Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (770)

Search Parameters:
Keywords = molecularly imprinted polymers

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 4361 KiB  
Article
Curcumin-Based Molecularly Imprinted Polymer Electropolymerized on Single-Use Graphite Electrode for Dipyridamole Analysis
by Daniel Preda, Gabriel Lucian Radu, Emilia-Elena Iorgulescu, Mihaela-Carmen Cheregi and Iulia Gabriela David
Molecules 2024, 29(19), 4630; https://doi.org/10.3390/molecules29194630 - 29 Sep 2024
Viewed by 221
Abstract
A new molecularly imprinted polymer (MIP)-based disposable electrochemical sensor for dipyridamole (DIP) determination was obtained. The sensor was rapidly prepared by potentiodynamic electrochemical polymerization on a pencil graphite electrode (PGE) using curcumin (CUR) as a functional monomer and DIP as a template molecule. [...] Read more.
A new molecularly imprinted polymer (MIP)-based disposable electrochemical sensor for dipyridamole (DIP) determination was obtained. The sensor was rapidly prepared by potentiodynamic electrochemical polymerization on a pencil graphite electrode (PGE) using curcumin (CUR) as a functional monomer and DIP as a template molecule. After the optimization of the conditions (pH, monomer–template ratio, scan rate, number of cyclic voltammetric cycles applied in the electro-polymerization process and extraction time of the template molecule) for MIP formation, DIP voltammetric behavior at the modified electrode (MIP_PGE) was investigated. DIP oxidation took place in a pH-dependent, irreversible mixed diffusion-adsorption controlled process. Differential pulse voltammetry (DPV) and adsorptive stripping differential pulse voltammetry (AdSDPV) were used to quantify DIP from pharmaceutical and tap water samples. Under optimized conditions (Britton–Robinson buffer at pH = 3.29), the obtained linear ranges were 5.00 × 10−8–1.00 × 10−5 mol/L and 5.00 × 10−9–1.00 × 10−7 mol/L DIP for DPV and AdSDPV, respectively. The limits of detection of the methods were 1.47 × 10−8 mol/L for DPV and 3.96 × 10−9 mol/L DIP for AdSDPV. Full article
(This article belongs to the Section Analytical Chemistry)
Show Figures

Figure 1

Figure 1
<p>Structural formula of dipyridamole (DIP).</p>
Full article ">Figure 2
<p>CV curves recorded at PGE for (<b>a</b>) 5.00 × 10<sup>−4</sup> mol/L CUR and (<b>b</b>) mixture of 5.00 × 10<sup>−4</sup> mol/L CUR and 2.50 × 10<sup>−5</sup> mol/L DIP in 0.2 mol/L NaOH solution, scan rate 0.100 V/s. Inset: expanded section of the potential window containing the CUR and DIP signals.</p>
Full article ">Figure 3
<p>Impedance spectra recorded and fitted for all the tested working electrodes. To obtain the Nyquist plot, 1.00 × 10<sup>−3</sup> mol/L [Fe(CN)<sub>6</sub>]<sup>4−</sup>/[Fe(CN<sub>6</sub>)]<sup>3−</sup> in acetate buffer solution with pH = 4.50 was used (DC potential of 0.230 V and frequency in the range 0.1 Hz–10.0 kHz). The equivalent circuit employed for the fitting curve is presented schematically above.</p>
Full article ">Figure 4
