Next Article in Journal
La-Doped ZnTiO3/TiO2 Nanocomposite Supported on Ecuadorian Diatomaceous Earth as a Highly Efficient Photocatalyst Driven by Solar Light
Next Article in Special Issue
Molecular Imprinting Using a Functional Chain Transfer Agent
Previous Article in Journal
The Ectodomains of rBAT and 4F2hc Are Fake or Orphan α-Glucosidases
Previous Article in Special Issue
A Note about Crosslinking Density in Imprinting Polymerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecularly Imprinted Polymers (MIPs) in Sensors for Environmental and Biomedical Applications: A Review

by
Abbas J. Kadhem
1,
Guillermina J. Gentile
2 and
Maria M. Fidalgo de Cortalezzi
1,*
1
Department of Civil and Environmental Engineering, University of Missouri, E2509 Lafferre Hall, Columbia, MO 65211, USA
2
Department of Chemical Engineering, Instituto Tecnológico de Buenos Aires, Lavardén 315, Buenos Aires C1437FBG, Argentina
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(20), 6233; https://doi.org/10.3390/molecules26206233
Submission received: 16 August 2021 / Revised: 6 October 2021 / Accepted: 12 October 2021 / Published: 15 October 2021

Abstract

:
Molecular imprinted polymers are custom made materials with specific recognition sites for a target molecule. Their specificity and the variety of materials and physical shapes in which they can be fabricated make them ideal components for sensing platforms. Despite their excellent properties, MIP-based sensors have rarely left the academic laboratory environment. This work presents a comprehensive review of recent reports in the environmental and biomedical fields, with a focus on electrochemical and optical signaling mechanisms. The discussion aims to identify knowledge gaps that hinder the translation of MIP-based technology from research laboratories to commercialization.

1. Introduction

The mechanism for the specific recognition of antibodies and antigens, enzymes and substrates, hormones and receptors inspired the development of synthetic materials that mimic nature’s ability to selectively capture chemical species from complex mixtures [1]. Molecularly imprinted materials are tailor-made polymers that present molecular recognition sites for a specific or closely-related target molecule [2]. Prior to polymerization, the target analyte, or template, is combined with a functional monomer to form a precursor structure by covalent [3], semi-covalent [4], or non-covalent [5,6] bonding. Then, they are polymerized in the presence of a crosslinker, along with an initiator in a porogenic solvent. Afterwards, the template is eluted, by extraction with a proper solvent or by chemical cleavage, to create empty recognition cavities in the polymer matrix, whose morphology and functionality are complementary to those of the template molecule [7,8].
The concept of molecular imprinting dates from 1930, but it was not until the description made by Wulff and Sarhan in 1972 [9] that research on molecularly imprinted polymers (MIPs) attracted scientific interest, driven by their promising characteristics: simplicity, robustness, stability, ease of preparation, and high affinity and selectivity towards the target molecule [10,11,12,13].
MIPs have been fabricated for solid phase extraction [14,15,16,17,18], chromatographic separation [19,20,21,22,23], catalysis [24,25,26,27,28], drug delivery [29,30,31,32,33], study of the structure and function of proteins [34,35,36,37,38], environmental and biomedical sensing [39,40,41,42,43], water and wastewater treatment [44,45,46,47,48], and membrane-based separations [49,50,51,52,53]. MIP use for purification purposes is the most commercially available application, particularly in analytical chemistry; other uses are still in need of further development [54].
The extensive literature on MIPs for sensing applications comprises a wide variety of fields. The transformative impact of MIP-based sensing for environmental and biomedical application is associated with their potential capacity to detect compounds at trace levels in complex matrices without pretreatment, which would open possibilities for contaminant monitoring in situ, as well as fast clinical analysis at the point of care for improved diagnosis and treatment. However, and although there is a genuine market need for such devices, MIP-based technology has remained mostly in the academic field.
This article aims to review advances in imprinted molecular technologies, particularly those applied to sensors in the environmental and biomedical fields. First, the most commonly used polymerization methods, physical forms, and materials are briefly described, followed by a comprehensive review of sensor fabrication reports of electrochemical and optical sensors. Given the simplicity and widespread availability of instruments for the detection of electrical and optical signals, these two mechanisms are the most promising for in situ testing and point of care diagnosis. Selected research is described in more detail for each mode of operation and application, to identify knowledge gaps and hurdles in the transition of the technology from laboratory development to commercial products.

2. Synthesis

In the synthesis process, the template molecule is covalently or non-covalently reversibly bonded to the functional monomer, with appropriate binding groups, and then polymerized with an excess of crosslinker [55]. The subsequent removal of the template originates microcavities, which are complementary to the shape, size, and spatially orientated functional groups of the template molecule [1,10]. Figure 1 presents a scheme of the imprinting process.
The functional monomer confers chemical stability before and during the polymerization and provides the MIP with the ability to interact with the target molecule, while the crosslinker offers mechanical stability and adequate porosity to the polymer, stabilizes the recognition sites, and determines the hydrophobicity [2]. A porogenic solvent brings all components into a homogenous system during the polymerization and creates the pores in the structure [56].
The choice of the functional monomer depends on the chemical structure of the template molecule and on the medium for which the MIP is designed (aqueous or organic). In environmental and biomedical applications, affinity with aqueous solutions is required, limiting the selection to hydrophilic materials. Strong template-monomer interactions enhance the ulterior affinity between the target analyte and the recognition sites. The molecule of the monomer has two units, one for recognition and the other for the polymerization. MIPs constituted by only one monomer have no more than two different kinds of binding interactions, which is sufficient for small molecules [57]. Instead, larger molecules with many functional groups require different specific bonds to achieve a desirable affinity and to prevent non-specific binding; thus, a combination of monomers may be selected: a neutral one as the backbone, along with another that is charged, hydrophobic, or capable of developing hydrogen bonds for constructing the imprinted cavities [58]. Biologically functional molecules that can specifically bind to the template molecule may also be combined with the monomer for enhancing selectivity and affinity; such is the case of aptamers, single-stranded oligonucleotides, or peptides with the ability to bind to proteins and nucleic acids [57].
The covalent imprinting route requires the formation of covalent bonds between the functional monomer and the template before polymerization, as well as between the MIP and the target molecule [59]. These bonds should be stable during the polymerization process and be cleaved without harming the MIP. The non-covalent approach, on the other hand, relies principally on hydrogen bonds, but also on hydrophobic, electrostatic, dipole-dipole, and ionic interactions between the functional monomer and the template and between the MIP and the target [60]. The non-covalent method is the most frequently used, due to ease of both preparation and template removal (by a simple wash in acidic or basic aqueous solution or with alcohol), as well as fast rebinding (with the target molecule) kinetics [61]. However, two limitations may arise: first, if the MIP is placed in a polar solvent, the interactions between the template and the functional monomer can be easily disrupted; and second, if the target molecule has only one point of interaction, the recognition properties are limited [59]. These limitations can be partially overcome in the semi-covalent imprinting, in which covalent bonds are formed between the monomer and the template and non-covalent interactions arise between the MIP and the analyte [4,62].
Covalent bonds lead to a sole organization of the functional groups in the cavities, in terms of number and orientation, whereas weaker non-covalent interactions result in less selectivity, since the target may enter the cavity in incorrect orientations, making it necessary to create an excess of binding sites to achieve the required orientation [55]. The template stoichiometrically attaches to the functional monomer in the covalent approach. Instead, the ratio of template to functional monomer usually used is 1:1 to 1:2 for the semi-covalent technique and 1:4 to 1:8 for the non-covalent, according to the affinity between them and the complexity of the template molecule [2].
Conductive polymers can be fabricated by chemical or electrochemical routes from aqueous solutions of their monomers, including enzymatic routes [63]. The imprinted sites are created based on the irreversible overoxidation that the polymers undergo during and after polymerization; the target is expelled from the polymer due to the overoxidation and, therefore, the template extraction procedure and its related complications are avoided [64]. Oxidative-chemical polymerization has been widely used due to its simplicity [63]. It is initiated by an oxidizing compound, such as FeCl3 or H2O2 [63], and is applied to the synthesis of polypyrrole, polyaniline, polythiophene, poly(1,10-phenanthroline-5,6-dione), poly(pyrrole-2-carboxylic acid), poly-9,10-phenanthrenequinone, polyphenanthroline, and some other conducting polymers. The most important electrochemical method of preparing conducting polymers is the anodic oxidation of suitable monomer species when the polymer formation and oxidation processes occur simultaneously [65]. Electrochemical polymerization has advantages over the chemical methods, as the overoxidation of the polymer creates oxygen containing groups that are useful to promote the recognition/attachment of the MIP target compounds [63]. On the other hand, cathodic electropolymerization has rarely been applied to the synthesis of conducting polymers [65]. Some redox enzymes (oxidases, such as glucose oxidase) and their substrates were used in a process similar to the chemical polymerization, due to their catalytic action that forms hydrogen peroxide. This process is conducted in an aqueous environment at neutral pH and room temperature, for maximal enzymatic activity, which, in turn, results in high biocompatibility of the polymers as desirable for biosensing applications [63]. Polymer deposition can be achieved by nucleation, growth, and other chemical steps in solid state conditions applying potentiostatic, potentiodynamic, or galvanostatic techniques to start and control these processes [65]. The selection of the deposition technique and the adjustment of the process parameters enable the formation of sensors with different characteristics. The process parameters most commonly adjusted are the applied voltage, potential pulse duration or potential sweep rate (cycling), and the electrical current [63].

2.1. Imprinting Techniques for Sensors

Ertürk and Mattiasson [66] describe bulk, epitope, and surface imprinting techniques that are especially used in the field of sensors.
Bulk imprinting requires the whole template molecule to be fully imprinted in the polymeric matrix, followed by polymerization and, finally, template removal. Then, the bulk polymer is crushed to obtain smaller particles. This is the preferred option in the case of small templates, since adsorption and release of the molecule are faster and reversible, with the consequent option of support reuse.
Epitope imprinting, on the other hand, relies on a small part of the template molecule being imprinted, making it useful for macromolecules, where only the imprinted fraction is able to represent the whole molecule, and reducing non-specific binding. Rachkov and Minoura [67] employed this concept to covalently bond only a peptide epitope, instead of the whole protein, to a surface, prior to polymerization. Removal of the support surface with the template allowed a thin film to be obtained, which proved effective in capturing the target protein.
Surface imprinting produces recognition sites only on the surface of the substrate by different routes, i.e., soft lithography, template immobilization, grafting, and emulsion polymerization. As the imprinting only takes place on the surface, a lower amount of template molecules is required, but sensitivity decreases because of the reduction in imprinting sites.
Other techniques have been applied less frequently to MIP fabrication, including soft lithography, template immobilization, and grafting.
Soft lithography (stamping) includes a step in which a pre-polymerization layer is coated on a transducer surface and where the template stamp is pressed. The obtained films are complementary to the template in structure, geometry, and chemistry.
Template immobilization is based on a molecule that is fixed onto a solid support by chemical bondings. For example, proteins are adsorbed on a support and are surrounded by the formation of a layer of disaccharides, followed by the formation of a thin plasma film. Dissolution and extraction of the template protein produce cavities on the surface that are complementary to the protein. Grafting first adsorbs the template to a support already grafted with the polymeric functional groups, obtaining affinity for the target.

2.2. Materials for MIP Fabrication

MIPs have been reported in a wide variety of functional monomers and crosslinkers, mainly determined by the type of application and target intended, and the initiators given by the polymerization reaction route.
Among the functional monomers, those containing carboxyl groups are preferred for the non-covalent technique, since they can be hydrogen donors and acceptors at the same time and are useful for the formation of hydrogen bonds. For instance, methacrylic acid (MAA) participates in non-covalent imprinting [54,57,58,60,68,69,70], forming ionic unions with amines and hydrogen bonds with amides, carbamates, and carboxylic groups [69]. Other functional monomers often used in the non-covalent technique include acrylamide (AAM) [54,58,60,70], acrylic acid (AA) [58,60,69], methyl methacrylate [54,69], 4-vinylpyridine [69,70], 2-vinylpyridine [69], and 1-vinylimidazole [69]. Pyrrole [54,68,71] offers good electrical conductivity, biocompatibility, and redox activity [72]. Aniline [54] forms polyaniline, a conducting polymer with a reversible redox system, environmental stability, and easy chemical or electrochemical polymerization [73]. In an extensive work, Chen et al. [61] mentioned the less often employed 2-acrylamido-2-methyl-1-propane sulfonic acid, 3-aminopropyltriethoxysilane (APTES), 4-ethylstyrene, glycidoxypropyltrimethoxysilane, 2-hydroxyethyl methacrylate (HEMA), itaconic acid (IA), methacrylamide (MAAM), 3-methylacryloxyprolyl trimethoxysilane, methylvinyldiethoxysilane, styrene, p-Vinylbenzoic acid, trans-3-(3-pyridyl)-acrylic acid, and trifluoromethyl acrylic acid. Otherwise, N-isopropylacrylamide (NIPAAm) [58,60,74], and o-phenylenediamine (o-PD) [75,76] may also be cited for the non-covalent approach. Vinylferrocene and ferrocenylmethyl methacrylate have been used as electroactive monomers, which allowed for the electrochemical detection of non-electroactive targets [77,78,79,80]. Fewer monomers can be selected for the semi-covalent or covalent approach, since specific bonds must develop; examples of these functional monomers are tert-butyl p-vinylphenol carbonate, 4-vinyl aniline, 4-vinyl benzaldehyde, and 4-vinyl benzene boric acid for covalent [61], whereas 3-isocyanatopropyltriethoxysilane [61] and m-aminophenylboronic acid [60,81] have been used for semi-covalent imprinting, although it is also highly dependent on the target molecule’s chemistry and interactions.
The two most often used crosslinkers are ethylene glycol dimethacrylate (EGDMA) [8,54,57,60,61,69] and divinylbenzene [8,60,61,69]. Both chemicals were applied in the non-covalent approach by means of free radical polymerization. Other compounds commonly reviewed for the same purpose are 3,5-bis(acryloylamido)benzoic acid, N,O-bisacryloyl-L-phenylalaninol, N,O-bis methacryloyl ethanolamine, pentaerythritol triacrylate, trimethylolpropane trimethacrylate [61,69], 2,6-bisacryloylamidopyridine, 1,4-diacryloyl piperazine, glycidil methacrylate, 1,3-diisopropenyl benzene, N,N’-methylenediacrylamide, N,N’-1,4-phenylenediacrylamine, and tetramethylene dimethacrylate [61]. For the free-radical polymerization of covalent complexes, bis-(1-(tert-butylperoxy)-1-methylethyl)-benzene, dicumyl peroxide, and triallyl isocyanurate have been mentioned [61]. Also, N,N-methylenebisacrylamide (MBA) [8,60,69], N,N-1,4-phenylenediacrylamide, and pentaerythritol tetraacrylate [69] were found as common examples of crosslinkers.
In free radical polymerization, the pre-polymerization monomer-target complex is subjected to heat or ultraviolet (UV) radiation in the presence of an initiator, such as 2,2′-azobis(isobutyronitrile) (AIBN) [54,61], 4,4′-azo(4-cyanovaleric acid), or azobis dimethylvaleronitrile [61]. UV light at the maximum absorption wavelength of the compound or high temperature enables the decomposition of the azo compounds and generates free radicals that start the chain reactions. Organic peroxides, e.g., benzoyl peroxide or benzyl dimethyl acetal, initiate the process by redox reactions, heat, or photochemically, and are especially suitable for aqueous matrices, since they are soluble in water as well as in organic solvents. Inorganic oxidants, such as potassium persulfate, have also been reported.
For the detection of organic, relatively hydrophobic pollutants in environmental applications, the porogenic solvent tends to be aprotic and non-polar, e.g., chloroform, toluene, acetonitrile, and tetrahydrofuran [59]. However, biomolecules require other kinds of solvents, since they are usually insoluble or unstable in organics. The polarity of the porogen is a key characteristic for its selection, particularly in non-covalent MIPs, as it affects the interactions between the template and the functional monomer, which, in turn, define the adsorption properties.

2.3. Physical Form

MIPs have been reported in different physical forms, such as blocks or monoliths, microspheres, nanospheres, thin films, nanocomposite membranes, and nanowires. Figure 2 shows a schematic illustration of these physical forms.
The choice of the physical form is often dictated by the application, and it defines the polymerization technique to be used [82].
Nano and microspheres have both surface and internal porosity, which results in desirable properties such as high surface area, low density, and a high loading capacity, and may find applications in analytical chemistry and biomedical applications: drug delivery, absorption and desorption, high speed chromatography, and tissue regeneration [83,84].
Different polymerization techniques may produce porous microspheres, such as suspension [85,86,87,88], precipitation [89,90,91,92,93], emulsion [94,95,96,97], grafting [98,99], and swelling [100,101,102], as well as a combination of two of these methods [103,104]. For example, in swelling polymerization, previously formed beaded seeds made of silica or polymers are used as scaffolds for the polymerization.
The suspension polymerization process is useful when both the monomer and the crosslinker are insoluble in the porogen, thus forming two phases, i.e., a liquid matrix that contains droplets of the monomer, inside which polymerization occurs [83]. The obtained spheres usually have a diameter between 10 and 100 μm [105]. It is often necessary to include an adequate stabilizer that covers the droplets in the form of a thin film and prevents coalescence, controling the size of the particles [106]. Common stabilizers include surfactants [107] and ionic liquids [108]. Solvents compatible with the stabilizers may be water [109], silicone oil [110], and polyvinyl alcohol [111]. As stated above, water may interfere with the non-covalent bonding between the functional monomer and the template; therefore, its use is limited to those cases where the stability is not affected, while the other solvents mentioned were able to overcome this drawback, but the monodispersivity was negatively affected [112].
In the precipitation or dispersion polymerization process, the components of the MIP are mixed with a much higher amount of solvent than required in the traditional polymerization. The polymer grows at a larger extent and precipitates when the chain is long enough to be insoluble [113]. The obtained particles are smaller, usually between 0.3 and 10 μm [114], with more uniform size, and are recovered by simple washing and centrifugation. It is possible to control the size and shape of the particles [115] without the aid of surfactants or stabilizers, which avoids potential contamination of the MIPs. This technique was first developed by Mosbach and collaborators [116] for the synthesis of MIP microspheres [117], nanospheres [118], core-shell particles [119], thin films [120], and nanocomposites [121].
Core-shell MIP particles are obtained by grafting or surface polymerization. All the MIP components are adsorbed on the surface of preformed beads, such as porous silica or spherical polymers, before the polymerization starts. Once the free-radical polymerization is over, the bead is removed, thus obtaining a spherical particle coated by a thin layer of MIP. It is important to limit the free-radical polymerization to the bead surface, for example, with the inclusion of an immobilized chemical compound whose functions are to initiate, transfer, and terminate the polymerization (iniferters) [122,123]. Advantages of the technique are a high surface density of polymer chains, high stability of the coated layer, and the ability to graft different polymers to the same substrate [124,125].
Emulsion polymerization can lead to both spherical and core-shell particles. The monomers are immersed in a solvent, usually water, in which they do not dissolve, and emulsification takes place. Then, the polymerization is started by an initiator soluble in one phase. When the process is finished, a fluid of milky consistency is obtained, which is usually referred as latex, synthetic latex, or polymer dispersion. In contrast with suspension polymerization, the droplets are inside micelles stabilized with the aid of a surfactant. To prepare core-shell particles, the core particles are produced and afterwards the shell layer is generated by emulsion polymerization. The type and amount of surfactant controls the particle size, usually in the range of 10–1000 nm [126,127]. The main disadvantage of this method is the use of water and surfactants that may precipitate interferences during polymerization between the template and the functional monomer, particularly in the non-covalent approach.
Block or monolith MIPs are obtained by bulk polymerization, also known as the traditional method. Bulk polymerization is the simplest and most widely used method, since the equipment is not sophisticated and no particular knowledge or mastery of organic chemistry is needed. The main components (template, functional, and cross-linking monomers) are mixed along with the initiator in a low volume of porogenic solvent, and the polymerization starts by heat or UV radiation. When the process is concluded, the monoliths are crushed, ground, and sieved to the desired particle size, reaching the micrometer range. Evident disadvantages of the grinding are that the obtained material is irregular in shape and size, nanosized particles cannot be produced, and many recognition sites are broken or inaccessible due to the lack of internal porosity.
In the field of analytical chemistry, Matsui et al. [128,129] introduced the monolithic imprinted polymerization technique that enables monolithic imprinted columns to be obtained for HPLC by a simple, one-step, free radical polymerization process that takes place within a chromatographic column. The template, functional monomer, cross-linking monomer, and initiator are first dissolved in the porogen and then the mixture is poured into the column. After the polymerization takes place, the template and the solvent are washed out of the column. An advantage of this route compared to traditional polymerization is that it is not necessary to crush, grind, sift, and pack the final material. Other advantages reported by the author are ease of preparation, good reproducibility, selectivity, sensitivity, high porosity, permeability, surface area, and fast mass transport.

3. Environmental and Biomedical Applications of MIPs

The specific binding properties of MIPs make them ideal materials for sensor fabrication. Although the literature on polymer formulation and synthesis methods is extensive, reports of effective use of MIPs on specific applications are scarce. Moreover, in order to use a MIP as a sensor, it needs to be coupled with a transducer or reading mechanism capable of determining the amount of target rebinding to the MIP after exposure to the test sample. The transducer technique should be simple, reliable, and not require external instruments or supplies. Based on these characteristics and their user-friendliness, this review focuses on electrochemical and optical sensors.
A shared objective of MIP-based sensors is to provide the market with simple, fast, and inexpensive methodologies for the detection and quantification of a chemical compound. Besides, researchers have focused their efforts on avoiding sample pretreatments, which are often time-consuming, tedious, or expensive, and in many cases involve toxic or hazardous solvents. Current protocols typically include a first pretreatment step, where the target of interest is removed from a complex matrix (e.g., by liquid extraction), followed by separation (e.g., by gas or liquid chromatography), enrichment, and detection.
A comprehensive review of environmental and biomedical sensors is presented below, organized by the sensing mechanism: electrochemical or optical. The limit of detection (LOD) and the limit of quantification (LOQ) are presented for each work, when reported by the authors and calculated as three and ten times the standard deviation, respectively, of the blank measured with the standards divided by the slope of calibration curve, as suggested by MacDougall et al. [130]. The units given by the authors were reproduced in this work to provide understandable and useful information, according to the specific target. Target molecules and the matrix in which they were tested are also listed for each work. Although all the cited reports focus on environmental or biomedical applications, not all of them employ a testing matrix that mimics real natural environments or the actual subjects of study (humans or animals). The advances and discoveries in the field are impressive and promising for the technology to transfer outside of research laboratories to commercial products. Yet, some shortcomings that need to be addressed are: limited detection, slow response, high LODs and LOQs, interferences, impaired performance in real complex mixtures, and reusability of the adsorbent to reduce waste, as well as devising, conceiving, and putting to work platforms to manufacture ready-to-use devices.

