3,060
Views
3
CrossRef citations to date
0
Altmetric
Reviews

An advantageous application of molecularly imprinted polymers in food processing and quality control

ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon & ORCID Icon

Abstract

In the global market era, food product control is very challenging. It is impossible to track and control all production and delivery chains not only for regular customers but also for the State Sanitary Inspections. Certified laboratories currently use accurate food safety and quality inspection methods. However, these methods are very laborious and costly. The present review highlights the need to develop fast, robust, and cost-effective analytical assays to determine food contamination. Application of the molecularly imprinted polymers (MIPs) as selective recognition units for chemosensors’ fabrication was herein explored. MIPs enable fast and inexpensive electrochemical and optical transduction, significantly improving detectability, sensitivity, and selectivity. MIPs compromise durability of synthetic materials with a high affinity to target analytes and selectivity of molecular recognition. Imprinted molecular cavities, present in MIPs structure, are complementary to the target analyte molecules in terms of size, shape, and location of recognizing sites. They perfectly mimic natural molecular recognition. The present review article critically covers MIPs’ applications in selective assays for a wide range of food products. Moreover, numerous potential applications of MIPs in the food industry, including sample pretreatment before analysis, removal of contaminants, or extraction of high-value ingredients, are discussed.

1. Introduction

There is no healthy diet without healthy food. Therefore, the modern lifestyle requires easy access to high-quality food products (Wijayaratne et al. Citation2018). On the one hand, this trend supports the sustainable development of agricultural production and returns it to the traditional cultivation methods (Antonelli and Viganò Citation2018). On the other hand, it boosts the import of food products from exotic destinations. Both trends significantly increase production and delivery costs, while market competition enforces low-price maintenance. That tempts dishonest producers and suppliers (Soon et al. Citation2019; Ali and Suleiman Citation2018). Therefore, food product quality should be controlled at various manufacturing and distributing stages.

This control is very challenging in the global market era (Soon et al. Citation2019; Ali and Suleiman Citation2018). It is nearly impossible to track and control all food production and delivery chains not only for regular customers but also for the State Sanitary Inspections. Therefore, food manufacturers would try to unfairly raise the quantity and quality of the manufactured products (Silva et al. Citation2021). Furthermore, different chemicals used during farming, i.e., chemical fertilizers, pesticides, herbicides, antibiotics, antifungals, hormones, and food additives used by the food industry, including food preservatives, dyes, and artificial fragrances, can be present in food in significant amounts. Moreover, food products may be accidentally contaminated with heavy metal ions and toxins of industrial or natural origin. These toxins may be produced by mold or bacteria during food rotting. All of the above have adverse effects on the consumers’ health.

Currently, certified laboratories are using accurate food quality and safety analysis methods. However, these methods are usually very laborious. They require well-trained and experienced employees and, frequently, costly instrumentation, e.g., high-performance liquid chromatography with mass spectrometry detection (HPLC-MS) (Silva et al. Citation2018). Therefore, the need to develop fast, robust and cost-effective analytical assays to determine food product contaminants should be highlighted. Moreover, newly developed analytical methods should easily be integrated with portable, hand-held devices, thus enabling on-spot sample analysis (Nelis et al. Citation2020).

Numerous electrochemical and optical methods of analytical signal transduction perfectly fulfill these requirements. However, their sensitivity and especially selectivity are far insufficient for successful sensors devising. Integrating selective recognizing units with the signal’s transducer is necessary to circumvent this deficiency (Khanmohammadi et al. Citation2020; Naresh and Lee Citation2021; Chen and Wang Citation2020). Applying biological recognizing units, including DNAs, aptamers, enzymes, monoclonal antibodies, cells, or even whole tissues, opened a route to devise various biosensors. Biosensors are sensitive and highly selective, and often specific. Therefore, abundant biosensors found their way to the market. Glucometers and pregnancy tests are the most widely offered biosensors. Despite their advantages, they suffer from several drawbacks. Their low durability originates from components of the biological origin. Therefore, they mostly serve as disposable devices. Alternately, e.g., multi-use glucometers require single-use sensing stripes with the glucose oxidase (GOx) or glucose dehydrogenase (GDH) enzyme immobilized on their surface. Such a cost-ineffective and high-waste approach may be accepted only in the healthcare field. More cost-effective sensors are in demand in food analysis, especially for controlling industrial-scale food production. For that purpose, chemically synthesized receptors may be immobilized on the transducer surface. Macrocyclic compounds, including cyclodextrins, crown ethers, and calixarenes, offer much better durability but at the expense of selectivity much lower than that of the bioreceptors.

Molecularly imprinted polymers (MIPs) compromise the durability of synthetic sensing materials with a high affinity to target analytes and selectivity of molecular recognition (Cieplak and Kutner Citation2016). Therefore, they effectively mimic natural molecular recognition. The present review critically covers the MIPs applications in chemosensors devised for selective food analysis using various transduction techniques and a wide range of tested products. Moreover, numerous potential applications of MIPs in the food industry, including sample pretreatment before analysis, removal of contaminants, or extraction of high-value ingredients, are discussed (Scheme 1).

Scheme 1. Outline of MIPs applications in the food industry and food quality control protocols.

Scheme 1. Outline of MIPs applications in the food industry and food quality control protocols.

2. Molecularly imprinted polymers (MIPs)

MIPs are an illustrative example of nature-inspired smart materials (Cieplak and Kutner Citation2016). In their matrices, MIPs contain molecular cavities imprinted during polymerization. For that purpose, template molecules are added to the solution for polymerization. Then, the polymer grows around these molecules during the polymerization. Later, these molecules’ removal results in molecular cavities in the polymer structure, resembling template molecules with their shape, size, and orientation of recognizing sites. These cavities can recognize and capture only these molecules that spatially fit them. Moreover, carefully selected monomers, called functional monomers, are used for polymerization to increase the selectivity of this molecular recognition. These monomers contain recognition sites, including functional groups, heteroatoms, π-π conjugated systems, etc., that interact with binding sites of template molecules via covalent bonds, hydrogen bonds, electrostatic attractions, π-π stacking, as well as hydrophobic and van der Walls interactions (Scheme 2). These monomers form stable pre-polymerization complexes with template molecules in solution; thus, they are built into the growing polymer during the polymerization. After template removal from the MIP, they stay as recognizing sites on the walls of the imprinted cavities, increasing the affinity of these cavities to the target analyte molecules and thus the binding selectivity.

Scheme 2. The flowchart of a general procedure of MIP synthesis. Adapted from Sharma et al. (Citation2013).

Scheme 2. The flowchart of a general procedure of MIP synthesis. Adapted from Sharma et al. (Citation2013).

Depending on their potential applications, MIPs are synthesized in many different forms, including bulk polymers, membranes, grains, micro- and nanoparticles (NPs), or thin films grafted on the solid support surfaces of diverse morphology. Several different polymerization procedures were employed for synthesizing MIPs, ranging from simple light-inducted free radical polymerization through emulsion polymerization and “living” polymerization to electrochemical polymerization. Details of MIPs synthesizing and characterizing have already been reported in various high-quality review articles (Uzun and Turner Citation2016; Wackerlig and Lieberzeit Citation2015; Niu, Chuong, and He Citation2016; Sharma et al. Citation2013; Beyazit et al. Citation2016). The readers who wish to explore the basics of molecular imprinting are encouraged to get acquainted with these articles. The present article focuses on the existing or potential applications of MIPs in food production and selective MIP chemosensors devised for food products analysis.

3. MIP sorbents application in food processing and control

The most common MIPs applications are separation materials where an unusually high degree of selectivity of the sorbents is utilized. For that, MIPs can be synthesized as irregular or spherical micro- or nanoparticles, membranes, or bulk materials that can be ground later. Prepared that way, MIPs may be used for dispersion extraction or as packing materials for solid-phase extraction (SPE) cartridges and high-performance liquid chromatography (HPLC) columns. The fast-growing field of MIP sorbents was reviewed in detail elsewhere (Turiel and Martin-Esteban Citation2010; Ashley et al. Citation2017). MIP-based separation materials are very convenient for sample preparation and purification. This opens up new opportunities for the food industry. Recently, several companies were established to develop and introduce MIP sorbents to the market. Only commercial examples are discussed in this chapter.

3.1. MIP SPE cartridges for sample pretreatment before food sample analysis

Developing new analytical methods for food product control and analysis may be a tremendous task. Usually, food products are inhomogeneous. They represent a complicated mixture of many different components in various concentration ranges. The analyte of interest may often be present in the analyzed sample at a concentration of several orders of magnitude lower than the interferences. Moreover, the HPLC system with a simple UV-vis or refractive index detection requires a relatively high analyte concentration. Therefore, food samples usually need laborious and time-consuming pretreatments to pre-concentrate the analyte and partially remove interferences that may foul and thus destroy the HPLC column. Hence, MIP sorbing materials appeared very useful in preliminary sample preparation procedures for HPLC assays (Scheme 3a). Recently, numerous MIP-SPE/HPLC resins were commercialized (). Mainly, they are dedicated to clinical analysis. However, examples of MIP-SPE/HPLC resins application to food analysis were also reported.

Scheme 3. (a) Food sample pretreatment before sample analysis. An example of ochratoxin A (OTA) HPLC determination in beer, red wine, and grape juice samples (Cao, Kong, et al. Citation2013) and GC-MS determination of bisphenol compounds in breast milk samples (Deceuninck et al. Citation2015). (b) Removal of food contaminants during processing. Purification of wine contaminated with pyrimethanil (Petcu Citation2013, Citation2015). (c) High-value compounds extraction. Molecularly imprinted SPE of oleanolic acid from grape pomace extract (Lu et al. Citation2018).

Scheme 3. (a) Food sample pretreatment before sample analysis. An example of ochratoxin A (OTA) HPLC determination in beer, red wine, and grape juice samples (Cao, Kong, et al. Citation2013) and GC-MS determination of bisphenol compounds in breast milk samples (Deceuninck et al. Citation2015). (b) Removal of food contaminants during processing. Purification of wine contaminated with pyrimethanil (Petcu Citation2013, Citation2015). (c) High-value compounds extraction. Molecularly imprinted SPE of oleanolic acid from grape pomace extract (Lu et al. Citation2018).

Table 1. Commercial solid-phase extraction (SPE) columns and their application in the food industry.

In the preceding decade, the NanoMyP® company introduced two types of MIP microparticles, commercialized as a dry powder, for selective extraction of ascorbic acid (Valero-Navarro et al. Citation2011) and tetracyclines via dispersion extraction. Moreover, the former resin was packed into an HPLC column and applied for selective ascorbic acid determination in fruit juice. Then, two more companies, namely POLYINTELL (AFFINISEP) and Merck (SupelMIP), offered several ready-to-use SPE cartridges. Their applications appeared to be useful for selectivity improvement in HPLC determination of mycotoxins (Catana et al. Citation2019; Cao, Zhou, et al. Citation2013; Cao, Kong, et al. Citation2013; Gonzalez-Salamo et al. Citation2015; Bryła et al. Citation2013; Lucci et al. Citation2010), antibiotics (Blasco and Pico Citation2012; Cirkva et al. Citation2019; Bustamante-Rangel, Rodríguez-Gonzalo, and Delgado-Zamareño Citation2022), bisphenols (Matejicek, Grycova, and Vlcek Citation2013; Deceuninck et al. Citation2015; Kubiak, Ciric, and Biesaga Citation2020; Bousoumah et al. Citation2015; Nicolucci et al. Citation2013), and other contaminants (Roszko, Szymczyk, and Jędrzejczak Citation2015; Hrobonova and Brokesova Citation2020) in water (Gonzalez-Salamo et al. Citation2015; Matejicek, Grycova, and Vlcek Citation2013), cider (Catana et al. Citation2019), wine (Cao, Kong, et al. Citation2013; Hrobonova and Brokesova Citation2020), beer (Cao, Kong, et al. Citation2013), grape juice (Cao, Kong, et al. Citation2013), cereal (Bryła et al. Citation2013; Lucci et al. Citation2010), pineapple (Kubiak, Ciric, and Biesaga Citation2020), lentil (Kubiak, Ciric, and Biesaga Citation2020), bean (Kubiak, Ciric, and Biesaga Citation2020), ginger (Cao, Zhou, et al. Citation2013), eggs (Blasco and Pico Citation2012), and meat products (Roszko, Szymczyk, and Jędrzejczak Citation2015; Cirkva et al. Citation2019).

3.2. Undesired ingredients removal from food products

Another possible application of MIP resins is the selective removal of undesired ingredients from food products. For that, a Ligar PL company was established in New Zealand to devise and fabricate MIPs to remove harmful food contaminants, including heavy metal ions or pesticides that may be extracted from liquid food products, i.e., drinking water, wine, cooking oil, or juice. Moreover, bitter, smoker, and unpleasant tastes or flavors may be removed from these products without significant loss of their desired properties. However, until now, this company has only presented successful removal of pyrimethanil fungicide from wine (Scheme 3b) (Petcu Citation2013, Citation2015). Ligar PL products have not yet been introduced to the worldwide market. But a high application potential of the purification procedure developed may be illustrated by MIP particles successfully devised by the K. Haupt group, in cooperation with the L’Oréal company, to remove odorous components of human sweat (Nestora et al. Citation2016; Mier et al. Citation2019).