<p>(<b>a</b>) DP voltammograms recorded at MIP_PGE for 5.00 × 10<sup>−5</sup> mol/L DIP in BRB solutions with different pH values and (<b>b</b>) the dependencies of DIP oxidation peak potential (E<sub>p</sub>)/current (I<sub>p</sub>) recorded by DPV at MIP_PGE on the pH of the supporting electrolyte.</p>
Full article ">Figure 5
<p>(<b>a</b>) CV curves recorded at different scan rates at MIP_PGE for 1.50 × 10<sup>−4</sup> mol/L DIP in BRB solution with pH = 3.29; and the dependencies (<b>b</b>) I<sub>p</sub> = f(v), (<b>c</b>) I<sub>p</sub> = f (v<sup>1/2</sup>) and (<b>d</b>) log I<sub>p</sub> = f (log v).</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) CV curves recorded at different scan rates at MIP_PGE for 1.50 × 10<sup>−4</sup> mol/L DIP in BRB solution with pH = 3.29; and the dependencies (<b>b</b>) I<sub>p</sub> = f(v), (<b>c</b>) I<sub>p</sub> = f (v<sup>1/2</sup>) and (<b>d</b>) log I<sub>p</sub> = f (log v).</p>
Full article ">Figure 6
<p>DP voltammograms recorded at MIP_PGE for BRB pH 3.29 solutions containing different DIP concentrations between (<b>a</b>) 5.00 × 10<sup>−8</sup>–5.00 × 10<sup>−6</sup> and (<b>b</b>) 1.00 × 10<sup>−5</sup>–1.00 × 10<sup>−4</sup> mol/L.</p>
Full article ">Figure 7
<p>AdSDP voltammograms recorded at MIP_PGE for different DIP concentrations in BRB solution with pH = 3.29; t<sub>acc</sub> 30 s; E<sub>acc</sub> −0.400 V.</p>
Full article ">Figure 8
<p>The variation of the anodic peak current recorded at MIP_PGE for 1.00 × 10<sup>−5</sup> mol/L DIP in BRB solution with pH = 3.29 at different periods of time, including 0, 24, 48 and 72 h after the sensor preparation.</p>
Full article ">Figure 9
<p>DPV peak currents recorded at MIP_PGE for 2.00 × 10<sup>−6</sup> mol/L DIP in BRB solution with pH = 3.29, without and with a 50-fold excess of different possible interfering species.</p>
Full article ">Figure 10
<p>(<b>a</b>) DP voltammograms for 10 mL DIPIRIDAMOL tablets working solution in BRB pH = 3.29, recorded at MIP_PGE. The initial sample and the 3 × 0.025 mL addition of 1.00 × 10<sup>−3</sup> mol/L DIP is also presented; (<b>b</b>) the dependence of the DIP oxidation signal on the C<sub>add</sub> DIP.</p>
Full article ">Scheme 1
<p>Possible mechanism for DIP electrooxidation at MIP_PGE in BRB solution with pH = 3.29.</p>
Full article ">Scheme 2
<p>Schematic representation of the steps involved in the preparation of the MIP_PGE.</p>
Full article ">
13 pages, 2609 KiB  
Article
Preparation and Utilization of a Highly Discriminative Absorbent Imprinted with Fetal Hemoglobin
by Ka Zhang, Tongchang Zhou, Cedric Dicko, Lei Ye and Leif Bülow
Polymers 2024, 16(19), 2734; https://doi.org/10.3390/polym16192734 - 27 Sep 2024
Viewed by 300
Abstract
Development in hemoglobin-based oxygen carriers (HBOCs) that may be used as alternatives to donated blood requires an extensive supply of highly pure hemoglobin (Hb) preparations. Therefore, it is essential to fabricate inexpensive, stable and highly selective absorbents for Hb purification. Molecular imprinting is [...] Read more.