3.1. Electrochemical Sensors in Environmental and Biomedical Applications

Electrochemical sensors include a cell with a working electrode of particular interest, accompanied by a reference and an auxiliary electrode. Depending on the measured electrical parameter, they can be classified into three categories: conductivity or capacitance sensors, potentiometric sensors, and voltammetric or amperometric sensors [131,132]. Conductivity or capacitance sensors measure the change in conductivity or the capacitive impedance over time as a function of the concentration of the target. Potentiometric sensors measure the potential of a redox reaction in order to determine a concentration, and voltammetric sensors measure the effect of the concentration of the target on the current-potential of the redox reaction. Among these last, the amperometric sensors are a subcategory that measure the current that results from a fixed potential that is applied to a redox system and is related to the concentration of the participating species.
Table 1 summarizes recent reports of MIP-based electrochemical sensors, grouped by quantifiable electrical output generated upon the rebinding of the target molecule.

3.1.1. MIP-Electrochemical Sensors in Environmental Applications

The detection and quantification of pollutants in environmental samples, such as natural waters and soils, as well as in aquatic organisms, is necessary to determine their fate and transport. Most environmental sensors have targeted endocrine disruptors, pesticides and pharmaceuticals. They occur in the environment at trace concentrations in complex matrices, which further challenges the analysis. A large number of efforts have been devoted to developing novel and economical determination and quantification alternatives. Some particularly illustrating examples are briefly described below.
A MIP sol-gel film formed by multi-walled carbon nanotubes (MWCNTs) and the conductive polymer Nafion was prepared for the determination of 2-nonylphenol, a xenobiotic endocrine disruptor [42]. The use of Nafion increased the homogeneity of the MIP sol-gel and improved the peak current of the electrochemical probe. APTES, phenyltrimethoxysilane, and a combination of these were used as monomers. A pH of 7 provided optimal electrostatic binding of the target in the MIP, due to the degree of ionization of 2-nonylphenol. The stability and repeatability tests showed a relative standard deviation of the MIP sol-gel response of only 3.6% for twenty successive measurements, and a higher current response of the electrode to the 2-nonylphenol when compared with its structural analogues. However, the effect of physical or chemical interactions in real samples, such as suspended solids, was not reported. Deng et al. [170] proposed a voltammetric sensor to detect bisphenol A, another endocrine disruptor of concern. An acetylene black paste electrode was prepared with acetylene black and paraffin and covered with 2 μL of a film of bisphenol A (10 mM), chitosan, and graphene oxide (0.25 mg/mL), using H2SO4 to crosslink the chitosan. Removal of the template was achieved by cyclic voltammetry (CV), and graphene oxide was reduced to obtain the final electrode. The best electrochemical determinations were obtained at pH 3 and with immersion of the electrode in the solution for 3 min. Six electrodes were produced to check reproducibility. The sensor was stably stored in H2SO4 at room temperature for 10 days. Repeatability was tested by using the same electrode ten times, removing the target between measurements by cyclic potential sweeps. Selectivity was good over other compounds, including various phenolic molecules, and no interference from ions was detected. Des Azevedo et al. [201] developed a MIP-hybrid electrochemical sensor for the detection of 17β-estradiol (E2) in water, based on N-phenylethylene diamine methacrylamide as bifunctional monomer, MAA as functional monomer, N,N′-diethyldithiocarbamic acid benzyl ester as initiator, EGDMA as crosslinker, and acetonitrile as porogen. However, the short linear range and low LOD would require improvements to be useful for real environmental samples. To test the sensor selectivity, 17α-estradiol (α-E2), progesterone, and (P4) estriol (E3) were used, since the three are structurally related to E2. No binding occurred between the sensor and α-E2, nor P4, showing good selectivity and specificity. However, E3 at high concentration was detected by the sensor, exhibiting a lack of selectivity due to its high structural similarity with E2.
Pesticides in environmental samples usually require arduous sample pretreatment followed by chromatographic techniques; therefore, MIPs have been proposed as a low cost, portable, and easy-to-use alternative. A MIP-based sensor for the organophosphorus pesticide methidathion was prepared using the template, the functional monomer MBA, and EGDMA in a molar ratio of 1:4:20, together with AIBN in dimethylformamide [134]. Bulk polymerization took place in an oil bath at 80 °C for 12 h, after which the monolith was ground and sieved, and the template was removed by Soxhlet extraction. The MIP particles were mixed with a TMOS sol-gel solution and deposited on a carbon screen-printed electrode (SPE). Good reproducibility and selectivity were obtained, since the sensor did not respond to other organophosphorus pesticides tested. Regeneration was achieved for five successive measurements by efficiently extracting the target with a mixture of methanol and acid. The sensor was evaluated in tap water, and although satisfactory, this relatively clean environment is far less challenging than those expected in natural water samples. Anirudhan and Alexander [145] reported a new synthesis to obtain a MIP-based potentiometric sensor to determine lindane (γ-hexachlorocyclohexane), an organochlorine pesticide, using MWCNTs. The potentiometric sensor was prepared using an active electrode of Cu and a calomel electrode as reference. First, MWCNTs were mixed with glycidyl methacrylate, which contains the epoxide functional groups essential to develop polarity in the nanotubes. Then, these MWCNTs were vinylated and polymerized with the template to obtain a covalently bonded complex matrix on their surface. The optimal pH was 3; the chlorine atoms found in the pesticide were electrostatically attracted to protonated carboxyl groups of the MWCNTs, increasing the sensitivity. The selectivity was tested using a potentiometric method and linear sweep voltammetry (LSV), with organophosphorus and organochlorine compounds, resulting in a high response compared to the other pesticides. The authors applied the sensor to the detection of lindane in fruits, vegetables, and water, but an extraction step from complex matrices was used. The sensor was able to detect the target, but the pretreatment added complexity and limited its application in situ.
Magnetic MIP nanoparticles, consisting of a magnetite core, were fabricated for the extraction, cleaning, and pre-concentration of the organophosphorus pesticide, methyl parathion in fish [178]. The nanoparticles were obtained by co-precipitation of Fe2+ and Fe3+, and a SiO2 shell, that were reacted with TEOS to acquire OH groups. In this way, the magnetic core-shell particles reacted with an acrylic group, obtaining active C=C groups that, in turn, were polymerized with the template in toluene. MAA, EGDMA, and AIBN were included as functional monomer, crosslinking agent, and initiator, respectively. The template was removed by Soxhlet extraction with methanol and acetic acid. The best working conditions were at pH 2 and maximum loading was reached after 1 h. Selectivity tests were performed with similar structures, confirming specific binding. The sensor could be reused for six measurements, though binding capacity was lost to some extent.
Significant efforts were directed towards the analysis of pollutants, in particular pharmaceuticals, in wastewater. Warwick et al. [138] proposed coupling MIPs with a capacitance sensor to offer a more economical alternative to the colorimetric method used in the detection of phosphates in wastewater. The selected template was phenylphosphonic acid, instead of HPO42− and H2PO4, to avoid solubility problems, since the synthesis required organic solvents. EGDMA was the crosslinker, AIBN the initiator, and N-allylthiourea the functional monomer. The template and the monomer were added in a molar ratio of 2:1. Polymerization was performed under UV light for 20 min, and the MIP was ground and sieved before Soxhlet extraction of the template molecules. The sensor exhibited good performance at pH 6.5–8, emulating that of domestic wastewaters. Selectivity was studied, finding out that sulphate, nitrate, and chloride did not interfere with the detection of phosphate, thus offering stability for longer times. Reusability of the MIP membrane was achieved for up to ten times. However, the LOD and LOQ were too high, above the typical concentrations of phosphate in wastewater. N-formylamphetamine, an intermediate and an indicator of amphetamine synthesis, was detected in wastewaters using MIP particles that were obtained on the gold surface of a wafer electrode, with a mixture of two functional monomers, HEMA and IA, along with EGDMA, in a reaction started with AIBN at 60 °C [136]. The monomers contained methylene and carbonyl groups to bond to the phenyl and amide groups of the template. The sensor was tested in buffer solutions, but no tests in real wastewaters were performed. Zhao et al. [162] developed a MIP modified boron-doped diamond (BDD) electrode to quantitatively determine the presence of the antibiotic sulfamethoxazole in surface waters. The MIP-BDD electrode was prepared by five electropolymerization cycles using pyrrole (40 mM) as a functional monomer on a BDD electrode in the presence of the template (20 mM) at pH 7.5. The selectivity of the sensor was high; however, occurrence of sulfamethoxazole in aquatic ecosystems [205] and wastewaters [206,207] is at much lower concentrations than the detection limit. A MIP for the detection of metronidazole, a drug to prevent parasites in fish and poultry, was designed by Gu et al. [173], combining molecular imprinting with mimetic enzymes. Melamine was both the functional monomer, capable of forming hydrogen bonds, and the molecular host of the mimetic enzyme. Cu was the active center, since its complexes present enzyme-like activity; Au nanoparticles amplified the signal and casted with chitosan on a glassy carbon electrode (GCE). In parallel, CuSO4 in acid and melamine in water were mixed until a complex between them was formed, and then the template was included along with NaCl. After the polymer was electrodeposited on the electrode, the template was removed by ten scan cycles in Britton–Robinson buffer. Recognition and catalytic activity were successfully achieved, as well as good reproducibility and stability. Selectivity over molecules with similar electrochemical response but different in shape, size, and functional groups was good thanks to the nature of the imprinting sites; however, when the tested compound had a similar structure, the interference was greater, evidencing the lack of specificity of the MIP.

3.1.2. MIP-Electrochemical Sensors in Biomedical Applications

A large number of sensors were designed with the intention of improving dose control or to measure pharmaceutical drugs in tablets, injections, or physiological fluids. However, most sensors were validated only in aqueous solutions or simulated environments, much simpler than the matrices they would encounter in biomedical applications.
Ji et al. [39] combined a MIP film with carboxylic functionalized MWCNTs GCE and Au nanoparticles to measure cholesterol concentrations. To prepare the MIP, the electrode was first immersed in a solution of the functional monomer, p-aminothiophenol, HAuCl4, and cholesterol to form the pre-polymerization complex, due to the strong interactions between the amino functional monomer and the acidic template. The polymer was formed through bonds between Au in the crosslinker and sulfur in the monomer, and the template was then removed by a solution of HCl in ethanol-water. Detection of the target was manifested by an increase in charge-transfer impedance, as well as a reduction in the differential pulse voltammetry current peak. The selectivity of this sensor was satisfactory and it remained stable after a month of storage at room temperature in HCl. Despite the promising results, the authors recognized that its application in clinical analysis/diagnosis would require further study. Rosy et al. [139] electropolymerized the functional monomer o-aminophenol on a GCE together with the target norepinephrine and NaClO4 for diagnosis and drug quality control. After the imprinting, the template was removed with H2SO4, capable of breaking the hydrogen bonds between o-aminophenol and the polymer. The sensor was tested in phosphate buffer solution (PBS) and selectivity, stability, and reproducibility were studied, with satisfactory results. A potentiometric sensor for the recognition of imidocarb dipropionate was synthesized by Rizk and coworkers [146], based on a potential difference between a MIP membrane sensor electrode and a reference electrode of Ag/AgCl. The prepolymerization solution was a mixture of the template, MAA, EGDMA, benzoyl peroxide, and acetonitrile that was bulk polymerized. Although the final application was to detect the target in the liver and kidney of animals, the sensor was only tested in aqueous solutions. A MIP sensor for the anticoagulant drug heparin was prepared using heparin as a template, MAA as a functional monomer, AIBN as an initiator, and EGDMA as a crosslinker [147]. The GCE was coated with the prepared solution and polymerization was performed under UV light. The effect of pH and prepolymerization solution formulation was tested, and the sensor was evaluated only in laboratory prepared solutions of the target. A MIP sensor for captopril, a drug used to treat hypertension, was designed using a GCE and a carbon paste electrode (CPE) [149]. The MIP particles were obtained by precipitation polymerization; captopril was dissolved in a mixture of acetonitrile and ethanol, then MAA was incorporated, followed by EGDMA and AIBN, and reacted for 12 h in an oil bath at 80 °C. The template was eluted with a solution of methanol and acetic acid. Good stability and reusability were obtained after twenty cycles of operation and selectivity over other interfering drugs was satisfactory, but tests were conducted in deionized water solutions. Li et al. [194] proposed a three-dimensional MIP modified voltammetry-based sensor for the detection of epinephrine. ZnO nanorods grew vertically on indium tin oxide (ITO) coated polyethylene terephthalate film by electrodeposition of pyrrole in the presence of the template and LiClO4; the template was eluted by immersion in KCl and PBS. Unfortunately, the saturation of imprinted sites prevented the linearity of the oxidation peak current vs. epinephrine in the range of 1–2000 μM. Although good selectivity facing structural analogues and repeatability were obtained, the response was not linear and the sensitivity was too low for physiologically relevant concentrations. Da Silva et al. [161] worked on a sensor to detect the antibacterial chemical norfloxacin in human urine. MWCNTs were deposited on the surface of a GCE, which was afterwards coated with a MIP film via cyclic voltammetry of polypyrrole. The human urine samples were spiked with the chemical and diluted 50% with sulfuric acid. The sensor was exposed to chemical structure analogs to the target, and interference was manifested when exposed to enrofloxacin. The MIPs were reused for thirty measurement cycles without significant change in the current response signal.
The rapid detection of biomarkers in a point of care setting is highly desirable for improved diagnosis and treatment and several authors reported efforts towards this goal. Electrochemical sensors have been reported for the detection of DNA [204] and proteins [203], although, in general, they were only tested in aqueous solutions and specificity and non-specific interactions were not explored. For example, Yola and Atar [189] developed a sensor to detect the cardiac biomarker Troponin-I in plasma. The template and pyrrole were imprinted on BN quantum dots coated GCE by cyclic voltammetry. No interference was detected due to the plasma; selectivity over other proteins in plasma, stability, and reproducibility were high. Moreira et al. [196] reported a point-of-care disposable sensor for myoglobin, another cardiac biomarker. The template and functional monomer (o-aminophenol) were adsorbed on a gold SPE and electropolymerized. The template was removed by digestion of the MIP in proteinase K that cleaved peptide bonds under mild conditions, hence avoiding alterations in the polymer. However, due to the small size of the protein, only those molecules on the outer surface could be removed, leaving the vast majority entrapped inside the film. The sensor was tested in MES buffer and synthetic urine; in this context, the results obtained were fast, sensitive, and selective. Other large molecules of interest for which MIP electrochemical biosensors have been reported are the clinical biomarkers for diagnosis of cancer and other cardiovascular diseases [208]. The imprinting of high molecular weight compounds, e.g., biomacromolecules, confronts particular additional challenges given by the size and complexity of the structure and conformation of the target, leading to binding sites with heterogeneous affinities, hindered target removal, and solvents that induce conformational changes in proteins (unfolding or denaturation) [209]. The epitope imprinting technique has been proposed to overcome these issues: in this approach, only a small but characteristic portion of the biomolecule is imprinted. However, the specificity problem may still arise if the epitope is not unique to the intended target [209].

3.2. Optical Sensors

Optical sensors rely on a change in an optical property, such as light absorption, fluorescence, light scattering, refractive index, or reflection, as the target rebinds to the MIP sites. This function is sufficient in the case of optically detectable targets, but, if the substance lacks optical properties, an indirect method of detection is needed. Alternatively, the change in color, fluorescence, etc., may occur after the complex formation with the MIP [210]. Different optical techniques can be used in these sensors, such as ultraviolet/visible spectroscopy (UV/Vis), fluorescence, chemiluminescence, surface plasmon resonance (SPR), and Raman scattering (RS) [211]. Conventional RS sensitivity is low because of its small cross-section, thus not attaining detection at trace level. In surface-enhanced Raman scattering (SERS), metallic nanoparticles are included (Ag, Au, etc.) to act as the active substrate where the target adsorbs with a notorious enhancement in the magnetic field [212]. Photoelectrochemical (PEC) sensors merge UV/Vis method with electrochemical sensors, by enabling amperometric detection thanks to photoirradiation. The measurements are based on electron transfer among an analyte, a semiconductor, and an electrode, coupled with photoirradiation [213]. Table 2 summarizes some reports of MIP-based optical sensors for environmental and biomedical applications.

3.2.1. MIP-Based Optical Sensors in Environmental Applications

A selection of innovative MIP optical sensors is presented below, with emphasis on two main contaminant groups for the environmental field: pesticides and industrial waste. Sensors that rely on a change of color of the material upon exposure to the target compound are particularly attractive for environmental applications, as they can offer naked-eye readings in the field. However, the usual low levels of the relevant pollutant and the complex and diverse matrices encountered are major obstacles.
Dyes present in textile and paper industrial effluents are toxic to humans and the environment and require monitoring. A MIP-based evanescent wave fiber sensor for the basic red 9 (BR9) was synthesized by bulk polymerization, using 2-acrylamido-2-methyl-1-propanesulfonic acid as a functional monomer and EGDMA as a crosslinker. The amount of dye in the samples was correlated to the absorbed light in aqueous solutions [217]. However, the device was not tested in real effluents. In another approach, Duan et al. [257] proposed a sensor based on chemiluminescence resonance energy transfer in CdTe quantum dots@ luminol nanomaterials combined with chitosan/graphene oxide-magnetite-MIP for detection of chrysoidine. The CdTe QDs@ luminol amplified the chemiluminescence signal, whereas chitosan and graphene oxide improved the adsorption. When tested in the presence of coexisting substances, the sensor showed decreased specificity. A carbon dot functionalized fluorescent MIP was fabricated for the detection of dinitrotoluene in groundwater with AA as the functional monomer, EDGMA, and AIBN [247]. The sensor was tested in spiked lake water and tap water samples, with overall acceptable performance, although organic matter interfered with the fluorescence signal. This drawback was partially overcome by a non-labeled photonic MIP sensor, with the optical active structure obtained by conducting the polymerization in the pore space of a sacrificial colloidal crystal: the method allowed the detection of 2-butoxyethanol wastewater from hydraulic fracking operations. The sensor performed well, but the polyacrylic acid polymer was severely damaged by the wastewater after each use and could not be recycled [220].
Bisphenol A (BPA) is another chemical of concern that was the target of several MIP-sensors. A fluorescence MIP sensor was fabricated combining silica-coated fluorescent carbon dots via sol-gel polymerization to be used in river water samples [239]. Xue et al. [279] fabricated surface-imprinted core-shell Au nanoparticles of BPA for detection by SERS in surface water and plastic bottled beverages. Both sensors showed good performance in laboratory prepared standard solutions, and acceptable measurements in a handful of real samples, although low pH beverages resulted in very low recoveries. A photonic sensor was proposed by Kadhem at al. [221] for the detection of testosterone in natural water, another example of endocrine disrupting chemicals in the environment. A mixture of AA, EDGMA, AIBN, and the target was polymerized inside a silica particles crystal that provided the optically active morphology. Rebinding of the target produced swelling of the polymer and consequent change in the wavelength of the reflected light. The sensor showed minimal non-specific adsorption and good reusability in laboratory-made test samples.
Several optical MIP-based sensors have been reported for the detection of pesticides and veterinary antibiotics. Zhao et al. [212] fabricated a MIP for atrazine extraction from apple juice by bulk polymerization of MAA, EGDMA, chloroform, and AIBN, reacted in an oil bath at 60 °C for 24 h. The obtained monolith was ground and sieved, the template removed by Soxhlet extraction, and particles were packed into a solid-phase-extraction cartridge. The pretreated solution was analyzed by a colorimetric method based on Au nanoparticles for rapid detection by SERS, but it did not reach a low LOD nor a linearity in the response. A sensor for the herbicide 2,4-dichlorophenoxyacetic acid was developed by Wagner et al. [231] using fluorescent core-shell MIP particles in a 3-dimensional microfluidic system for droplet extraction from the water matrix and mixture with the MIP, that reached a LOD below the drinking water guideline. However, nonspecific binding due to matrix effects were observed. Electrochemiluminescent graphene quantum dots were proposed for the detection of the herbicide 2-methyl-4-chlorophenoxyacetic acid [254]. A layer of hybrid nanocomposite of graphene quantum dots and MoS2, in a mass ratio of 2:3, was coated on a GCE upon which the MIP was synthetized by cyclic voltammetry with 2-methyl-4-chlorophenoxyacetic acid as the selected template and the functional monomer o-PD. The template removal was accomplished by shaking in methanol and acetic acid. Samples were subjected to an extensive pre-treatment for the extraction of the target from water and food samples, and redissolution in PBS (pH 7.4). Competitive adsorption tests demonstrated good selectivity as well as good stability.