3.3. Desired compounds extraction from natural sources

The under-estimated but presumably potentially the most profitable application of MIP particles is a selective extraction of high-value compounds from natural sources. Toward that, MIP particles can be dispersed in natural mixtures to absorb desired ingredients and then readily separated by sedimentation, filtration, centrifugation, or by applying a magnetic field in the case of magnetic MIP NPs (da Fonseca Alves et al. 2021; Aylaz et al. Citation2021). Therefore, this extraction may be performed under continuous-flow conditions, even on an industrial scale. In one important example, the G. Szekely group demonstrated extensive optimization of MIP particles synthesis and regeneration procedures (Kupai et al. Citation2017). Different L-phenylalanine methyl ester imprinted MIP particles were subjected to 100 adsorption-regeneration cycles. There was no loss of binding capacity within these 100 cycles for all examined MIP particles, even if the process was performed at elevated temperature (65 °C) and if methanol was used as the solvent for particle regeneration. Moreover, if divinyl benzene was applied as the cross-linking monomer, MIP particles’ binding capacity was unaffected by regeneration neither in acidic nor basic aqueous solution during all 100 cycles. However, MIP particles synthesized with ethylene glycol dimethacrylate or N,N’-methylenebis(acrylamide) cross-linking monomers deteriorated after the first 20 sorption-regeneration cycles. Therefore, applying MIP particles for selective extraction seems economically justified if considering high extraction selectivity and potentially low cost of the MIP particles’ synthesis. Accordingly, extraction of rosmarinic acid (Zahara et al. Citation2021) and quercetin (Pakade et al. Citation2013), as well as kaempferol (Pakade et al. Citation2013) from 3 kg of dried Salvia hypoleuca (Zahara et al. Citation2021) and 5 g of dried Moringa oleifera leaves (Pakade et al. Citation2013) were demonstrated, respectively. Moreover, in another example oleanolic acid imprinted polymer was applied to extract this acid from 150 mg grape pomace extract (Scheme 3c) (Lu et al. Citation2018). After the purification using MIP loaded column, the content of oleanolic acid increased from 13.4% to 93.2%. Unfortunately, these procedures were performed only on the laboratory scale and have not been scaled up yet. Recently, Ligar PL has established a new daughter company, Amber Purification Ltd., devoted to developing a large-scale system to purify cannabinoid extracts from hemp. However, their products have not been commercialized yet.

4. MIP chemosensors for food products analysis

The HPLC assays mentioned above enable very selective and reproducible contaminants determination in food samples, thus ensuring highly reliable food quality control. However, these assays suffer from several disadvantages. They use high quantities of expensive solvents of very high purity. With the unit price exceeding 100,000 EUR, an HPLC system equipped with a mass spectrometry detector is mainly supplied to sanitary inspection laboratories and industrial-scale food manufacturers. It is inaccessible for small food manufacturers, not to mention individual farmers and food market customers. Importantly, these systems are unsuitable for in-field determinations. Therefore, all examined samples must be collected and transported to specialized laboratories. That highlights the urge to design inexpensive, hand-held chemosensors for fast and straightforward analyte determinations. The following sections describe the MIP chemosensors’ application for various food sample testing, including determining particular contaminants or other relevant compounds. The most critical analysis details, the techniques used, and the food sample types are summarized in and .

Table 2. Examples of MIPs application for optical sensing in food samples.

Table 3. Examples of MIPs application for non-optical sensing in food samples.

4.1. MIP chemosensors for optical assays

Optical sensors are robust and inexpensive. For many procedures, optical assays are designed to give a readout that can be recorded by the naked eye of the operator, i.e., by observing color changes caused by the characteristic color reaction. MIPs may provide analyte preconcentration and selectivity enhancement for these procedures (Ye et al. Citation2018; Wu et al. Citation2018; Zhao et al. Citation2019; Feng et al. Citation2017). Fluorescence transduction application results in assays of enhanced sensitivity. Accordingly, a competitive fluorescent assay for histamine in the fish extract was reported (Mattsson et al. Citation2018). MIP particles and a fluorescence-tagged histamine derivative were added to the sample solutions examined. Then, the histamine content in the solution was indirectly determined by evaluating the amount of fluorescent derivative remaining in the solution after mixing with MIP particles. In another report, a catalytically active AgNPs@MOF@MIP (where MOF denotes a metal-organic framework) nanocomposite was proposed (Bagheri et al. Citation2018). In the H2O2 presence, terephthalic acid oxidation to fluorescent 2-hydroxyterephthalic acid was catalyzed. The patulin analyte inhibited this reaction by binding to MIP cavities, thus lowering the recorded fluorescence intensity. Moreover, a very fast and robust ELISA-like competitive assay was devised by depositing 17β-estradiol imprinted silica on the surface of a filter paper chemosensor (Scheme 4c) (Xiao et al. Citation2017). The target analyte competition with estradiol-labeled horseradish peroxidase (HRP) allowed for naked-eye 17β-estradiol detection by soaking the sensor in a solution of a colored reaction substrate. For milk samples containing 17β-estradiol, the color change was much less pronounced.

Scheme 4. (a I) A typical structure of MIP nanoparticles synthesized for optical assays. (a II and a III) An example of a fluorescent MIP optical assay. Carbon QD@MIP core-shell fluorescent NPs applied for tartrazine determination in saffron tea. Adapted from Zoughi et al. (Citation2021); (b) Colorimetric determination of pyrethroid metabolite in fruit juice and beverages based on adsorption-desorption on imprinted silica NPs and color reaction with KMnO4, adapted from Ye et al. (Citation2018). A competitive assay for naked-eye semi-quantitative determination of 17β-E2 hormone in milk using MIP-coated nylon membrane and horseradish peroxidase (HRP) labeled 17β-E2, adapted from Xiao et al. (Citation2017); (d) Photonic structures of magnetic MIP particles assembled in a magnetic field applied for colorimetric determination of melamine, adapted from You, Cao, and Cao (Citation2016); (e) Color changes of inverse-opal MIP film to methyl anthranilate concentration changes allow semi-quantitative detection of the target analyte in wine samples, adapted from Wu et al. (Citation2019).

Scheme 4. (a I) A typical structure of MIP nanoparticles synthesized for optical assays. (a II and a III) An example of a fluorescent MIP optical assay. Carbon QD@MIP core-shell fluorescent NPs applied for tartrazine determination in saffron tea. Adapted from Zoughi et al. (Citation2021); (b) Colorimetric determination of pyrethroid metabolite in fruit juice and beverages based on adsorption-desorption on imprinted silica NPs and color reaction with KMnO4, adapted from Ye et al. (Citation2018). A competitive assay for naked-eye semi-quantitative determination of 17β-E2 hormone in milk using MIP-coated nylon membrane and horseradish peroxidase (HRP) labeled 17β-E2, adapted from Xiao et al. (Citation2017); (d) Photonic structures of magnetic MIP particles assembled in a magnetic field applied for colorimetric determination of melamine, adapted from You, Cao, and Cao (Citation2016); (e) Color changes of inverse-opal MIP film to methyl anthranilate concentration changes allow semi-quantitative detection of the target analyte in wine samples, adapted from Wu et al. (Citation2019).

The above procedures require using costly and environmentally unfriendly chemicals. Moreover, most of these chemicals are being dumped into waste after use. Therefore, they are cost-ineffective and non-ecological. It is much more reasonable to synthesize MIP particles to serve as dyes themselves (Scheme 4a, I). Toward that, MIP particles containing co-polymerized fluorescent monomers were synthesized (Gao, Li, et al. Citation2014; Ashley, Feng, and Sun Citation2018; Li, Yin, et al. Citation2015). Moreover, an MIP film can be grafted on the surface of fluorescent quantum dots (QDs) (Sun et al. Citation2018; Wang, Fang, et al. Citation2017; Jalili et al. Citation2020; Wu, Lin, et al. Citation2017; Li, Jiao, et al. Citation2018; Fang et al. Citation2019; Cui et al. Citation2020; Shirani et al. Citation2021; Zoughi et al. Citation2021; Chen, Fu, et al. Citation2022; Sa-nguanprang, Phuruangrat, and Bunkoed Citation2022) or luminescent upconverting NPs (Liu et al. Citation2017). Those particles can be collected and regenerated after the assay and re-used many times. Recently, dual-emission MIP fluorescent particles were invented (Jalili et al. Citation2020). They were synthesized in two steps in a one-pot reaction. First, silica core particles containing (blue light)-emitting carbon QDs were synthesized. Then, a penicillin G imprinted mesoporous silica film containing (yellow light)-emitting carbon QDs was grafted as a shell. Due to spatial separation of (blue light)-emitting QDs, the penicillin G analyte binding in imprinted molecular cavities quenched the fluorescence of only (yellow light)-emitting carbon QDs. As tested on milk samples, a pronounced color change in the penicillin G analyte presence was observed even with the naked eye. In another report, using the sonication encapsulation method, luminescent and magnetic NPs were entrapped in MIP nanocomposites (Li and Wang Citation2013). Magnetic field-driven MIP nanocomposites separation from the sample solution decreased interference from other polycyclic aromatic hydrocarbons (PAHs), and only the phenanthrene target analyte significantly quenched the MIP luminescence. MIP nanocomposites emitted red light (λ = 620 nm). Therefore, the naked eye readily observed luminescence intensity changes due to phenanthrene presence.

Another procedure involved devising label-free optical assays. For that, MIPs were deposited on the surface of gold-layered SPR chips as thin films (Scheme 6a, I and V) (Jiang et al. Citation2015; Zhang et al. Citation2018), or MIP NPs’ monolayers (Ashley et al. Citation2018; Yao et al. Citation2016; Çimen, Bereli, and Denizli Citation2022). In this case, analyte binding in MIP caused a change in the electric permittivity of the film that was in contact with an ultra-thin gold film deposited on the SPR chip surface. This binding shifts the evanescing light angle and wavelength, at which resonance with surface plasmon occurs. Thus, light is being absorbed. However, this approach is usually dedicated to determining macromolecular compounds, and SPR determination sensitivity to small-molecule compounds is relatively low. Therefore, Au NPs (Altintas Citation2018) and magnetic MIP NPs (Yao et al. Citation2013) were applied to enhance the sensitivity of SPR chemosensors. Moreover, the SPR spectrometer readout depends on the angle of the light evanescence. It makes chemosensors susceptible to mechanical vibrations and thus useless for in-field assays. Therefore, an MIP film was deposited by electropolymerization on the surface of an optical fiber to overcome this deficiency (Li, Zheng, et al. Citation2018). To this end, fibers were unclothed using a sharp blade, and then the cladding of sensing sections was entirely removed by immersing them in an HF solution. Finally, a 5-nm thick Cr underlayer and a 50-nm thick Au layer were consecutively sputtered on the sensing section of the optical fiber sensor. This Au layer served as the working electrode during deposition by electropolymerization of the MIP film, and a transducer sensitive to analyte binding to the MIP imprinted cavities. In another example, an MIP film was deposited on a 100-nm diameter Au nanodisks array (Guerreiro et al. Citation2017). Thus, changes in localized surface plasmon resonance were monitored.

Scheme 6. (a) Examples of the most common MIP chemosensors including (I) a general procedure of thin MIP film deposition on the sensors’ surface. Simplified schemes of (II) electrochemical, (III) EG-FET, (IV) QCM, and (V) SPR MIP chemosensors, adapted respectively, from (I) (Sharma et al. Citation2012); (II) (Pieta et al. Citation2013); (III) (Iskierko et al. Citation2016), and (IV) (Dabrowski et al. Citation2016); (b) Microfluidic devise for simultaneous purification and electrochemical determination of carbofuran in fruit and vegetable samples. Adapted from Li, Li, et al. (Citation2018); (c) ECL assay based on MIP film and upconverting NPs for clenbuterol determination in meat samples. Adapted from Jin et al. (Citation2018).

Scheme 6. (a) Examples of the most common MIP chemosensors including (I) a general procedure of thin MIP film deposition on the sensors’ surface. Simplified schemes of (II) electrochemical, (III) EG-FET, (IV) QCM, and (V) SPR MIP chemosensors, adapted respectively, from (I) (Sharma et al. Citation2012); (II) (Pieta et al. Citation2013); (III) (Iskierko et al. Citation2016), and (IV) (Dabrowski et al. Citation2016); (b) Microfluidic devise for simultaneous purification and electrochemical determination of carbofuran in fruit and vegetable samples. Adapted from Li, Li, et al. (Citation2018); (c) ECL assay based on MIP film and upconverting NPs for clenbuterol determination in meat samples. Adapted from Jin et al. (Citation2018).

Very robust label-free optical chemosensors based on MIP photonic structures were proposed. To this end, a magnetic field-assisted colloidal crystal of magnetic MIP NPs was deposited (Scheme 4d) (You, Cao, and Cao Citation2016; You et al. Citation2017). In another approach, MIP films of the inverse opal structures were synthesized using silica beads as sacrificial molds (Scheme 4e) (Yang, Peng, et al. Citation2017; Li et al. Citation2019; Wu et al. Citation2019; Qiu et al. Citation2020). In both cases, analyte binding in the MIP film caused this film to swell and or to change its electric permittivity. In turn, that generated a significant change in the film color originating from Bragg diffraction. That way, toxins in various food samples were determined. The MIP chemosensors based on the colloidal crystals or the inverse-opal structures seem promising candidates for hand-held portable device fabrication applications. That is because of their robustness, independence from any external power supplies, and the possibility to detect analytes just by naked eye observation of the color change.

Furthermore, surface-enhanced Raman spectroscopy (SERS) transduction was combined with MIP recognition. The advantages of SERS sensors, including high sensitivity and qualitative analyte identification, were improved by MIPs selectivity. MIPs were applied for sample pretreatment in the most robust approach before the SERS assay to pre-concentrate and purify the target analyte (Feng et al. Citation2017; Feng et al. Citation2013; Wu et al. Citation2016; Hua et al. Citation2018; Feng et al. Citation2018; Zhao et al. Citation2019). Moreover, an Au NPs suspension was applied for extraction to collect the analyte, namely, histamine, accumulated on the MIP SPE column (Gao, Grant, and Lu Citation2015). In more advanced procedures, the bulk MIPs were decorated with Au (Xie et al. Citation2017; Wang et al. Citation2020) and Ag (Hu and Lu Citation2016) NPs. Alternately, a thin MIP film was grafted on the surface of Au NPs (Zhou et al. Citation2020; Yin, Wu, et al. Citation2018) or ZnO@TiO2@Ag NPs (Chen, Wang, et al. Citation2022). Similarly, a filter paper was coated with carbon ink on one side and decorated with silver dendrites on the other (Scheme 5a) (Zhao, Liu, et al. Citation2020). Then, an MIP was deposited by electropolymerization inside this paper. Ag NPs were synthesized on top of the MIP to enhance the SERS signal. In these procedures, analyte extraction and SERS determination were performed simultaneously. Various vegetable samples were tested in this manner. In another report, MIP microparticles served as a thin-layer chromatography (TLC) stationary phase (Gao et al. Citation2015). After developing and drying, the TLC plate was decorated with Au NPs, and the SERS signal was recorded (Scheme 5b). That enabled rapid Sudan I determination in paprika powder with minimal sample pretreatment.

