot, C.; Toury, B.; Chagnon, M.-C. In Vitro Toxicity Assessment of Extracts Derived from Sol–Gel Coatings on Polycarbonate Intended to Be Used in Food Contact Applications. Food Chem. Toxicol. 2016, 93, 51–57. [Google Scholar] [CrossRef]
  20. Bitar, M.; Lafarge, C.; Sok, N.; Cayot, P.; Bou-Maroun, E. Molecularly Imprinted Sol-Gel Polymers for the Analysis of Iprodione Fungicide in Wine: Synthesis in Green Solvent. Food Chem. 2019, 293, 226–232. [Google Scholar] [CrossRef]
  21. Mamo, S.K.; Gonzalez-Rodriguez, J. Development of a Molecularly Imprinted Polymer-Based Sensor for the Electrochemical Determination of Triacetone Triperoxide (TATP). Sensors 2014, 14, 23269–23282. [Google Scholar] [CrossRef] [PubMed]
  22. Lu, Z.; Du, X.; Sun, M.; Zhang, Y.; Li, Y.; Wang, X.; Wang, Y.; Du, H.; Yin, H.; Rao, H. Novel Dual-Template Molecular Imprinted Electrochemical Sensor for Simultaneous Detection of CA and TPH Based on Peanut Twin-like NiFe2O4/CoFe2O4/NCDs Nanospheres: Fabrication, Application and DFT Theoretical Study. Biosens. Bioelectron. 2021, 190, 113408. [Google Scholar] [CrossRef] [PubMed]
  23. Choi, D.Y.; Yang, J.C.; Hong, S.W.; Park, J. Molecularly Imprinted Polymer-Based Electrochemical Impedimetric Sensors on Screen-Printed Carbon Electrodes for the Detection of Trace Cytokine IL-1β. Biosens. Bioelectron. 2022, 204, 114073. [Google Scholar] [CrossRef] [PubMed]
  24. Rao, H.; Lu, Z.; Ge, H.; Liu, X.; Chen, B.; Zou, P.; Wang, X.; He, H.; Zeng, X.; Wang, Y. Electrochemical Creatinine Sensor Based on a Glassy Carbon Electrode Modified with a Molecularly Imprinted Polymer and a Ni@polyaniline Nanocomposite. Microchim. Acta 2017, 184, 261–269. [Google Scholar] [CrossRef]
  25. Tong, Y.; Li, H.; Guan, H.; Zhao, J.; Majeed, S.; Anjum, S.; Liang, F.; Xu, G. Electrochemical Cholesterol Sensor Based on Carbon Nanotube@molecularly Imprinted Polymer Modified Ceramic Carbon Electrode. Biosens. Bioelectron. 2013, 47, 553–558. [Google Scholar] [CrossRef]
  26. Rebelo, P.; Pacheco, J.G.; Voroshylova, I.V.; Melo, A.; Cordeiro, M.N.D.S.; Delerue-Matos, C. A Simple Electrochemical Detection of Atorvastatin Based on Disposable Screen-Printed Carbon Electrodes Modified by Molecularly Imprinted Polymer: Experiment and Simulation. Anal. Chim. Acta 2022, 1194, 339410. [Google Scholar] [CrossRef]
  27. Alanazi, K.; Garcia Cruz, A.; Di Masi, S.; Voorhaar, A.; Ahmad, O.S.; Cowen, T.; Piletska, E.; Langford, N.; Coats, T.J.; Sims, M.R.; et al. Disposable Paracetamol Sensor Based on Electroactive Molecularly Imprinted Polymer Nanoparticles for Plasma Monitoring. Sens. Actuators B Chem. 2021, 329, 129128. [Google Scholar] [CrossRef]
  28. Yang, G.; Zhao, F. Electrochemical Sensor for Chloramphenicol Based on Novel Multiwalled Carbon Nanotubes@molecularly Imprinted Polymer. Biosens. Bioelectron. 2015, 64, 416–422. [Google Scholar] [CrossRef]
  29. Chen, D.; Deng, J.; Liang, J.; Xie, J.; Hu, C.; Huang, K. A Core–Shell Molecularly Imprinted Polymer Grafted onto a Magnetic Glassy Carbon Electrode as a Selective Sensor for the Determination of Metronidazole. Sens. Actuators B Chem. 2013, 183, 594–600. [Google Scholar] [CrossRef]