3.2.2. MIP-Based Optical Sensors in Biomedical Applications

Efforts in the biomedical field are primarily aimed at developing point-of-care devices that provide non-invasive, safe, and fast detection, as well as quantification of drugs for dose control, especially when serious side effects may appear.
The detection of proteins by MIP-based sensors has been reported by fluorescence, surface plasmon resonance, and changes in the Bragg diffraction of optically active imprinted hydrogels. SPR showed good resistance to fouling and the consequent non-specific binding in biological matrices, but the technique requires relatively more expensive equipment than the measurement of Bragg diffraction. On the other hand, the response of photonic hydrogels can be affected by ionic strength or pH (buffers), possibly limiting their application to protein sensing [302].
Sensors for the biomarker α-fetoprotein were developed by Tan et al. [223] based on fluorescence and Ye et al. [278] on SERS. The fluorescent sensor was a combination of ZnS quantum dots and MIPs of functional monomers methyl methacrylate and 4-vinylphenylboronic. The two functional monomers were chosen so the boronic acid group would form a covalent bond with the template, giving a cyclic ester in alkaline medium; γ-methacryloxypropyl linked both the organic and inorganic phases, enabling the sol-gel polymerization. Serum samples were added to PBS and mixed with the MIP particles, a carbonate buffer, and, finally, diluted with water. Although high loading capacity and selectivity were obtained, the synthesis process was rather complex and the samples needed pretreatment. The SERS methodology involved Ag nanoparticles labeled MIPs with boronate affinity [278]. Specificity for α-fetoprotein was studied as well as cross-reactivity, finding out that the highest values were obtained for glycoproteins of similar molecular weight as the target. The glycoprotein RNase B was detected by a photoelectrochemical approach, given by three-dimensional anatase hierarchically cactus-like arrays vertically grown on a FTO substrate for PEC detection [291]. The electrochromic indicator employed was a Prussian blue electrode that discolored to Prussian white, as a function of the target concentration. The MIP was fabricated on TiO2 arrays amino-functionalized with APTES, then immersed in a solution of 2,4-difluoro-3-formylphenylboronic acid and NaBH3CN, washed with water and, finally, washed in a solution of NH4HCO3 at pH 8.5 containing the templates. The electrode was washed with NH4HCO3 solution and subsequently imprinted in ethanol where TEOS was added. Finally, washing with ethanol-acetic acid removed the template. The PEC measurement was performed by allowing the sample to be bound to the modified electrode, and then connecting to the Prussian blue electrode in PBS (pH 7.4). The discolored electrode was taken out and the absorbance measured after light irradiation for 10 s. Stability, selectivity, and reproducibility were studied with acceptable results.
Duan et al. [252] used chemiluminescence in the detection of dopamine, useful in the diagnosis of Parkinsonism. This work is based on silanized Fe3O4 magnetic nanoparticle-graphene oxide MIP. The magnetic graphene oxide was included in an ethanol solution of dopamine and acrylamide with EGDMA and AIBN. A solution of methanol and acetic acid was used to extract the template. A major inconvenience was caused by epinephrine, that attached to the imprinted cavities. The detection of phenylalanine in urine for diagnosis purposes was conducted by magnetic MIP nanoparticles with fluorescence spectrophotometry and RS [285]. Iron oxide nanoparticles were added to ethylene-co-vinyl alcohol-dimethylsulfoxide solution, and thereafter mixed with phenylalanine for non-covalent imprinting; the template was removed by ethanol and acetic acid. Tests on urine samples showed cross-reactivity with structurally similar compounds, in particular tyrosine and L-3,4-dihydroxyphenylalanine.
A sensor for the antipsychotic drug thioridazine was made from MIP-coated fluorescent ZnO quantum dots [240]. The quantum dots were obtained by a core precipitation from Zn(CH3COO)2 with NaOH and a silica shell. The MIP was prepared on the quantum dots by reverse microemulsion with TEOS, and NH4OH to hydrolyze the monomer; afterwards, the template and APTES were introduced into the solution and acetone to break the emulsion. The dots were then subjected to centrifugation and the precipitate washed with a mixture of ethanol and acetonitrile to remove the template. Tests were performed on plasma samples with a pretreatment to remove proteins and resuspension in a suitable buffer solution. Selectivity over the remaining compounds was demonstrated for the tested samples.
Significant efforts were devoted to the detection of antibiotics. A sensor for the detection of ornidazole, combining graphene quantum dots and silica MIPs, was fabricated by citric acid pyrolysis with APTES, followed by sol-gel polymerization of the target and silica matrix; methanol was used to remove the target molecules [243]. The sensor was evaluated in human plasma, pretreated to separate the proteins, and adjusted to pH 9. Repeatability, selectivity, and reproducibility were satisfactory, as well as the low interference by ions commonly found in serum. Fluorescent quantum dots were employed for the detection of sulfasalazine [244]. An amino-functionalized glass slide was incubated with APTES, covered with semi-conductor CdSeS/ZnS quantum dots, and, finally, functionalized with methacryloxypropyl trimethoxysilane to enhance adhesion. To prepare the MIP, the glass was immersed in a mixture of sulfasalazine, MAA, EGDMA, and AIBN in acetonitrile and toluene. Secondly, it was heated to 60 °C for 2 h and washed with methanol and acetic acid to remove unreacted monomers and the template. Plasma and urine samples were centrifuged and the supernatant diluted prior measurement. Reusability, selectivity over structural analogues, and reproducibility were satisfactory.

4. Technical Barriers to Commercialization of MIP Sensors and Devices

Molecular imprinted polymers are a promising technology in the environmental and biomedical sectors. The reported LODs comprise environmental and toxicological relevant concentrations of many chemicals of concern. Their stability and simplicity of use in comparison with more established analytical techniques make them particularly attractive for field measurements or contamination monitoring in remote areas without access to traditional chemical laboratory facilities. In the medical field, point of care diagnostics using biosensing based on enzymes and antibodies have introduced significant improvements in glucose monitoring for diabetics or pregnancy tests, providing convenience, privacy, and lower costs. MIPs have the potential to introduce similar benefits to many other conditions, since they can be fabricated for a wide range of targets, including biomarkers for cancer, infectious, and inflammatory diseases. The approach will be transformative for both medical and environmental fields; however, and in spite of the large number of patents filed worldwide on MIPs, the technology is still mostly restricted to academic laboratories [303]. Nevertheless, there is an unquestionable market need for such a device. We have conducted an extensive customer discovery study funded by the US National Science Foundation that highlighted the demand for a point of care analytical device for hormone analysis in blood from all stakeholders: health care providers, clinical laboratories, and patients.
The translation of MIP technology from the laboratory to final product has been hindered by technological challenges in two main areas: the device design and fabrication process and the scale up of the manufacturing process. A useful device should not only capture the target compound specifically from diverse and complex matrices, where most of the current research efforts have been directed, but should also provide the user with a system to read and store the measured data. The system should be small, preferably handheld, user friendly, fast, and low cost. Recent advances in electronics and optical interfaces have contributed to the solution of this problem, and interfaces for smart phones to work with MIP-based sensors have been suggested for diagnostics/analysis platforms. A fully developed and calibrated MIP-based system is still lacking, and a significant research investment will be required for its full development.
The manufacturing process, and in particular the scale up from the academic laboratory to large batches of MIP materials, is, in our view, the main obstacle to commercialization. The lack of reproducibility of MIPs between batches, both with respect to their morphological and chemical binding properties, is the main drawback. The issue seems to be related to insufficient control of fabrication parameters, which becomes extremely challenging given the complex synthesis of some of the proposed sensors. In addition, academic laboratory protocols for manufacturing usually need to be completely reengineered when shifting to large-scale manufacturing in an industrial setting, adding more research and development tasks before a successful prototype can be obtained and requiring large investments in the early stages of product development.
Finally, mass production of MIPs will consume considerable amounts of template, which can be either unavailable or economically unfeasible. A process for target recovery and purification after elution from the newly synthesized MIPs will be needed in order to reuse the target molecule in different batches.

5. Summary and Conclusions

The search for better sensitivity and selectivity has driven significant research in the field of materials science and engineering, and many sensors have been reported with diverse nanomaterials and sophisticated manufacturing processes. The new materials are well characterized and calibrated in laboratory made solutions, with some examples of assays in real samples. Although these attempts may improve the analytical performance, they also complicate commercialization efforts due to cost and fabrication complexity.
Currently, commercialization of MIPs is focused on niche markets in biotechnology, as well as analytical and separation chemistry. Several start-ups derived from academic laboratories are currently commercializing the technology. Semorex (Fanwood, NJ, US) specializes in protein-imprinted polymers for the elimination of specific proteins from the gastrointestinal tract in the therapy of Crohn’s disease. MIP Technologies AB (Lund, Sweden) offers tailored purification resins. AFFINISEP (Petit Couronne, France) has developed a range of solid-phase-extraction phases used in food and environment analysis, life sciences, and pharmaceuticals. MIP Diagnostics Ltd. (Bedford, UK) commercializes different types of tailor-made MIPs for in vitro diagnostics. Biotage (Cardiff, UK) designs resins for the removal of low-level contaminants, or extraction of high value desirables, from any process, particularly for the food, beverage, flavor, and fragrance industries. In addition, the life science technologies and specialty chemicals company Sigma-Aldrich (St Louis, MO, US) offers solid-phase-extraction materials based on MIP technology.
There are significant obstacles to the introduction of MIP-based sensors to the consumer market. First, the interference of structural analog compounds to the target molecule is a general problem in the literature, and some of the complex materials and synthesis processes proposed to avoid this issue are not viable for large-scale production due to the extra cost and manufacturing restrictions. A reengineering of the fabrication methods will be required in most cases, as laboratory bench protocols are converted to industrial manufacturing processes. We have reviewed numerous examples of MIP-based sensors that achieved the required LOD and linear range as needed for biological and environmental use; however, the majority of these sensors are fabricated following costly and complicated methods. In order for these materials to reach the consumer, the optimization of the production method is of utmost importance so they can be efficiently mass manufactured. Secondly, while tests in real samples are included in most of the reviewed articles, they are limited to just a few promising results. Natural waters and wastewaters are the most common matrices in environmental sensing, and their composition can vary widely in pH, dissolved solids concentration, and organic matter content. Clinical trials of MIP-based sensors are the necessary first step for regulatory approval and validation of any biomedical device. Large scale testing is an expensive and time-consuming endeavor, and constitutes one of the most important roadblocks for the advancement of the technology. Although the challenges are significant, the promises of MIP technology continue to attract numerous application-minded researchers to the field, working towards the achievement of its full potential.