Scheme 5. (a) Multilayer paper SERS chemosensor based on star-shaped silver dendrites, MIP film, and Ag NPs for imidacloprid determination in the cucumber, chives, and soybean samples. Adapted from Zhao, Liu, et al. (Citation2020); (b) Combining TLC on the MIP film-coated plate with SERS assay for selective Sudan I determination in the paprika extract samples. Adapted from Gao et al. (Citation2015).

Scheme 5. (a) Multilayer paper SERS chemosensor based on star-shaped silver dendrites, MIP film, and Ag NPs for imidacloprid determination in the cucumber, chives, and soybean samples. Adapted from Zhao, Liu, et al. (Citation2020); (b) Combining TLC on the MIP film-coated plate with SERS assay for selective Sudan I determination in the paprika extract samples. Adapted from Gao et al. (Citation2015).

4.2. Electrochemical MIP chemosensors

Electrochemical sensors are gaining more and more interest because of their robustness, easy operation, and highly reproducible determinations. However, they usually suffer from low selectivity. The biological receptor immobilization on the electrode surface enables circumventing this disadvantage. Despite high selectivity and sensitivity, prepared that way, electrochemical biosensors reveal many deficiencies, mainly originating from the fragility of the biological recognition units used for their fabrication. Therefore, electrodes coated with selective MIP films (Scheme 6a, I and 6a, II) have recently attracted more and more interest. So far, numerous examples of MIP film-coated electrodes have been reported for possible application in food quality control. Charged compounds can easily be determined with potentiometry (Shirzadmehr, Afkhami, and Madrakian Citation2015; Anirudhan and Alexander Citation2015). This technique usually covers a broad concentration range. But, by its nature, the potentiometric sensor’s response linearly depends on the logarithm of concentration. Hence, it is insensitive to small changes in the analyte concentration.

If the target analyte is electroactive, it can be determined by recording its oxidation or reduction current changes with time by chronoamperometry at a selected constant potential applied (Lian et al. Citation2013; Turco, Corvaglia, and Mazzotta Citation2015; Turco et al. Citation2018; Amatatongchai et al. Citation2018). But there is a substantial limitation. That is, the target analyte must be the only electroactive component of the sample in the studied potential range. Voltammetric techniques enable partial overcoming of this deficiency. However, the faradaic currents originating from electrode reactions are overlapped by interfering capacity currents in a simple representation (Yang, Zhao, and Zeng Citation2016; Zhang et al. Citation2017; Deng, Xu, and Kuang Citation2014; Hassan et al. Citation2019; Li, Liu, et al. Citation2015). These undesired currents are subtracted in advanced voltammetry techniques, e.g., differential pulse voltammetry (DPV). Therefore, DPV sensitivity is very high, and the limit of detection (LOD) is low (Liu et al. Citation2022). Importantly, these techniques are dedicated to determining electroactive analytes (Yang, Zhao, and Zeng Citation2016; Zhang et al. Citation2017; Deng, Xu, and Kuang Citation2014; Hassan et al. Citation2019; Li, Liu, et al. Citation2015). If the target analyte is electroinactive, MIP chemosensors can signify their advantage. Electroinactive analytes may be determined using the so-called “gate effect” (Sharma et al. Citation2019). For that, a redox probe is added to the sample solution, and changes in the faradaic current caused by the changes in MIP film properties incurred by analyte binding are recorded (Sharma et al. Citation2019). Interestingly, redox probes can also be immobilized inside the MIP film (Lach et al. Citation2021). Moreover, such advanced techniques as electrochemical impedance spectroscopy (EIS) not only provide insight into the mechanism of the electrode processes (Sharma et al. Citation2019) but may also serve as a sensitive transduction tool for analyte determination (Lach et al. Citation2017; Lach et al. Citation2019; Ayerdurai, Cieplak, et al. Citation2021; Munawar et al. Citation2020; Shamsipur, Moradi, and Pashabadi Citation2018). In another example, capacitive impedimetry at MIP film-coated electrode was applied to determine cancerogenic aromatic amines in meat samples (Ayerdurai, Garcia-Cruz, et al. Citation2021).

Electrochemical redox processes can be combined with optical transduction to enhance the sensitivity and selectivity of determinations involving MIPs. For that, quenching of electrochemiluminescence (ELC) resulting from analyte binding by an MIP was applied (Li, Liu, et al. Citation2017; Zhang et al. Citation2022; Wang et al. Citation2022; Zhao et al. Citation2022; Jin et al. Citation2018). MIP thin films were deposited by electropolymerization on the electrodes decorated with carbon QDs (Li, Liu, et al. Citation2017; Zhang et al. Citation2022), Cu nanoclusters (Wang et al. Citation2022), and Ru(bpy)32+ decorated Fe2O3 microfibers (Zhao et al. Citation2022) or upconverting NPs (Jin et al. Citation2018; Scheme 6c). Au NPs, GO, and rGO were deposited on the electrodes to increase the electrode surface area and conductivity. Thus, reagents enabling electron transfer in their excited states were electrochemically generated on NPs more efficiently. Analyte molecules binding in MIP cavities disturb this process. Therefore, it was possible to determine target analytes at very low concentrations.

4.3. Other types of MIP chemosensors

Molecular imprinting was also combined with several other transduction techniques. For instance, MIP films were successfully deposited on the surface of quartz crystal resonators (Scheme 6a, IV) (Lach et al. Citation2017; Pietrzyk et al. Citation2009; Liu et al. Citation2014; Ebarvia, Ubando, and Sevilla Citation2015; Fang et al. Citation2017; Lin et al. Citation2018; Dayal et al. Citation2019; Zhao, He, et al. Citation2020; Ceylan Cömert et al. Citation2022). Then, the MIP film mass changes caused by analyte binding were measured by piezoelectric microgravimetry using a quartz crystal microbalance (QCM). This microgravimetry is a very sensitive technique enabling mass change measurements down to nanograms, i.e., below a monolayer coverage. However, QCM is very difficult to miniaturize into a hand-held device. On the contrary, sensors based on field-effect transistors (FETs), especially the extended-gate field-effect transistors (EG-FETs), seem to ease miniaturizing because of their robustness. Moreover, the high sensitivity and selectivity of EG-FET chemosensors make them attractive candidates for portable analytical devices useful for in-field measurements. For instance, rapid gluten determination in semolina flour was reported (Scheme 6a, III) (Iskierko et al. Citation2019).

Noteworthy, a microfluidic MIP-based device for carbofuran determination in fruit and vegetable samples was fabricated (Scheme 6b) (Li, Li, et al. Citation2018). This device consisted of two compartments connected with microchannels. One compartment contained the MIP that ensured sample purification and analyte preconcentration, while the other served as an electrochemical cell with a working electrode decorated with the DNA aptamer targeted to carbofuran. This aptamer ensured additional enhancement in both sensitivity and selectivity of the chemosensor. In another procedure, microfluidic chips containing microreactors connected to miniaturized thermistors were reported (Athikomrattanakul, Gajovic-Eichelmann, and Scheller Citation2011; Cornelis et al. Citation2019). These microreactors contained MIP NPs (Athikomrattanakul, Gajovic-Eichelmann, and Scheller Citation2011) or a surface imprinted MIP film (Cornelis et al. Citation2019). Then, the thermistors determined the heat released because of the strong interactions of the target analyte (Athikomrattanakul, Gajovic-Eichelmann, and Scheller Citation2011) or bacteria (Cornelis et al. Citation2019) with MIP molecular cavities.

5. Applications of MIP chemosensors in food products analysis

Although MIP-based chemosensors are still unavailable commercially, numerous possible applications in food product analysis were reported in the literature. The most evident applications include detecting and quantizing toxic contaminants in food products. Most of these contaminants originate from incorrect methods of plant cultivation and animal breeding. These contaminants include pesticides (Yao et al. Citation2013; Ye et al. Citation2018; Shirani et al. Citation2021; Amatatongchai et al. Citation2018; Capoferri et al. Citation2017; Fang et al. Citation2017; Lin et al. Citation2018; Dayal et al. Citation2019), insecticides (Wu et al. Citation2018; Feng et al. Citation2017; Zhao, Liu, et al. Citation2020; Zhang et al. Citation2017; Li, Liu, et al. Citation2016; Shi et al. Citation2017; Li, Li, et al. Citation2018), and herbicides (Zhao et al. Citation2019) in plants. In the case of meat products, mostly antibiotics (Tarannum, Khatoon, and Dzantiev Citation2020; Wang et al. Citation2022), hormones (Yao et al. Citation2016), fertilizers (J. H. Li, Xu, et al. Citation2017; Zhao et al. Citation2022), and introduced during high-temperature processing carcinogens (Lach et al. Citation2017; Ayerdurai, Garcia-Cruz, et al. Citation2021) are being determined. Fish products are even more at risk of harmful contamination than regular meat products. Accordingly, such toxins as heavy metal ions (Shirzadmehr, Afkhami, and Madrakian Citation2015), insecticides (Li, Jiao, et al. Citation2018), herbicides (Zhang et al. Citation2022), antibiotics (Liu et al. Citation2017; Li, Liu, et al. Citation2015; Wang, Yao, et al. Citation2017; Liu et al. Citation2019; Wang et al. Citation2022), bisphenols (Zhang et al. Citation2015), and hormones (Futra et al. Citation2016) were determined in fish using MIP-based chemosensors. Milk and honey, similarly to fish, belong to a group of food products of most concern. Therefore, numerous MIP sensors were devised to determine hormones (Bai et al. Citation2017; Xiao et al. Citation2017), herbicides (Hua et al. Citation2018), antibiotics (Xie et al. Citation2017; Yang, Peng, et al. Citation2017; Zhang et al. Citation2018; Altintas Citation2018; Lian et al. Citation2013; Turco, Corvaglia, and Mazzotta Citation2015; Turco et al. Citation2018; Yang and Zhao Citation2015; Ebarvia, Ubando, and Sevilla Citation2015; Sa-nguanprang, Phuruangrat, and Bunkoed Citation2022), and chemical contaminants (Li and Wang Citation2013; Zhao, He, et al. Citation2020; Chen, Fu, et al. Citation2022) including bisphenols (Yin, Wu, et al. Citation2018; Dadkhah et al. Citation2016) in milk. Similarly, MIP chemosensors for pesticides (Gao, Li, et al. Citation2014) and antibiotics (Lian et al. Citation2013; Song et al. Citation2014; Yang and Zhao Citation2015; Ebarvia, Ubando, and Sevilla Citation2015) determination in honey were proposed. All toxins mentioned above were introduced to food products unwittingly by producers. However, there are also examples of intentional contamination of food. This contamination is incurred because of food product falsification. In most cases, food products are being artificially colored. Therefore, several MIP chemosensors for dyes in fish (Wu, Lin, et al. Citation2017), tea (Zoughi et al. Citation2021), soft drinks (Li, Wang, et al. Citation2016; Wang et al. Citation2020; Yin, Cheng, et al. Citation2018), jelly (Yin, Cheng, et al. Citation2018), ice cream (Yin, Cheng, et al. Citation2018), and candy (Yin, Cheng, et al. Citation2018) were devised. Milk falsification is much more serious misconduct. That is, falsified milk products are purposefully contaminated with melamine (Li, Song, and Wen Citation2019). The reason is a false increase in detected protein levels in low-quality milk products. Consumption of melamine-containing food products has severely adverse and even lethal effects on humans (Li, Song, and Wen Citation2019). Therefore, melamine determination in milk products is so important. For that purpose, numerous MIP chemosensors were devised (Hu and Lu Citation2016; You, Cao, and Cao Citation2016; Li, Zheng, et al. Citation2018; Shang, Zhao, and Zeng Citation2014; Xu et al. Citation2018; Shamsipur, Moradi, and Pashabadi Citation2018; Pietrzyk et al. Citation2009; Ceylan Cömert et al. Citation2022).

Another parameter influencing food safety is its freshness. Stale or rotten food not only loses its texture and taste but may harm consumers’ health. Therefore, MIP chemosensors for determination of biogenic amines signaling food rottenness, namely histamine (Gao, Grant, and Lu Citation2015; Mattsson et al. Citation2018; Wang, Fang, et al. Citation2017; Jiang et al. Citation2015; Hassan et al. Citation2019; Chen, Wang, et al. Citation2022), and tyramine (Ayerdurai, Cieplak, et al. Citation2021) were applied for analysis of fish (Gao, Grant, and Lu Citation2015; Mattsson et al. Citation2018; Q. H. Wang, Fang, et al. Citation2017; Jiang et al. Citation2015; Hassan et al. Citation2019), prawns (Chen, Wang, et al. Citation2022), cheese (Ayerdurai, Cieplak, et al. Citation2021), and vinegar (Chen, Wang, et al. Citation2022) samples. Moreover, mycotoxins (Bagheri et al. Citation2018; Munawar et al. Citation2020; Guo et al. Citation2017; Pacheco et al. Citation2015; Gao, Cao, et al. Citation2014; Qiu et al. Citation2020; Çimen, Bereli, and Denizli Citation2022) and bacterial enterotoxins (Liu et al. Citation2014) indicate the presence of fungi and bacteria, respectively, and bacteria E. coli cells by themselves (Cornelis et al. Citation2019), were determined using MIP chemosensors.

5.1. Miscellaneous applications

Interestingly, MIP chemosensors were applied not only for undesired contaminants determination in food samples but also for food quality assessment. Accordingly, wine astringency was estimated. For that, the LSPR MIP chemosensor imprinted with saliva proteins was designed to study the interactions of these proteins with wine samples (Guerreiro et al. Citation2017). Presumably, wine ingredients bound saliva proteins, thus causing these proteins to shear and feel dryness. The wine astringency estimated by the MIP chemosensor and expressed in pentagalloyl glucose units (PGG) agreed well with the wine evaluation by a professional taster.