  30. Sundhoro, M.; Agnihotra, S.R.; Amberger, B.; Augustus, K.; Khan, N.D.; Barnes, A.; BelBruno, J.; Mendecki, L. An Electrochemical Molecularly Imprinted Polymer Sensor for Rapid and Selective Food Allergen Detection. Food Chem. 2021, 344, 128648. [Google Scholar] [CrossRef] [PubMed]
  31. Amatatongchai, M.; Sroysee, W.; Sodkrathok, P.; Kesangam, N.; Chairam, S.; Jarujamrus, P. Novel Three-Dimensional Molecularly Imprinted Polymer-Coated Carbon Nanotubes (3D-CNTs@MIP) for Selective Detection of Profenofos in Food. Anal. Chim. Acta 2019, 1076, 64–72. [Google Scholar] [CrossRef]
  32. Bakas, I.; Hayat, A.; Piletsky, S.; Piletska, E.; Chehimi, M.M.; Noguer, T.; Rouillon, R. Electrochemical Impedimetric Sensor Based on Molecularly Imprinted Polymers/Sol–Gel Chemistry for Methidathion Organophosphorous Insecticide Recognition. Talanta 2014, 130, 294–298. [Google Scholar] [CrossRef]
  33. Sadriu, I.; Bouden, S.; Nicolle, J.; Podvorica, F.I.; Bertagna, V.; Berho, C.; Amalric, L.; Vautrin-Ul, C. Molecularly Imprinted Polymer Modified Glassy Carbon Electrodes for the Electrochemical Analysis of Isoproturon in Water. Talanta 2020, 207, 120222. [Google Scholar] [CrossRef]
  34. Ma, Y.; Shen, X.-L.; Zeng, Q.; Wang, H.-S.; Wang, L.-S. A Multi-Walled Carbon Nanotubes Based Molecularly Imprinted Polymers Electrochemical Sensor for the Sensitive Determination of HIV-P24. Talanta 2017, 164, 121–127. [Google Scholar] [CrossRef] [PubMed]
  35. Wu, Y.; Deng, P.; Tian, Y.; Ding, Z.; Li, G.; Liu, J.; Zuberi, Z.; He, Q. Rapid Recognition and Determination of Tryptophan by Carbon Nanotubes and Molecularly Imprinted Polymer-Modified Glassy Carbon Electrode. Bioelectrochemistry 2020, 131, 107393. [Google Scholar] [CrossRef] [PubMed]
  36. Alizadeh, T.; Atashi, F.; Ganjali, M.R. Molecularly Imprinted Polymer Nano-Sphere/Multi-Walled Carbon Nanotube Coated Glassy Carbon Electrode as an Ultra-Sensitive Voltammetric Sensor for Picomolar Level Determination of RDX. Talanta 2019, 194, 415–421. [Google Scholar] [CrossRef]
  37. Arvand, M.; Zamani, M.; Sayyar Ardaki, M. Rapid Electrochemical Synthesis of Molecularly Imprinted Polymers on Functionalized Multi-Walled Carbon Nanotubes for Selective Recognition of Sunset Yellow in Food Samples. Sens. Actuators B Chem. 2017, 243, 927–939. [Google Scholar] [CrossRef]
  38. Wei, X.; Xu, X.; Qi, W.; Wu, Y.; Wang, L. Molecularly Imprinted Polymer/Graphene Oxide Modified Glassy Carbon Electrode for Selective Detection of Sulfanilamide. Prog. Nat. Sci. Mater. Int. 2017, 27, 374–379. [Google Scholar] [CrossRef]
  39. Dadkhah, S.; Ziaei, E.; Mehdinia, A.; Baradaran Kayyal, T.; Jabbari, A. A Glassy Carbon Electrode Modified with Amino-Functionalized Graphene Oxide and Molecularly Imprinted Polymer for Electrochemical Sensing of Bisphenol A. Microchim. Acta 2016, 183, 1933–1941. [Google Scholar] [CrossRef]