Development in hemoglobin-based oxygen carriers (HBOCs) that may be used as alternatives to donated blood requires an extensive supply of highly pure hemoglobin (Hb) preparations. Therefore, it is essential to fabricate inexpensive, stable and highly selective absorbents for Hb purification. Molecular imprinting is an attractive technology for preparing such materials for targeted molecular recognition and rapid separations. In this case study, we developed human fetal hemoglobin (HbF)-imprinted polymer beads through the fusion of surface imprinting and Pickering emulsion polymerization. HbF was firstly covalently coupled to silica nanoparticles through its surface-exposed amino groups. The particle-supported HbF molecules were subsequently employed as templates for the synthesis of molecularly imprinted polymers (MIPs) with high selectivity for Hb. After removing the silica support and HbF, the resulting MIPs underwent equilibrium and kinetic binding experiments with both adult Hb (HbA) and HbF. These surface-imprinted MIPs exhibited excellent selectivity for both HbA and HbF, facilitating the one-step isolation of recombinant Hb from crude biological samples. The saturation capacities of HbA and HbF were found to be 15.4 and 17.1 mg/g polymer, respectively. The present study opens new possibilities for designed resins for tailored protein purification, separation and analysis. Full article
(This article belongs to the Section Polymer Applications)
Show Figures

Figure 1

Figure 1
<p>FTIR spectra of silica nanoparticles with different functional groups.</p>
Full article ">Figure 2
<p>Alignment of the β and γ chains, respectively, using the CLUSTAL program [<a href="#B53-polymers-16-02734" class="html-bibr">53</a>]. The identity between the two chains is 73.5%.</p>
Full article ">Figure 3
<p>Analysis of the obtained kinetic binding data using a pseudo-second-order kinetic model. The inset shows the kinetics of HbA and HbF binding with 5 mg of MIP particles. The initial Hb concentration was 6 μM.</p>
Full article ">Figure 4
<p>Analysis of the equilibrium adsorption data using the Langmuir model. The inset shows the equilibrium HbA and HbF binding with 5 mg of MIP particles. The incubation time was fixed at 10 min for equilibrium adsorption.</p>
Full article ">Figure 5
<p>Displacement of GFP-HbF from 5 mg of MIP particles by HbA and HbF. Bound and Bound<sub>0</sub> are the amount of the bound GFP-HbF measured in the presence and absence of competing HbA/HbF, respectively.</p>
Full article ">Figure 6
<p>SDS-PAGE of the protein solutions collected from the binding test. Lane M: molecular weight markers; Lane HbA-1/HbF-1: solution of thawed recombinant HbA/HbF before adding the MIPs; Lane HbA-2/HbF-2: solution of unbound recombinant HbA/HbF after depletion with the MIPs; Lane HbA-3/HbF-3: solution of recombinant HbA/HbF eluted from the MIPs.</p>
Full article ">Figure 7
<p>Elution profiles of recombinant HbA and HbF in <span class="html-italic">E. coli</span> crude cell extracts from the MIP column using a linear gradient of buffer A (20 mM sodium phosphate buffer, pH 6.0) mixed with buffer B (20 mM sodium phosphate buffer, pH 8.0). The inset is the SDS-PAGE of the peak fractions. Lane M: molecular weight marker; Lane HbA/HbF: the HbA/HbF crude extracts; Lane P1/P2: the elution peaks of HbA/HbF from the MIP column.</p>
Full article ">Scheme 1
<p>Schematic illustration of the MIP particle preparation using the Pickering emulsion technique. HbF-immobilized silica particles were used to stabilize the Pickering emulsion. Following the polymerization, the solid beads were collected and subjected to washing. The morphology of the obtained MIP was also checked by SEM.</p>
Full article ">
54 pages, 3730 KiB  
Review
Hazardous Materials from Threats to Safety: Molecularly Imprinted Polymers as Versatile Safeguarding Platforms
by Ana-Mihaela Gavrila, Aurel Diacon, Tanta-Verona Iordache, Traian Rotariu, Mariana Ionita and Gabriela Toader
Polymers 2024, 16(19), 2699; https://doi.org/10.3390/polym16192699 - 24 Sep 2024
Viewed by 667
Abstract
Hazards associated with highly dangerous pollutants/contaminants in water, air, and land resources, as well as food, are serious threats to public health and the environment. Thus, it is imperative to detect or decontaminate, as risk-control strategies, the possible harmful substances sensitively and efficiently. [...] Read more.