Moreover, flavors (Yang, Zhao, and Zeng Citation2016; W. H. Wu, Yang, et al. Citation2017), caffeine (Santos et al. Citation2012), quercetin (Sun et al. Citation2013; Yang, Xu, et al. Citation2017), and necessary nutrients, including vitamins (Feng et al. Citation2013) or L-phenylalanine (Zhou et al. Citation2020) were determined with MIP chemosensors to quantify food products quality.

6. Challenges and future prospectives

Polymer synthesis is easily scalable and can be implemented on an industrial scale. Therefore, MIPs can be readily synthesized in the form of bulk polymers, sponges, membranes, micro-, and nanoparticles. Recently, MIP-based SPE columns were introduced to the market. Because of a relatively high price, commercial applications of these MIPs are limited to sample pretreatment in clinical analysis. However, with the increase in the production volume, MIP resins price should drop so much that they will find applications in the food industry. Possibly, first in veterinary, later in food quality control, and, finally, in food products processing.

However, several important issues should be resolved before implementing the MIPs synthesis in the industry. One deficiency is the necessity of using templates to synthesize MIPs. These templates are removed from MIPs within the final steps of their synthesis and then they are being dumped to waists. Because toxic compounds, i.e., heavy metal ions, mycotoxins, pesticides, antibiotics, and hormones, are usually used as templates, producing waists containing these compounds may constitute a severe risk to human health and the environment. This issue may be resolved by the approach proposed by Piletsky’s (Canfarotta et al. Citation2016). and Haupt’s (Ambrosini et al. Citation2013; Xu et al. Citation2016) research groups. Both groups have developed protocols for automated MIP NPs synthesis on solid supports. In this approach, template molecules are immobilized on a glass bead (solid support) surface. Next, MIP NPs are grown around these immobilized templates, and then washed out. Thus, the template stays on the solid support surface and can be used multiple times for MIP NPs synthesis. Moreover, the toxic template does not contaminate synthesized MIP, nor toxic wastes are produced during the synthesis.

The most critical challenge in introducing MIPs to the food processing industry seems to be preventing microplastic contamination. This contamination accompanies any production and application of polymers nowadays. That is, MIPs may be a source of not only microplastic in the environment but also may contaminate processed food products. Microplastic contamination is not only a severe environmental burden but also has significant adverse effects on human health (Rainieri and Barranco Citation2019; De-la-Torre Citation2020; Kwon et al. Citation2020). One solution to this problem is the synthesis of magnetic MIP NPs. Those MIP NPs can be readily removed from samples and collected by simply applying a magnetic field (da Fonseca Alves et al. 2021; Aylaz et al. Citation2021; Siciliano et al. Citation2022). Another approach may be to use biodegradable polymers for MIP synthesis. For example, the imprinted chitosan (Zouaoui et al. Citation2020; Bagheri and Ghaedi Citation2019), cellulose (Wen et al. Citation2022; Cao et al. Citation2021), and crosslinked poly(lactide-co-glycolide) dendrimers (Kumar, Jha, and Panda Citation2019; Gagliardi, Bertero, and Bifone Citation2017) have already been reported. Moreover, dimeric vanillin derivatives were recently applied as monomers for synthesizing a photodegradable polymer (Singathi et al. Citation2022). Importantly, this polymer decomposed in a controlled manner into vanillin dimers upon irradiation with UV light of λ = 300 nm. If it were applied for MIP NPs synthesis, these MIP NPs would be stable until irradiating with the light of the above wavelength. Then, they would decompose to harmless vanillin dimers. Similarly, NPs of polycoumarin were synthesized by irradiation with UV light (Avó, Lima, and Jorge Parola 2019). Importantly, photodimerization of coumarin is a reversible process if it is irradiated with light of a defined wavelength (Wolff and Görner Citation2010). Therefore, it would also be possible to photodegrade such NPs. However, none of these two polymers has yet been applied in MIP synthesis. Moreover, it is possible to synthesize MIPs from materials that are biocompatible and harmless to the environment. Namely, imprinted polydopamine (Palladino, Bettazzi, and Scarano Citation2019; Siciliano et al. Citation2022), polyscopoletin (Bognár et al. Citation2022; Jetzschmann et al. Citation2019; Di Giulio, Mazzotta, and Malitesta Citation2020), and silica (Susanti and Hasanah Citation2021; Susanti, Mutakin, and Hasanah Citation2022) were reported. These materials occur in the natural environment and seem to have no adverse influence on living organisms.

Presumably, most of the above challenges are solvable. Notably, the most significant advantage of MIPs fabrication consists in their versatility. That is, if an MIP-based product, e.g., an SPE cartridge, HPLC column, chemosensor, etc., is implemented into mass production, only limited optimization is needed to implement other analogous products selective for other analytes. Therefore, most implementation costs must be covered for the first product in the manufacturer’s offer. But if this investment pays off, extending the range of products will be less costly. Therefore, we assume that if MIP-based products mentioned in the present article prove to be a commercial success, within the next few years, the offer of MIP-based analytical tools will be extended, and their price will be significantly reduced. That will open the field for the widespread application of these products.

7. Conclusions

Molecularly imprinted polymers (MIPs) exemplify the idea of smart materials. As selective sorbents, MIPs have recently been introduced to the market. With the increasing number of applications, their cost will decrease, and their availability will increase. Due to their unique properties, including high selectivity and durability, MIPs may find numerous applications in food manufacturing, food safety, and food quality control. MIP-based chemosensors fabrication is a constantly growing field. Several examples of robust, very sensitive, and selective MIP chemosensors reported in the literature suggest that the MIP-using technology is sufficiently mature to enter the market even within the current decade. Hand-held devices, especially those based on visual readout with naked-eye, may find interest from both sides, i.e., end-user customers and farmers who would like to control quality of their food products during production. It is easy to envision that in the not-too-distant future, customers will come to a food market equipped with hand-held analytical devices of the size of a mobile phone, capable of testing the quality of food products on the spot by themselves.

Abbreviations
AChE=

Acetylcholinesterase

ATRP=

Atom transfer radical polymerization

bpy=

2,2’-Bipyridine

CLB=

Clenbuterol

DPV=

Differential pulse voltammetry

ECL=

Electrochemiluminescence

EG-FET=

Extended-gate field-effect transistor

EIS=

Electrochemical impedance spectroscopy

ELISA=

Enzyme-linked immunosorbent assay

GCE=

Glassy carbon electrode

GO=

Graphene oxide

GDH=

Glucose dehydrogenase

GOx=

Glucose oxidase

HPLC=

High-performance liquid chromatography

HPLC-MS=

High-performance liquid chromatography with mass spectrometry detection

HRP=

Horseradish peroxidase

ITO=

Indium-tin oxide

LSPR=

Localized surface plasmon resonance

MIP=

Molecularly imprinted polymer

MOF=

Metal-organic framework

MWCNT=

Multi-walled carbon nanotube

NP=

Nanoparticle

PAH=

Polycyclic aromatic hydrocarbon

PDMS=

Polydimethylsiloxane

PGG=

pentagalloyl glucose unit

PVC=

Poly(vinyl chloride)

rGO=

Reduced graphene oxide

QCM=

Quartz crystal microbalance

QCR=

Quartz crystal resonator

QD=

Quantum dot

SAM=

Self-assembled monolayer

SERS=

Surface-enhanced Raman spectroscopy

SPE=

Solid-phase extraction

SPR=

Surface plasmon resonance

SWCNT=

Single-walled carbon nanotube

TLC=

Thin-layer chromatography

UCNPs=

Upconversion nanoparticles

Disclosure statement

We confirm that none of the coauthors has any conflict of interest to be declared.

Funding

The National Science Center of Poland financially supported the present research (Grant SONATA no. 2018/31/D/ST5/02890 to M.C.). Moreover, the present scientific work was partially funded from the financial resources for science in 2017–2021, awarded by the Polish Ministry of Science and Higher Education for implementing an international co-financed project. Furthermore, the present publication is part of a project that has received funding from the European Union’s Horizon 2020 research and innovation program under the Marie Sklodowska-Curie grant agreement No. 711859.