Author Contributions

Conceptualization, A.J.K. and M.M.F.d.C.; methodology, A.J.K.; writing—original draft preparation, A.J.K.; writing—review and editing, A.J.K., G.J.G. and M.M.F.d.C.; visualization, A.J.K. and G.J.G.; supervision, M.M.F.d.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding; A.J.K. is grateful to The Higher Committee for Education Development (HCED) and Environmental Research Center, University of Technology, Baghdad, Iraq, for a doctoral fellowship.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Haupt, K.; Linares, A.V.; Bompart, M.; Bui, B.T.S. Molecularly Imprinted Polymers. Top. Curr. Chem. 2011, 325, 1–28. [Google Scholar] [CrossRef]
  2. Beltran, A.; Borrull, F.; Marcé-Recasens, R.M.; Cormack, P. Molecularly-imprinted polymers: Useful sorbents for selective extractions. TrAC Trends Anal. Chem. 2010, 29, 1363–1375. [Google Scholar] [CrossRef]
  3. Wulff, G.; Grobe-Einsler, R.; Vesper, W.; Sarhan, A. Enzyme-analogue built polymers, 5. On the specificity distribution of chiral cavities prepared in synthetic polymers. Makromol. Chem. 1977, 178, 2817–2825. [Google Scholar] [CrossRef]
  4. Whitcombe, M.J.; Rodriguez, M.E.; Villar, P.; Vulfson, E.N. A New Method for the Introduction of Recognition Site Functionality into Polymers Prepared by Molecular Imprinting: Synthesis and Characterization of Polymeric Receptors for Cholesterol. J. Am. Chem. Soc. 1995, 117, 7105–7111. [Google Scholar] [CrossRef]
  5. Vlatakis, G.; Andersson, L.I.; Muller, R.S.; Mosbach, K. Drug assay using antibody mimics made by molecular imprinting. Nature 1993, 361, 645–647. [Google Scholar] [CrossRef] [PubMed]
  6. Arshady, R.; Mosbach, K. Synthesis of substrate-selective polymers by host-guest polymerization. Makromol. Chem. 1981, 182, 687–692. [Google Scholar] [CrossRef]
  7. Alexander, C.; Andersson, H.; Andersson, L.I.; Ansell, R.J.; Kirsch, N.; Nicholls, I.A.; O’Mahony, J.; Whitcombe, M.J. Molecular imprinting science and technology: A survey of the literature for the years up to and including 2003. J. Mol. Recognit. 2006, 19, 106–180. [Google Scholar] [CrossRef]
  8. Kupai, J.; Razali, M.; Buyuktiryaki, S.; Kecili, R.; Szekely, G. Long-term stability and reusability of molecularly imprinted polymers. Polym. Chem. 2016, 8, 666–673. [Google Scholar] [CrossRef] [Green Version]
  9. Wulff, G.; Sarhan, A. Über die Anwendung von enzymanalog gebauten Polymeren zur Racemattrennung. Angew. Chem. 1972, 84, 364. [Google Scholar] [CrossRef]
  10. Vasapollo, G.; del Sole, R.; Mergola, L.; Lazzoi, M.R.; Scardino, A.; Scorrano, S.; Mele, G.; Vasapollo, G.; del Sole, R.; Mergola, L.; et al. Molecularly Imprinted Polymers: Present and Future Prospective. Int. J. Mol. Sci. 2011, 12, 5908–5945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Asman, S.; Mohamad, S.; Sarih, N.M. Exploiting β-Cyclodextrin in Molecular Imprinting for Achieving Recognition of Benzylparaben in Aqueous Media. Int. J. Mol. Sci. 2015, 16, 3656–3676. [Google Scholar] [CrossRef]
  12. Zaidi, S.A. Molecular imprinted polymers as drug delivery vehicles. Drug Deliv. 2014, 23, 2262–2271. [Google Scholar] [CrossRef] [PubMed]
  13. Pardo, A.; Mespouille, L.; Dubois, P.; Blankert, B.; Duez, P. Molecularly Imprinted Polymers: Compromise between Flexibility and Rigidity for Improving Capture of Template Analogues. Chem. A Eur. J. 2014, 20, 3500–3509. [Google Scholar] [CrossRef] [PubMed]
  14. Caro, E.; Marce, R.; Borrull, F.; Cormack, P.; Sherrington, D. Application of molecularly imprinted polymers to solid-phase extraction of compounds from environmental and biological samples. TrAC Trends Anal. Chem. 2006, 25, 143–154. [Google Scholar] [CrossRef]
  15. Sanagi, M.M.; Salleh, S.; Ibrahim, W.A.W.; Abu Naim, A.; Hermawan, D.; Miskam, M.; Hussain, I.; Aboul-Enein, H.Y. Molecularly imprinted polymer solid-phase extraction for the analysis of organophosphorus pesticides in fruit samples. J. Food Compos. Anal. 2013, 32, 155–161. [Google Scholar] [CrossRef]
  16. Bitar, M.; Cayot, P.; Bou-Maroun, E. Molecularly imprinted polymer solid phase extraction of fungicides from wine samples. Anal. Methods 2014, 6, 6467–6472. [Google Scholar] [CrossRef]
  17. Wang, G.N.; Yang, K.; Liu, H.Z.; Feng, M.X.; Wang, J.P. Molecularly imprinted polymer-based solid phase extraction combined high performance liquid chromatography for determination of fluoroquinolones in milk. Anal. Methods 2016, 8, 5511–5518. [Google Scholar] [CrossRef]
  18. Yang, Y.; Yu, J.; Yin, J.; Shao, B.; Zhang, J. Molecularly Imprinted Solid-Phase Extraction for Selective Extraction of Bisphenol Analogues in Beverages and Canned Food. J. Agric. Food Chem. 2014, 62, 11130–11137. [Google Scholar] [CrossRef]
  19. Baltrons, O.; López-Mesas, M.; Palet, C.; Le Derf, F.; Portet-Koltalo, F. Molecularly imprinted polymer-liquid chromatography/fluorescence for the selective clean-up of hydroxylated polycyclic aromatic hydrocarbons in soils. Anal. Methods 2013, 5, 6297–6305. [Google Scholar] [CrossRef]
  20. Panahi, H.A.; Mahabadi, S.A. Application of Molecularly Imprinted Polymer for Extraction and Determination of Nalidixic Acid by High-Performance Liquid Chromatography. Sep. Sci. Technol. 2014, 50, 683–689. [Google Scholar] [CrossRef]
  21. Liu, Q.; Wan, J.; Cao, X. Synthesis of core-shell molecularly imprinted polymers (MIP) for spiramycin I and their application in MIP chromatography. Process. Biochem. 2018, 70, 168–178. [Google Scholar] [CrossRef]
  22. Yang, S.; Wang, Y.; Jiang, Y.; Li, S.; Liu, W. Molecularly Imprinted Polymers for the Identification and Separation of Chiral Drugs and Biomolecules. Polymers 2016, 8, 216. [Google Scholar] [CrossRef]
  23. Li, Z.W.; Jia, X.; Xu, C.M.; Liu, L.; Fu, D.C. Chiral Separation of Amlodipine and its Enantiomer on a Molecularly Imprinted Polymer-Based Stationary Phase. Adv. Mater. Res. 2013, 706-708, 36–39. [Google Scholar] [CrossRef]
  24. Orozco, J.; Cortés, A.; Cheng, G.; Sattayasamitsathit, S.; Gao, W.; Feng, X.; Shen, Y.; Wang, J. Molecularly Imprinted Polymer-Based Catalytic Micromotors for Selective Protein Transport. J. Am. Chem. Soc. 2013, 135, 5336–5339. [Google Scholar] [CrossRef] [Green Version]
  25. Li, S.; Zhu, M.; Whitcombe, M.J.; Piletsky, S.A.; Turner, A.P.F. Molecularly Imprinted Polymers for Enzyme-like Catalysis: Principle, Design, and Applications. In Molecularly Imprinted Catalysts; Li, S., Cao, S., Piletsky, S.A., Turner, A.P.F., Eds.; Elsevier: Amsterdam, The Netherlands, 2016; pp. 1–17. [Google Scholar]
  26. Czulak, J.; Jakubiak-Marcinkowska, A.; Trochimczuk, A. Polymer Catalysts Imprinted with Metal Ions as Biomimics of Metalloenzymes. Adv. Mater. Sci. Eng. 2013, 2013, 464265. [Google Scholar] [CrossRef] [Green Version]
  27. Wei, W.; Zhou, T.; Wu, S.; Shen, X.; Zhu, M.; Li, S. An enzyme-like imprinted-polymer reactor with segregated quantum confinements for a tandem catalyst. RSC Adv. 2018, 8, 1610–1620. [Google Scholar] [CrossRef] [Green Version]
  28. Li, S.; Ge, Y.; Turner, A.P.F. A Catalytic and Positively Thermosensitive Molecularly Imprinted Polymer. Adv. Funct. Mater. 2011, 21, 1194–1200. [Google Scholar] [CrossRef]
  29. Kurczewska, J.; Cegłowski, M.; Pecyna, P.; Ratajczak, M.; Gajecka, M.; Schroeder, G. Molecularly imprinted polymer as drug delivery carrier in alginate dressing. Mater. Lett. 2017, 201, 46–49. [Google Scholar] [CrossRef]
  30. Fareghi, A.R.; Moghadam, P.N.; Khalafy, J.; Bahram, M.; Moghtader, M. Preparation of a new molecularly imprinted polymer based on self-crosslinkable cellulose acrylate in aqueous solution: A drug delivery system for furosemide. J. Appl. Polym. Sci. 2017, 134, 45581. [Google Scholar] [CrossRef]
  31. Anirudhan, T.; Divya, P.; Nima, J. Silylated montmorillonite based molecularly imprinted polymer for the selective binding and controlled release of thiamine hydrochloride. React. Funct. Polym. 2013, 73, 1144–1155. [Google Scholar] [CrossRef]
  32. Mohajeri, S.A.; Tabassi, S.A.S.; Moghadam, M.H. Preparation of a pH-sensitive pantoprazole-imprinted polymer and evaluation of its drug-binding and -releasing properties. Sci. China Ser. B Chem. 2014, 57, 857–865. [Google Scholar] [CrossRef]
  33. Kempe, H.; Pujolràs, A.P.; Kempe, M. Molecularly Imprinted Polymer Nanocarriers for Sustained Release of Erythromycin. Pharm. Res. 2014, 32, 375–388. [Google Scholar] [CrossRef]
  34. Shumyantseva, V.V.; Bulko, T.V.; Baychorov, I.H.; Archakov, A.I. Molecularly imprinted polymers (MIP) in electroanalysis of proteins. Biochem. Suppl. Ser. B Biomed. Chem. 2016, 10, 145–151. [Google Scholar] [CrossRef]
  35. Liu, J.; Deng, Q.; Tao, D.; Yang, K.; Zhang, L.; Liang, Z.; Zhang, Y. Preparation of protein imprinted materials by hierarchical imprinting techniques and application in selective depletion of albumin from human serum. Sci. Rep. 2014, 4, 5487. [Google Scholar] [CrossRef]
  36. Kuwata, T.; Uchida, A.; Takano, E.; Kitayama, Y.; Takeuchi, T. Molecularly Imprinted Polymer Arrays as Synthetic Protein Chips Prepared by Transcription-type Molecular Imprinting by Use of Protein-Immobilized Dots as Stamps. Anal. Chem. 2015, 87, 11784–11791. [Google Scholar] [CrossRef] [PubMed]
  37. Rossetti, C.; Świtnicka-Plak, M.A.; Halvorsen, T.G.; Cormack, P.A.; Sellergren, B.; Reubsaet, L. Automated Protein Biomarker Analysis: Online extraction of clinical samples by Molecularly Imprinted Polymers. Sci. Rep. 2017, 7, srep44298. [Google Scholar] [CrossRef]
  38. Demir, E.F.; Özçalışkan, E.; Karakaş, H.; Uygun, M.; Uygun, D.A.; Akgöl, S.; Denizli, A. Synthesis and characterization of albumin imprinted polymeric hydrogel membranes for proteomic studies. J. Biomater. Sci. Polym. Ed. 2018, 29, 2218–2236. [Google Scholar] [CrossRef]
  39. Ji, J.; Zhou, Z.; Zhao, X.; Sun, J.; Sun, X. Electrochemical sensor based on molecularly imprinted film at Au nanoparticles-carbon nanotubes modified electrode for determination of cholesterol. Biosens. Bioelectron. 2014, 66, 590–595. [Google Scholar] [CrossRef]
  40. Wang, F.-R.; Lee, G.-J.; Haridharan, N.; Wu, J.J. Electrochemical Sensor Using Molecular Imprinting Polymerization Modified Electrodes to Detect Methyl Parathion in Environmental Media. Electrocatalysis 2017, 9, 1–9. [Google Scholar] [CrossRef]
  41. Zhang, B.; Lu, L.; Huang, F.; Lin, Z. [Ru(bpy)3] 2+-mediated photoelectrochemical detection of bisphenol A on a molecularly imprinted polypyrrole modified SnO2 electrode. Anal. Chim. Acta 2015, 887, 59–66. [Google Scholar] [CrossRef]
  42. Zhang, J.; Niu, Y.; Li, S.; Luo, R.; Wang, C. A molecularly imprinted electrochemical sensor based on sol–gel technology and multiwalled carbon nanotubes–Nafion functional layer for determination of 2-nonylphenol in environmental samples. Sens. Actuators B Chem. 2014, 193, 844–850. [Google Scholar] [CrossRef]
  43. Zheng, W.; Wu, H.; Jiang, Y.; Xu, J.; Li, X.; Zhang, W.; Qiu, F. A molecularly-imprinted-electrochemical-sensor modified with nano-carbon-dots with high sensitivity and selectivity for rapid determination of glucose. Anal. Biochem. 2018, 555, 42–49. [Google Scholar] [CrossRef] [PubMed]
  44. Schreiber, T.; Weber, A.; Niedergall, K.; Riegler, J.; Bryniok, D.; Hirth, T.; Tovar, G.E.M. Water treatment by molecularly imprinted polymer nanoparticles. MRS Proc. 2009, 1169. [Google Scholar] [CrossRef]
  45. Shen, X.; Xu, C.; Ye, L. Molecularly Imprinted Polymers for Clean Water: Analysis and Purification. Indian Eng. Chem. Res. 2012, 52, 13890–13899. [Google Scholar] [CrossRef]
  46. Martínez, L.D.D.L.; Rodríguez-Aguilar, M.; Perez, R.O.; Gutiérrez-Hernández, J.M.; Díaz-Barriga, F.; Batres-Esquivel, L.; Flores-Ramírez, R. Synthesis and Evaluation of a Molecularly Imprinted Polymer for the Determination of Metronidazole in Water Samples. Bull. Environ. Contam. Toxicol. 2018, 100, 395–401. [Google Scholar] [CrossRef]
  47. Dai, C.; Zhang, J.; Zhang, Y.; Zhou, X.; Liu, S. Application of Molecularly Imprinted Polymers to Selective Removal of Clofibric Acid from Water. PLoS ONE 2013, 8, e78167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Okutucu, B.; Sanlıer, S.H. Decolorization of textile wastewater by dye-imprinted polymer. Desalin. Water Treat. 2015, 57, 21577–21584. [Google Scholar] [CrossRef]
  49. Yusof, N.A.; Zakaria, N.D.; Maamor, N.A.M.; Abdullah, A.H.; Haron, J. Synthesis and Characterization of Molecularly Imprinted Polymer Membrane for the Removal of 2,4-Dinitrophenol. Int. J. Mol. Sci. 2013, 14, 3993–4004. [Google Scholar] [CrossRef]
  50. Kashani, T.; Jahanshahi, M.; Rahimpour, A.; Peyravi, M. Nanopore Molecularly Imprinted Polymer Membranes for Environmental Usage: Selective Separation of 2,4-Dichlorophenoxyacetic Acid as a Toxic Herbicide from Water. Polym. Technol. Eng. 2016, 55, 1700–1712. [Google Scholar] [CrossRef]
  51. Ghasemi, S.; Nematollahzadeh, A. Molecularly imprinted ultrafiltration polysulfone membrane with specific nano-cavities for selective separation and enrichment of paclitaxel from plant extract. React. Funct. Polym. 2018, 126, 9–19. [Google Scholar] [CrossRef]
  52. Gao, B.; Li, Y.; Cui, K. Molecularly imprinted membrane with innovative structure and high performance for chiral separation of amino acids. Int. J. Polym. Mater. 2017, 67, 517–527. [Google Scholar] [CrossRef]
  53. Ng, M.H.K.; Leo, C.P.; Abdullah, A.Z. Selective removal of dyes by molecular imprinted TiO2 nanoparticles in polysulfone ultrafiltration membrane. J. Environ. Chem. Eng. 2017, 5, 3991–3998. [Google Scholar] [CrossRef]
  54. BelBruno, J.J. Molecularly Imprinted Polymers. Chem. Rev. 2018, 119, 94–119. [Google Scholar] [CrossRef] [PubMed]
  55. Wulff, G. The role of binding-site interactions in the molecular imprinting of polymers. Trends Biotechnol. 1993, 11, 85–87. [Google Scholar] [CrossRef]
  56. Cormack, P.A.; Elorza, A.Z. Molecularly imprinted polymers: Synthesis and characterisation. J. Chromatogr. B 2004, 804, 173–182. [Google Scholar] [CrossRef]
  57. Yemiş, F.; Alkan, P.; Yenigül, B.; Yenigül, M. Molecularly Imprinted Polymers and Their Synthesis by Different Methods. Polym. Polym. Compos. 2013, 21, 145–150. [Google Scholar] [CrossRef]
  58. Lin, Z.; DeMarr, V.; Bao, J.; Wu, T. Molecularly Imprinted Polymer-Based Biosensors: For the Early, Rapid Detection of Pathogens, Biomarkers, and Toxins in Clinical, Environmental, or Food Samples. IEEE Nanotechnol. Mag. 2018, 12, 6–13. [Google Scholar] [CrossRef]
  59. Yan, H.; Row, K.H. Characteristic and Synthetic Approach of Molecularly Imprinted Polymer. Int. J. Mol. Sci. 2006, 7, 155–178. [Google Scholar] [CrossRef] [Green Version]
  60. Lee, S.H.; Doong, R.A. Design of Size-Tunable Molecularly Imprinted Polymer for Selective Adsorption of Pharmaceuticals and Biomolecules. J. Biosens. Bioelectron. 2016, 7, 4. [Google Scholar] [CrossRef] [Green Version]
  61. Chen, L.; Wang, X.; Lu, W.; Wu, X.; Li, J. Molecular imprinting: Perspectives and applications. Chem. Soc. Rev. 2016, 45, 2137–2211. [Google Scholar] [CrossRef]
  62. Zaidi, S.A. Molecular imprinting polymers and their composites: A promising material for diverse applications. Biomater. Sci. 2017, 5, 388–402. [Google Scholar] [CrossRef]
  63. Ramanavicius, S.; Ramanavicius, A. Conducting Polymers in the Design of Biosensors and Biofuel Cells. Polymers 2020, 13, 49. [Google Scholar] [CrossRef]
  64. Blanco-López, M.; Lobo-Castañón, M.J.; Miranda-Ordieres, A.; Tuñón-Blanco, P. Electrochemical sensors based on molecularly imprinted polymers. Trends Anal. Chem. 2004, 23, 36–48. [Google Scholar] [CrossRef]
  65. Heinze, J.; Frontana-Uribe, B.A.; Ludwigs, S. Electrochemistry of Conducting Polymers—Persistent Models and New Concepts. Chem. Rev. 2010, 110, 4724–4771. [Google Scholar] [CrossRef]
  66. Ertürk, G.; Mattiasson, B. Molecular Imprinting Techniques Used for the Preparation of Biosensors. Sensors 2017, 17, 288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Rachkov, A.; Minoura, N. Towards molecularly imprinted polymers selective to peptides and proteins. The epitope approach. Biochim. Biophys. Acta Protein Struct. Mol. Enzym. 2001, 1544, 255–266. [Google Scholar] [CrossRef]
  68. Justino, C.I.; Freitas, A.; Pereira, R.; Duarte, A.C.; Santos, T.A.R. Recent developments in recognition elements for chemical sensors and biosensors. Trends Anal. Chem. 2015, 68, 2–17. [Google Scholar] [CrossRef]
  69. Algieri, C.; Drioli, E.; Guzzo, L.; Donato, L. Bio-Mimetic Sensors Based on Molecularly Imprinted Membranes. Sensors 2014, 14, 13863–13912. [Google Scholar] [CrossRef] [Green Version]
  70. Sanjuán, A.M.; Ruiz, J.A.R.; García, F.C.; García, J.M. Recent developments in sensing devices based on polymeric systems. React. Funct. Polym. 2018, 133, 103–125. [Google Scholar] [CrossRef]
  71. Dhanjai; Sinha, A.; Wu, L.; Lu, X.; Chen, J.; Jain, R. Advances in sensing and biosensing of bisphenols: A review. Anal. Chim. Acta 2018, 998, 1–27. [Google Scholar] [CrossRef]
  72. Malinauskas, A. Electrochemical response of ascorbic acid at conducting and electrogenerated polymer modified electrodes for electroanalytical applications: A review. Talanta 2004, 64, 121–129. [Google Scholar] [CrossRef]
  73. Rahmadhani, S.; Setiyanto, H.; Zulfikar, M.A. Electropolymerized of aniline as a new molecularly imprinted polymer for determination of phenol: A study for phenol sensor. In Proceedings of the 2017 International Seminar on Sensors, Instrumentation, Measurement and Metrology (ISSIMM), Surabaya, Indonesia, 25–26 August 2017; pp. 124–128. [Google Scholar] [CrossRef]
  74. Nguyen, T.H.; Ansell, R.J. N-isopropylacrylamide as a functional monomer for noncovalent molecular imprinting. J. Mol. Recognit. 2011, 25, 1–10. [Google Scholar] [CrossRef] [PubMed]
  75. Shumyantseva, V.V.; Bulko, T.V.; Suprun, E.V.; Kuzikov, A.V.; Agafonova, L.E.; Archakov, A.I. Electrochemical methods in biomedical studies. Biochem. Suppl. Ser. B Biomed. Chem. 2015, 9, 228–243. [Google Scholar] [CrossRef]
  76. Li, X.; He, Y.; Zhao, F.; Zhang, W.; Ye, Z. Molecularly imprinted polymer-based sensors for atrazine detection by electropolymerization of o-phenylenediamine. RSC Adv. 2015, 5, 56534–56540. [Google Scholar] [CrossRef]
  77. Mazzotta, E.; Turco, A.; Chianella, I.; Guerreiro, A.; Piletsky, S.A.; Malitesta, C. Solid-phase synthesis of electroactive nanoparticles of molecularly imprinted polymers. A novel platform for indirect electrochemical sensing applications. Sens. Actuators B Chem. 2016, 229, 174–180. [Google Scholar] [CrossRef] [Green Version]
  78. Udomsap, D.; Brisset, H.; Culioli, G.; Dollet, P.; Laatikainen, K.; Siren, H.; Branger, C. Electrochemical molecularly imprinted polymers as material for pollutant detection. Mater. Today Commun. 2018, 17, 458–465. [Google Scholar] [CrossRef]
  79. Rebocho, S.; Cordas, C.; Viveiros, R.; Casimiro, T. Development of a ferrocenyl-based MIP in supercritical carbon dioxide: Towards an electrochemical sensor for bisphenol A. J. Supercrit. Fluids 2018, 135, 98–104. [Google Scholar] [CrossRef]
  80. Ekomo, V.M.; Branger, C.; Gavrila, A.-M.; Sarbu, A.; Koutsouras, D.A.; Stolz, C.; Malliaras, G.G.; Brisset, H. Electrochemical molecularly imprinted polymers in microelectrode devices. MRS Commun. 2020, 10, 324–331. [Google Scholar] [CrossRef]