  40. dos Santos Moretti, E.; de Fátima Giarola, J.; Kuceki, M.; Prete, M.C.; Pereira, A.C.; Tarley, C.R.T. A Nanocomposite Based on Multi-Walled Carbon Nanotubes Grafted by Molecularly Imprinted Poly(Methacrylic Acid–Hemin) as a Peroxidase-like Catalyst for Biomimetic Sensing of Acetaminophen. RSC Adv. 2016, 6, 28751–28760. [Google Scholar] [CrossRef]
  41. dos Santos Moretti, E.; de Oliveira, F.M.; Scheel, G.L.; DalĺAntônia, L.H.; Borsato, D.; Kubota, L.T.; Segatelli, M.G.; Tarley, C.R.T. Synthesis of Surface Molecularly Imprinted Poly(Methacrylic Acid-Hemin) on Carbon Nanotubes for the Voltammetric Simultaneous Determination of Antioxidants from Lipid Matrices and Biodiesel. Electrochim. Acta 2016, 212, 322–332. [Google Scholar] [CrossRef]
  42. Gholivand, M.B.; Karimian, N. Fabrication of a Highly Selective and Sensitive Voltammetric Ganciclovir Sensor Based on Electropolymerized Molecularly Imprinted Polymer and Gold Nanoparticles on Multiwall Carbon Nanotubes/Glassy Carbon Electrode. Sens. Actuators B Chem. 2015, 215, 471–479. [Google Scholar] [CrossRef]
  43. Yuan, L.; Jiang, L.; Hui, T.; Jie, L.; Bingbin, X.; Feng, Y.; Yingchun, L. Fabrication of Highly Sensitive and Selective Electrochemical Sensor by Using Optimized Molecularly Imprinted Polymers on Multi-Walled Carbon Nanotubes for Metronidazole Measurement. Sens. Actuators B Chem. 2015, 206, 647–652. [Google Scholar] [CrossRef]
  44. Shekarchizadeh, H.; Ensafi, A.A.; Kadivar, M. Selective Determination of Sucrose Based on Electropolymerized Molecularly Imprinted Polymer Modified Multiwall Carbon Nanotubes/Glassy Carbon Electrode. Mater. Sci. Eng. C 2013, 33, 3553–3561. [Google Scholar] [CrossRef] [PubMed]
  45. Jiang, Z.; Li, G.; Zhang, M. A Novel Sensor Based on Bifunctional Monomer Molecularly Imprinted Film at Graphene Modified Glassy Carbon Electrode for Detecting Traces of Moxifloxacin. RSC Adv. 2016, 6, 32915–32921. [Google Scholar] [CrossRef]
  46. Dechtrirat, D.; Sookcharoenpinyo, B.; Prajongtat, P.; Sriprachuabwong, C.; Sanguankiat, A.; Tuantranont, A.; Hannongbua, S. An Electrochemical MIP Sensor for Selective Detection of Salbutamol Based on a Graphene/PEDOT:PSS Modified Screen Printed Carbon Electrode. RSC Adv. 2017, 8, 206–212. [Google Scholar] [CrossRef]
  47. Dong, C.; Li, X.; Guo, Z.; Qi, J. Development of a Model for the Rational Design of Molecular Imprinted Polymer: Computational Approach for Combined Molecular Dynamics/Quantum Mechanics Calculations. Anal. Chim. Acta 2009, 647, 117–124. [Google Scholar] [CrossRef] [PubMed]
  48. Liu, Z.; Xu, Z.; Wang, D.; Yang, Y.; Duan, Y.; Ma, L.; Lin, T.; Liu, H. A Review on Molecularly Imprinted Polymers Preparation by Computational Simulation-Aided Methods. Polymers 2021, 13, 2657. [Google Scholar] [CrossRef] [PubMed]
  49. Xie, L.; Xiao, N.; Li, L.; Xie, X.; Li, Y. An Investigation of the Intermolecular Interactions and Recognition Properties of Molecular Imprinted Polymers for Deltamethrin through Computational Strategies. Polymers 2019, 11, 1872. [Google Scholar] [CrossRef]