References

  • Ali, M. H., and N. Suleiman. 2018. Eleven shades of food integrity: A halal supply chain perspective. Trends in Food Science & Technology 71:216–24. doi: 10.1016/j.tifs.2017.11.016.
  • Altintas, Z. 2018. Surface plasmon resonance based sensor for the detection of glycopeptide antibiotics in milk using rationally ­designed nanoMIPs. Scientific Reports 8:11222. doi. 10.1038/s41598-018-29585-2.
  • Amatatongchai, M., W. Sroysee, P. Jarujamrus, D. Nacapricha, and P. A. Lieberzeit. 2018. Selective amperometric flow-injection analysis of carbofuran using a molecularly-imprinted polymer and gold-coated-magnetite modified carbon nanotube-paste electrode. Talanta 179:700–9. doi: 10.1016/j.talanta.2017.11.064.
  • Ambrosini, S., S. Beyazit, K. Haupt, and B. T. S. Bui. 2013. Solid-phase synthesis of molecularly imprinted nanoparticles for protein recognition. Chemical Communications (Cambridge, England) 49 (60):6746–8.
  • Anirudhan, T. S., and S. Alexander. 2015. Design and fabrication of molecularly imprinted polymer-based potentiometric sensor from the surface modified multiwalled carbon nanotube for the determination of lindane (gamma-hexachlorocyclohexane), an organochlorine pesticide. Biosensors & Bioelectronics 64:586–93. doi: 10.1016/j.bios.2014.09.074.
  • Antonelli, G., and E. Viganò. 2018. Global challenges in traditional food production and consumption. In Case studies in the traditional food sector, by A. Cavicchi and C. Santini, 25–46. Duxford: Woodhead Publishing. doi. 10.1016/B978-0-08-101007-5.00003-8.
  • Ashley, J., X. T. Feng, and Y. Sun. 2018. A multifunctional molecularly imprinted polymer-based biosensor for direct detection of doxycycline in food samples. Talanta 182:49–54. doi: 10.1016/j.talanta.2018.01.056.
  • Ashley, J., M. A. Shahbazi, K. Kant, V. A. Chidambara, A. Wolff, D. D. Bang, and Y. Sun. 2017. Molecularly imprinted polymers for sample preparation and biosensing in food analysis: Progress and perspectives. Biosensors & Bioelectronics 91:606–15. doi: 10.1016/j.bios.2017.01.018.
  • Ashley, J., Y. Shukor, R. D’Aurelio, L. Trinh, T. L. Rodgers, J. Temblay, M. Pleasants, and I. E. Tothill. 2018. Synthesis of molecularly imprinted polymer nanoparticles for alpha-casein detection using surface plasmon resonance as a milk allergen sensor. ACS Sensors 3 (2):418–24. doi: 10.1021/acssensors.7b00850.
  • Athikomrattanakul, U., N. Gajovic-Eichelmann, and F. W. Scheller. 2011. Thermometric sensing of nitrofurantoin by noncovalently imprinted polymers containing two complementary functional monomers. Analytical Chemistry 83 (20):7704–11. doi: 10.1021/ac201099h.
  • Avó, J., J. C. Lima, and A. Jorge Parola. 2019. Photo-controlled growth of polymeric submicron-sized particles. Photochemical & Photobiological Sciences 18 (5):993–6. doi: 10.1039/c9pp00086k.
  • Ayerdurai, V., M. Cieplak, K. R. Noworyta, M. Gajda, A. Ziminska, M. Sosnowska, J. Piechowska, P. Borowicz, W. Lisowski, S. Shao, et al. 2021. Electrochemical sensor for selective tyramine determination, amplified by a molecularly imprinted polymer film. Bioelectrochemistry 138:107695. doi: 10.1016/j.bioelechem.2020.107695.
  • Ayerdurai, V., A. Garcia-Cruz, J. Piechowska, M. Cieplak, P. Borowicz, K. R. Noworyta, G. Spolnik, W. Danikiewicz, W. Lisowski, A. Pietrzyk-Le, et al. 2021. Selective impedimetric chemosensing of carcinogenic heterocyclic aromatic amine in pork by dsDNA-mimicking molecularly imprinted polymer film-coated electrodes. Journal of Agricultural and Food Chemistry 69 (48):14689–98. doi: 10.1021/acs.jafc.1c05084.
  • Aylaz, G., J. Kuhn, E. C. H. T. Lau, C. Yeung, V. A. L. Roy, M. Duman, and H. H. P. Yiu. 2021. Recent developments on magnetic molecular imprinted polymers (MMIPs) for sensing, capturing, and monitoring pharmaceutical and agricultural pollutants. Journal of Chemical Technology & Biotechnology 96 (5):1151–60. doi: 10.1002/jctb.6681.
  • Bagheri, A. R., and M. Ghaedi. 2019. Synthesis of chitosan based molecularly imprinted polymer for pipette-tip solid phase extraction of Rhodamine B from chili powder samples. International Journal of Biological Macromolecules 139:40–8. doi: 10.1016/j.ijbiomac.2019.07.196.
  • Bagheri, N., A. Khataee, B. Habibi, and J. Hassanzadeh. 2018. Mimetic Ag nanoparticle/Zn-based MOF nanocomposite (AgNPs@ZnMOF) capped with molecularly imprinted polymer for the selective detection of patulin. Talanta 179:710–8. doi: 10.1016/j.talanta.2017.12.009.
  • Bai, J. L., X. Y. Zhang, Y. Peng, X. D. Hong, Y. Y. Liu, S. Y. Jiang, B. A. Ning, and Z. X. Gao. 2017. Ultrasensitive sensing of diethylstilbestrol based on AuNPs/MWCNTs-CS composites coupling with sol-gel molecularly imprinted polymer as a recognition element of an electrochemical sensor. Sensors and Actuators B: Chemical 238:420–6. doi: 10.1016/j.snb.2016.07.035.
  • Beyazit, S., B. T. S. Bui, K. Haupt, and C. Gonzato. 2016. Molecularly imprinted polymer nanomaterials and nanocomposites by controlled/living radical polymerization. Progress in Polymer Science 62:1–21. doi: 10.1016/j.progpolymsci.2016.04.001.
  • Blasco, C., and Y. Pico. 2012. Development of an improved method for trace analysis of quinolones in eggs of laying hens and wildlife species using molecularly imprinted polymers. Journal of Agricultural and Food Chemistry 60 (44):11005–14. doi: 10.1021/jf303222a.
  • Bognár, Z., E. Supala, A. Yarman, X. Zhang, F. F. Bier, F. W. Scheller, and R. E. Gyurcsányi. 2022. Peptide epitope-imprinted polymer microarrays for selective protein recognition. Application for SARS-CoV-2 RBD protein. Chemical Science 13 (5):1263–9. doi: 10.1039/d1sc04502d.
  • Bousoumah, R., J. P. Antignac, V. Camel, M. Grimaldi, P. Balaguer, F. Courant, E. Bichon, M. L. Morvan, and B. L. Bizec. 2015. Development of a molecular recognition based approach for multi-residue extraction of estrogenic endocrine disruptors from biological fluids coupled to liquid chromatography-tandem mass spectrometry measurement. Analytical and Bioanalytical Chemistry 407 (29):8713–23. doi: 10.1007/s00216-015-9024-4.
  • Bryła, M., R. Jędrzejczak, M. Roszko, K. Szymczyk, M. W. Obiedziński, J. Sękul, and M. Rzepkowska. 2013. Application of molecularly imprinted polymers to determine B-1, B-2, and B-3 fumonisins in cereal products. Journal of Separation Science 36 (3):578–84. doi: 10.1002/jssc.201200753.
  • Bustamante-Rangel, M., E. Rodríguez-Gonzalo, and M. M. Delgado-Zamareño. 2022. Evaluation of the selectivity of molecularly imprinted polymer cartridges for nitroimidazoles. Application to the simultaneous extraction of nitroimidazoles and benzimidazoles from samples of animal origin. Microchemical Journal 172:107000. doi: 10.1016/j.microc.2021.107000.
  • Canfarotta, F., A. Poma, A. Guerreiro, and S. Piletsky. 2016. Solid-phase synthesis of molecularly imprinted nanoparticles. Nature Protocols 11 (3):443–55. doi: 10.1038/nprot.2016.030.
  • Cao, J. L., W. J. Kong, S. J. Zhou, L. H. Yin, L. Wan, and M. H. Yang. 2013. Molecularly imprinted polymer-based solid phase clean-up for analysis of ochratoxin A in beer, red wine, and grape juice. Journal of Separation Science 36 (7):1291–7. doi: 10.1002/jssc.201201055.
  • Cao, J. L., S. J. Zhou, W. J. Kong, M. H. Yang, L. Wan, and S. H. Yang. 2013. Molecularly imprinted polymer-based solid phase clean-up for analysis of ochratoxin A in ginger and LC-MS/MS confirmation. Food Control 33 (2):337–43. doi: 10.1016/j.foodcont.2013.03.023.
  • Cao, J., C. Shen, X. Wang, Y. Zhu, S. Bao, X. Wu, and Y. Fu. 2021. A porous cellulose-based molecular imprinted polymer for specific recognition and enrichment of resveratrol. Carbohydrate Polymers 251:117026. doi: 10.1016/j.carbpol.2020.117026.
  • Capoferri, D., M. Del Carlo, N. Ntshongontshi, E. I. Iwuoha, M. Sergi, F. D. Ottavio, and D. Compagnone. 2017. MIP-MEPS based sensing strategy for the selective assay of dimethoate. Application to wheat flour samples. Talanta 174:599–604. doi: 10.1016/j.talanta.2017.06.062.
  • Catana, M., L. Catana, E. Iorga, A. C. Asanica, A. G. Lazar, M. A. Lazar, N. Belc, and G. Pirvu. 2019. Internal validation of rapid and performance method for patulin determination in apple cider by high-performance liquid chromatography. Revista de Chimie 70 (11):3921–5. doi: 10.37358/RC.19.11.7673.
  • Ceylan Cömert, Ş., E. Özgür, L. Uzun, and M. Odabaşı. 2022. The creation of selective imprinted cavities on quartz crystal microbalance electrode for the detection of melamine in milk sample. Food Chemistry 372:131254. doi: 10.1016/j.foodchem.2021.131254.
  • Chen, C., and J. S. Wang. 2020. Optical biosensors: An exhaustive and comprehensive review. The Analyst 145 (5):1605–28. doi: 10.1039/c9an01998g.
  • Chen, C., X. Wang, G. I. N. Waterhouse, X. Qiao, and Z. Xu. 2022. A surface-imprinted surface-enhanced Raman scattering sensor for histamine detection based on dual semiconductors and Ag nanoparticles. Food Chemistry 369:130971. doi: 10.1016/j.foodchem.2021.130971.
  • Chen, S., J. Fu, S. Zhou, P. Zhao, X. Wu, S. Tang, and Z. Zhang. 2022. Rapid recognition of di-n-butyl phthalate in food samples with a near infrared fluorescence imprinted sensor based on zeolite imidazolate framework-67. Food Chemistry 367:130505. doi: 10.1016/j.foodchem.2021.130505.
  • Cieplak, M., and W. Kutner. 2016. Artificial biosensors: How can molecular imprinting mimic biorecognition? Trends in Biotechnology 34 (11):922–41. doi: 10.1016/j.tibtech.2016.05.011.
  • Çimen, D., N. Bereli, and A. Denizli. 2022. Patulin imprinted nanoparticles decorated surface plasmon resonance chips for patulin detection. Photonic Sensors 12 (2):117–29. doi: 10.1007/s13320-021-0638-1.
  • Cirkva, A., I. Malkova, M. Rejtharova, E. Vernerova, A. Hera, and J. Bures. 2019. Residue study of nitroimidazoles depletion in chicken feathers in comparison with some other selected matrixes. Food Additives & Contaminants, Part A, Chemistry, Analysis, Control, Exposure & Risk Assessment 36 (8):1206–17. doi: 10.1080/19440049.2019.1627000.
  • Cornelis, P., S. Givanoudi, D. Yongabi, H. Iken, S. Duwé, O. Deschaume, J. Robbens, P. Dedecker, C. Bartic, M. Wübbenhorst, et al. 2019. Sensitive and specific detection of E. coli using biomimetic receptors in combination with a modified heat-transfer method. Biosensors & Bioelectronics 136:97–105. doi: 10.1016/j.bios.2019.04.026.
  • Cui, Z. M., Z. Y. Li, Y. T. Jin, T. T. Ren, J. A. Chen, X. H. Wang, K. L. Zhong, L. J. Tang, Y. W. Tang, and M. R. Cao. 2020. Novel magnetic fluorescence probe based on carbon quantum dots-doped molecularly imprinted polymer for AHLs signaling molecules sensing in fish juice and milk. Food Chemistry 328:127063. doi: 10.1016/j.foodchem.2020.127063.
  • da Fonseca Alves, R., L. Neres Chagas da Silva, G. M. Neto, I. F. Ierick, T. L. Ferreira and M. D. P. T. Sotomayor. 2021. Magnetic MIPs: Synthesis and applications. In Molecularly imprinted polymers, Methods in Molecular Biology, by A. Martín-Esteban, 85–96. New York: Humana.
  • Dabrowski, M., P. S. Sharma, Z. Iskierko, K. Noworyta, M. Cieplak, W. Lisowski, S. Oborska, A. Kuhn, and W. Kutner. 2016. Early diagnosis of fungal infections using piezomicrogravimetric and electric chemosensors based on polymers molecularly imprinted with D-arabitol. Biosensors & Bioelectronics 79:627–35. doi: 10.1016/j.bios.2015.12.088.
  • Dadkhah, S., E. Ziaei, A. Mehdinia, T. B. Kayyal, and A. Jabbari. 2016. A glassy carbon electrode modified with amino-functionalized graphene oxide and molecularly imprinted polymer for electrochemical sensing of bisphenol A. Microchimica Acta 183 (6):1933–41. doi: 10.1007/s00604-016-1824-5.
  • Dayal, H., W. Y. Ng, X. H. Lin, and S. F. Y. Li. 2019. Development of a hydrophilic molecularly imprinted polymer for the detection of hydrophilic targets using quartz crystal microbalance. Sensors and Actuators B: Chemical 300:127044. doi: 10.1016/j.snb.2019.127044.
  • De-la-Torre, G. E. 2020. Microplastics: An emerging threat to food security and human health. Journal of Food Science and Technology 57 (5):1601–8. doi: 10.1007/s13197-019-04138-1.
  • Deceuninck, Y., E. Bichon, P. Marchand, C. Y. Boquien, A. Legrand, C. Boscher, J. P. Antignac, and B. L. Bizec. 2015. Determination of bisphenol A and related substitutes/analogues in human breast milk using gas chromatography-tandem mass spectrometry. Analytical and Bioanalytical Chemistry 407 (9):2485–97. doi: 10.1007/s00216-015-8469-9.
  • Deng, P. H., Z. F. Xu, and Y. F. Kuang. 2014. Electrochemical determination of bisphenol A in plastic bottled drinking water and canned beverages using a molecularly imprinted chitosan-graphene composite film modified electrode. Food Chemistry 157:490–7. doi: 10.1016/j.foodchem.2014.02.074.