  81. Wang, H.; Li, W.; He, X.; Chen, L.; Zhang, Y. m-Aminophenylboronic acid as a functional monomer for fabricating molecularly imprinted polymer for the recognition of bovine serum albumin. React. Funct. Polym. 2008, 68, 1291–1296. [Google Scholar] [CrossRef] [Green Version]
  82. Haupt, K. Peer Reviewed: Molecularly Imprinted Polymers: The Next Generation. Anal. Chem. 2003, 75, 376A–383A. [Google Scholar] [CrossRef] [Green Version]
  83. Yuan, W.; Cai, Y.; Chen, Y.; Hong, X.; Liu, Z. Porous microsphere and its applications. Int. J. Nanomed. 2013, 8, 1111–1120. [Google Scholar] [CrossRef] [Green Version]
  84. Wackerlig, J.; Lieberzeit, P.A. Molecularly imprinted polymer nanoparticles in chemical sensing—Synthesis, characterisation and application. Sen. Actuators B Chem. 2015, 207, 144–157. [Google Scholar] [CrossRef]
  85. He, J.; Tang, H.; You, L.; Yuan, L.; Liu, Z.; Zhu, J.; Lu, K.; Chen, X. Synthesis of Fragment-Imprinted Microspheres of 2,6-Dichloropyrimidine as Templates and Determination of Sulfonamides in Milk Samples. Chromatographia 2013, 76, 959–965. [Google Scholar] [CrossRef]
  86. Lv, Y.-K.; Zhao, C.-X.; Li, P.; He, Y.-D.; Yang, Z.-R.; Sun, H.-W. Preparation of doxycycline-imprinted magnetic microspheres by inverse-emulsion suspension polymerization for magnetic dispersion extraction of tetracyclines from milk samples. J. Sep. Sci. 2013, 36, 2656–2663. [Google Scholar] [CrossRef]
  87. Su, X.; Li, X.; Li, J.; Liu, M.; Lei, F.; Tan, X.; Li, P.; Luo, W. Synthesis and characterization of core–shell magnetic molecularly imprinted polymers for solid-phase extraction and determination of Rhodamine B in food. Food Chem. 2014, 171, 292–297. [Google Scholar] [CrossRef] [PubMed]
  88. Mayes, A.G.; Mosbach, K. Molecularly Imprinted Polymer Beads: Suspension Polymerization Using a Liquid Perfluorocarbon as the Dispersing Phase. Anal. Chem. 1996, 68, 3769–3774. [Google Scholar] [CrossRef] [PubMed]
  89. Xia, Q.; Yun, Y.; Li, Q.; Huang, Z.; Liang, Z. Preparation and characterization of monodisperse molecularly imprinted polymer microspheres by precipitation polymerization for kaempferol. Des. Monomers Polym. 2016, 20, 201–209. [Google Scholar] [CrossRef] [Green Version]
  90. Kitabatake, T.; Tabo, H.; Matsunaga, H.; Haginaka, J. Preparation of monodisperse curcumin-imprinted polymer by precipitation polymerization and its application for the extraction of curcuminoids from Curcuma longa L. Anal. Bioanal. Chem. 2013, 405, 6555–6561. [Google Scholar] [CrossRef]
  91. Yu, Z.; Su, Q.; Tang, Y.; Xu, Z. Preparation and evaluation of aconitine imprinted microspheres and its application to body fluid samples. J. Appl. Polym. Sci. 2012, 128, 3425–3431. [Google Scholar] [CrossRef]
  92. Zhao, C.; Dai, J.; Zhou, Z.; Dai, X.; Zou, Y.; Yu, P.; Zou, T.; Li, C.; Yan, Y. One-pot method for obtaining hydrophilic tetracycline-imprinted particles via precipitation polymerization in ethanol. J. Appl. Polym. Sci. 2013, 131, 1–11. [Google Scholar] [CrossRef]
  93. Puoci, F.; Hampel, S.; Parisi, O.I.; Hassan, A.; Cirillo, G.; Picci, N. Imprinted microspheres doped with carbon nanotubes as novel electroresponsive drug-delivery systems. J. Appl. Polym. Sci. 2013, 130, 829–834. [Google Scholar] [CrossRef]
  94. Zhou, T.; Shen, X.; Chaudhary, S.; Ye, L. Molecularly imprinted polymer beads prepared by pickering emulsion polymerization for steroid recognition. J. Appl. Polym. Sci. 2013, 131, 1–7. [Google Scholar] [CrossRef]
  95. Hang, H.; Li, C.; Pan, J.; Li, L.; Dai, J.; Dai, X.; Yu, P.; Feng, Y. Selective separation of lambdacyhalothrin by porous/magnetic molecularly imprinted polymers prepared by Pickering emulsion polymerization. J. Sep. Sci. 2013, 36, 3285–3294. [Google Scholar] [CrossRef]
  96. Zhou, T.; Kamra, T.; Ye, L. Preparation of diclofenac-imprinted polymer beads for selective molecular separation in water. J. Mol. Recognit. 2017, 31, e2608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Pan, J.; Li, L.; Hang, H.; Wu, R.; Dai, X.; Shi, W.; Yan, Y. Fabrication and Evaluation of Magnetic/Hollow Double-Shelled Imprinted Sorbents Formed by Pickering Emulsion Polymerization. Langmuir 2013, 29, 8170–8178. [Google Scholar] [CrossRef]
  98. Li, X.-J.; Zhong, S.-A.; Li, C.-E. Synthesis of novel photoresponsive molecularly imprinted polymer microspheres with special binding properties. J. Appl. Polym. Sci. 2013, 130, 869–876. [Google Scholar] [CrossRef]
  99. Zheng, X.-F.; Lian, Q.; Yang, H.; Wang, X. Surface Molecularly Imprinted Polymer of Chitosan Grafted Poly(methyl methacrylate) for 5-Fluorouracil and Controlled Release. Sci. Rep. 2016, 6, 21409. [Google Scholar] [CrossRef] [Green Version]
  100. Zhang, Y.-J.; Feng, R.; Qi, N.; Zhang, Y.-P.; Bai, L.-Y.; Huang, M.-X. Synthesis and Evaluation of Cyromazine Molecularly Imprinted Polymeric Microspheres by Two-Step Seed Swelling Polymerization. Asian J. Chem. 2013, 25, 8329–8332. [Google Scholar] [CrossRef]
  101. Chen, J.; Bai, L.; Zhang, Y.; Chen, N.; Zhang, Y. Fabrication of Atrazine Molecularly Imprinted Polymer Microsphere by Two Step Seed Swelling Polymerization Method. J. Chin. Chem. Soc. 2012, 59, 1493–1499. [Google Scholar] [CrossRef]
  102. Qi, N.; Chen, J.; Zhang, Y.P.; Zhang, Y.J.; Bai, L.Y. Preparation of Melamine Molecularly Imprinted Polymeric Microspheres by Two-Step Seed Swelling Polymerization. Adv. Mater. Res. 2013, 668, 128–131. [Google Scholar] [CrossRef]
  103. Zhang, Y.; Yao, X. Preparation of molecularly imprinted polymer for vanillin via seed swelling and suspension polymerization. Polym. Sci. Ser. B 2014, 56, 538–545. [Google Scholar] [CrossRef]
  104. Wolska, J.; Bryjak, M. Removal of Bisphenol A from Aqueous Solution by Molecularly Imprinted Polymers. Sep. Sci. Technol. 2014, 49, 1643–1653. [Google Scholar] [CrossRef]
  105. Shi, X.; Wu, A.; Zheng, S.; Li, R.; Zhang, D. Molecularly imprinted polymer microspheres for solid-phase extraction of chloramphenicol residues in foods. J. Chromatogr. B 2007, 850, 24–30. [Google Scholar] [CrossRef]
  106. Brooks, B. Suspension Polymerization Processes. Chem. Eng. Technol. 2010, 33, 1737–1744. [Google Scholar] [CrossRef] [Green Version]
  107. Nabavi, S.A.; Vladisavljević, G.T.; Eguagie, E.M.; Li, B.; Georgiadou, S.; Manović, V. Production of spherical mesoporous molecularly imprinted polymer particles containing tunable amine decorated nanocavities with CO2 molecule recognition properties. Chem. Eng. J. 2016, 306, 214–225. [Google Scholar] [CrossRef] [Green Version]
  108. Qian, L.-W.; Hu, X.-L.; Guan, P.; Gao, B.; Wang, D.; Wang, C.-L.; Li, J.; Du, C.-B.; Song, W.-Q. Thermal preparation of lysozyme-imprinted microspheres by using ionic liquid as a stabilizer. Anal. Bioanal. Chem. 2014, 406, 7221–7231. [Google Scholar] [CrossRef]
  109. Zhang, Y.; Ding, J.; Gong, S. Preparation of molecularly imprinted polymers for vanillin via reversible addition-fragmentation chain transfer suspension polymerization. J. Appl. Polym. Sci. 2012, 128, 2927–2932. [Google Scholar] [CrossRef]
  110. Suwanwong, Y.; Kulkeratiyut, S.; Prachayasittikul, V.; Boonpangrak, S. Effects of Polymerization Methods and Functional Monomers on Curcumin Imprinted Polymer Properties. Sep. Sci. Technol. 2014, 49, 1086–1095. [Google Scholar] [CrossRef]
  111. Sun, H.; Lai, J.-P.; Chen, F.; Zhu, D.-R. Molecularly imprinted microspheres synthesized by a simple, fast, and universal suspension polymerization for selective extraction of the topical anesthetic benzocaine in human serum and fish tissues. Anal. Bioanal. Chem. 2015, 407, 1745–1752. [Google Scholar] [CrossRef]
  112. Moreno Bondi, M.C.; Urraca, J.L.; Carrasco, S.; Navarro-Villoslada, F. Preparation of Molecularly Imprinted Polymers. In Handbook of Molecularly Imprinted Polymers; Álvarez-Lorenzo, C., Concheiro, Á., Eds.; Smithers Rapra Technology Ltd.: Billingham, UK, 2013. [Google Scholar]
  113. Li, G.L.; Möhwald, H.; Shchukin, D.G. Precipitation polymerization for fabrication of complex core–shell hybrid particles and hollow structures. Chem. Soc. Rev. 2013, 42, 3628–3646. [Google Scholar] [CrossRef]
  114. Cacho, C.; Turiel, E.; Pérez-Conde, C. Molecularly imprinted polymers: An analytical tool for the determination of benzimidazole compounds in water samples. Talanta 2009, 78, 1029–1035. [Google Scholar] [CrossRef]
  115. Pardeshi, S.; Dhodapkar, R.; Kumar, A. Molecularly imprinted microspheres and nanoparticles prepared using precipitation polymerisation method for selective extraction of gallic acid from Emblica officinalis. Food Chem. 2013, 146, 385–393. [Google Scholar] [CrossRef]
  116. Ye, L.; Cormack, P.A.G.; Mosbach, K. Molecularly imprinted monodisperse microspheres for competitive radioassay. Anal. Commun. 1999, 36, 35–38. [Google Scholar] [CrossRef]
  117. Liu, M.; Li, Y.; Han, J.; Dong, X. Synthesis of tetracycline-imprinted polymer microspheres by reversible addition-fragmentation chain-transfer precipitation polymerization using polyethylene glycol as a coporogen. J. Sep. Sci. 2014, 37, 1118–1125. [Google Scholar] [CrossRef]
  118. Gao, F.-X.; Ma, X.-T.; He, X.-W.; Li, W.-Y.; Zhang, Y.-K. Smart surface imprinting polymer nanospheres for selective recognition and separation of glycoprotein. Colloids Surf. A Physicochem. Eng. Asp. 2013, 433, 191–199. [Google Scholar] [CrossRef]
  119. Azodi-Deilami, S.; Abdouss, M.; Kordestani, D.; Shariatinia, Z. Preparation of N,N-p-phenylene bismethacryl amide as a novel cross-link agent for synthesis and characterization of the core–shell magnetic molecularly imprinted polymer nanoparticles. J. Mater. Sci. Mater. Electron. 2013, 25, 645–656. [Google Scholar] [CrossRef]
  120. Yao, T.; Gu, X.; Li, T.; Li, J.; Li, J.; Zhao, Z.; Wang, J.; Qin, Y.; She, Y. Enhancement of surface plasmon resonance signals using a MIP/GNPs/rGO nano-hybrid film for the rapid detection of ractopamine. Biosens. Bioelectron. 2015, 75, 96–100. [Google Scholar] [CrossRef]
  121. Tan, F.; Zhao, Q.; Teng, F.; Sun, D.; Gao, J.; Quan, X.; Chen, J. Molecularly imprinted polymer/mesoporous carbon nanoparticles as electrode sensing material for selective detection of ofloxacin. Mater. Lett. 2014, 129, 95–97. [Google Scholar] [CrossRef]
  122. Sellergren, B.; Rückert, B.; Hall, A. Layer-by-Layer Grafting of Molecularly Imprinted Polymers via Iniferter Modified Supports. Adv. Mater. 2002, 14, 1204–1208. [Google Scholar] [CrossRef]
  123. Rückert, B.; Hall, A.J.; Sellergren, B. Molecularly imprinted composite materials via iniferter-modified supports. J. Mater. Chem. 2002, 12, 2275–2280. [Google Scholar] [CrossRef]
  124. Barahona, F.; Turiel, E.; Cormack, P.A.G.; Martín-Esteban, A. Chromatographic performance of molecularly imprinted polymers: Core-shell microspheres by precipitation polymerization and grafted MIP films via iniferter-modified silica beads. J. Polym. Sci. Part A Polym. Chem. 2010, 48, 1058–1066. [Google Scholar] [CrossRef]
  125. Minko, S. Grafting on Solid Surfaces: “Grafting to”and “Grafting from” Methods. In Polymer Surfaces and Interfaces: Characterization, Modification and Applications; Stamm, M., Ed.; Springer: Berlin/Heidelberg, Germany, 2008; pp. 215–234. [Google Scholar]
  126. Chern, C. Emulsion polymerization mechanisms and kinetics. Prog. Polym. Sci. 2006, 31, 443–486. [Google Scholar] [CrossRef]
  127. Yamak, H.B. Emulsion Polymerization: Effects of Polymerization Variables on the Properties of Vinyl Acetate Based Emulsion Polymers. In Polymer Science; Yilmaz, F., Ed.; InTech: London, UK, 2013. [Google Scholar]
  128. Matsui, J.; Kato, T.; Takeuchi, T.; Suzuki, M.; Yokoyama, K.; Tamiya, E.; Karube, I. Molecular recognition in continuous polymer rods prepared by a molecular imprinting technique. Anal. Chem. 1993, 65, 2223–2224. [Google Scholar] [CrossRef]
  129. Matsui, J.; Miyoshi, Y.; Matsui, R.; Takeuchi, T. Rod-Type Affinity Media for Liquid Chromatography Prepared by in-situ-Molecular Imprinting. Anal. Sci. 1995, 11, 1017–1019. [Google Scholar] [CrossRef] [Green Version]
  130. MacDougall, D.; Francis, J.A.; Cox, G.V.; Grosby, D.G.; Estes, F.L.; Freeman, D.H.; Gibbs, W.E.; Gordon, G.E.; Keith, L.H.; Lal, J.; et al. Guidelines for data acquisition and data quality evaluation in environmental chemistry. Anal. Chem. 1980, 52, 2242–2249. [Google Scholar] [CrossRef]
  131. Alvarez-Lorenzo, C. Handbook of Molecularly Imprinted Polymers; Smithers Information Ltd.: Akron, OH, USA, 2013. [Google Scholar]
  132. Liu, C.-C. Electrochemical Sensors. In The Biomedical Engineering Handbook Medical Devices and Systems, 3rd ed.; Bronzino, J.D., Ed.; CRC Press Taylor & Francis Group: Boca Raton, FL, USA, 2006. [Google Scholar]
  133. Jolly, P.; Tamboli, V.; Harniman, R.L.; Estrela, P.; Allender, C.J.; Bowen, J.L. Aptamer–MIP hybrid receptor for highly sensitive electrochemical detection of prostate specific antigen. Biosens. Bioelectron. 2016, 75, 188–195. [Google Scholar] [CrossRef] [Green Version]
  134. 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]
  135. Zhang, W.; Xiong, H.; Chen, M.; Zhang, X.; Wang, S. Surface-enhanced molecularly imprinted electrochemiluminescence sensor based on Ru@SiO 2 for ultrasensitive detection of fumonisin B1. Biosens. Bioelectron. 2017, 96, 55–61. [Google Scholar] [CrossRef]
  136. Graniczkowska, K.; Pütz, M.; Hauser, F.M.; De Saeger, S.; Beloglazova, N.V. Capacitive sensing of N-formylamphetamine based on immobilized molecular imprinted polymers. Biosens. Bioelectron. 2017, 92, 741–747. [Google Scholar] [CrossRef] [Green Version]
  137. Lenain, P.; de Saeger, S.; Mattiasson, B.; Hedström, M. Affinity sensor based on immobilized molecular imprinted synthetic recognition elements. Biosens. Bioelectron. 2015, 69, 34–39. [Google Scholar] [CrossRef]
  138. Warwick, C.; Guerreiro, A.; Gomez-Caballero, A.; Wood, E.; Kitson, J.; Robinson, J.; Soares, A. Conductance based sensing and analysis of soluble phosphates in wastewater. Biosens. Bioelectron. 2013, 52, 173–179. [Google Scholar] [CrossRef] [Green Version]
  139. Rosy; Chasta, H.; Goyal, R.N. Molecularly imprinted sensor based on o-aminophenol for the selective determination of norepinephrine in pharmaceutical and biological samples. Talanta 2014, 125, 167–173. [Google Scholar] [CrossRef]
  140. Uygun, Z.O.; Dilgin, Y. A novel impedimetric sensor based on molecularly imprinted polypyrrole modified pencil graphite electrode for trace level determination of chlorpyrifos. Sens. Actuators B Chem. 2013, 188, 78–84. [Google Scholar] [CrossRef]
  141. Ratautaite, V.; Janssens, S.; Haenen, K.; Nesládek, M.; Ramanaviciene, A.; Baleviciute, I.; Ramanavicius, A. Molecularly Imprinted Polypyrrole Based Impedimentric Sensor for Theophylline Determination. Electrochim. Acta 2014, 130, 361–367. [Google Scholar] [CrossRef]
  142. Gurtova, O.; Ye, L.; Chmilenko, F. Potentiometric propranolol-selective sensor based on molecularly imprinted polymer. Anal. Bioanal. Chem. 2012, 405, 287–295. [Google Scholar] [CrossRef]
  143. Bagheri, H.; Shirzadmehr, A.; Rezaei, M. Designing and fabrication of new molecularly imprinted polymer-based potentiometric nano-graphene/ionic liquid/carbon paste electrode for the determination of losartan. J. Mol. Liq. 2015, 212, 96–102. [Google Scholar] [CrossRef]
  144. Basozabal, I.; Guerreiro, A.; Gomez-Caballero, A.; Goicolea, M.A.; Barrio, R.J. Direct potentiometric quantification of histamine using solid-phase imprinted nanoparticles as recognition elements. Biosens. Bioelectron. 2014, 58, 138–144. [Google Scholar] [CrossRef] [PubMed]
  145. Anirudhan, T.S.; Alexander, S. Design and fabrication of molecularly imprinted polymer-based potentiometric sensor from the surface modified multiwalled carbon nanotube for the determination of lindane (γ-hexachlorocyclohexane), an organochlorine pesticide. Biosens. Bioelectron. 2015, 64, 586–593. [Google Scholar] [CrossRef]
  146. Rizk, M.; Toubar, S.S.; Sayour, H.E.E.-D.; Mohamed, D.; Touny, R.M. A new potentiometric sensor based on molecularly imprinted polymer for analysis of a veterinary drug imidocarb dipropionate. Eur. J. Chem. 2014, 5, 18–23. [Google Scholar] [CrossRef] [Green Version]
  147. Li, L.; Liang, Y.; Liu, Y. Designing of molecularly imprinted polymer-based potentiometric sensor for the determination of heparin. Anal. Biochem. 2013, 434, 242–246. [Google Scholar] [CrossRef]
  148. Kou, L.-J.; Liang, R.-N.; Wang, X.-W.; Chen, Y.; Qin, W. Potentiometric sensor for determination of neutral bisphenol A using a molecularly imprinted polymer as a receptor. Anal. Bioanal. Chem. 2013, 405, 4931–4936. [Google Scholar] [CrossRef]
  149. Zarezadeh, A.; Rajabi, H.R.; Sheydaei, O.; Khajehsharifi, H. Application of a nano-structured molecularly imprinted polymer as an efficient modifier for the design of captopril drug selective sensor: Mechanism study and quantitative determination. Mater. Sci. Eng. C 2018, 94, 879–885. [Google Scholar] [CrossRef]
  150. 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]
  151. Liu, B.; Xiao, B.; Cui, L.; Wang, M. Molecularly imprinted electrochemical sensor for the highly selective and sensitive determination of melamine. Mater. Sci. Eng. C 2015, 55, 457–461. [Google Scholar] [CrossRef]
  152. Tan, X.; Hu, Q.; Wu, J.; Li, X.; Li, P.; Yu, H.; Li, X.; Lei, F. Electrochemical sensor based on molecularly imprinted polymer reduced graphene oxide and gold nanoparticles modified electrode for detection of carbofuran. Sens. Actuators B Chem. 2015, 220, 216–221. [Google Scholar] [CrossRef]
  153. Zhao, L.; Zeng, B.; Zhao, F. Electrochemical determination of tartrazine using a molecularly imprinted polymer—Multiwalled carbon nanotubes—Ionic liquid supported Pt nanoparticles composite film coated electrode. Electrochim. Acta 2014, 146, 611–617. [Google Scholar] [CrossRef]
  154. Pacheco, J.; Castro, M.; Machado, S.; Barroso, M.F.; Nouws, H.; Delerue-Matos, C. Molecularly imprinted electrochemical sensor for ochratoxin A detection in food samples. Sens. Actuators B Chem. 2015, 215, 107–112. [Google Scholar] [CrossRef]
  155. Nezhadali, A.; Mojarrab, M. Fabrication of an electrochemical molecularly imprinted polymer triamterene sensor based on multivariate optimization using multi-walled carbon nanotubes. J. Electroanal. Chem. 2015, 744, 85–94. [Google Scholar] [CrossRef]
  156. Lei, R.; Guo, C.; Xiong, H.; Dong, C.; Zhang, X.; Wang, S. A Novel Electrochemical Sensor for β2-Agonists with High Sensitivity and Selectivity Based on Surface Molecularly Imprinted Sol-gel Doped with Antimony-Doped Tin Oxide. Electroanalysis 2014, 26, 1004–1012. [Google Scholar] [CrossRef]
  157. Gholivand, M.; Torkashvand, M. The fabrication of a new electrochemical sensor based on electropolymerization of nanocomposite gold nanoparticle-molecularly imprinted polymer for determination of valganciclovir. Mater. Sci. Eng. C 2016, 59, 594–603. [Google Scholar] [CrossRef]
  158. Li, Y.; Song, H.; Zhang, L.; Zuo, P.; Ye, B.-C.; Yao, J.; Chen, W. Supportless electrochemical sensor based on molecularly imprinted polymer modified nanoporous microrod for determination of dopamine at trace level. Biosens. Bioelectron. 2016, 78, 308–314. [Google Scholar] [CrossRef] [PubMed]