Hazards associated with highly dangerous pollutants/contaminants in water, air, and land resources, as well as food, are serious threats to public health and the environment. Thus, it is imperative to detect or decontaminate, as risk-control strategies, the possible harmful substances sensitively and efficiently. In this context, due to their capacity to be specifically designed for various types of hazardous compounds, the synthesis and use of molecularly imprinted polymers (MIPs) have become widespread. By molecular imprinting, affinity sites with complementary shape, size, and functionality can be created for any template molecule. MIPs' unique functions in response to external factors have attracted researchers to develop a broad range of MIP-based sensors with increased sensitivity, specificity, and selectivity of the recognition element toward target hazardous compounds. Therefore, this paper comprehensively reviews the very recent progress of MIPs and smart polymer applications for sensing or decontamination of hazardous compounds (e.g., drugs, explosives, and biological or chemical agents) in various fields from 2020 to 2024, providing researchers with a rapid tool for investigating the latest research status. Full article
(This article belongs to the Section Polymer Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of the synthesis of an MIP by covalent, semi-covalent, and non-covalent bonds (adapted from [<a href="#B13-polymers-16-02699" class="html-bibr">13</a>]).</p>
Full article ">Figure 2
<p>Classification of the main CWAs based on their chemical structure.</p>
Full article ">Figure 3
<p>Classification of explosives based on their chemical structure.</p>
Full article ">Figure 3 Cont.
<p>Classification of explosives based on their chemical structure.</p>
Full article ">Figure 4
<p>MIP structure and interaction for picric acid detection, adapted from Huynh et al. [<a href="#B110-polymers-16-02699" class="html-bibr">110</a>].</p>
Full article ">Figure 5
<p>Classification chart of the main illicit drugs, depending upon the drug’s major effects and origin, as well the existent Drug Scheduling on the most hazardous (Schedule I and II, according to US Drug Enforcement Administration).</p>
Full article ">Figure 6
<p>MIP-based cell recognition strategies.</p>
Full article ">Figure 7
<p>Graphic illustration of MIP preparation, <span class="html-italic">Ps. aeruginosa</span> recognition, and photothermal inactivation—adapted from [<a href="#B254-polymers-16-02699" class="html-bibr">254</a>].</p>
Full article ">
21 pages, 4395 KiB  
Review
Developments and Applications of Molecularly Imprinted Polymer-Based In-Tube Solid Phase Microextraction Technique for Efficient Sample Preparation
by Hiroyuki Kataoka, Atsushi Ishizaki, Keita Saito and Kentaro Ehara
Molecules 2024, 29(18), 4472; https://doi.org/10.3390/molecules29184472 - 20 Sep 2024
Viewed by 576
Abstract
Despite advancements in the sensitivity and performance of analytical instruments, sample preparation remains a bottleneck in the analytical process. Currently, solid-phase extraction is more widely used than traditional organic solvent extraction due to its ease of use and lower solvent requirements. Moreover, various [...] Read more.
Despite advancements in the sensitivity and performance of analytical instruments, sample preparation remains a bottleneck in the analytical process. Currently, solid-phase extraction is more widely used than traditional organic solvent extraction due to its ease of use and lower solvent requirements. Moreover, various microextraction techniques such as micro solid-phase extraction, dispersive micro solid-phase extraction, solid-phase microextraction, stir bar sorptive extraction, liquid-phase microextraction, and magnetic bead extraction have been developed to minimize sample size, reduce solvent usage, and enable automation. Among these, in-tube solid-phase microextraction (IT-SPME) using capillaries as extraction devices has gained attention as an advanced “green extraction technique” that combines miniaturization, on-line automation, and reduced solvent consumption. Capillary tubes in IT-SPME are categorized into configurations: inner-wall-coated, particle-packed, fiber-packed, and rod monolith, operating either in a draw/eject system or a flow-through system. Additionally, the developments of novel adsorbents such as monoliths, ionic liquids, restricted-access materials, molecularly imprinted polymers (MIPs), graphene, carbon nanotubes, inorganic nanoparticles, and organometallic frameworks have improved extraction efficiency and selectivity. MIPs, in particular, are stable, custom-made polymers with molecular recognition capabilities formed during synthesis, making them exceptional “smart adsorbents” for selective sample preparation. The MIP fabrication process involves three main stages: pre-arrangement for recognition capability, polymerization, and template removal. After forming the template-monomer complex, polymerization creates a polymer network where the template molecules are anchored, and the final step involves removing the template to produce an MIP with cavities complementary to the template molecules. This review is the first paper to focus on advanced MIP-based IT-SPME, which integrates the selectivity of MIPs into efficient IT-SPME, and summarizes its recent developments and applications. Full article