  • Di Giulio, T., E. Mazzotta, and C. Malitesta. 2020. Molecularly imprinted polyscopoletin for the electrochemical detection of the chronic disease marker lysozyme. Biosensors 11 (1):3. doi: 10.3390/bios11010003.
  • Ebarvia, B. S., I. E. Ubando, and F. B. Sevilla. 2015. Biomimetic piezoelectric quartz crystal sensor with chloramphenicol-imprinted polymer sensing layer. Talanta 144:1260–5. doi: 10.1016/j.talanta.2015.08.001.
  • Fang, G. Z., Y. K. Yang, H. D. Zhu, Y. Qi, J. M. Liu, H. L. Liu, and S. Wang. 2017. Development and application of molecularly imprinted quartz crystal microbalance sensor for rapid detection of metolcarb in foods. Sensors and Actuators B: Chemical 251:720–8. doi: 10.1016/j.snb.2017.05.094.
  • Fang, M., L. Zhou, H. Zhang, L. Liu, and Z. Y. Gong. 2019. A molecularly imprinted polymers/carbon dots-grafted paper sensor for 3-monochloropropane-1,2-diol determination. Food Chemistry 274:156–61. doi: 10.1016/j.foodchem.2018.08.133.
  • Feng, J. Y., Y. X. Hu, E. Grant, and X. N. Lu. 2018. Determination of thiabendazole in orange juice using an MISPE-SERS chemosensor. Food Chemistry 239:816–22. doi: 10.1016/j.foodchem.2017.07.014.
  • Feng, S. L., F. Gao, Z. W. Chen, E. Grant, D. D. Kitts, S. Wang, and X. N. Lu. 2013. Determination of alpha-tocopherol in vegetable oils using a molecularly imprinted polymers-surface-enhanced Raman spectroscopic biosensor. Journal of Agricultural and Food Chemistry 61 (44):10467–75. doi: 10.1021/jf4038858.
  • Feng, S. L., Y. X. Hu, L. Y. Ma, and X. N. Lu. 2017. Development of molecularly imprinted polymers-surface-enhanced Raman spectroscopy/colorimetric dual sensor for determination of chlorpyrifos in apple juice. Sensors and Actuators B: Chemical 241:750–7. doi: 10.1016/j.snb.2016.10.131.
  • Futra, D., L. Y. Heng, M. Z. Jaapar, A. Ulianas, K. Saeedfar, and T. L. Ling. 2016. A novel electrochemical sensor for 17 beta-estradiol from molecularly imprinted polymeric microspheres and multi-walled carbon nanotubes grafted with gold nanoparticles. Analytical Methods 8 (6):1381–9. doi: 10.1039/C5AY02796A.
  • Gagliardi, M., A. Bertero, and A. Bifone. 2017. Molecularly imprinted biodegradable nanoparticles. Scientific Reports 7:40046. doi: 10.1038/srep40046.
  • Gao, F., E. Grant, and X. N. Lu. 2015. Determination of histamine in canned tuna by molecularly imprinted polymers-surface enhanced Raman spectroscopy. Analytica Chimica Acta 901:68–75. doi: 10.1016/j.aca.2015.10.025.
  • Gao, F., Y. X. Hu, D. Chen, E. C. Y. Li-Chan, E. Grant, and X. N. Lu. 2015. Determination of Sudan I in paprika powder by molecularly imprinted polymers-thin layer chromatography-surface enhanced Raman spectroscopic biosensor. Talanta 143:344–52. doi: 10.1016/j.talanta.2015.05.003.
  • Gao, L., X. Y. Li, Q. Zhang, J. D. Dai, X. Wei, Z. L. Song, Y. S. Yan, and C. X. Li. 2014. Molecularly imprinted polymer microspheres for optical measurement of ultra trace nonfluorescent cyhalothrin in honey. Food Chemistry 156:1–6. doi: 10.1016/j.foodchem.2013.12.065.
  • Gao, X., W. Y. Cao, M. M. Chen, H. Y. Xiong, X. H. Zhang, and S. F. Wang. 2014. A high sensitivity electrochemical sensor based on Fe3+-ion molecularly imprinted film for the detection of T-2 toxin. Electroanalysis 26 (12):2739–46. doi: 10.1002/elan.201400237.
  • Gonzalez-Salamo, J., B. Socas-Rodriguez, J. Hernandez-Borges, M. D. Afonso, and M. A. Rodriguez-Delgado. 2015. Evaluation of two molecularly imprinted polymers for the solid-phase extraction of natural, synthetic and mycoestrogens from environmental water samples before liquid chromatography with mass spectrometry. Journal of Separation Science 38:2692–9.
  • Guerreiro, J. R. L., N. Teixeira, V. De Freitas, M. G. F. Sales, and D. S. Sutherland. 2017. A saliva molecular imprinted localized surface plasmon resonance biosensor for wine astringency estimation. Food Chemistry 233:457–66. doi: 10.1016/j.foodchem.2017.04.051.
  • Guo, W., F. W. Pi, H. X. Zhang, J. D. Sun, Y. Z. Zhang, and X. L. Sun. 2017. A novel molecularly imprinted electrochemical sensor modified with carbon dots, chitosan, gold nanoparticles for the determination of patulin. Biosensors & Bioelectronics 98:299–304. doi: 10.1016/j.bios.2017.06.036.
  • Hassan, A. H. A., L. Sappia, S. L. Moura, F. H. M. Ali, W. A. Moselhy, M. D. T. Sotomayor, and M. I. Pividori. 2019. Biomimetic magnetic sensor for electrochemical determination of scombrotoxin in fish. Talanta 194:997–1004. doi: 10.1016/j.talanta.2018.10.066.
  • Hrobonova, K., and E. Brokesova. 2020. Comparison of different types of sorbents for extraction of coumarins. Food Chemistry 332:127404.
  • Hu, Y. X., and X. N. Lu. 2016. Rapid detection of melamine in tap water and milk using conjugated "One-Step" molecularly imprinted polymers-surface enhanced Raman spectroscopic sensor. Journal of Food Science 81 (5):N1272–N1280. doi: 10.1111/1750-3841.13283.
  • Hua, M. Z., S. L. Feng, S. Wang, and X. N. Lu. 2018. Rapid detection and quantification of 2,4-dichlorophenoxyacetic acid in milk using molecularly imprinted polymers-surface-enhanced Raman spectroscopy. Food Chemistry 258:254–9. doi: 10.1016/j.foodchem.2018.03.075.
  • Iskierko, Z., P. S. Sharma, K. R. Noworyta, P. Borowicz, M. Cieplak, W. Kutner, and A. M. Bossi. 2019. Selective PQQPFPQQ gluten epitope chemical sensor with a molecularly imprinted polymer recognition unit and an extended-gate field-effect transistor transduction unit. Analytical Chemistry 91 (7):4537–43. doi: 10.1021/acs.analchem.8b05557.
  • Iskierko, Z., P. S. Sharma, D. Prochowicz, K. Fronc, F. D’Souza, D. Toczydłowska, F. Stefaniak, and K. Noworyta. 2016. Molecularly imprinted polymer (MIP) film with improved surface area developed by using metal-organic framework (MOF) for sensitive lipocalin (NGAL) determination. ACS Applied Materials & Interfaces 8 (31):19860–5. doi: 10.1021/acsami.6b05515.
  • Jalili, R., A. Khataee, M. R. Rashidi, and A. Razmjou. 2020. Detection of penicillin G residues in milk based on dual-emission carbon dots and molecularly imprinted polymers. Food Chemistry 314:126172. doi: 10.1016/j.foodchem.2020.126172.
  • Jetzschmann, K. J., S. Tank, G. Jágerszki, R. E. Gyurcsányi, U. Wollenberger, and F. W. Scheller. 2019. Bio‐electrosynthesis of vectorially imprinted polymer nanofilms for Cytochrome P450cam. ChemElectroChem 6 (6):1818–23. doi: 10.1002/celc.201801851.
  • Jiang, S. Y., Y. Peng, B. A. Ning, J. L. Bai, Y. Y. Liu, N. Zhang, and Z. X. Gao. 2015. Surface plasmon resonance sensor based on molecularly imprinted polymer film for detection of histamine. Sensors and Actuators B: Chemical 221:15–21. doi: 10.1016/j.snb.2015.06.058.
  • Jin, X. C., G. Z. Fang, M. F. Pan, Y. K. Yang, X. Y. Bai, and S. Wang. 2018. A molecularly imprinted electrochemiluminescence sensor based on upconversion nanoparticles enhanced by electrodeposited rGO for selective and ultrasensitive detection of clenbuterol. Biosensors & Bioelectronics 102:357–64. doi: 10.1016/j.bios.2017.11.016.
  • Khanmohammadi, A., A. Aghaie, E. Vahedi, A. Qazvini, M. Ghanei, A. Afkhami, A. Hajian, and H. Bagheri. 2020. Electrochemical biosensors for the detection of lung cancer biomarkers: A review. Talanta 206:120251. doi: 10.1016/j.talanta.2019.120251.
  • Kubiak, A., A. Ciric, and M. Biesaga. 2020. Dummy molecularly imprinted polymer (DMIP) as a sorbent for bisphenol S and bisphenol F extraction from food samples. Microchemical Journal 156:104836. doi: 10.1016/j.microc.2020.104836.
  • Kumar, R., D. Jha, and A. K. Panda. 2019. Antimicrobial therapeutics delivery systems based on biodegradable polylactide/polylactide-co-glycolide particles. Environmental Chemistry Letters 17 (3):1237–49. doi: 10.1007/s10311-019-00871-3.
  • Kupai, J., M. Razali, S. Buyuktiryaki, R. Kecili, and G. Szekely. 2017. Long-term stability and reusability of molecularly imprinted polymers. Polymer Chemistry 8 (4):666–73. doi: 10.1039/c6py01853j.
  • Kwon, J.-H., J.-W. Kim, T. D. Pham, A. Tarafdar, S. Hong, S.-H. Chun, S.-H. Lee, D.-Y. Kang, J.-Y. Kim, S.-B. Kim, et al. 2020. Microplastics in food: A review on analytical methods and challenges. International Journal of Environmental Research and Public Health 17 (18):6710. doi: 10.3390/ijerph17186710.
  • Lach, P., M. Cieplak, M. Majewska, K. R. Noworyta, P. S. Sharma, and W. Kutner. 2019. “Gate effect" in p-synephrine electrochemical sensing with a molecularly imprinted polymer and redox probes. Analytical Chemistry 91 (12):7546–53. doi: 10.1021/acs.analchem.8b05512.
  • Lach, P., M. Cieplak, K. R. Noworyta, P. Pieta, W. Lisowski, J. Kalecki, R. Chitta, F. D’Souza, W. Kutner, and P. S. Sharma. 2021. Self-reporting molecularly imprinted polymer with the covalently immobilized ferrocene redox probe for selective electrochemical sensing of p-synephrine. Sensors and Actuators B: Chemical 344:130276. doi: 10.1016/j.snb.2021.130276.
  • Lach, P., P. S. Sharma, K. Golebiewska, M. Cieplak, F. D’Souza, and W. Kutner. 2017. Molecularly imprinted polymer chemosensor for selective determination of an N-nitroso-L-proline food toxin. Chemistry (Weinheim an Der Bergstrasse, Germany) 23 (8):1942–9. doi: 10.1002/chem.201604799.
  • Li, H., and L. Y. Wang. 2013. Highly selective detection of polycyclic aromatic hydrocarbons using multifunctional magnetic-luminescent molecularly imprinted polymers. ACS Applied Materials & Interfaces 5 (21):10502–9. doi: 10.1021/am4020605.
  • Li, J. B., X. J. Wang, H. M. Duan, Y. H. Wang, Y. N. Bu, and C. N. Luo. 2016. Based on magnetic graphene oxide highly sensitive and selective imprinted sensor for determination of sunset yellow. Talanta 147:169–76. doi: 10.1016/j.talanta.2015.09.056.
  • Li, J. H., Z. F. Xu, M. Q. Liu, P. H. Deng, S. P. Tang, J. B. Jiang, H. B. Feng, D. Qian, and L. Z. He. 2017. Ag/N-doped reduced graphene oxide incorporated with molecularly imprinted polymer: An advanced electrochemical sensing platform for salbutamol determination. Biosensors & Bioelectronics 90:210–6. doi: 10.1016/j.bios.2016.11.016.
  • Li, L., Z. Z. Lin, Z. Y. Huang, and A. H. Peng. 2019. Rapid detection of sulfaguanidine in fish by using a photonic crystal molecularly imprinted polymer. Food Chemistry 281:57–62. doi: 10.1016/j.foodchem.2018.12.073.
  • Li, Q., P. Song, and J. Wen. 2019. Melamine and food safety: A 10-year review. Current Opinion in Food Science 30:79–84. doi: 10.1016/j.cofs.2019.05.008.
  • Li, S. H., J. P. Li, J. H. Luo, Z. Xu, and X. H. Ma. 2018. A microfluidic chip containing a molecularly imprinted polymer and a DNA aptamer for voltammetric determination of carbofuran. Microchimica Acta 185 (6):295.
  • Li, S. H., C. H. Liu, G. H. Yin, J. H. Luo, Z. S. Zhang, and Y. X. Xie. 2016. Supramolecular imprinted electrochemical sensor for the neonicotinoid insecticide imidacloprid based on double amplification by Pt-In catalytic nanoparticles and a Bromophenol blue doped molecularly imprinted film. Microchimica Acta 183 (12):3101–9. doi: 10.1007/s00604-016-1962-9.
  • Li, S. H., C. H. Liu, G. H. Yin, Q. Zhang, J. H. Luo, and N. C. Wu. 2017. Aptamer-molecularly imprinted sensor base on electrogenerated chemiluminescence energy transfer for detection of lincomycin. Biosensors & Bioelectronics 91:687–91. doi: 10.1016/j.bios.2017.01.038.
  • Li, S. H., G. H. Yin, Q. Zhang, C. L. Li, J. H. Luo, Z. Xu, and A. L. Qin. 2015. Selective detection of fenaminosulf via a molecularly imprinted fluorescence switch and silver nano-film amplification. Biosensors & Bioelectronics 71:342–7. doi: 10.1016/j.bios.2015.04.066.
  • Li, W., Y. P. Zheng, T. W. Zhang, S. J. Wu, J. Zhang, and J. Fang. 2018. A surface plasmon resonance-based optical fiber probe fabricated with electropolymerized molecular imprinting film for melamine detection. Sensors 18 (3):828. doi: 10.3390/s18030828.
  • Li, X. J., H. F. Jiao, X. Z. Shi, A. L. Sun, X. J. Wang, J. Y. Chai, D. X. Li, and J. Chen. 2018. Development and application of a novel fluorescent nanosensor based on FeSe quantum dots embedded silica molecularly imprinted polymer for the rapid optosensing of cyfluthrin. Biosensors and Bioelectronics 99:268–73. doi: 10.1016/j.bios.2017.07.071.
  • Li, Y. C., Y. Liu, Y. Yang, F. Yu, J. Liu, H. Song, J. Liu, H. Tang, B. C. Ye, and Z. P. Sun. 2015. Novel electrochemical sensing platform based on a molecularly imprinted polymer decorated 3D nanoporous nickel skeleton for ultrasensitive and selective determination of metronidazole. ACS Applied Materials & Interfaces 7 (28):15474–80. doi: 10.1021/acsami.5b03755.
  • Lian, W. J., S. Liu, J. H. Yu, J. Li, M. Cui, W. Xu, and J. D. Huang. 2013. Electrochemical sensor using neomycin-imprinted film as recognition element based on chitosan-silver nanoparticles/graphene-multiwalled carbon nanotubes composites modified electrode. Biosensors & Bioelectronics 44:70–6. doi: 10.1016/j.bios.2013.01.002.