  159. Yang, G.; Zhao, F. Electrochemical sensor for dimetridazole based on novel gold nanoparticles@molecularly imprinted polymer. Sens. Actuators B Chem. 2015, 220, 1017–1022. [Google Scholar] [CrossRef]
  160. Dadkhah, S.; Ziaei, E.; Mehdinia, A.; Kayyal, T.B.; 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]
  161. Silva, H.; Pacheco, J.; Silva, J.; Viswanathan, S.; Delerue-Matos, C. Molecularly imprinted sensor for voltammetric detection of norfloxacin. Sens. Actuators B Chem. 2015, 219, 301–307. [Google Scholar] [CrossRef]
  162. Zhao, Y.; Yuan, F.; Quan, X.; Yu, H.; Chen, S.; Zhao, H.; Liu, Z.; Hilal, N. An electrochemical sensor for selective determination of sulfamethoxazole in surface water using a molecularly imprinted polymer modified BDD electrode. Anal. Methods 2015, 7, 2693–2698. [Google Scholar] [CrossRef]
  163. Kumar, N.; Rosy; Goyal, R.N. A melamine based molecularly imprinted sensor for the determination of 8-hydroxydeoxyguanosine in human urine. Talanta 2017, 166, 215–222. [Google Scholar] [CrossRef]
  164. Toro, M.J.U.; Marestoni, L.D.; Sotomayor, M.D.P.T. A new biomimetic sensor based on molecularly imprinted polymers for highly sensitive and selective determination of hexazinone herbicide. Sens. Actuators B Chem. 2015, 208, 299–306. [Google Scholar] [CrossRef]
  165. Rao, H.; Chen, M.; Ge, H.; Lu, Z.; Liu, X.; Zou, P.; Wang, X.; He, H.; Zeng, X.; Wang, Y. A novel electrochemical sensor based on Au@PANI composites film modified glassy carbon electrode binding molecular imprinting technique for the determination of melamine. Biosens. Bioelectron. 2016, 87, 1029–1035. [Google Scholar] [CrossRef]
  166. Cai, R.; Rao, W.; Zhang, Z.; Long, F.; Yin, Y. An imprinted electrochemical sensor for bisphenol A determination based on electrodeposition of a graphene and Ag nanoparticle modified carbon electrode. Anal. Methods 2014, 6, 1590–1597. [Google Scholar] [CrossRef]
  167. Akhoundian, M.; Rüter, A.; Shinde, S. Ultratrace Detection of Histamine Using a Molecularly-Imprinted Polymer-Based Voltammetric Sensor. Sensors 2017, 17, 645. [Google Scholar] [CrossRef] [Green Version]
  168. Prasad, B.B.; Singh, K. An electroconducting copper (II) imprinted sensor using algae as cheap substitute of multiwalled carbon nanotubes. Electrochim. Acta 2016, 187, 193–203. [Google Scholar] [CrossRef]
  169. Tan, Y.; Jin, J.; Zhang, S.; Shi, Z.; Wang, J.; Zhang, J.; Pu, W.; Yang, C. Electrochemical Determination of Bisphenol A Using a Molecularly Imprinted Chitosan-acetylene Black Composite Film Modified Glassy Carbon Electrode. Electroanalysis 2015, 28, 189–196. [Google Scholar] [CrossRef]
  170. Deng, P.; Xu, Z.; Kuang, Y. Electrochemical determination of bisphenol A in plastic bottled drinking water and canned beverages using a molecularly imprinted chitosan–graphene composite film modified electrode. Food Chem. 2014, 157, 490–497. [Google Scholar] [CrossRef] [PubMed]
  171. Motia, S.; Tudor, I.A.; Popescu, L.M.C.; Piticescu, R.M.; Bouchikhi, B.; El Bari, N. Development of a novel electrochemical sensor based on electropolymerized molecularly imprinted polymer for selective detection of sodium lauryl sulfate in environmental waters and cosmetic products. J. Electroanal. Chem. 2018, 823, 553–562. [Google Scholar] [CrossRef]
  172. Li, Y.; Liu, J.; Zhang, Y.; Gu, M.; Wang, D.; Dang, Y.-Y.; Ye, B.-C.; Li, Y. A robust electrochemical sensing platform using carbon paste electrode modified with molecularly imprinted microsphere and its application on methyl parathion detection. Biosens. Bioelectron. 2018, 106, 71–77. [Google Scholar] [CrossRef] [PubMed]
  173. Gu, Y.; Yan, X.; Li, C.; Zheng, B.; Li, Y.; Liu, W.; Zhang, Z.; Yang, M. Biomimetic sensor based on molecularly imprinted polymer with nitroreductase-like activity for metronidazole detection. Biosens. Bioelectron. 2015, 77, 393–399. [Google Scholar] [CrossRef] [PubMed]
  174. Chen, H.-J.; Zhang, Z.-H.; Cai, R.; Kong, X.-Q.; Chen, X.; Liu, Y.-N.; Yao, S.-Z. Molecularly imprinted electrochemical sensor based on a reduced graphene modified carbon electrode for tetrabromobisphenol A detection. Analyst 2013, 138, 2769–2776. [Google Scholar] [CrossRef]
  175. Chen, H.; Zhang, Z.; Cai, R.; Rao, W.; Long, F. Molecularly imprinted electrochemical sensor based on nickel nanoparticles-graphene nanocomposites modified electrode for determination of tetrabromobisphenol A. Electrochim. Acta 2014, 117, 385–392. [Google Scholar] [CrossRef]
  176. Zhang, Z.; Cai, R.; Long, F.; Wang, J. Development and application of tetrabromobisphenol A imprinted electrochemical sensor based on graphene/carbon nanotubes three-dimensional nanocomposites modified carbon electrode. Talanta 2015, 134, 435–442. [Google Scholar] [CrossRef]
  177. Li, J.; Xu, Z.; Liu, M.; Deng, P.; Tang, S.; Jiang, J.; Feng, H.; Qian, D.; He, L. Ag/N-doped reduced graphene oxide incorporated with molecularly imprinted polymer: An advanced electrochemical sensing platform for salbutamol determination. Biosens. Bioelectron. 2017, 90, 210–216. [Google Scholar] [CrossRef]
  178. Hassan, A.H.; Moura, S.L.; Ali, F.; Moselhy, W.A.; Sotomayor, M.; Pividori, M.I. Electrochemical sensing of methyl parathion on magnetic molecularly imprinted polymer. Biosens. Bioelectron. 2018, 118, 181–187. [Google Scholar] [CrossRef] [Green Version]
  179. Anirudhan, T.; Athira, V.; Sekhar, V.C. Electrochemical sensing and nano molar level detection of Bisphenol-A with molecularly imprinted polymer tailored on multiwalled carbon nanotubes. Polymer 2018, 146, 312–320. [Google Scholar] [CrossRef]
  180. Nezhadali, A.; Bonakdar, G.A. Multivariate optimization of mebeverine analysis using molecularly imprinted polymer electrochemical sensor based on silver nanoparticles. J. Food Drug Anal. 2018, 27, 305–314. [Google Scholar] [CrossRef] [Green Version]
  181. Tan, F.; Cong, L.; Li, X.; Zhao, Q.; Zhao, H.; Quan, X.; Chen, J. An electrochemical sensor based on molecularly imprinted polypyrrole/graphene quantum dots composite for detection of bisphenol A in water samples. Sens. Actuators B Chem. 2016, 233, 599–606. [Google Scholar] [CrossRef]
  182. Karimian, N.; Zavar, M.H.A.; Chamsaz, M.; Turner, A.; Tiwari, A. On/off-switchable electrochemical folic acid sensor based on molecularly imprinted polymer electrode. Electrochem. Commun. 2013, 36, 92–95. [Google Scholar] [CrossRef]
  183. Chen, Z.; Tang, C.; Zeng, Y.; Liu, H.; Yin, Z.; Li, L. Determination of Bisphenol a Using an Electrochemical Sensor Based on a Molecularly Imprinted Polymer-Modified Multiwalled Carbon Nanotube Paste Electrode. Anal. Lett. 2014, 47, 996–1014. [Google Scholar] [CrossRef]
  184. Saksena, K.; Shrivastava, A.; Kant, R. Chiral analysis of ascorbic acid in bovine serum using ultrathin molecular imprinted polyaniline/graphite electrode. J. Electroanal. Chem. 2017, 795, 103–109. [Google Scholar] [CrossRef]
  185. Deng, P.; Xu, Z.; Li, J.; Kuang, Y. Acetylene black paste electrode modified with a molecularly imprinted chitosan film for the detection of bisphenol A. Microchim. Acta 2013, 180, 861–869. [Google Scholar] [CrossRef]
  186. Nie, D.; Han, Z.; Yu, Y.; Shi, G. Composites of multiwalled carbon nanotubes/polyethyleneimine (MWCNTs/PEI) and molecularly imprinted polymers for dinitrotoluene recognition. Sens. Actuators B Chem. 2016, 224, 584–591. [Google Scholar] [CrossRef]
  187. Huang, X.; Wei, S.; Yao, S.; Zhang, H.; He, C.; Cao, J. Development of molecularly imprinted electrochemical sensor with reduced graphene oxide and titanium dioxide enhanced performance for the detection of toltrazuril in chicken muscle and egg. J. Pharm. Biomed. Anal. 2018, 164, 607–614. [Google Scholar] [CrossRef] [PubMed]
  188. Jaiswal, S.; Singh, R.; Singh, K.; Fatma, S.; Prasad, B.B. Enantioselective analysis of D- and l- Serine on a layer-by-layer imprinted electrochemical sensor. Biosens. Bioelectron. 2018, 124–125, 176–183. [Google Scholar] [CrossRef]
  189. Yola, M.L.; Atar, N. Development of cardiac troponin-I biosensor based on boron nitride quantum dots including molecularly imprinted polymer. Biosens. Bioelectron. 2018, 126, 418–424. [Google Scholar] [CrossRef]
  190. Sun, Y.; Xu, L.; Waterhouse, G.I.; Wang, M.; Qiao, X.; Xu, Z. Novel three-dimensional electrochemical sensor with dual signal amplification based on MoS2 nanosheets and high-conductive NH2-MWCNT@COF for sulfamerazine determination. Sens. Actuators B Chem. 2018, 281, 107–114. [Google Scholar] [CrossRef]
  191. 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 2018, 194, 415–421. [Google Scholar] [CrossRef]
  192. Yu, R.; Zhou, H.; Li, M.; Song, Q. Rational selection of the monomer for molecularly imprinted polymer preparation for selective and sensitive detection of 3-methylindole in water. J. Electroanal. Chem. 2018, 832, 129–136. [Google Scholar] [CrossRef]
  193. Liu, Y.; Liang, Y.; Yang, R.; Li, J.; Qu, L. A highly sensitive and selective electrochemical sensor based on polydopamine functionalized graphene and molecularly imprinted polymer for the 2,4-dichlorophenol recognition and detection. Talanta 2018, 195, 691–698. [Google Scholar] [CrossRef]
  194. Li, H.-H.; Wang, H.-H.; Li, W.-T.; Fang, X.-X.; Guo, X.-C.; Zhou, W.-H.; Cao, X.; Kou, D.-X.; Zhou, Z.-J.; Wu, S.-X. A novel electrochemical sensor for epinephrine based on three dimensional molecularly imprinted polymer arrays. Sens. Actuators B Chem. 2016, 222, 1127–1133. [Google Scholar] [CrossRef]
  195. Singh, A.K.; Singh, M. QCM sensing of melphalan via electropolymerized molecularly imprinted polythiophene films. Biosens. Bioelectron. 2015, 74, 711–717. [Google Scholar] [CrossRef] [PubMed]
  196. Moreira, F.; Sharma, S.; Dutra, R.A.; Noronha, J.P.; Cass, A.E.; Sales, M.G.F. Protein-responsive polymers for point-of-care detection of cardiac biomarker. Sens. Actuators B Chem. 2014, 196, 123–132. [Google Scholar] [CrossRef] [Green Version]
  197. Cardoso, A.R.; Marques, A.C.; Santos, L.; Carvalho, A.; Costa, F.M.; Martins, R.; Sales, M.G.F.; Fortunato, E. Molecularly-imprinted chloramphenicol sensor with laser-induced graphene electrodes. Biosens. Bioelectron. 2018, 124–125, 167–175. [Google Scholar] [CrossRef] [PubMed]
  198. Xue, C.; Han, Q.; Wang, Y.; Wu, J.; Wen, T.; Wang, R.; Hong, J.; Zhou, X.; Jiang, H. Amperometric detection of dopamine in human serumby electrochemical sensor based on gold nanoparticles doped molecularly imprinted polymers. Biosens. Bioelectron. 2013, 49, 199–203. [Google Scholar] [CrossRef]
  199. Yarman, A.; Scheller, F.W. The First Electrochemical MIP Sensor for Tamoxifen. Sensors 2014, 14, 7647–7654. [Google Scholar] [CrossRef] [Green Version]
  200. Wang, Z.; Li, J.; Liu, X.; Yang, J.; Lu, X. Preparation of an amperometric sensor for norfloxacin based on molecularly imprinted grafting photopolymerization. Anal. Bioanal. Chem. 2013, 405, 2525–2533. [Google Scholar] [CrossRef]
  201. Azevedo, S.D.; Lakshmi, D.; Chianella, I.; Whitcombe, M.; Karim, K.; Ivanova-Mitseva, P.K.; Subrahmanyam, S.; Piletsky, S. Molecularly Imprinted Polymer-Hybrid Electrochemical Sensor for the Detection of β-Estradiol. Indian Eng. Chem. Res. 2013, 52, 13917–13923. [Google Scholar] [CrossRef]
  202. Zhao, H.; Wang, H.; Quan, X.; Tan, F. Amperometric Sensor for Tetracycline Determination Based on Molecularly Imprinted Technique. Proced. Environ. Sci. 2013, 18, 249–257. [Google Scholar] [CrossRef] [Green Version]
  203. Ramanaviciene, A.; Ramanavicius, A. Molecularly imprinted polypyrrole-based synthetic receptor for direct detection of bovine leukemia virus glycoproteins. Biosens. Bioelectron. 2004, 20, 1076–1082. [Google Scholar] [CrossRef]
  204. Ratautaite, V.; Topkaya, S.N.; Mikoliunaite, L.; Ozsoz, M.; Oztekin, Y.; Ramanaviciene, A.; Ramanavicius, A. Molecularly Imprinted Polypyrrole for DNA Determination. Electroanalysis 2013, 25, 1169–1177. [Google Scholar] [CrossRef]
  205. Hirsch, R.; Ternes, T.; Haberer, K.; Kratz, K.-L. Occurrence of antibiotics in the aquatic environment. Sci. Total. Environ. 1999, 225, 109–118. [Google Scholar] [CrossRef]
  206. Al Qarni, H.; Collier, P.; O’Keeffe, J.; Akunna, J. Investigating the removal of some pharmaceutical compounds in hospital wastewater treatment plants operating in Saudi Arabia. Environ. Sci. Pollut. Res. 2016, 23, 13003–13014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. A Ternes, T.; Stüber, J.; Herrmann, N.; McDowell, D.; Ried, A.; Kampmann, M.; Teiser, B. Ozonation: A tool for removal of pharmaceuticals, contrast media and musk fragrances from wastewater? Water Res. 2003, 37, 1976–1982. [Google Scholar] [CrossRef]
  208. 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]
  209. Dietl, S.; Sobek, H.; Mizaikoff, B. Epitope-imprinted polymers for biomacromolecules: Recent strategies, future challenges and selected applications. TrAC Trends Anal. Chem. 2021, 143, 116414. [Google Scholar] [CrossRef]
  210. Al-Kindy, S.; Badía, R.; Suárez-Rodríguez, J.L.; Díaz-García, M.E. Molecularly Imprinted Polymers and Optical Sensing Applications. Crit. Rev. Anal. Chem. 2000, 30, 291–309. [Google Scholar] [CrossRef]
  211. Henry, O.Y.F.; Cullen, D.; Piletsky, S.A. Optical interrogation of molecularly imprinted polymers and development of MIP sensors: A review. Anal. Bioanal. Chem. 2005, 382, 947–956. [Google Scholar] [CrossRef] [PubMed]
  212. Zhao, B.; Feng, S.; Hu, Y.; Wang, S.; Lu, X. Rapid determination of atrazine in apple juice using molecularly imprinted polymers coupled with gold nanoparticles-colorimetric/SERS dual chemosensor. Food Chem. 2018, 276, 366–375. [Google Scholar] [CrossRef] [PubMed]
  213. Wang, P.; Ge, L.; Li, M.; Li, W.; Li, L.; Wang, Y.; Yu, J. Photoelectrochemical Sensor Based on Molecularly Imprinted Polymer-Coated TiO2 Nanotubes for Lindane Specific Recognition and Detection. J. Inorg. Organomet. Polym. Mater. 2013, 23, 703–711. [Google Scholar] [CrossRef]
  214. Kong, Q.; Wang, Y.; Zhang, L.; Ge, S.; Yu, J. A novel microfluidic paper-based colorimetric sensor based on molecularly imprinted polymer membranes for highly selective and sensitive detection of bisphenol A. Sens. Actuators B Chem. 2017, 243, 130–136. [Google Scholar] [CrossRef]
  215. Sergeyeva, T.; Chelyadina, D.S.; Gorbach, L.A.; Brovko, O.O.; Piletska, E.V.; Piletsky, S.; Sergeeva, L.M.; El’Skaya, A.V. Colorimetric biomimetic sensor systems based on molecularly imprinted polymer membranes for highly-selective detection of phenol in environmental samples. Biopolym. Cell 2014, 30, 209–215. [Google Scholar] [CrossRef] [Green Version]
  216. Huang, K.; Chen, Y.; Zhou, F.; Zhao, X.; Liu, J.; Mei, S.; Zhou, Y.; Jing, T. Integrated ion imprinted polymers-paper composites for selective and sensitive detection of Cd(II) ions. J. Hazard. Mater. 2017, 333, 137–143. [Google Scholar] [CrossRef]
  217. Foguel, M.V.; Ton, X.-A.; Zanoni, M.V.; Sotomayor, M.; Haupt, K.; Bui, B.T.S. A molecularly imprinted polymer-based evanescent wave fiber optic sensor for the detection of basic red 9 dye. Sens. Actuators B Chem. 2015, 218, 222–228. [Google Scholar] [CrossRef] [Green Version]
  218. Liu, X.; Yu, D.; Yu, Y.; Ji, S. Preparation of a magnetic molecularly imprinted polymer for selective recognition of rhodamine B. Appl. Surf. Sci. 2014, 320, 138–145. [Google Scholar] [CrossRef]
  219. Yan, K.; Yang, Y.; Zhang, J. A self-powered sensor based on molecularly imprinted polymer-coupled graphitic carbon nitride photoanode for selective detection of bisphenol A. Sens. Actuators B Chem. 2018, 259, 394–401. [Google Scholar] [CrossRef]
  220. Dai, J.; Vu, D.; Nagel, S.; Lin, C.-H.; De Cortalezzi, M.F. Colloidal crystal templated molecular imprinted polymer for the detection of 2-butoxyethanol in water contaminated by hydraulic fracturing. Microchim. Acta 2017, 185, 32. [Google Scholar] [CrossRef] [PubMed]
  221. Kadhem, A.J.; Xiang, S.; Nagel, S.; Lin, C.-H.; de Cortalezzi, M.F. Photonic Molecularly Imprinted Polymer Film for the Detection of Testosterone in Aqueous Samples. Polymers 2018, 10, 349. [Google Scholar] [CrossRef] [Green Version]
  222. Du, Q.; Zhang, Y.; Yu, L.; He, H. Surface molecularly imprinted polymers fabricated by differential UV–vis spectra and reverse prediction method for the enrichment and determination of sterigmatocystin. Food Chem. 2021, 367, 130715. [Google Scholar] [CrossRef]
  223. Tan, L.; Chen, K.; Huang, C.; Peng, R.; Luo, X.; Yang, R.; Cheng, Y.; Tang, Y. A fluorescent turn-on detection scheme for α-fetoprotein using quantum dots placed in a boronate-modified molecularly imprinted polymer with high affinity for glycoproteins. Microchim. Acta 2015, 182, 2615–2622. [Google Scholar] [CrossRef]
  224. Xu, L.; Fang, G.; Pan, M.; Wang, X.; Wang, S. One-pot synthesis of carbon dots-embedded molecularly imprinted polymer for specific recognition of sterigmatocystin in grains. Biosens. Bioelectron. 2016, 77, 950–956. [Google Scholar] [CrossRef]
  225. Ren, X.; Chen, L. Preparation of molecularly imprinted polymer coated quantum dots to detect nicosulfuron in water samples. Anal. Bioanal. Chem. 2015, 407, 8087–8095. [Google Scholar] [CrossRef]
  226. Xiong, Y.; Ye, Z.; Xu, J.; Liu, Y.; Zhang, H. A microvolume molecularly imprinted polymer modified fiber-optic evanescent wave sensor for bisphenol A determination. Anal. Bioanal. Chem. 2014, 406, 2411–2420. [Google Scholar] [CrossRef]
  227. Wang, J.; Cheng, R.; Wang, Y.; Sun, L.; Chen, L.; Dai, X.; Pan, J.; Pan, G.; Yan, Y. Surface-imprinted fluorescence microspheres as ultrasensitive sensor for rapid and effective detection of tetracycline in real biological samples. Sens. Actuators B Chem. 2018, 263, 533–542. [Google Scholar] [CrossRef]
  228. Lu, X.; Yang, Y.; Zeng, Y.; Li, L.; Wu, X. Rapid and reliable determination of p-nitroaniline in wastewater by molecularly imprinted fluorescent polymeric ionic liquid microspheres. Biosens. Bioelectron. 2018, 99, 47–55. [Google Scholar] [CrossRef]
  229. Wei, X.; Xu, G.; Gong, C.; Qin, F.; Gong, X.; Li, C. Fabrication and evaluation of sulfanilamide-imprinted composite sensors by developing a custom-tailored strategy. Sens. Actuators B Chem. 2018, 255, 2697–2703. [Google Scholar] [CrossRef]
  230. Wei, X.; Hao, T.; Xu, Y.; Lu, K.; Li, H.; Yan, Y.; Zhou, Z. Facile polymerizable surfactant inspired synthesis of fluorescent molecularly imprinted composite sensor via aqueous CdTe quantum dots for highly selective detection of λ-cyhalothrin. Sens. Actuators B Chem. 2016, 224, 315–324. [Google Scholar] [CrossRef]
  231. Wagner, S.; Bell, J.; Biyikal, M.; Gawlitza, K.; Rurack, K. Integrating fluorescent molecularly imprinted polymer (MIP) sensor particles with a modular microfluidic platform for nanomolar small-molecule detection directly in aqueous samples. Biosens. Bioelectron. 2018, 99, 244–250. [Google Scholar] [CrossRef]
  232. Xu, S.; Lu, H. Mesoporous structured MIPs@CDs fluorescence sensor for highly sensitive detection of TNT. Biosens. Bioelectron. 2016, 85, 950–956. [Google Scholar] [CrossRef] [PubMed]
  233. Wang, J.; Qiu, H.; Shen, H.; Pan, J.; Dai, X.; Yan, Y.; Pan, G.; Sellergren, B. Molecularly imprinted fluorescent hollow nanoparticles as sensors for rapid and efficient detection λ-cyhalothrin in environmental water. Biosens. Bioelectron. 2016, 85, 387–394. [Google Scholar] [CrossRef] [PubMed]