  50. Terracina, J.J.; Bergkvist, M.; Sharfstein, S.T. Computational Investigation of Stoichiometric Effects, Binding Site Heterogeneities, and Selectivities of Molecularly Imprinted Polymers. J. Mol. Model. 2016, 22, 139. [Google Scholar] [CrossRef] [PubMed]
  51. Zhao, W.; Liu, J.; Tang, S.; Jin, R. Theoretical Research of Molecular Imprinted Polymers Formed from Formaldehyde and Methacrylic Acid. J. Mol. Model. 2020, 26, 88. [Google Scholar] [CrossRef]
  52. Bitar, M.; Bou-Maroun, E.; Lerbret, A.; Ouaini, N.; Cayot, P. Binding Characteristics of Molecularly Imprinted Polymers Based on Fungicides in Hydroalcoholic Media. J. Sep. Sci. 2015, 38, 3607–3614. [Google Scholar] [CrossRef]
  53. He, Q.; Liang, J.-J.; Chen, L.-X.; Chen, S.-L.; Zheng, H.-L.; Liu, H.-X.; Zhang, H.-J. Removal of the Environmental Pollutant Carbamazepine Using Molecular Imprinted Adsorbents: Molecular Simulation, Adsorption Properties, and Mechanisms. Water Res. 2020, 168, 115164. [Google Scholar] [CrossRef]
  54. Ao, J.; Gu, J.; Yuan, T.; Li, D.; Ma, Y.; Shen, Z. Applying Molecular Modelling and Experimental Studies to Develop Molecularly Imprinted Polymer for Domoic Acid Enrichment from Both Seawater and Shellfish. Chemosphere 2018, 199, 98–106. [Google Scholar] [CrossRef] [PubMed]
  55. Xiong, H.; Zhao, Y.; Liu, P.; Zhang, X.; Wang, S. Electrochemical Properties and the Determination of Nicotine at a Multi-Walled Carbon Nanotubes Modified Glassy Carbon Electrode. Microchim. Acta 2010, 168, 31–36. [Google Scholar] [CrossRef]
  56. Smart, A.; Crew, A.; Pemberton, R.; Hughes, G.; Doran, O.; Hart, J.P. Screen-Printed Carbon Based Biosensors and Their Applications in Agri-Food Safety. TrAC Trends Anal. Chem. 2020, 127, 115898. [Google Scholar] [CrossRef]
  57. Crapnell, R.D.; Dempsey-Hibbert, N.C.; Peeters, M.; Tridente, A.; Banks, C.E. Molecularly Imprinted Polymer Based Electrochemical Biosensors: Overcoming the Challenges of Detecting Vital Biomarkers and Speeding up Diagnosis. Talanta Open 2020, 2, 100018. [Google Scholar] [CrossRef]
  58. Wang, Q.; Paim, L.L.; Zhang, X.; Wang, S.; Stradiotto, N.R. An Electrochemical Sensor for Reducing Sugars Based on a Glassy Carbon Electrode Modified with Electropolymerized Molecularly Imprinted Poly-o-Phenylenediamine Film. Electroanalysis 2014, 26, 1612–1622. [Google Scholar] [CrossRef]
  59. Rayanasukha, Y.; Pratontep, S.; Porntheeraphat, S.; Bunjongpru, W.; Nukeaw, J. Non-Enzymatic Urea Sensor Using Molecularly Imprinted Polymers Surface Modified Based-on Ion-Sensitive Field Effect Transistor (ISFET). Surf. Coat. Technol. 2016, 306, 147–150. [Google Scholar] [CrossRef]
  60. Karrat, A.; Lamaoui, A.; Amine, A.; Palacios-Santander, J.M.; Cubillana-Aguilera, L. Applications of Chitosan in Molecularly and Ion Imprinted Polymers. Chem. Afr. 2020, 3, 513–533. [Google Scholar] [CrossRef]
  61. Kan, X.; Zhao, Y.; Geng, Z.; Wang, Z.; Zhu, J.-J. Composites of Multiwalled Carbon Nanotubes and Molecularly Imprinted Polymers for Dopamine Recognition. J. Phys. Chem. C 2008, 112, 4849–4854. [Google Scholar] [CrossRef]