(This article belongs to the Special Issue Applications of Solid-Phase Microextraction and Related Techniques)
Show Figures

Figure 1

Figure 1
<p>Two operating systems of automated online IT-SPME coupled with HPLC. (<b>A</b>,<b>C</b>) are the steps of extracting compounds from the sample solution into the capillary stationary phase, and (<b>B</b>,<b>D</b>) are the steps of desorbing the compounds extracted into the capillary. The green and blue lines indicate the flow of the sample solution and the mobile phase, respectively. Reproduced from Figure 5 of Ref. [<a href="#B3-molecules-29-04472" class="html-bibr">3</a>] with permission from Elsevier.</p>
Full article ">Figure 2
<p>Configurations of capillary tubes for IT-SPME. (<b>A</b>) Inner-surface-coated capillary, (<b>B</b>) particle-packed capillary, (<b>C</b>) fiber-packed capillary, (<b>D</b>) monolithic capillary. Reproduced from Figure 4 of Ref. [<a href="#B3-molecules-29-04472" class="html-bibr">3</a>] with permission from Elsevier.</p>
Full article ">Figure 3
<p>Fabrication process of molecularly imprinted polymer.</p>
Full article ">Figure 4
<p>Chromatograms of steroid hormones obtained by HPLC-UV. (A) Direct injection, (B) IT-SPME using host capillary, (C) IT-SPME using NMIP, (D) IT-SPME using MIP. HPLC conditions: column, Eclipse SDB-C8 (150 × 4.6 mm ID, 5 μm particle size, Agilent Technologies, Santa Clara, CA, USA); gradient elution, acetonitrile/H<sub>2</sub>O (45/55) 1 mL min<sup>−1</sup> → acetonitrile/H<sub>2</sub>O (65/35) 1.8 mL min<sup>−1</sup> (8 min); column temperature, 40 °C; detection, UV at 200 and 245 nm. Peaks: 1 = estriol, 2 = β-estradiol, 3 = ethynylestradiol, 4 = diethylesilbestrol, 5 = corticosterone, 6 = testosterone, 7 = estrone, 8 = progesterone.</p>
Full article ">Figure 5
<p>Open-tubular MIP-capillary preparation. (<b>a</b>) Both the tips of the glass-capillary were coned with flame to the diameter size of the desired metal rod. (<b>b</b>) The prepared assembly was placed in a bigger capillary that contained polymer mixture. (<b>c</b>) After the polymerization, the metal rod was removed from the polymer. (<b>d</b>) Magnified cross section of the polymer tube inside the 20 μL capillary glass. Reproduced from Figures 1 and 3 of Ref. [<a href="#B88-molecules-29-04472" class="html-bibr">88</a>] with permission from Elsevier.</p>
Full article ">Figure 6
<p>MIP fibers and fiber-packed tubes that recognize fluoroquinolones. (<b>a</b>) The chemical structure of fluoroquinolones and the schematic diagram of the resultant MIP structure. (<b>b</b>) Micrograph of the multiple-fiber-packed tube. (<b>c</b>,<b>d</b>) SEM images of the ofloxacin MIP fiber. Reproduced from Figures 1–3 of Ref. [<a href="#B94-molecules-29-04472" class="html-bibr">94</a>] with permission from Elsevier.</p>
Full article ">Figure 7
<p>Schematic representation of (<b>A</b>) one-pot synthesis of protein-imprinted hybrid monolithic column and (<b>B</b>) the recognition mechanism between template protein and functional monomers. Reproduced from Figure S1 of Ref. [<a href="#B98-molecules-29-04472" class="html-bibr">98</a>] with permission from Elsevier.</p>
Full article ">Figure 8
<p>SEM images of the Lyz-MIP (<b>a</b>–<b>c</b>) and NIP (<b>d</b>–<b>f</b>) hybrid monolithic columns. (<b>a</b>,<b>d</b>) 3000×, (<b>b</b>,<b>e</b>) 10,000×, (<b>c</b>,<b>f</b>) 30,000×. Reproduced from Figure 1 of Ref. [<a href="#B98-molecules-29-04472" class="html-bibr">98</a>] with permission from Elsevier.</p>
Full article ">
15 pages, 2784 KiB  
Article
Green Synthesis of Molecularly Imprinted Polymers for Selective Extraction of Protocatechuic Acid from Mango Juice
by Liping Zhang, Xin Song, Yuxiao Dong and Xiyan Zhao
Foods 2024, 13(18), 2955; https://doi.org/10.3390/foods13182955 - 18 Sep 2024
Viewed by 444
Abstract
A novel and environmentally friendly molecularly imprinted polymer (PCA-MIP) was successfully synthesized in an aqueous solution for the selective extraction of protocatechuic acid (PCA). In this study, a deep eutectic solvent (DES, choline chloride/methacrylic acid, 1:2, mol/mol) and chitosan were employed as the [...] Read more.