  234. Li, H.; Li, N.; Jiang, J.; Chen, D.; Xu, Q.; Li, H.; He, J.; Lu, J. Molecularly imprinted magnetic microparticles for the simultaneous detection and extraction of Rhodamine B. Sens. Actuators B Chem. 2017, 246, 286–292. [Google Scholar] [CrossRef]
  235. Amjadi, M.; Jalili, R. Molecularly imprinted mesoporous silica embedded with carbon dots and semiconductor quantum dots as a ratiometric fluorescent sensor for diniconazole. Biosens. Bioelectron. 2017, 96, 121–126. [Google Scholar] [CrossRef] [PubMed]
  236. Wu, X.; Zhang, Z.; Li, J.; You, H.; Li, Y.; Chen, L. Molecularly imprinted polymers-coated gold nanoclusters for fluorescent detection of bisphenol A. Sens. Actuators B Chem. 2015, 211, 507–514. [Google Scholar] [CrossRef]
  237. Wu, Y.-T.; Liu, Y.-J.; Gao, X.; Gao, K.-C.; Xia, H.; Luo, M.-F.; Wang, X.-J.; Ye, L.; Shi, Y.; Lu, B. Monitoring bisphenol A and its biodegradation in water using a fluorescent molecularly imprinted chemosensor. Chemosphere 2015, 119, 515–523. [Google Scholar] [CrossRef]
  238. Ren, X.; Chen, L. Quantum dots coated with molecularly imprinted polymer as fluorescence probe for detection of cyphenothrin. Biosens. Bioelectron. 2015, 64, 182–188. [Google Scholar] [CrossRef]
  239. Liu, G.; Chen, Z.; Jiang, X.; Feng, D.-Q.; Zhao, J.; Fan, D.; Wang, W. In-situ hydrothermal synthesis of molecularly imprinted polymers coated carbon dots for fluorescent detection of bisphenol A. Sens. Actuators B Chem. 2016, 228, 302–307. [Google Scholar] [CrossRef]
  240. Ensafi, A.A.; Zakery, M.; Rezaei, B. An optical sensor with specific binding sites for the detection of thioridazine hydrochloride based on ZnO-QDs coated with molecularly imprinted polymer. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 206, 460–465. [Google Scholar] [CrossRef]
  241. Fang, M.; Zhou, L.; Zhang, H.; Liu, L.; Gong, Z.-Y. A molecularly imprinted polymers/carbon dots-grafted paper sensor for 3-monochloropropane-1,2-diol determination. Food Chem. 2018, 274, 156–161. [Google Scholar] [CrossRef]
  242. Feng, J.; Tao, Y.; Shen, X.; Jin, H.; Zhou, T.; Zhou, Y.; Hu, L.; Luo, D.; Mei, S.; Lee, Y.-I. Highly sensitive and selective fluorescent sensor for tetrabromobisphenol-A in electronic waste samples using molecularly imprinted polymer coated quantum dots. Microchem. J. 2018, 144, 93–101. [Google Scholar] [CrossRef]
  243. Mehrzad-Samarin, M.; Faridbod, F.; Ganjali, M.R. A luminescence nanosensor for Ornidazole detection using graphene quantum dots entrapped in silica molecular imprinted polymer. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 206, 430–436. [Google Scholar] [CrossRef] [PubMed]
  244. Ahmadpour, H.; Hosseini, S.M.M. A solid-phase luminescence sensor based on molecularly imprinted polymer-CdSeS/ZnS quantum dots for selective extraction and detection of sulfasalazine in biological samples. Talanta 2018, 194, 534–541. [Google Scholar] [CrossRef] [PubMed]
  245. Shirani, M.P.; Rezaei, B.; Ensafi, A.A. A novel optical sensor based on carbon dots embedded molecularly imprinted silica for selective acetamiprid detection. Spectrochim. Acta Part. A Mol. Biomol. Spectrosc. 2018, 210, 36–43. [Google Scholar] [CrossRef]
  246. Zheng, L.; Zheng, Y.; Liu, Y.; Long, S.; Du, L.; Liang, J.; Huang, C.; Swihart, M.; Tan, K. Core-shell quantum dots coated with molecularly imprinted polymer for selective photoluminescence sensing of perfluorooctanoic acid. Talanta 2018, 194, 1–6. [Google Scholar] [CrossRef]
  247. Dai, J.; Dong, X.; de Cortalezzi, M.F. Molecularly imprinted polymers labeled with amino-functionalized carbon dots for fluorescent determination of 2,4-dinitrotoluene. Microchim. Acta 2017, 184, 1369–1377. [Google Scholar] [CrossRef]
  248. Dai, J.; de Cortalezzi, M.F. Influence of pH, ionic strength and natural organic matter concentration on a MIP-Fluorescent sensor for the quantification of DNT in water. Heliyon 2019, 5, e01922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Sa-Nguanprang, S.; Phuruangrat, A.; Bunkoed, O. An optosensor based on a hybrid sensing probe of mesoporous carbon and quantum dots embedded in imprinted polymer for ultrasensitive detection of thiamphenicol in milk. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2021, 264, 120324. [Google Scholar] [CrossRef]
  250. Cui, Y.; Su, A.; Feng, J.; Dong, W.; Li, J.; Wang, H.; Ni, X.; Jiang, Y. Development of silica molecularly imprinted polymer on carbon dots as a fluorescence probe for selective and sensitive determination of cetirizine in saliva and urine. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2021, 264, 120293. [Google Scholar] [CrossRef] [PubMed]
  251. Liu, W.; Guo, Y.; Luo, J.; Kou, J.; Zheng, H.; Li, B.; Zhang, Z. A molecularly imprinted polymer based a lab-on-paper chemiluminescence device for the detection of dichlorvos. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 141, 51–57. [Google Scholar] [CrossRef] [PubMed]
  252. Duan, H.; Li, L.; Wang, X.; Wang, Y.; Li, J.; Luo, C. A sensitive and selective chemiluminescence sensor for the determination of dopamine based on silanized magnetic graphene oxide-molecularly imprinted polymer. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 139, 374–379. [Google Scholar] [CrossRef]
  253. Qiu, H.; Fan, L.; Li, X.; Li, L.; Sun, M.; Luo, C. A microflow chemiluminescence sensor for indirect determination of dibutyl phthalate by hydrolyzing based on biological recognition materials. J. Pharm. Biomed. Anal. 2013, 75, 123–129. [Google Scholar] [CrossRef]
  254. Yang, Y.; Fang, G.; Wang, X.; Zhang, F.; Liu, J.; Zheng, W.; Wang, S. Electrochemiluminescent graphene quantum dots enhanced by MoS2 as sensing platform: A novel molecularly imprinted electrochemiluminescence sensor for 2-methyl-4-chlorophenoxyacetic acid assay. Electrochim. Acta 2017, 228, 107–113. [Google Scholar] [CrossRef]
  255. Yang, Y.; Fang, G.; Wang, X.; Liu, G.; Wang, S. Imprinting of molecular recognition sites combined with π-donor–acceptor interactions using bis-aniline-crosslinked Au–CdSe/ZnS nanoparticles array on electrodes: Development of electrochemiluminescence sensor for the ultrasensitive and selective detection of 2-methyl-4-chlorophenoxyacetic acid. Biosens. Bioelectron. 2015, 77, 1134–1143. [Google Scholar] [CrossRef]
  256. Wang, S.; Ge, L.; Li, L.; Yan, M.; Ge, S.; Yu, J. Molecularly imprinted polymer grafted paper-based multi-disk micro-disk plate for chemiluminescence detection of pesticide. Biosens. Bioelectron. 2013, 50, 262–268. [Google Scholar] [CrossRef]
  257. Duan, H.; Li, L.; Wang, X.; Wang, Y.; Li, J.; Luo, C. CdTe quantum dots@luminol as signal amplification system for chrysoidine with chemiluminescence-chitosan/graphene oxide-magnetite-molecularly imprinting sensor. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2016, 153, 535–541. [Google Scholar] [CrossRef]
  258. Jiang, Q.; Zhang, D.; Cao, Y.; Gan, N. An antibody-free and signal-on type electrochemiluminescence sensor for diethylstilbestrol detection based on magnetic molecularly imprinted polymers-quantum dots labeled aptamer conjugated probes. J. Electroanal. Chem. 2017, 789, 1–8. [Google Scholar] [CrossRef]
  259. Hu, L.; Zhou, T.; Feng, J.; Jin, H.; Tao, Y.; Luo, D.; Mei, S.; Lee, Y.-I. A rapid and sensitive molecularly imprinted electrochemiluminescence sensor for Azithromycin determination in biological samples. J. Electroanal. Chem. 2018, 813, 1–8. [Google Scholar] [CrossRef]
  260. Wang, D.; Jiang, S.; Liang, Y.; Wang, X.; Zhuang, X.; Tian, C.; Luan, F.; Chen, L. Selective detection of enrofloxacin in biological and environmental samples using a molecularly imprinted electrochemiluminescence sensor based on functionalized copper nanoclusters. Talanta 2021, 236, 122835. [Google Scholar] [CrossRef]
  261. Cennamo, N.; D’Agostino, G.; Pesavento, M.; Zeni, L. High selectivity and sensitivity sensor based on MIP and SPR in tapered plastic optical fibers for the detection of l-nicotine. Sens. Actuators B Chem. 2014, 191, 529–536. [Google Scholar] [CrossRef]
  262. Cennamo, N.; De Maria, L.; D’Agostino, G.; Zeni, L.; Pesavento, M. Monitoring of Low Levels of Furfural in Power Transformer Oil with a Sensor System Based on a POF-MIP Platform. Sensors 2015, 15, 8499–8511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Shrivastav, A.M.; Mishra, S.K.; Gupta, B.D. Surface Plasmon Resonance-Based Fiber Optic Sensor for the Detection of Ascorbic Acid Utilizing Molecularly Imprinted Polyaniline Film. Plasmonics 2015, 10, 1853–1861. [Google Scholar] [CrossRef]
  264. Shrivastav, A.; Usha, S.P.; Gupta, B.D. Fiber optic profenofos sensor based on surface plasmon resonance technique and molecular imprinting. Biosens. Bioelectron. 2015, 79, 150–157. [Google Scholar] [CrossRef]
  265. Ayankojo, A.G.; Reut, J.; Öpik, A.; Furchner, A.; Syritski, V. Hybrid molecularly imprinted polymer for amoxicillin detection. Biosens. Bioelectron. 2018, 118, 102–107. [Google Scholar] [CrossRef]
  266. Jiang, S.; Peng, Y.; Ning, B.; Bai, J.; Liu, Y.; Zhang, N.; Gao, Z. Surface plasmon resonance sensor based on molecularly imprinted polymer film for detection of histamine. Sens. Actuators B Chem. 2015, 221, 15–21. [Google Scholar] [CrossRef]
  267. Rahtuvanoğlu, A.; Akgönüllü, S.; Karacan, S.; Denizli, A. Biomimetic Nanoparticles Based Surface Plasmon Resonance Biosensors for Histamine Detection in Foods. Chem. Select 2020, 5, 5683–5692. [Google Scholar] [CrossRef]
  268. Sun, T.; Zhang, Y.; Zhao, F.; Xia, N.; Liu, L. Self-assembled biotin-phenylalanine nanoparticles for the signal amplification of surface plasmon resonance biosensors. Microchim. Acta 2020, 187, 1–7. [Google Scholar] [CrossRef]
  269. Çakır, O.; Baysal, Z. Pesticide analysis with molecularly imprinted nanofilms using surface plasmon resonance sensor and LC-MS/MS: Comparative study for environmental water samples. Sens. Actuators B Chem. 2019, 297, 126764. [Google Scholar] [CrossRef]
  270. Akgönüllü, S.; Yavuz, H.; Denizli, A. SPR nanosensor based on molecularly imprinted polymer film with gold nanoparticles for sensitive detection of aflatoxin B1. Talanta 2020, 219, 121219. [Google Scholar] [CrossRef] [PubMed]
  271. Özgür, E.; Topçu, A.A.; Yılmaz, E.; Denizli, A. Surface plasmon resonance based biomimetic sensor for urinary tract infections. Talanta 2020, 212, 120778. [Google Scholar] [CrossRef] [PubMed]
  272. Kamra, T.; Zhou, T.; Montelius, L.; Schnadt, J.; Ye, L. Implementation of Molecularly Imprinted Polymer Beads for Surface Enhanced Raman Detection. Anal. Chem. 2015, 87, 5056–5061. [Google Scholar] [CrossRef] [PubMed]
  273. Kamra, T.; Xu, C.; Montelius, L.; Schnadt, J.; Wijesundera, S.A.; Yan, M.; Ye, L. Photoconjugation of Molecularly Imprinted Polymer Nanoparticles for Surface-Enhanced Raman Detection of Propranolol. ACS Appl. Mater. Interfaces 2015, 7, 27479–27485. [Google Scholar] [CrossRef] [PubMed]
  274. Hu, Y.; Feng, S.; Gao, F.; Li-Chan, E.C.; Grant, E.; Lu, X. Detection of melamine in milk using molecularly imprinted polymers–surface enhanced Raman spectroscopy. Food Chem. 2015, 176, 123–129. [Google Scholar] [CrossRef] [PubMed]
  275. Chen, S.N.; Li, X.; Han, S.; Liu, J.H.; Zhao, Y.Y. Synthesis of surface-imprinted Ag nanoplates for detecting organic pollutants in water environments based on surface enhanced Raman scattering. RSC Adv. 2015, 5, 99914–99919. [Google Scholar] [CrossRef]
  276. Guo, Y.; Kang, L.; Chen, S.; Li, X. High performance surface-enhanced Raman scattering from molecular imprinting polymer capsulated silver spheres. Phys. Chem. Chem. Phys. 2015, 17, 21343–21347. [Google Scholar] [CrossRef]
  277. Chang, L.; Ding, Y.; Li, X. Surface molecular imprinting onto silver microspheres for surface enhanc24 June 2013ed Raman scattering applications. Biosens. Bioelectron. 2013, 50, 106–110. [Google Scholar] [CrossRef]
  278. Ye, J.; Chen, Y.; Liu, Z. A Boronate Affinity Sandwich Assay: An Appealing Alternative to Immunoassays for the Determination of Glycoproteins. Angew. Chem. Int. Ed. 2014, 53, 10386–10389. [Google Scholar] [CrossRef]
  279. Xue, J.-Q.; Li, D.-W.; Qu, L.; Long, Y. Surface-imprinted core–shell Au nanoparticles for selective detection of bisphenol A based on surface-enhanced Raman scattering. Anal. Chim. Acta 2013, 777, 57–62. [Google Scholar] [CrossRef]
  280. Guo, Z.; Chen, L.; Lv, H.; Yu, Z.; Zhao, B. Magnetic imprinted surface enhanced Raman scattering (MI-SERS) based ultrasensitive detection of ciprofloxacin from a mixed sample. Anal. Methods 2013, 6, 1627–1632. [Google Scholar] [CrossRef]
  281. Yin, W.; Wu, L.; Ding, F.; Li, Q.; Wang, P.; Li, J.; Lu, Z.; Han, H. Surface-imprinted SiO2@Ag nanoparticles for the selective detection of BPA using surface enhanced Raman scattering. Sens. Actuators B Chem. 2017, 258, 566–573. [Google Scholar] [CrossRef]
  282. Ren, X.; Li, X. Flower-like Ag coated with molecularly imprinted polymers as a surface-enhanced Raman scattering substrate for the sensitive and selective detection of glibenclamide. Anal. Methods 2020, 12, 2858–2864. [Google Scholar] [CrossRef]
  283. Li, H.; Wang, Y.; Li, Y.; Qiao, Y.; Liu, L.; Wang, Q.; Che, G. High-sensitive molecularly imprinted sensor with multilayer nanocomposite for 2,6-dichlorophenol detection based on surface-enhanced Raman scattering. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 228, 117784. [Google Scholar] [CrossRef]
  284. Zhou, J.; Sheth, S.; Zhou, H.; Song, Q. Highly selective detection of l-Phenylalanine by molecularly imprinted polymers coated Au nanoparticles via surface-enhanced Raman scattering. Talanta 2020, 211, 120745. [Google Scholar] [CrossRef]
  285. Hsu, C.-Y.; Lee, M.-H.; Thomas, J.L.; Shih, C.-P.; Hung, T.-L.; Whang, T.-J.; Lin, H.-Y. Optical sensing of phenylalanine in urine via extraction with magnetic molecularly imprinted poly(ethylene-co-vinyl alcohol) nanoparticles. Nanotechnology 2015, 26, 305502. [Google Scholar] [CrossRef] [PubMed]
  286. Sun, G.; Wang, P.; Ge, S.; Ge, L.; Yu, J.; Yan, M. Photoelectrochemical sensor for pentachlorophenol on microfluidic paper-based analytical device based on the molecular imprinting technique. Biosens. Bioelectron. 2014, 56, 97–103. [Google Scholar] [CrossRef] [PubMed]
  287. Fang, T.; Yang, X.; Zhang, L.; Gong, J. Ultrasensitive photoelectrochemical determination of chromium(VI) in water samples by ion-imprinted/formate anion-incorporated graphitic carbon nitride nanostructured hybrid. J. Hazard. Mater. 2016, 312, 106–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Wang, P.; Dai, W.; Ge, L.; Yan, M.; Ge, S.; Yu, J. Visible light photoelectrochemical sensor based on Au nanoparticles and molecularly imprinted poly(o-phenylenediamine)-modified TiO2nanotubes for specific and sensitive detection chlorpyrifos. Analyst 2012, 138, 939–945. [Google Scholar] [CrossRef]
  289. Lu, B.; Liu, M.; Shi, H.; Huang, X.; Zhao, G. A Novel Photoelectrochemical Sensor for Bisphenol A with High Sensitivity and Selectivity Based on Surface Molecularly Imprinted Polypyrrole Modified TiO2Nanotubes. Electroanalysis 2013, 25, 771–779. [Google Scholar] [CrossRef]
  290. Sun, X.; Gao, C.; Zhang, L.; Yan, M.; Yu, J.; Ge, S. Photoelectrochemical sensor based on molecularly imprinted film modified hierarchical branched titanium dioxide nanorods for chlorpyrifos detection. Sens. Actuators B Chem. 2017, 251, 1–8. [Google Scholar] [CrossRef]
  291. Gao, C.; Wang, Y.; Yuan, S.; Xue, J.; Cao, B.; Yu, J. Engineering anatase hierarchically cactus-like TiO 2 arrays for photoelectrochemical and visualized sensing platform. Biosens. Bioelectron. 2016, 90, 336–342. [Google Scholar] [CrossRef] [PubMed]
  292. Wang, Y.; Zang, D.; Ge, S.; Ge, L.; Yu, J.; Yan, M. A novel microfluidic origami photoelectrochemical sensor based on CdTe quantum dots modified molecularly imprinted polymer and its highly selective detection of S-fenvalerate. Electrochim. Acta 2013, 107, 147–154. [Google Scholar] [CrossRef]
  293. Liu, M.; Ding, X.; Yang, Q.; Wang, Y.; Zhao, G.; Yang, N. A pM leveled photoelectrochemical sensor for microcystin-LR based on surface molecularly imprinted TiO2 @CNTs nanostructure. J. Hazard. Mater. 2017, 331, 309–320. [Google Scholar] [CrossRef] [PubMed]
  294. Dong, R.; Li, J.; Xiong, H.; Lu, W.; Peng, H.; Chen, L. Thermosensitive molecularly imprinted polymers on porous carriers: Preparation, characterization and properties as novel adsorbents for bisphenol A. Talanta 2014, 130, 182–191. [Google Scholar] [CrossRef]
  295. Li, X.; Wang, X.; Fang, T.; Zhang, L.; Gong, J. Disposable photoelectrochemical sensing strip for highly sensitive determination of perfluorooctane sulfonyl fluoride on functionalized screen-printed carbon electrode. Talanta 2018, 181, 147–153. [Google Scholar] [CrossRef]
  296. Li, R.; Liu, Y.; Cheng, L.; Yang, C.; Zhang, J. Photoelectrochemical Aptasensing of Kanamycin Using Visible Light-Activated Carbon Nitride and Graphene Oxide Nanocomposites. Anal. Chem. 2014, 86, 9372–9375. [Google Scholar] [CrossRef]
  297. Yang, X.; Li, X.; Zhang, L.; Gong, J. Electrospun template directed molecularly imprinted nanofibers incorporated with BiOI nanoflake arrays as photoactive electrode for photoelectrochemical detection of triphenyl phosphate. Biosens. Bioelectron. 2017, 92, 61–67. [Google Scholar] [CrossRef] [Green Version]
  298. Tran, T.; Li, J.; Feng, H.; Cai, J.; Yuan, L.; Wang, N.; Cai, Q. Molecularly imprinted polymer modified TiO2 nanotube arrays for photoelectrochemical determination of perfluorooctane sulfonate (PFOS). Sens. Actuators B Chem. 2014, 190, 745–751. [Google Scholar] [CrossRef]
  299. Wang, R.; Yan, K.; Wang, F.; Zhang, J. A highly sensitive photoelectrochemical sensor for 4-aminophenol based on CdS-graphene nanocomposites and molecularly imprinted polypyrrole. Electrochim. Acta 2014, 121, 102–108. [Google Scholar] [CrossRef]
  300. Gong, J.; Fang, T.; Peng, D.; Li, A.; Zhang, L. A highly sensitive photoelectrochemical detection of perfluorooctanic acid with molecularly imprined polymer-functionalized nanoarchitectured hybrid of AgI–BiOI composite. Biosens. Bioelectron. 2015, 73, 256–263. [Google Scholar] [CrossRef] [PubMed]
  301. Viter, R.; Kunene, K.; Genys, P.; Jevdokimovs, D.; Erts, D.; Sutka, A.; Bisetty, K.; Viksna, A.; Ramanaviciene, A.; Ramanavicius, A. Photoelectrochemical Bisphenol S Sensor Based on ZnO-Nanoroads Modified by Molecularly Imprinted Polypyrrole. Macromol. Chem. Phys. 2019, 221, 1900232. [Google Scholar] [CrossRef]
  302. Whitcombe, M.J.; Chianella, I.; Larcombe, L.; Piletsky, S.A.; Noble, J.; Porter, R.; Horgan, A. The rational development of molecularly imprinted polymer-based sensors for protein detection. Chem. Soc. Rev. 2010, 40, 1547–1571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Lowdon, J.W.; Diliën, H.; Singla, P.; Peeters, M.; Cleij, T.J.; van Grinsven, B.; Eersels, K. MIPs for commercial application in low-cost sensors and assays—An overview of the current status quo. Sens. Actuators B Chem. 2020, 325, 128973. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the molecularly imprinting process.
Figure 1. Schematic diagram of the molecularly imprinting process.
Molecules 26 06233 g001