  • Lin, X. H., S. X. L. Aik, J. Angkasa, Q. H. Le, K. S. Chooi, and S. F. Y. Li. 2018. Selective and sensitive sensors based on molecularly imprinted poly(vinylidene fluoride) for determination of pesticides and chemical threat agent simulants. Sensors and Actuators B: Chemical 258:228–37. doi: 10.1016/j.snb.2017.11.070.
  • Liu, J., Y. Xu, S. Liu, S. Yu, Z. Yu, and S. S. Low. 2022. Application and progress of chemometrics in voltammetric biosensing. Biosensors 12 (7):494. doi: 10.3390/bios12070494.
  • Liu, N., X. L. Li, X. H. Ma, G. R. Ou, and Z. X. Gao. 2014. Rapid and multiple detections of staphylococcal enterotoxins by two-dimensional molecularly imprinted film-coated QCM sensor. Sensors and Actuators B: Chemical 191:326–31. doi: 10.1016/j.snb.2013.09.086.
  • Liu, X. Y., J. Ren, L. H. Su, X. Gao, Y. W. Tang, T. Ma, L. J. Zhu, and J. R. Li. 2017. Novel hybrid probe based on double recognition of aptamer-molecularly imprinted polymer grafted on upconversion nanoparticles for enrofloxacin sensing. Biosensors & Bioelectronics 87:203–8. doi: 10.1016/j.bios.2016.08.051.
  • Liu, Z. Y., Y. Zhang, J. H. Feng, Q. Z. Han, and Q. Wei. 2019. Ni(OH)(2) nanoarrays based molecularly imprinted polymer electrochemical sensor for sensitive detection of sulfapyridine. Sensors and Actuators B: Chemical 287:551–6. doi: 10.1016/j.snb.2019.02.079.
  • Lu, C., Z. Tang, C. Liu, and X. Ma. 2018. Surface molecularly imprinted polymers prepared by two-step precipitation polymerization for the selective extraction of oleanolic acid from grape pomace extract. Journal of Separation Science 41 (17):3496–502. doi: 10.1002/jssc.201800474.
  • Lucci, P., D. Derrien, F. Alix, C. Perollier, and S. Bayoudh. 2010. Molecularly imprinted polymer solid-phase extraction for detection of zearalenone in cereal sample extracts. Analytica Chimica Acta 672 (1–2):15–9. doi: 10.1016/j.aca.2010.03.010.
  • Matejicek, D., A. Grycova, and J. Vlcek. 2013. The use of molecularly imprinted polymers for the multicomponent determination of endocrine-disrupting compounds in water and sediment. Journal of Separation Science 36:1097–103.
  • Mattsson, L., J. J. Xu, C. Preininger, B. T. S. Bui, and K. Haupt. 2018. Competitive fluorescent pseudo-immunoassay exploiting molecularly imprinted polymers for the detection of biogenic amines in fish matrix. Talanta 181:190–6. doi: 10.1016/j.talanta.2018.01.010.
  • Mier, A., S. Nestora, P. X. M. Rangel, Y. Rossez, K. Haupt, and B. T. S. Bui. 2019. Cytocompatibility of molecularly imprinted polymers for deodorants: Evaluation on human keratinocytes and axillary-hosted bacteria. ACS Applied Bio Materials 2 (8):3439–47. doi: 10.1021/acsabm.9b00388.
  • Munawar, H., A. Garcia-Cruz, M. Majewska, K. Karim, W. Kutner, and S. A. Piletsky. 2020. Electrochemical determination of fumonisin B1 using a chemosensor with a recognition unit comprising molecularly imprinted polymer nanoparticles. Sensors and Actuators B: Chemical 321:128552. doi: 10.1016/j.snb.2020.128552.
  • Naresh, V., and N. Lee. 2021. A review on biosensors and recent development of nanostructured materials-enabled biosensors. Sensors 21 (4):1109. doi: 10.3390/s21041109.
  • Nelis, J. L. D., A. S. Tsagkaris, M. J. Dillon, J. Hajslova, and C. T. Elliott. 2020. Smartphone-based optical assays in the food safety field. Trends in Analytical Chemistry: TRAC 129:115934. doi: 10.1016/j.trac.2020.115934.
  • Nestora, S., F. Merlier, S. Beyazit, E. Prost, L. Duma, B. Baril, A. Greaves, K. Haupt, and B. T. S. Bui. 2016. Plastic antibodies for cosmetics: Molecularly imprinted polymers scavenge precursors of malodors. Angewandte Chemie (International ed. in English) 55 (21):6252–6. doi: 10.1002/anie.201602076.
  • Nicolucci, C., S. Rossi, C. Menale, E. M. del Giudice, L. Perrone, P. Gallo, D. G. Mita, and N. Diano. 2013. A high selective and sensitive liquid chromatography-tandem mass spectrometry method for quantization of BPA urinary levels in children. Analytical and Bioanalytical Chemistry 405 (28):9139–48.
  • Niu, M. C., P. H. Chuong, and H. He. 2016. Core-shell nanoparticles coated with molecularly imprinted polymers: A review. Microchimica Acta 183 (10):2677–95. doi: 10.1007/s00604-016-1930-4.
  • Pacheco, J. G., M. Castro, S. Machado, M. F. Barroso, H. P. A. Nouws, and C. Delerue-Matos. 2015. Molecularly imprinted electrochemical sensor for ochratoxin A detection in food samples. Sensors and Actuators B: Chemical 215:107–12. doi: 10.1016/j.snb.2015.03.046.
  • Pakade, V., E. Cukrowska, S. Lindahl, C. Turner, and L. Chimuka. 2013. Molecular imprinted polymer for solid-phase extraction of flavonol aglycones from Moringa oleifera extracts. Journal of Separation Science 36 (3):548–55. doi: 10.1002/jssc.201200576.
  • Palladino, P., F. Bettazzi, and S. Scarano. 2019. Polydopamine: Surface coating, molecular imprinting, and electrochemistry—successful applications and future perspectives in (bio)analysis. Analytical and Bioanalytical Chemistry 411 (19):4327–38. doi: 10.1007/s00216-019-01665-w.
  • Petcu, M. 2013. Polymer and method of use. WO 2013190506, Jun 21.
  • Petcu, M. 2015. Polymer and method of use. US 20150361203, Dec 17.
  • Pieta, P., I. Obraztsov, J. W. Sobczak, O. Chernyayeva, S. K. Das, F. D’Souza, and W. Kutner. 2013. A versatile material for a symmetrical electric energy storage device: A composite of the polymer of the ferrocene adduct of C-60 and single-wall carbon nanotubes exhibiting redox conductivity at both positive and negative potentials. The Journal of Physical Chemistry C 117 (5):1995–2007. doi: 10.1021/jp210450y.
  • Pietrzyk, A., W. Kutner, R. Chitta, M. E. Zandler, F. D’Souza, F. Sannicolo, and P. R. Mussini. 2009. Melamine acoustic chemosensor based on molecularly imprinted polymer film. Analytical Chemistry 81 (24):10061–70. doi: 10.1021/ac9020352.
  • Qi, P., J. Wang, Z. Wang, X. Wang, X. Wang, X. Xu, H. Xu, S. Di, H. Zhang, Q. Wang, et al. 2018. Construction of a probe-immobilized molecularly imprinted electrochemical sensor with dual signal amplification of thiol graphene and gold nanoparticles for selective detection of tebuconazole in vegetable and fruit samples. Electrochimica Acta 274:406–14. doi: 10.1016/j.electacta.2018.04.128.
  • Qin, R. Z., Q. L. Wang, C. Q. Ren, X. Z. Dai, and H. J. Han. 2017. Development of a molecularly imprinted electrochemical sensor for tert-butylhydroquinone recognition. International Journal of Electrochemical Science 12:8953–62. doi: 10.20964/2017.10.04.
  • Qiu, X. Z., W. M. Chen, Y. L. Luo, Y. L. Wang, Y. L. Wang, and H. S. Guo. 2020. Highly sensitive alpha-amanitin sensor based on molecularly imprinted photonic crystals. Analytica Chimica Acta 1093:142–9. doi: 10.1016/j.aca.2019.09.066.
  • Rainieri, S., and A. Barranco. 2019. Microplastics, a food safety issue? Trends in Food Science and Technology 84:55–7. doi: 10.1016/j.tifs.2018.12.009.
  • Roszko, M., K. Szymczyk, and R. Jędrzejczak. 2015. Simultaneous separation of chlorinated/brominated dioxins, polychlorinated biphenyls, polybrominated diphenyl ethers and their methoxylated derivatives from hydroxylated analogues on molecularly imprinted polymers prior to gas/liquid chromatography and mass spectrometry. Talanta 144:171–83. doi: 10.1016/j.talanta.2015.04.070.
  • Sa-nguanprang, S., A. Phuruangrat, and O. Bunkoed. 2022. 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. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 264:120324. doi: 10.1016/j.saa.2021.120324.
  • Santos, W. D. R., M. Santhiago, I. V. P. Yoshida, and L. T. Kubota. 2012. Electrochemical sensor based on imprinted sol-gel and nanomaterial for determination of caffeine. Sensors and Actuators B: Chemical 166–167:739–45. doi: 10.1016/j.snb.2012.03.051.
  • Shamsipur, M., N. Moradi, and A. Pashabadi. 2018. Coupled electrochemical-chemical procedure used in construction of molecularly imprinted polymer-based electrode: A highly sensitive impedimetric melamine sensor. Journal of Solid State Electrochemistry 22 (1):169–80. doi: 10.1007/s10008-017-3731-z.
  • Shang, L., F. Q. Zhao, and B. Z. Zeng. 2014. 3D Porous graphene-porous PdCu alloy nanoparticles-molecularly imprinted poly(para-aminobenzoic acid) composite for the electrocatalytic assay of melamine. ACS Applied Materials & Interfaces 6 (21):18721–7. doi: 10.1021/am504276g.
  • Sharma, P. S., M. Dabrowski, F. D’Souza, and W. Kutner. 2013. Surface development of molecularly imprinted polymer films to enhance sensing signals. TrAC: Trends in Analytical Chemistry 51:146–57. doi: 10.1016/j.trac.2013.07.006.
  • Sharma, P. S., A. Garcia-Cruz, M. Cieplak, K. R. Noworyta, and W. Kutner. 2019. Gate effect’ in molecularly imprinted polymers: The current state of understanding. Current Opinion in Electrochemistry 16:50–6. doi: 10.1016/j.coelec.2019.04.020.
  • Sharma, P. S., A. Pietrzyk-Le, F. D’Souza, and W. Kutner. 2012. Electrochemically synthesized polymers in molecular imprinting for chemical sensing. Analytical and Bioanalytical Chemistry 402 (10):3177–204. doi: 10.1007/s00216-011-5696-6.
  • Shi, X. J., J. X. Lu, H. Z. Yin, X. G. Qiao, and Z. X. Xu. 2017. A biomimetic sensor with signal enhancement of ferriferrous oxide-reduced graphene oxide nanocomposites for ultratrace levels quantification of methamidophos or omethoate in vegetables. Food Analytical Methods 10 (4):910–20. doi: 10.1007/s12161-016-0641-0.
  • Shirani, M. P., B. Rezaei, A. A. Ensafi, and M. Ramezani. 2021. Development of an eco-friendly fluorescence nanosensor based on molecularly imprinted polymer on silica-carbon quantum dot for the rapid indoxacarb detection. Food Chemistry 339:127920. doi: 10.1016/j.foodchem.2020.127920.
  • Shirzadmehr, A., A. Afkhami, and T. Madrakian. 2015. A new nano-composite potentiometric sensor containing an Hg2+-ion imprinted polymer for the trace determination of mercury ions in different matrices. Journal of Molecular Liquids 204:227–35. doi: 10.1016/j.molliq.2015.01.014.
  • Siciliano, G., A. G. Monteduro, A. Turco, E. Primiceri, S. Rizzato, N. Depalo, M. L. Curri, and G. Maruccio. 2022. Polydopamine-coated magnetic iron oxide nanoparticles: From design to applications. Nanomaterials 12 (7):1145. doi: 10.3390/nano12071145.
  • Silva, P., J. Freitas, F. M. Nunes, and J. S. Camara. 2021. A predictive strategy based on volatile profile and chemometric analysis for traceability and authenticity of sugarcane honey on the global market. Foods 10 (7):1559. doi: 10.3390/foods10071559.
  • Silva, P., C. L. Silva, R. Perestrelo, F. M. Nunes, and J. S. Camara. 2018. Fingerprint targeted compounds in authenticity of sugarcane honey - An approach based on chromatographic and statistical data. LWT 96:82–9. doi: 10.1016/j.lwt.2018.04.076.
  • Singathi, R., R. Raghunathan, R. Krishnan, S. Kumar Rajendran, S. Baburaj, Mukund, P. Sibi, D. C. Webster, and J. Sivaguru. 2022. Towards upcycling biomass‐derived crosslinked polymers with light. Angewandte Chemie International Edition 22:e202203353.
  • Song, B., Y. S. Zhou, H. Jin, T. Jing, T. T. Zhou, Q. L. Hao, Y. K. Zhou, S. R. Mei, and Y. I. Lee. 2014. Selective and sensitive determination of erythromycin in honey and dairy products by molecularly imprinted polymers based electrochemical sensor. Microchemical Journal 116:183–90. doi: 10.1016/j.microc.2014.05.010.
  • Soon, J. M., S. C. Krzyzaniak, Z. Shuttlewood, M. Smith, and L. Jack. 2019. Food fraud vulnerability assessment tools used in food industry. Food Control 101:225–32. doi: 10.1016/j.foodcont.2019.03.002.
  • Sun, A., J. Chai, T. Xiao, X. Shi, X. Li, Q. Zhao, D. Li, and J. Chen. 2018. Development of a selective fluorescence nanosensor based on molecularly imprinted-quantum dot optosensing materials for saxitoxin detection in shellfish samples. Sensors and Actuators B: Chemical 258:408–14. doi: 10.1016/j.snb.2017.11.143.
  • Sun, S., M. Q. Zhang, Y. J. Li, and X. W. He. 2013. A molecularly imprinted polymer with incorporated graphene oxide for electrochemical determination of quercetin. Sensors (Basel, Switzerland) 13 (5):5493–506. doi: 10.3390/s130505493.
  • Susanti, I., and A. N. Hasanah. 2021. How to develop molecularly imprinted mesoporous silica for selective recognition of analytes in pharmaceutical, environmental, and food samples. Polymers for Advanced Technologies 32 (5):1965–80. doi: 10.1002/pat.5251.
  • Susanti, I., M. Mutakin, and A. N. Hasanah. 2022. Factors affecting the analytical performance of molecularly imprinted mesoporous silica. Polymers for Advanced Technologies 33 (2):469–83. doi: 10.1002/pat.5545.
  • Tarannum, N., S. Khatoon, and B. B. Dzantiev. 2020. Perspective and application of molecular imprinting approach for antibiotic detection in food and environmental samples: A critical review. Food Control 118:107381. doi: 10.1016/j.foodcont.2020.107381.