A novel and environmentally friendly molecularly imprinted polymer (PCA-MIP) was successfully synthesized in an aqueous solution for the selective extraction of protocatechuic acid (PCA). In this study, a deep eutectic solvent (DES, choline chloride/methacrylic acid, 1:2, mol/mol) and chitosan were employed as the eco-friendly functional monomers. These two components interacted with PCA through hydrogen bonding, integrating a multitude of recognition sites within the PCA-MIP. Thus, the resulting PCA-MIP exhibited outstanding adsorption performance, rapid adsorption rate, and better selectivity, with a maximum binding capacity of 30.56 mg/g and an equilibrium time of 30 min. The scanning electron microscope (SEM) and Brunauer–Emmett–Teller (BET) analyses revealed that the synthesized polymers possessed a uniform morphology and substantial surface areas, which were conducive to their adsorption properties. Moreover, the PCA-MIP integrated with HPLC demonstrated its efficacy as an adsorbent for the selective extraction of PCA from mango juice. The PCA-MIP presented itself as an exemplary adsorbent, offering a highly effective and eco-friendly method for the enrichment of PCA from complex matrices. Full article
(This article belongs to the Topic Advances in Analysis of Food and Beverages)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the procedure for the preparation of PCA-MIP.</p>
Full article ">Figure 2
<p>Effect of (<b>a</b>) content of DES and (<b>b</b>) volume of Glu on absorption capacity (Q) and imprinting factor (IF) of PCA-MIP and PCA-NIP. The pink bar represents PCA-MIP, the blue bar represents PCA-NIP, and the purple line is the imprinting factor.</p>
Full article ">Figure 3
<p>The SEM images of PCA-MIP (<b>a</b>,<b>a1</b>) and PCA-NIP (<b>b</b>,<b>b1</b>) with magnifications of 10,000× and 30,000×.</p>
Full article ">Figure 4
<p>(<b>a</b>) FTIR spectra and (<b>b</b>) nitrogen adsorption–desorption isotherms of PCA-MIP and PCA-NIP. Inset in picture b are the pore parameters of PCA-MIP and PCA-NIP.</p>
Full article ">Figure 5
<p>Effect of pH of incubation solution on absorption capacity (Q) and imprinting factor (IF) of PCA-MIP and PCA-NIP. The pink bar represents PCA-MIP, the blue bar represents PCA-NIP, and the purple line is the imprinting factor.</p>
Full article ">Figure 6
<p>(<b>a</b>) Equilibrium adsorption curves, (<b>b</b>) linear fitting curves of Freundlich model, (<b>c</b>) adsorption kinetics curves, and (<b>d</b>) linear fitting curves of a pseudo-second-order kinetic model for PCA-MIP and PCA-NIP.</p>
Full article ">Figure 7
<p>(<b>a</b>) Selective adsorption capacities of PCA-MIP and PCA-NIP for PCA and its analogs. (<b>b</b>) Reusability analysis of PCA-MIP via five sequential cycles of adsorption–desorption.</p>
Full article ">Figure 8
<p>Chromatograms of (<b>a</b>) mango juice crude extract, (<b>b</b>) supernatant of mango juice crude extract from PCA-MIP, (<b>c</b>) elution of mango juice crude extract from DYM-MIP, and (<b>d</b>) PCA standard.</p>
Full article ">
17 pages, 3509 KiB  
Article
Assessment of Acrylamide Levels by Advanced Molecularly Imprinted Polymer-Imprinted Surface Plasmon Resonance (SPR) Sensor Technology and Sensory Quality in Homemade Fried Potatoes
by Betül Karslıoğlu, Bahar Bankoğlu Yola, İlknur Polat, Harun Yiğit Alkan and Mehmet Lütfi Yola
Foods 2024, 13(18), 2927; https://doi.org/10.3390/foods13182927 - 15 Sep 2024
Viewed by 514
Abstract
This study evaluated acrylamide (AA) levels and various quality parameters in homemade fried potatoes prepared in different sizes by integrating principles from the Slow Food Movement with advanced sensor technology. To this aim, a surface plasmon resonance (SPR) sensor based on a molecularly [...] Read more.