Figure 2. MIP physical forms: (a) blocks or monoliths, (b) microspheres, (c) nanospheres, (d) thin films, (e) nanocomposite membranes, and (f) nanowires.
Figure 2. MIP physical forms: (a) blocks or monoliths, (b) microspheres, (c) nanospheres, (d) thin films, (e) nanocomposite membranes, and (f) nanowires.
Molecules 26 06233 g002
Table 1. MIP-based electrochemical sensors in environmental and biomedical applications.
Table 1. MIP-based electrochemical sensors in environmental and biomedical applications.
SensorFunctional MonomerElectrodeTargetSampleLinear RangeRef.
LOD
Capacitance
Aptamer-MIPDopamineAuProstate specific antigenTBST buffer0.1–100 ng/mL[133]
1 pg/mL
Sol-gel-MIPN,N-methylene bis acrylamide (MBAA)CMethidathionWastewater40–200 μg/L[134]
5.14 μg/L
Silica NP-Au NP-MIP-ChitosanMAAGCEFumosin B1Maize, milk0.001–100 ng/mL[135]
0.35 pg/mL
MIPMAAAuN-formylamphetamineUltrapure waterVariable[136]
10 μM
MIPMAAAuMetergolinePBS1–50 μM[137]
1 μM
MIPN-allylthiourea (thiourea)PtPhosphateWastewater0.66–8 mg P/L[138]
0.16 mg P/L
MIPo-aminophenolGCENorepinephrineHuman plasma and urine, pharmaceuticals5 × 10−8–10−5 M[139]
4.9 × 10−10 M
MIPPyrrolePGEChlorpyrifosTap water, non-agricultural soil, corn leaves20–300 μg/L[140]
4.5 μg/L
MIPPyrroleB-doped nanocrystalline diamondTheophyllinePBSNI[141]
NI
Potentiometry
MIPMAAPVCPropranololWater, pharmaceuticals10−4–10−1/10−5–10−1 M *[142]
10−4/10−5 M *
MIP-Nanographene-ILMAACPELosartanUrine, pharmaceuticals3 × 10−9–10−2 M[143]
1.82 × 10−9 M
MIP NPMAAPVCHistamineWine, fish1.12 × 10−6–10−2 M[144]
1.12 × 10−6 M
MIP-MWCNTMAACuLindaneGround, tap, and sea water, orange, grape, tomato, cabbage10−9–10−5; 10−5–10−3 M[145]
10−10 M
MIPMAAPVCImidocarb dipropionateBovine liver and kidney10−5–10−2 M[146]
2 × 10−6 M
MIPMAAGCEHeparineHeparine sodium injection3 × 10−9–7 × 10−7 M[147]
10−9 M
MIP4-VPPVCBisphenol APolycarbonate drinking water bottle in water solution0.1–1 μM[148]
0.02 μM
Nanostructured MIP particlesMAAGCE
CPE
CaptoprilCaptopril tablet10−6–10−1 M (GCE); 3 × 10−9–10−1 M (CPE)[149]
NI (GCE); 10−9 M (CPE)
Voltammetry
MIPPyrroleGCETriacetone triperoxideAcetonitrile82–44,300 μg/L[150]
26.9 μg/L
MIP-C NT/ILPyrroleGCEMelamineMilk0.4–9.2 μM[151]
0.11 μM
MIP-RGO-Au NPMAAGCECarbofuranVegetables5 × 10−8–2 × 10−5 M[152]
2 × 10−8 M
MIP-MWCNT-IL-Pt4-vinylpyridineGCETartrazineOrange powder and soft drinks0.03–5; 5–20 μM[153]
8 nM
MIP-MWCNTPyrrole GCEOchratoxin ABeer, wine5 × 10−8–10−6 M[154]
4.1 × 10−9 M
MIP-MWCNTPyrrole PGETriamtereneHuman serum, pharmaceuticals0.08–265 μM[155]
3.35 nM
Sol-gel-MIPAntimony-doped tin oxide (ATO) Ptβ2-agonistsHuman serum5.5 nM–6.3 μM[156]
1.7 nM
MIP-MWCNT2,2′-dithiodianiline GCEValganciclovirHuman serum, tablet1–500; 500–2000 nM[157]
0.3 nM
MIPo-aminophenol Au-AgDopamineRabbit serum, rat brain tissue2 × 10−13–2 × 10−8 M[158]
7.63 × 10−14 M
Au NP-MIPAu NPs@IL, ionic liquid (IL, i.e., 3-propyl-1-vinylimidazolium bromide)GCEDimetridazoleMilk, honey2 × 10−9–2.5 × 10−7; 2.5 × 10−7–3 × 10−6 M[159]
5 × 10−10 M
MIP-GOAPTESGCEBisphenol AMilk, mineralized water0.006–0.1; 0.2–20 μM[160]
0.003 μM
MIP-MWCNTPyrroleGCENorfloxacinHuman urine10−7–8 × 10−6 M[161]
4.6 × 10−8 M
MIPPyrroleBDDSulfamethoxazoleSurface water0.1–100 μM[162]
24.1 nM
MIP-MelamineMelamineEPPG8-hydroxydeoxyguanosineHuman urine2 × 10−8–3 × 10−6 M[163]
3 × 10−9 M
MIP-Chitosan-C dots3-aminobenzeneboronic acidGCEGlucoseHuman serum0.5–40; 50–600 μM[43]
0.09 μM
MIP2-vinylpyridine, methacrylic acid, and acrylamide CPEHexazinoneRiver water1.19 × 10−11–1.1 × 10−9 M[164]
2.6 × 10−12 M
MIP-Au-PolyanilineMAAGCEMelamineMilk, food10−5–10 μM[165]
1.39 × 10−6 μM
MIP-GO-Ag NPPyrroleCBisphenol APVC and soil in water solution10−11–10−8 M[166]
3.2 × 10−12 M
MIPMAACPEHistamineHuman serum10−10–7 × 10−9; 7 × 10−9–4 × 10−7 M[167]
7.4 × 10−11 M
MIP-AlgaeN-methacryloyl glutamic acidCCu(II)Human serum, soil, lake water, pharmaceutical tabletDepending on the sample[168]
MIP-MWNT-Au NPP-aminothiophenol (p-ATP)GCECholesterolEthanol10−13–10−9 M[39]
3.3 × 10−14 M
MIP-Chitosan-Acetylene blackChitosanGCEBisphenol ADrinking water0.005–0.2; 0.5–10 mM[169]
2 nM
MIP-Chitosan-GrapheneChitosan (CHI)ABPEBisphenol ABottled water, canned beverages0.008–1; 1–20 μM[170]
0.006 μM
MIP2-aminothiophenol (2-ATP)AuSodium lauryl sulfateRiver and wastewater, personal hygiene products0.1–1000 pg/mL[171]
0.18 pg/mL
MIPMAACPEMethyl parathionSoil, vegetables10−12–8 × 10−9 M[172]
3.4 × 10−13 M
MIPo-phenylenediamine (o-PD) AuAtrazineDeionized water5 × 10−9–1.4 × 10−7 M[76]
10−9 M
MIP-Cu-MelamineMelamineGCEMetronidazoleStandard injection0.5–1000 μM[173]
0.12 μM
MIP-GraphenePyrroleCTetrabromobisphenol ARain and lake water0.5–4.5 nM[174]
0.23 nM
Ni NP-MIP-GraphenePyrroleCTetrabromobisphenol ARain, lake, and tap water0.5–104 nM[175]
0.13 nM
MIP-Graphene-C NTPyrroleCTetrabromobisphenol AFish10−11–10−8 M[176]
3.7 × 10−12 M
Ag-N-RGO-MIPo-phenylenediamine (o-PD)GCESalbutamolHuman urine, pork0.03–20 μM[177]
7 nM
Magnetite NP-MIPMAAmGECMethyl parathionFishNI[178]
1.22 × 10−6 mg/L
MIP-MWNTMAA and vinylpyridineGCEBisphenol ABaby feeding bottle in PBS0.1 nM–400 μM[179]
0.02 nM
MIP-Ag NPPyrrolePGEMebeverineHuman serum, capsule10−8–10−6; 10−5–10−3 M[180]
8.6 × 10−9 M
MIP-Graphene QDPyrroleGCEBisphenol ATap and sea water0.1–50 μM[181]
0.04 μM
MIPAmine terminated poly(N-isopropylacrylamide)
(PNIPAAm) and o-phenylenediamine (o-PD)
AuFolic acidPBS1–200 μM[182]
0.9 μM
Au NP-MIPAcetate buffer (pH 6.0), quercetin, resorcinol, methyl parathion,
KClO4
GCEMethyl parathionWater, tangerine juice, sweet potato leaves0.05–15 μM[40]
0.01 μM
MIP-MWCNT4-vinylpyridineCPEBisphenol ARiver, tap, and pure water0.08–100 μM[183]
0.022 μM
MIPAnilinePGEL-ascorbic acidBovine serum, water1–100 μM[184]
1 μM
MIP-ChitosanChitosanABPEBisphenol APolycarbonate baby feeding and water bottle, PVC bottle, PVC food package in water solution80 nM–10 μM[185]
60 nM
MWCNT-Polyethyleneimine-MIPAcrylamideGCE2,4-dinitrotolueneRiver and wastewater2.2 × 10−9–10−6 M[186]
10−9 M
MIP-RGO-TiO2Carboxymethyl-β-cyclodextrin (CM-β-CD)PtToltrazurilChicken muscles, egg0.43–42.54 μg/L[187]
0.21 μg/L
MIP-RGOAcrylamide PGED-, L-serineBlood serum, cerebrospinal fluid, water, pharmaceuticsDepending on the sample[188]
0.24 ng/mL
MIP-BN QDPyrroleGCECardiac Troponin-IHuman plasma0.01–5 ng/mL[189]
0.0005 ng/mL
Conductive NH2-MWCNT-MoS2-MIPp-aminobenzoic acidGCESulfamerazinePork, chicken3 × 10−7–2 × 10−4 M[190]
1.1 × 10−7 M
Nano MIP-MWCNTMAAGCETrinitroperhydro-1,3,5-triazineTap, river, and sea water0.1–10 nM; 0.01–1 μM[191]
20 pM
MIPo-Phenylenediamine (o-PD)GCE3-methylindoleTap and lake water0.01–1.2 μM[192]
4 nM
MIP-Polydopamine-RGOo-phenylenediamine (o-PD)GCE2,4-dichlorophenolLake water2–10; 10–100 nM[193]
0.8 nM
MIP-ZnOPyrroleITO/PETEpinephrineEpinephrine hydrochloride injections1–10; 10–800 μM[194]
1 μM
MIP3-thiophene acetic acid (3-TAA)AuMelphalanPharmaceutical tablet0.01–0.07 mM[195]
NI
MIPVinylferrocene, 4-vinylpyridineCPEBenzo(a)pyreneUltrapure water0–8; 2–16 μM[78]
0.93 μM
MIPFerrocenylmethyl methacrylateC SPEBisphenol APBS4.7–8 nM[79]
3.2 nM
MIP microbeadsFerrocenylmethyl methacrylateCPEBisphenol AUltrapure waterNI[80]
NI
Capacitance. Voltammetry
MIPAminophenol (AP)Au SPEMyoglobinSynthetic human serum1.5–4 (EIS); 0.8–3.5 (SWV) μg/mL[196]
1.5 (EIS); 0.8 (SWV) μg/mL
MIPEriochrome black T (EBT)Graphene SPE
C SPE
ChloramphenicolAcetonitrile buffer1 nM–10 mM[197]
0.62 nM
Amperometry
MIP-MWCNT-NafionAPTES and PTMSGCE2-nonylphenolTap and river water, soil0.2–360 μM[42]
0.06 μM
Au NP-MIPFunctionalized AuNPs (F-AuNPs) AuDopamineHuman serum0.02–0.54 μM[198]
7.8 nM
MIPO-phenylenediamine‒resorcinolGCETamoxifenAcetate buffer1–100 nM[199]
NI
MIPAcrylamideAuNorfloxacinHuman urine10−9–10−3 M[200]
10−10 M
Conducting MIPN-phenylethylene diamine methacrylamide (NPEDMA)Au17β-estradiolEthanol/PBS solutions10−7–8 × 10−7 M[201]
6.89 × 10−7 M
MIP-PtMAATiTetracyclineUltrapure water0.1–10 mg/L[202]
0.026 mg/L
MIP NPVinylferrocene, ferrocenylmethyl methacrylateGCEVancomycinTRIS buffer83–410 μM[77]
NI
MIPPyrrolePt wire sealed in glassBovine leukemia virus glycoproteinsUltrapure waterNI[203]
NI
MIPPyrrolePGEDNAAcetate bufferNI[204]
NI
*: depending on the MIP content. LOD: limit of detection. NI: not informed. ABPE: acetylene black paste. BDD: boron doped diamond. C: carbon. CPE: carbon paste electrode. EIS: electrochemical impedance spectroscopy. EPPG: edge plane pyrolytic graphite. GCE: glassy carbon electrode. GO: graphene oxide. IL: ionic liquid. ITO: indium tin oxide. ITO/PET: indium tin oxide coated polyethylene terephthalate. mGEC: magneto-electrode based on graphite–epoxy composite. MWCNT: multi-walled carbon nanotube. NP: nanoparticle. NT: nanotube. PBS: phosphate buffer solution. PGE: pencil graphite electrode. QD: quantum dot. RGO: reduced graphene oxide. SPE: Screen-printed electrode. SWV: square wave voltammetry.
Table 2. MIP-based optical sensors in environmental and biomedical applications.
Table 2. MIP-based optical sensors in environmental and biomedical applications.
Sensor Form or Electrode Functional MonomerTargetSampleLinear RangeRef.
LOD
UV/Visible spectroscopy
ZnFe2O4/MIPMembraneAcrylamide (AM)Bisphenol AUltrapure water10–1000 nM[214]
6.18 nM
MIPMembraneItaconic acid (IA)PhenolDrinking, natural, and wastewater50 nM–10 mM[215]
50 nM
MIPPaperMAA + polyethylenimine (PEI),Cd(II)Lake water1–100 ng/mL[216]
0.4 ng/mL
MIPParticles2-acrylamido-2-methyl-1-propanesulfonic acid (AMPSA)Basic red 9Ultrapure waterNI[217]
NI
Magnetite-MIPMicrospheresMAARhodamine BWine0.04–1.4 μg/mL[218]
1.05 μg/L
MIP-Graphitic C3N4FTO4-vinylpyridine (4-VP)Bisphenol ABottled water5–200 μM[219]
1.3 μM
MIPFilmAcrylic acid (AA)2-butoxyethanolHydraulic fracking wastewater10 ppb–100 ppm[220]
3.4 ppb
MIPFilmAATestosteroneUltrapure water5–100 ppb[221]
4.2 ppb
Magnetic MIPNPTriallyl isocyanurateSterigmatocystinWheat1.8–25 ng/g[222]
0.63 ng/g
Fluorescence
MIP/Mn-ZnS QDNP4-vinylphenylboronic acid and methyl methacrylate α-fetoproteinHuman serum0.05–10 μg/L[223]
48 ng/L
C dots-MIPNPMAASterigmatocystinMillet, rice, corn0.05–2 mg/L[224]
0.019 mg/L
MIP/Mn-ZnS QDNP3-aminopropyltriethoxysilane (APTES) NicosulfuronRiver water12–6000 nM[225]
1.1 nM
MIP/POFCapillary tubeMAABisphenol AMineral water bottle in ethanol and water solutions3 × 10−9–5 × 10−6 g/mL[226]
1.7 × 10−9 g/mL
Allyl fluorescein-MIPMicrospheres4-vinylphenylboronic acid+ MAA (VPBA/MAA)TetracyclineEtracycline hydrochloride, human serum, swine urine4.26–150 nM[227]
4.26 nM
MIP-ILMicrospheres3-(anthracen-9-ylmethyl)-1-vinyl-1H-imidazol-3-ium chloride (Fluorescent IL monomer) p-nitroanilineRiver, lake, and tap water10 nM–10 M[228]
9 nM
MIP-SiO2-CdTe QDCompositeAcrylamide (AM)SulfanilamideRiver water2–30 μM[229]
0.17 μM
CdTe QD-MIPCompositeAMλ-cyhalothrinRiver water0.1–16 μM[230]
0.03 μM
MIPCore-shell particlesmethacrylamide2,4-diclorophenoxyacetic acidRiver water20 nM–5 μM[231]
20 nM
MIP-C dotsMesoporous structureAmino-CDsTNTTap water, soil0.5–20 μM[232]
17 nM
MIPHollow NPAMλ-cyhalothrinCanal water10.26–160 nM[233]
10.26 nM
Fe3O4-SiO2-MIPCore-shell magnetic NPnitrobenzoxadiazole fluorophore (NBD-MA)Rhodamine BMethanolNI[234]
10−8 M
CdTe/CdS QD-MIPMesoporous structure3-Aminopropyltriethoxysilane (APTES) DiniconazoleSoil, river and wastewater20–160 μg/L[235]
6.4 μg/L
MIPAu NCAPTESBisphenol ASeawater0.1–13 μM[236]
0.1 μM
MIPMicroplateDansyl methacrylateBisphenol ATap, river and distilled water10–2000 μg/L[237]
3 μg/L
QD-MIPCompositeAcrylamide (AM)CyphenothrinRiver water0.1–80 μM[238]
9 nM
MIP-C dotsNanocompositeAPTESBisphenol ARiver water10−7–4.2 × 10−6 M[239]
3 × 10−8 M
MIP-ZnO QDNanocompositeAPTESThioridazine hydrochlorideHuman plasma4–120 nM[240]
0.43 nM
MIP-C dotsPaperMAA3-monochloropropane-1,2-diolSoy sauce1–150 ng/mL[241]
0.6 ng/mL
MIP-QDCompositeAPTESTetrabromobisphenol-AElectronic waste1–60 ng/mL[242]
3.6 ng/mL
MIP-Graphene QDComposite3-aminopropyltriethoxysilane (APTS)OmidazoleHuman plasma0.75–30 μM[243]
0.24 μM
MIP-CdSeS/ZnS QDGlass slideMAASulfasalazineHuman plasma and urine0.02–1.5 μM[244]
0.0071 μM
SiH4-C dots-MIPNanocompositeAPTESAcetamipridWastewater, apple7–107 nM[245]
2 nM
QD-MIPCore-shellAPTESPerfluorooctanoic acidTap and river water0.25–15 μM[246]
25 nM
MIP-C dotsFilmPoly(methyl acrylate-co-acrylic acid)2,4- dinitrotolueneUltrapure water1–15 ppm[247]
0.28 ppm
MIP-C dotsFilmAcrylic acid(AA) + methyl acrylate (MA) 2,4- dinitrotolueneLake and tap water1–15 ppm[248]
0.28 ppm
MIP-QDNanocompositeAPTESThiamphenicolMilk0.10–100 μM[249]
0.04 μM
MIP-C dotsFilmAPTESCetricineUrine, saliva0.5–500 ng/mL[250]
0.41 ng/mL
Chemiluminescence
MIP/Cromatography paperPaper stripMAADichlorvosCabbage leaves, tomato skin0.003–10 μg/mL[251]
0.8 ng/mL
Silanized magnetic graphene-MIP/CapillaryCapillary tubeAcrylamide (AM) DopamineUrine, dopamine hydrochloride injection8–200 ng/mL[252]
1.5 ng/mL
MIP-Magnetic NPNPMAADibutyl phthalateJuice3.84 × 10−8–2.08 × 10−5 M[253]
2.09 × 10−9 M
MoS2-Graphene QD-MIPGCEo-phenyl-enediamine (o–PD) 2-methyl-4-chlorophenoxyacetic acidOat, tap, and lake water10 pM–0.1 μM[254]
5.5 pM
MIP-Au NP-CdSe/ZnS QDAu electrodeThioanilin2-methyl-4-chlorophenoxyacetic acidTap, lake and river water, oat, rice10 pM–50 μM[255]
2.2 nM
MIP/Chromatography paperPaper diskAM2,4-dichlorophenoxyacetic acidTap and lake water5 pM–10 μM[256]
1 pM
CdTe QD-Chitosan-GO-Magnetite-MIPNPMAAChrysoidinePaper, fabric10−7–10−5 M[257]
3.2 × 10−9 M
MIP-Aptamer-CdS QDSPEDopamineDiethylstilbestrolFish0.3–100,000 pg/mL[258]
0.1 pg/mL
MIPCPEMAAAzithromycinBlood serum, urine10−10–4 × 10−7 M[259]
2.3 × 10−11 M
MIP-Cu nanoclustersGCEo-phenylenediamineEnrofloxacinMeat, lake water, bovine serum, human urine0.1 nM–1 μM[260]
27 pM
Surface plasmon resonance
MIP/POFOptical fiberMAAL-nicotineUltrapure water1.86 × 10−4–10−3 M[261]
1.86 × 10−4 M
MIP/POFOptical fiberMAAFurfuralTransformer oil9–30 ppb[262]
9 ppb
MIP/Ag-POFOptical fiberPolyanilineAscorbic acidUltrapure water10−8–10−7; 10−6–10−4 M[263]
1.28 × 10−10 M
MIP/Ag-POFOptical fiberMAAProfenofosPBS10−4 μg/L[264]
2.5 × 10−6 μg/L
MIPFilmMethacrylamideAmoxicillinTap water, PBS0.1–2.6 nM[265]
73 pM
MIPFilmMAAHistamineFish25–1000 μg/L[266]
25 μg/L
MIPNPN-methacryloyl-(L)-histidine methyl esterHistamineCanned tuna, cheese0.001–10 μg/L[267]
0.58 ng/L
MIPNPBiotinylated phenylalanineProstate-specific antigenBlood0.001–0.2 ng/mL[268]
1 pg/mL
MIPNanofilmN-methacryloyl-(L)-tryptophan methyl esterCarbofuran, dimethoateRiver water10–1000 ng/L[269]
7.11 (carbofuran); 8.37 (dimethoate) ng/L
MIP-Au NPFilmN-methacryloyl-(L)-phenylalanineAflatoxin B1Corn, peanut0.0001–10 ng/mL[270]
1.04 pg/mL
MIP-Ag NPFilmN-methacryloyl-(L)-histidine methyl esterEscherichia coliMimic urine15–1,500,000 CFU/mL[271]
0.57 CFU/mL
Raman scattering
MIP microspheres/Au-Klarite substrateMicrosphresMAANicotineAcetonitrileNI[272]
NI
MIP/Au-Disulfide-derivatized perfluorophenylazide-Klarite substrateNPMAAPropranololHuman urineNI[273]
7.7 × 10−4 M
MIP/Ag dentrite nanostructure substrateFine particlesMAAMelamineMilk0.005–0.05 mM[274]
0.012 mM
Ag-MIPCore-shell nanoplatesMethacrylamide Rhodamine BUltrapure waterNI[275]
10−12 M
Ag-MIPCore-shellAMRhodamine 6GWater10−12–10−6 M[276]
10−14 M
Ag-MIPCore-shellMAA4-mercaptobenzoic acidWaterNI[277]
10−15 M
Boronate affinity MIP-Boronate affinity SERSSandwich assayAPTES α-fetoproteinHuman serum0.001–10 μg/mL[278]
0.1 ng/mL
MIP-Au NPCore-shell3-(triethoxysilyl)propyl isocyanate (TEPIC) Bisphenol ASurface water, plastic-bottled beverages2.2 × 10−6–10−4 M[279]
5.37 × 10−7 M
MIP-Magnetic NPCore-shellMAACiprofloxacinFetal bovine serum10−7–10−4 M[280]
10−7 M
MIP-SiO2-AgNPCore-shelltetraethyl orthosilicate (TEOS) Bisphenol ATap and lake water, milk1.75 × 10−11–1.75 × 10−6 M[281]
1.46 × 10−11 M
MIP-AgCore-shellAcrylamideGlibenclamideWater1 ng/mL–100 μg/mL[282]
1 ng/mL
SiO2/rGO/Au-MIPNPMethacrylic acid, acrylamide2,6-dichlorophenolWater1–100 nM[283]
0.02 nM
MIP-Au NPCore-shellPhenyltrimethoxysilaneL-PhenylalanineBovine serum10−8–10−4 M[284]
10−9 M
UV/visible spectroscopy. Raman scattering
MIP-Au NPFine particlesMAAAtrazineApple juiceNA (Color); 0.005–1 mg/L (SERS)[212]
0.005 (Color); 0.0012 (SERS) mg/L
Fluorescence. Raman scattering
Magnetic MIPNPPoly(ethylene-co-vinyl alcohol) PhenylalanineHuman urine7–100 (F); 5–800 μg/mL (RS)[285]
NI (F); 0.4 μg/mL (RS)
Photoelectrochemical
MIPITOPyrrole Bisphenol ARiver and tap water2–500 nM[41]
1.2 nM
MIP-Au NP-ZnO NPPaperPyrrolePentacholorophenolDrinking and river water0.01–100 ng/mL[286]
4 pg/mL
F graphitic C3N4-MIPFTO4-vinylpyridine (4-VP) Cr(VI)Tap and river water0.01–100 ppb[287]
0.006 ppb
MIP/TiO2 NTThin filmo-phenyl-enediamine (o–PD) LindaneDrinking and river water0.1–10 μM[213]
0.03 μM
MIP/Au NP-TiO2 NTThin filmo-PD ChlorpyrifosGreen vegetables0.05–10 μM[288]
0.96 nM
MIP/TiO2 NTVertical NTPyrroleBisphenol ADrinking, river, and tap water, domestic sewage4.5–108 nM[289]
2 nM
MIP-TiO2 NanorodsP-aminothiophenol (ATP)ChlorpyrifosDrinking and river water0.01–100 ng/mL[290]
7.4 pg/mL
MIP-TiO2FTO glass substrateAPTESRNase BHuman serum0.5 pM–2 μM[291]
0.12 pM
MIP-CdTe QD-Au NPPaper/SPEAAMS-fenvalerateApple, pear, tomato, cucumber10−8–10−6 M[292]
3.2 × 10−9 M
MIP-TiO2-MWCNTCore-shellTiO2Microcystin-LRTap, pond, and river water1 pM–3 nM[293]
0.4 pM
MIPPorous carrier4-vinylpyridine (VP) and N-isopropylacrylamide (NIPAM)Bisphenol ASeawater, yogurtNI[294]
11.14 μg/L
MIPStrip/SPEAMPerfluorooctane sulfonyl fluorideTap, river, and lake water0.05–500 ppb[295]
0.01 ppb
Aptamer-MIP-GO-C3N4FTOAptamerKanamycinNa2SO4 solution1–230 nM[296]
0.2 nM
MIP-BiIO nanoflake arrayNanofibersChitosanTriphenyl PhosphateTap, river, and lake water0.01–500 ng/mL[297]
0.008 ng/mL
MIP/TiO2 NTVertical NTAMPerfluorooctane sulfonateTap, river, and mountain water0.5–10 μM[298]
86 ng/mL
CdS dots-Graphene-MIPFTOPyrrole4-aminophenolLake water5 × 10−8–3.5 × 10−6 M[299]
2.3 × 10−8 M
MIP-AgI NP-BiIO nanoflake arrayFTOAMPerfluorooctanic acidTap and river water0.02–1000 ppb[300]
0.01 ppb
Glass/ZnO/MIPNanorodsPyrroleBisphenol SPBS2.5–12.5 μM[301]
0.7 μM
LOD: limit of detection. NA: not achieved. NI: not informed. CPE: carbon paste electrode. FTO: F-doped SnO2. IL: ionic liquid. ITO: indium tin oxide. GCE: glassy carbon electrode. GO: graphene oxide. MWCNT: multi-walled carbon nanotube. NC: nanocluster. NP: nanoparticle. NT: nanotube. PBS: phosphate buffer solution. POF: plastic optical fiber. QD: quantum dot. rGO: reduced graphene oxide. SERS: surface enhanced Raman scattering. SPE: Screen-printed electrode.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kadhem, A.J.; Gentile, G.J.; Fidalgo de Cortalezzi, M.M. Molecularly Imprinted Polymers (MIPs) in Sensors for Environmental and Biomedical Applications: A Review. Molecules 2021, 26, 6233. https://doi.org/10.3390/molecules26206233

AMA Style

Kadhem AJ, Gentile GJ, Fidalgo de Cortalezzi MM. Molecularly Imprinted Polymers (MIPs) in Sensors for Environmental and Biomedical Applications: A Review. Molecules. 2021; 26(20):6233. https://doi.org/10.3390/molecules26206233

Chicago/Turabian Style

Kadhem, Abbas J., Guillermina J. Gentile, and Maria M. Fidalgo de Cortalezzi. 2021. "Molecularly Imprinted Polymers (MIPs) in Sensors for Environmental and Biomedical Applications: A Review" Molecules 26, no. 20: 6233. https://doi.org/10.3390/molecules26206233

Article Metrics

Back to TopTop