  • Turco, A., S. Corvaglia, and E. Mazzotta. 2015. Electrochemical sensor for sulfadimethoxine based on molecularly imprinted polypyrrole: Study of imprinting parameters. Biosensors & Bioelectronics 63:240–7. doi: 10.1016/j.bios.2014.07.045.
  • Turco, A., S. Corvaglia, E. Mazzotta, P. P. Pompa, and C. Malitesta. 2018. Preparation and characterization of molecularly imprinted mussel inspired film as antifouling and selective layer for electrochemical detection of sulfamethoxazole. Sensors and Actuators B: Chemical 255:3374–83. doi: 10.1016/j.snb.2017.09.164.
  • Turiel, E., and A. Martin-Esteban. 2010. Molecularly imprinted polymers for sample preparation: A review. Analytica Chimica Acta 668 (2):87–99.
  • Uzun, L., and A. P. F. Turner. 2016. Molecularly-imprinted polymer sensors: Realising their potential. Biosensors & Bioelectronics 76:131–44. doi: 10.1016/j.bios.2015.07.013.
  • Valero-Navarro, A., M. Gomez-Romero, J. F. Fernandez-Sanchez, P. A. G. Cormack, A. Segura-Carretero, and A. Fernandez-Gutierrez. 2011. Synthesis of caffeic acid molecularly imprinted polymer microspheres and high-performance liquid chromatography evaluation of their sorption properties. Journal of Chromatography A 1218 (41):7289–96. doi: 10.1016/j.chroma.2011.08.043.
  • Wackerlig, J., and P. A. Lieberzeit. 2015. Molecularly imprinted polymer nanoparticles in chemical sensing - Synthesis, characterisation and application. Sensors and Actuators B: Chemical 207:144–57. doi: 10.1016/j.snb.2014.09.094.
  • Wang, D., S. Jiang, Y. Liang, X. Wang, X. Zhuang, C. Tian, F. Luan, and L. Chen. 2022. Selective detection of enrofloxacin in biological and environmental samples using a molecularly imprinted electrochemiluminescence sensor based on functionalized copper nanoclusters. Talanta 236:122835. doi: 10.1016/j.talanta.2021.122835.
  • Wang, H. W., S. Yao, Y. Q. Liu, S. L. Wei, J. W. Su, and G. X. Hu. 2017. Molecularly imprinted electrochemical sensor based on Au nanoparticles in carboxylated multi-walled carbon nanotubes for sensitive determination of olaquindox in food and feedstuffs. Biosensors & Bioelectronics 87:417–21. doi: 10.1016/j.bios.2016.08.092.
  • Wang, J., J. Y. Li, C. Zeng, Q. Qu, M. F. Wang, W. Qi, R. X. Su, and Z. M. He. 2020. Sandwich-like sensor for the highly specific and reproducible detection of Rhodamine 6G on a surface-enhanced Raman scattering platform. ACS Applied Materials & Interfaces 12 (4):4699–706. doi: 10.1021/acsami.9b16773.
  • Wang, Q. H., G. Z. Fang, Y. Y. Liu, D. D. Zhang, J. M. Liu, and S. Wang. 2017. Fluorescent sensing probe for the sensitive detection of histamine based on molecular imprinting ionic liquid-modified quantum dots. Food Analytical Methods 10 (7):2585–92. doi: 10.1007/s12161-017-0795-4.
  • Wen, Z., D. Gao, J. Lin, S. Li, K. Zhang, Z. Xia, and D. Wang. 2022. Magnetic porous cellulose surface-imprinted polymers synthetized with assistance of deep eutectic solvent for specific recognition and purification of bisphenols. International Journal of Biological Macromolecules 216:374–87. doi: 10.1016/j.ijbiomac.2022.06.187.
  • Wijayaratne, S. P., M. Reid, K. Westberg, A. Worsley, and F. Mavondo. 2018. Food literacy, healthy eating barriers and household diet. European Journal of Marketing 52 (12):2449–77. doi: 10.1108/EJM-10-2017-0760.
  • Wolff, T., and H. Görner. 2010. Photocleavage of dimers of coumarin and 6-alkylcoumarins. Journal of Photochemistry and Photobiology A: Chemistry 209 (2–3):219–23. doi: 10.1016/j.jphotochem.2009.11.018.
  • Wu, L., Z. Z. Lin, H. P. Zhong, X. M. Chen, and Z. Y. Huang. 2017. Rapid determination of malachite green in water and fish using a fluorescent probe based on CdTe quantum dots coated with molecularly imprinted polymer. Sensors and Actuators B: Chemical 239:69–75. doi: 10.1016/j.snb.2016.07.166.
  • Wu, M., H. Y. Deng, Y. J. Fan, Y. C. Hu, Y. P. Guo, and L. W. Xie. 2018. Rapid colorimetric detection of cartap residues by AgNP sensor with magnetic molecularly imprinted microspheres as recognition elements. Molecules 23 (6):1443. doi: 10.3390/molecules23061443.
  • Wu, W. H., L. T. Yang, F. Q. Zhao, and B. Z. Zeng. 2017. A vanillin electrochemical sensor based on molecularly imprinted poly(1-vinyl-3-octylimidazole hexafluoride phosphorus)-multi-walled carbon nanotubes@polydopamine-carboxyl single-walled carbon nanotubes composite. Sensors and Actuators B: Chemical 239:481–7. doi: 10.1016/j.snb.2016.08.041.
  • Wu, W. Z., M. X. Huang, Q. D. Huang, C. H. Lyu, J. P. Lai, and H. Sun. 2019. Molecularly imprinted photonic hydrogels for visual detection of methylanthranilate in wine. Chinese Journal of Analytical Chemistry 47 (9):1330–6. doi: 10.1016/S1872-2040(19)61188-6.
  • Wu, Z. Z., E. B. Xu, J. P. Li, J. Long, A. Q. Jiao, and Z. Y. Jin. 2016. Highly sensitive determination of ethyl carbamate in alcoholic beverages by surface-enhanced Raman spectroscopy combined with a molecular imprinting polymer. RSC Advances 6 (111):109442–52. doi: 10.1039/C6RA23165A.
  • Xiao, L., Z. Zhang, C. C. Wu, L. Y. Han, and H. Y. Zhang. 2017. Molecularly imprinted polymer grafted paper-based method for the detection of 17 beta-estradiol. Food Chemistry 221:82–6. doi: 10.1016/j.foodchem.2016.10.062.
  • Xie, Y., M. Zhao, Q. Hu, Y. Cheng, Y. Guo, H. Qian, and W. Yao. 2017. Selective detection of chloramphenicol in milk based on a molecularly imprinted polymer-surface-enhanced Raman spectroscopic nanosensor. Journal of Raman Spectroscopy 48 (2):204–10. doi: 10.1002/jrs.5034.
  • Xu, G., Y. Chi, L. Li, S. Liu, and X. Kan. 2015. Imprinted propyl gallate electrochemical sensor based on graphene/single walled carbon nanotubes/sol-gel film. Food Chemistry 177:37–42. doi: 10.1016/j.foodchem.2014.12.097.
  • Xu, J. J., S. Ambrosini, E. Tamahkar, C. Rossi, K. Haupt, and B. T. S. Bui. 2016. Toward a universal method for preparing molecularly imprinted polymer nanoparticles with antibody-like affinity for proteins. Biomacromolecules 17 (1):345–53. doi: 10.1021/acs.biomac.5b01454.
  • Xu, S., G. Y. Lin, W. Zhao, Q. Wu, J. Luo, W. Wei, X. Y. Liu, and Y. Zhu. 2018. Necklace-like molecularly imprinted nanohybrids based on polymeric nanoparticles decorated multiwalled carbon nanotubes for highly sensitive and selective melamine detection. ACS Applied Materials & Interfaces 10 (29):24850–9. doi: 10.1021/acsami.8b08558.
  • Yang, G. M., and F. Q. Zhao. 2015. Electrochemical sensor for dimetridazole based on novel gold nanoparticles@molecularly imprinted polymer. Sensors and Actuators B: Chemical 220:1017–22. doi: 10.1016/j.snb.2015.06.051.
  • Yang, L. T., B. J. Xu, H. L. Ye, F. Q. Zhao, and B. Z. Zeng. 2017. A novel quercetin electrochemical sensor based on molecularly imprinted poly(para-aminobenzoic acid) on 3D Pd nanoparticles-porous graphene-carbon nanotubes composite. Sensors and Actuators B: Chemical 251:601–8. doi: 10.1016/j.snb.2017.04.006.
  • Yang, L. T., F. Q. Zhao, and B. Z. Zeng. 2016. Electrochemical determination of eugenol using a three-dimensional molecularly imprinted poly (p-aminothiophenol-co-p-aminobenzoic acids) film modified electrode. Electrochimica Acta 210:293–300. doi: 10.1016/j.electacta.2016.05.167.
  • Yang, Q., H. L. Peng, J. H. Li, Y. B. Li, H. Xiong, and L. X. Chen. 2017. Label-free colorimetric detection of tetracycline using analyte-responsive inverse-opal hydrogels based on molecular imprinting technology. New Journal of Chemistry 41 (18):10174–80. doi: 10.1039/C7NJ02368E.
  • Yao, G. H., R. P. Liang, C. F. Huang, Y. Wang, and J. D. Qiu. 2013. Surface plasmon resonance sensor based on magnetic molecularly imprinted polymers amplification for pesticide recognition. Analytical Chemistry 85 (24):11944–51. doi: 10.1021/ac402848x.
  • Yao, T., X. Gu, T. Li, J. Li, J. Li, Z. Zhao, J. Wang, Y. Qin, and Y. She. 2016. Enhancement of surface plasmon resonance signals using a MIP/GNPs/rGO nano-hybrid film for the rapid detection of ractopamine. Biosensors & Bioelectronics 75:96–100. doi: 10.1016/j.bios.2015.08.027.
  • Ye, T., W. Yin, N. Zhu, M. Yuan, H. Cao, J. Yu, Z. Gou, X. Wang, H. Zhu, A. Reyihanguli, et al. 2018. Colorimetric detection of pyrethroid metabolite by using surface molecularly imprinted polymer. Sensors and Actuators B: Chemical 254:417–23. doi: 10.1016/j.snb.2017.07.132.
  • Yin, W. M., L. Wu, F. Ding, Q. Li, P. Wang, J. J. Li, Z. C. Lu, and H. Y. Han. 2018. Surface-imprinted SiO2@Ag nanoparticles for the selective detection of BPA using surface enhanced Raman scattering. Sensors and Actuators B: Chemical 258:566–73. doi: 10.1016/j.snb.2017.11.141.
  • Yin, Z.-Z., S.-W. Cheng, L.-B. Xu, H.-Y. Liu, K. Huang, L. Li, Y.-Y. Zhai, Y.-B. Zeng, H.-Q. Liu, Y. Shao, et al. 2018. Highly sensitive and selective sensor for sunset yellow based on molecularly imprinted polydopamine-coated multi-walled carbon nanotubes. Biosensors & Bioelectronics 100:565–70. doi: 10.1016/j.bios.2017.10.010.
  • You, A. M., Y. H. Cao, and G. Q. Cao. 2016. Colorimetric sensing of melamine using colloidal magnetically assembled molecularly imprinted photonic crystals. RSC Advances 6 (87):83663–7. doi: 10.1039/C6RA18617C.
  • You, A. M., X. J. Ni, Y. H. Cao, and G. Q. Cao. 2017. Colorimetric chemosensor for chloramphenicol based on colloidal magnetically assembled molecularly imprinted photonic crystals. Journal of the Chinese Chemical Society 64 (10):1235–41. doi: 10.1002/jccs.201700126.
  • Zahara, S., M. A. Minhas, H. Shaikh, M. S. Ali, M. I. Bhanger, and M. I. Malik. 2021. Molecular imprinting-based extraction of rosmarinic acid from Salvia hypoleuca extract. Reactive and Functional Polymers 166:104984. doi: 10.1016/j.reactfunctpolym.2021.104984.
  • Zhang, L. L., C. C. Zhu, C. B. Chen, S. H. Zhu, J. Zhou, M. L. Wang, and P. P. Shang. 2018. Determination of kanamycin using a molecularly imprinted SPR sensor. Food Chemistry 266:170–4. doi: 10.1016/j.foodchem.2018.05.128.
  • Zhang, M., H. T. Zhao, T. J. Xie, X. Yang, A. J. Dong, H. Zhang, J. Wang, and Z. Y. Wang. 2017. Molecularly imprinted polymer on graphene surface for selective and sensitive electrochemical sensing imidacloprid. Sensors and Actuators B: Chemical 252:991–1002. doi: 10.1016/j.snb.2017.04.159.
  • Zhang, R.-R., X.-T. Gan, J.-J. Xu, Q.-F. Pan, H. Liu, A.-L. Sun, X.-Z. Shi, and Z.-M. Zhang. 2022. Ultrasensitive electrochemiluminescence sensor based on perovskite quantum dots coated with molecularly imprinted polymer for prometryn determination. Food Chemistry 370:131353. doi: 10.1016/j.foodchem.2021.131353.
  • Zhang, Z. H., R. Cai, F. Long, and J. Wang. 2015. Development and application of tetrabromobisphenol A imprinted electrochemical sensor based on graphene/carbon nanotubes three-dimensional nanocomposites modified carbon electrode. Talanta 134:435–42. doi: 10.1016/j.talanta.2014.11.040.
  • Zhao, B. W., S. L. Feng, Y. X. Hu, S. Wang, and X. N. Lu. 2019. Rapid determination of atrazine in apple juice using molecularly imprinted polymers coupled with gold nanoparticles-colorimetric/SERS dual chemosensor. Food Chemistry 276:366–75. doi: 10.1016/j.foodchem.2018.10.036.
  • Zhao, P. N., H. Y. Liu, L. N. Zhang, P. H. Zhu, S. G. Ge, and J. H. Yu. 2020. Paper-based SERS sensing platform based on 3D silver dendrites and molecularly imprinted identifier sandwich hybrid for neonicotinoid quantification. ACS Applied Materials & Interfaces 12 (7):8845–54. doi: 10.1021/acsami.9b20341.
  • Zhao, X. L., Y. He, Y. N. Wang, S. Wang, and J. P. Wang. 2020. Hollow molecularly imprinted polymer based quartz crystal microbalance sensor for rapid detection of methimazole in food samples. Food Chemistry 309:125787. doi: 10.1016/j.foodchem.2019.125787.
  • Zhao, Y., L. Tian, X. Zhang, Z. Sun, X. Shan, Q. Wu, R. Chen, and J. Lu. 2022. A novel molecularly imprinted polymer electrochemiluminescence sensor based on Fe2O3@Ru(bpy)32+ for determination of clenbuterol. Sensors and Actuators B: Chemical 350:130822. doi: 10.1016/j.snb.2021.130822.
  • Zhou, J. Y., S. Sheth, H. F. Zhou, and Q. J. Song. 2020. Highly selective detection of L-Phenylalanine by molecularly imprinted polymers coated Au nanoparticles via surface-enhanced Raman scattering. Talanta 211:120745. doi: 10.1016/j.talanta.2020.120745.
  • Zouaoui, F., S. Bourouina-Bacha, M. Bourouina, N. Jaffrezic-Renault, N. Zine, and A. Errachid. 2020. Electrochemical sensors based on molecularly imprinted chitosan: A review. TrAC: Trends in Analytical Chemistry 130:115982. doi: 10.1016/j.trac.2020.115982.
  • Zoughi, S., F. Faridbod, A. Amiri, and M. R. Ganjali. 2021. Detection of tartrazine in fake saffron containing products by a sensitive optical nanosensor. Food Chemistry 350:129197. doi: 10.1016/j.foodchem.2021.129197.