  62. Dourandish, Z.; Beitollahi, H. Electrochemical Sensing of Isoproterenol Using Graphite Screen-Printed Electrode Modified with Graphene Quantum Dots. Anal. Bioanal. Electrochem. 2018, 10, 192–202. [Google Scholar]
  63. Sharma, A.; Arya, S.; Chauhan, D.; Solanki, P.R.; Khajuria, S.; Khosla, A. Synthesis of Au–SnO2 Nanoparticles for Electrochemical Determination of Vitamin B12. J. Mater. Res. Technol. 2020, 9, 14321–14337. [Google Scholar] [CrossRef]
  64. Rai, V.; Deng, J.; Toh, C.-S. Electrochemical Nanoporous Alumina Membrane-Based Label-Free DNA Biosensor for the Detection of Legionella Sp. Talanta 2012, 98, 112–117. [Google Scholar] [CrossRef] [PubMed]
  65. 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]
  66. Lisdat, F.; Schäfer, D. The Use of Electrochemical Impedance Spectroscopy for Biosensing. Anal. Bioanal. Chem. 2008, 391, 1555–1567. [Google Scholar] [CrossRef] [PubMed]
  67. ErtuğruL, H.D.; Uygun, Z.O.; ErtuğruL, H.D.; Uygun, Z.O. Impedimetric Biosensors for Label-Free and Enzymless Detection. In State of the Art in Biosensors—General Aspects; IntechOpen: London, UK, 2013; ISBN 978-953-51-1004-0. [Google Scholar]
  68. Laschuk, N.O.; Easton, E.B.; Zenkina, O.V. Reducing the Resistance for the Use of Electrochemical Impedance Spectroscopy Analysis in Materials Chemistry. RSC Adv. 2021, 11, 27925–27936. [Google Scholar] [CrossRef] [PubMed]
  69. Skalová, Š.; Vyskočil, V.; Barek, J.; Navrátil, T. Model Biological Membranes and Possibilities of Application of Electrochemical Impedance Spectroscopy for Their Characterization. Electroanalysis 2018, 30, 207–219. [Google Scholar] [CrossRef]
  70. Martins, J.C.; Neto, J.C.D.M.; Passos, R.R.; Pocrifka, L.A. Electrochemical Behavior of Polyaniline: A Study by Electrochemical Impedance Spectroscopy (EIS) in Low-Frequency. Solid State Ion. 2020, 346, 115198. [Google Scholar] [CrossRef]
  71. Washe, A.P.; Lozano-Sánchez, P.; Bejarano-Nosas, D.; Katakis, I. Facile and Versatile Approaches to Enhancing Electrochemical Performance of Screen Printed Electrodes. Electrochim. Acta 2013, 91, 166–172. [Google Scholar] [CrossRef]
  72. Mazzaracchio, V.; Tomei, M.R.; Cacciotti, I.; Chiodoni, A.; Novara, C.; Castellino, M.; Scordo, G.; Amine, A.; Moscone, D.; Arduini, F. Inside the Different Types of Carbon Black as Nanomodifiers for Screen-Printed Electrodes. Electrochim. Acta 2019, 317, 673–683. [Google Scholar] [CrossRef]
  73. Niu, X.; Chen, C.; Zhao, H.; Tang, J.; Li, Y.; Lan, M. Porous Screen-Printed Carbon Electrode. Electrochem. Commun. 2012, 22, 170–173. [Google Scholar] [CrossRef]
  74. Araújo, D.A.G.; Camargo, J.R.; Pradela-Filho, L.A.; Lima, A.P.; Muñoz, R.A.A.; Takeuchi, R.M.; Janegitz, B.C.; Santos, A.L. A Lab-Made Screen-Printed Electrode as a Platform to Study the Effect of the Size and Functionalization of Carbon Nanotubes on the Voltammetric Determination of Caffeic Acid. Microchem. J. 2020, 158, 105297. [Google Scholar] [CrossRef]