This study evaluated acrylamide (AA) levels and various quality parameters in homemade fried potatoes prepared in different sizes by integrating principles from the Slow Food Movement with advanced sensor technology. To this aim, a surface plasmon resonance (SPR) sensor based on a molecularly imprinted polymer (MIP) was first developed for the determination of AA in homemade fried potatoes at low levels, and the AA levels in the samples were established. First of all, monolayer formation of allyl mercaptane on the SPR chip surface was carried out to form double bonds that could polymerize on the chip surface. AA-imprinted SPR chip surfaces modified with allyl mercaptane were prepared via UV polymerization using ethylene glycol dimethacrylate (EGDMA) as a cross-linker, N,N′-azobisisobutyronitrile (AIBN) as an initiator, and methacryloylamidoglutamicacid (MAGA) as a monomer. The prepared AA-imprinted and nonimprinted surfaces were characterized by atomic force microscopy (AFM) and Fourier transform infrared (FTIR) spectroscopy methods. The SPR sensor indicated linearity in the range of 1.0 × 10−9–5.0 × 10−8 M with a detection limit (LOD) of 3.0 × 10−10 M in homemade fried potatoes, and the SPR sensor demonstrated high selectivity and repeatability in terms of AA detection. Additionally, the highest AA level was observed in the potato sample belonging to the T1 group, at 15.37 nM (p < 0.05), and a strong and positive correlation was found between AA levels and sensory parameters, the a* value, the ΔE value, and the browning index (BI) (p < 0.05). Full article
(This article belongs to the Section Food Toxicology)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) FTIR spectra of MIP/SPR chip; AFM images of (<b>B</b>) bare SPR chip and (<b>C</b>) MIP/SPR chip.</p>
Full article ">Figure 2
<p>(<b>A</b>) SPR sensorgrams for 10.0 nM AA at different pHs of PBS. (<b>B</b>) Effect of pH on MIP/SPR chip: (a) adsorption; (b) desorption; (c) regeneration.</p>
Full article ">Figure 3
<p>Effect of AA concentration on MIP/SPR chip. Inset: Calibration curve of AA concentrations of MIP/SPR chip in the presence of pH 6.0 of PBS (from 1.0 nM to 50.0 nM AA): (a) adsorption; (b) desorption; (c) regeneration.</p>
Full article ">Figure 4
<p>Selectivity tests: SPR sensorgrams of (<b>A</b>) MIP/SPR chip and (<b>B</b>) NIP/SPR chip in 10.0 nM AA, 1000.0 nM MA, 1000.0 nM PA, 1000.0 nM DL-ALA, 1000.0 nM L-ASP: (a) adsorption; (b) desorption; (c) regeneration.</p>
Full article ">Figure 5
<p>Repeatability of MIP/SPR chip in 10.0 nM AA. (a) adsorption; (b) desorption; (c) regeneration.</p>
Full article ">Figure 6
<p>Pearson correlation analysis results. AA: Acrylamide, Overall acceptability: Overall acceptability, Ph: pH.</p>
Full article ">
13 pages, 2521 KiB  
Article
Sensitive Coatings Based on Molecular-Imprinted Polymers for Triazine Pesticides’ Detection
by Usman Latif, Sadaf Yaqub and Franz L. Dickert
Sensors 2024, 24(18), 5934; https://doi.org/10.3390/s24185934 - 13 Sep 2024
Viewed by 361
Abstract
Triazine pesticide (atrazine and its derivatives) detection sensors have been developed to thoroughly check for the presence of these chemicals and ultimately prevent their exposure to humans. Sensitive coatings were designed by utilizing molecular imprinting technology, which aims to create artificial receptors for [...] Read more.