  75. Pellicer, C.; Gomez-Caballero, A.; Unceta, N.; Goicolea, M.A.; Barrio, R.J. Using a Portable Device Based on a Screen-Printed Sensor Modified with a Molecularly Imprinted Polymer for the Determination of the Insecticide Fenitrothion in Forest Samples. Anal. Methods 2010, 2, 1280–1285. [Google Scholar] [CrossRef]
  76. Kumar, D.; Prasad, B.B. Multiwalled Carbon Nanotubes Embedded Molecularly Imprinted Polymer-Modified Screen Printed Carbon Electrode for the Quantitative Analysis of C-Reactive Protein. Sens. Actuators B Chem. 2012, 171–172, 1141–1150. [Google Scholar] [CrossRef]
  77. Radi, A.-E.; El-Naggar, A.-E.; Nassef, H.M. Molecularly Imprinted Polymer Based Electrochemical Sensor for the Determination of the Anthelmintic Drug Oxfendazole. J. Electroanal. Chem. 2014, 729, 135–141. [Google Scholar] [CrossRef]
  78. Lopes, F.; Pacheco, J.G.; Rebelo, P.; Delerue-Matos, C. Molecularly Imprinted Electrochemical Sensor Prepared on a Screen Printed Carbon Electrode for Naloxone Detection. Sens. Actuators B Chem. 2017, 243, 745–752. [Google Scholar] [CrossRef]
  79. Ekomo, V.M.; Branger, C.; Bikanga, R.; Florea, A.-M.; Istamboulie, G.; Calas-Blanchard, C.; Noguer, T.; Sarbu, A.; Brisset, H. Detection of Bisphenol A in Aqueous Medium by Screen Printed Carbon Electrodes Incorporating Electrochemical Molecularly Imprinted Polymers. Biosens. Bioelectron. 2018, 112, 156–161. [Google Scholar] [CrossRef]
  80. Pacheco, J.G.; Rebelo, P.; Cagide, F.; Gonçalves, L.M.; Borges, F.; Rodrigues, J.A.; Delerue-Matos, C. Electrochemical Sensing of the Thyroid Hormone Thyronamine (T0AM) via Molecular Imprinted Polymers (MIPs). Talanta 2019, 194, 689–696. [Google Scholar] [CrossRef]
  81. Couto, R.A.S.; Costa, S.S.; Mounssef, B.; Pacheco, J.G.; Fernandes, E.; Carvalho, F.; Rodrigues, C.M.P.; Delerue-Matos, C.; Braga, A.A.C.; Moreira Gonçalves, L.; et al. Electrochemical Sensing of Ecstasy with Electropolymerized Molecularly Imprinted Poly(o-Phenylenediamine) Polymer on the Surface of Disposable Screen-Printed Carbon Electrodes. Sens. Actuators B Chem. 2019, 290, 378–386. [Google Scholar] [CrossRef]
  82. 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]
  83. Roushani, M.; Jalilian, Z.; Nezhadali, A. Screen Printed Carbon Electrode Sensor with Thiol Graphene Quantum Dots and Gold Nanoparticles for Voltammetric Determination of Solatol. Heliyon 2019, 5, e01984. [Google Scholar] [CrossRef]
  84. Khosrokhavar, R.; Motaharian, A.; Milani Hosseini, M.R.; Mohammadsadegh, S. Screen-Printed Carbon Electrode (SPCE) Modified by Molecularly Imprinted Polymer (MIP) Nanoparticles and Graphene Nanosheets for Determination of Sertraline Antidepressant Drug. Microchem. J. 2020, 159, 105348. [Google Scholar] [CrossRef]
  85. 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]
  86. Abo-Elmagd, I.F.; Mahmoud, A.M.; Al-Ghobashy, M.A.; Nebsen, M.; El Sayed, N.S.; Nofal, S.; Soror, S.H.; Todd, R.; Elgebaly, S.A. Impedimetric Sensors for Cyclocreatine Phosphate Determination in Plasma Based on Electropolymerized Poly(o-Phenylenediamine) Molecularly Imprinted Polymers. ACS Omega 2021, 6, 31282–31291. [Google Scholar] [CrossRef] [PubMed]
  87. Ben Hassine, A.; Raouafi, N.; Moreira, F.T.C. Novel Electrochemical Molecularly Imprinted Polymer-Based Biosensor for Tau Protein Detection. Chemosensors 2021, 9, 238. [Google Scholar] [CrossRef]
  88. Jesadabundit, W.; Jampasa, S.; Patarakul, K.; Siangproh, W.; Chailapakul, O. Enzyme-Free Impedimetric Biosensor-Based Molecularly Imprinted Polymer for Selective Determination of L-Hydroxyproline. Biosens. Bioelectron. 2021, 191, 113387. [Google Scholar] [CrossRef] [PubMed]
  89. Leepheng, P.; Limthin, D.; Onlaor, K.; Tunhoo, B.; Thiwawong, T.; Suramitr, S.; Phromyothin, D. Selective Electrochemical Determination Based on Magnetic Molecularly Imprinted Polymers for Albumin Detection. Jpn. J. Appl. Phys. 2022, 61, SD1009. [Google Scholar] [CrossRef]
  90. 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] [PubMed]
  91. Tang, X.; Catanante, G.; Huang, X.; Marty, J.-L.; Wang, H.; Zhang, Q.; Li, P. Screen-Printed Electrochemical Immunosensor Based on a Novel Nanobody for Analyzing Aflatoxin M1 in Milk. Food Chem. 2022, 383, 132598. [Google Scholar] [CrossRef]
  92. Zidarič, T.; Majer, D.; Maver, T.; Finšgar, M.; Maver, U. The Development of an Electropolymerized, Molecularly Imprinted Polymer (MIP) Sensor for Insulin Determination Using Single-Drop Analysis. Analyst 2023, 148, 1102–1115. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The principle of MIP synthesis and the phenomenon of molecular recognition [15].
Figure 1. The principle of MIP synthesis and the phenomenon of molecular recognition [15].
Chemosensors 11 00548 g001
Figure 2. Schematic representation of synthesis of nanocomposite based on multi-walled carbon nanotubes grafted by poly(methacrylic acid-hemin). Reproduced from [40] with the permission of Royal Society of Chemistry.
Figure 2. Schematic representation of synthesis of nanocomposite based on multi-walled carbon nanotubes grafted by poly(methacrylic acid-hemin). Reproduced from [40] with the permission of Royal Society of Chemistry.
Chemosensors 11 00548 g002
Figure 3. Synthesis route of MWCNTs-MIPs showing the chemical grafting of an MIP at the surface of MWCNts. 4-DMAP: (4-dimethylamine) pyridine; DA: Dopamine; AIBN: azobisisobutyronitrile. Reproduced from [61] with the permission of ACS Publications.
Figure 3. Synthesis route of MWCNTs-MIPs showing the chemical grafting of an MIP at the surface of MWCNts. 4-DMAP: (4-dimethylamine) pyridine; DA: Dopamine; AIBN: azobisisobutyronitrile. Reproduced from [61] with the permission of ACS Publications.
Chemosensors 11 00548 g003
Figure 4. Schematic illustration of the fabrication and operation of MIP/SPCE biosensor. Reproduced from [88] with the permission of Elsevier.
Figure 4. Schematic illustration of the fabrication and operation of MIP/SPCE biosensor. Reproduced from [88] with the permission of Elsevier.
Chemosensors 11 00548 g004
Table 2. Advantages and drawbacks of the functionalization methods.
Table 2. Advantages and drawbacks of the functionalization methods.
Functionalization MethodAdvantagesDrawbacks
ElectropolymerizationEasy, simple
Soft polymerization in comparison with UV
Control of the layer thickness
Covalent attachment of the monomer
Suitable for large scale production
Limited to aqueous medium
Drop castingFast
Does not require specific equipment
Freedom in the choice of MIP formulation
Weak interactions between the MIP and the electrode
The MIP synthesis is time consuming
Not suitable for large scale production
Use of immobilization matrix: PEG, chitosan
Chemical graftingCovalent attachment of MIPTime consuming
Not suitable for large scale production
Layer thickness difficult to control
Table 3. Examples of functionalization of SPCE with MIP.
Table 3. Examples of functionalization of SPCE with MIP.
AnalyteMatrixMIP Deposition MethodDetection MethodLODReference
Fenitrothion insecticideForest sample (leaves)EPSWV222 µg/L[75]
C-reactive proteinBlood serumSCDPV0.04 mg/L[76]
Oxfendazole drugMilkEPDPV, SWV8.0 µg/kg[77]
Methidathion insecticideWaste waterDCEIS5.1 µg/L[32]
Naloxone drugUrine/human serumEPDPV65 µg/L[78]
Salbutamol drugSwine meat
feed samples
EPDPV23.9 ng/L[46]
Bisphenol A
Plastic component
water/acetonitrile (99/1)SCDPV then CV13.7 ng/L[79]
Thyroid hormone
Thyronamine
PBS bufferEPSWV19 µg/L[80]
MDMA ecstasyHuman blood serum and urineEPSWV0.15 mg/L[81]
Cefquinome antibioticPhosphate bufferEPSWV26.4 µg/L[82]
Solatol drugTablet and human blood serumDCCV9.53 µg/L[83]
Sertraline drugHuman serumDCDPV609 ng/L[84]
Azithromycin antibioticEnvironmental waterEPDPV59.9 µg/L[85]
Cyclocreatine drugPlasma sampleEPEIS55.1 ng/L[86]
Tau protein, biomarker of Alzheimer’s diseaseSerum sampleEPEIS1.1 ng/L[87]
Genistein allergenFood productsEPDPV100 µg/kg[30]
L-hydroxyproline, biomarker of bone diseaseHuman serumEPEIS0.13 mg/L[88]
Paracetamol drugPlasmaCGDPV7.56 mg/L[27]
Atorvastatin drugWaterEPDPV0.56 µg/L[26]
Cytokine interleukin 1-β
protein
Human serumEPEIS0.23 ng/L[23]
Albumin allergenPBS bufferDCCV, Am180 mg/L[89]
Trazodone drugTap water samples
Human serum
EPDPV595 µg/L[90]
Aflatoxin M1MilkCGChronoamperometry0.09 µg/L[91]
Insulin hormonePharmaceutical sampleEPSWV11.0 µg/L[92]
EP: Electropolymerization. CG: Chemical grafting. SC: Spin coating DC: Drop casting. SP: Screen-Printing. Am: Amperometry.
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

Bou-Maroun, E. Carbon Electrode Modified with Molecularly Imprinted Polymers for the Development of Electrochemical Sensor: Application to Pharmacy, Food Safety, Environmental Monitoring, and Biomedical Analysis. Chemosensors 2023, 11, 548. https://doi.org/10.3390/chemosensors11110548

AMA Style

Bou-Maroun E. Carbon Electrode Modified with Molecularly Imprinted Polymers for the Development of Electrochemical Sensor: Application to Pharmacy, Food Safety, Environmental Monitoring, and Biomedical Analysis. Chemosensors. 2023; 11(11):548. https://doi.org/10.3390/chemosensors11110548

Chicago/Turabian Style

Bou-Maroun, Elias. 2023. "Carbon Electrode Modified with Molecularly Imprinted Polymers for the Development of Electrochemical Sensor: Application to Pharmacy, Food Safety, Environmental Monitoring, and Biomedical Analysis" Chemosensors 11, no. 11: 548. https://doi.org/10.3390/chemosensors11110548

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