Triazine pesticide (atrazine and its derivatives) detection sensors have been developed to thoroughly check for the presence of these chemicals and ultimately prevent their exposure to humans. Sensitive coatings were designed by utilizing molecular imprinting technology, which aims to create artificial receptors for the detection of chlorotriazine pesticides with gravimetric transducers. Initially, imprinted polymers were developed, using acrylate and methacrylate monomers containing hydrophilic and hydrophobic side chains, specifically for atrazine, which shares a basic heterocyclic triazine structure with its structural analogs. By adjusting the ratio of the acid to the cross-linker and introducing acrylate ester as a copolymer, optimal non-covalent interactions were achieved with the hydrophobic core of triazine molecules and their amino groups. A maximum sensor response of 546 Hz (frequency shift/layer height equal to 87.36) was observed for a sensitive coating composed of 46% methacrylic acid and 54% ethylene glycol dimethacrylate, with a demonstrated layer height of 250 nm (6.25 kHz). The molecularly imprinted copolymer demonstrated fully reversible sensor responses, not only for atrazine but also for its metabolites, like des-ethyl atrazine, and structural analogs, such as propazine and terbuthylazine. The efficiency of modified molecularly imprinted polymers for targeted analytes was tested by combining them with a universally applicable quartz crystal microbalance transducer. The stable selectivity pattern of the developed sensor provides an excellent basis for a pattern recognition procedure. Full article
(This article belongs to the Section Chemical Sensors)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of triazine pesticides: (<b>a</b>) ATR, (<b>b</b>) PRO, (<b>c</b>) TBA, and (<b>d</b>) DEA.</p>
Full article ">Figure 2
<p>ATR-IR spectra of atrazine imprinted and reference polymers before and after washing with methanol.</p>
Full article ">Figure 3
<p>AFM of pesticide-imprinted acrylate coating. Methacrylic acid (MAA) as the monomer, ethylene glycol dimethacrylate (EGDMA) as the cross-linker, and atrazine as the template were used to prepare the atrazine-imprinted polymer.</p>
Full article ">Figure 4
<p>Sensor response of atrazine-imprinted polymer layer to different concentrations of the templated analyte.</p>
Full article ">Figure 5
<p>Sensor response of PRO-imprinted coating against different concentrations of propazine ranging from 0.35 mg/L to 7 mg/L.</p>
Full article ">Figure 6
<p>Sensor responses of terbuthylazine-imprinted coatings to different concentrations of TBA ranging from 0.35 mg/L to 7 mg/L.</p>
Full article ">Figure 7
<p>Sensor responses of DEA-imprinted coatings against different concentrations of DEA ranging from 0.35 mg/L to 7 mg/L.</p>
Full article ">Figure 8
<p>Cross-sensitivity responses of pesticide-imprinted coatings of PRO, TBA, ATR, and DEA against their templated analytes and interfering species. Frequency shifts were normalized for layer thickness, and frequency shifts of reference electrodes were subtracted in each case.</p>
Full article ">
12 pages, 8166 KiB  
Article
Paper-Based Fluorescent Sensor for Rapid Multi-Channel Detection of Tetracycline Based on Graphene Quantum Dots Coated with Molecularly Imprinted Polymer
by Linzhe Wang, Jingfang Hu, Wensong Wei, Yu Song, Yansheng Li, Guowei Gao, Chunhui Zhang and Fangting Fu
Polymers 2024, 16(17), 2540; https://doi.org/10.3390/polym16172540 - 8 Sep 2024
Viewed by 517
Abstract
In this paper, we developed a paper-based fluorescent sensor using functional composite materials composed of graphene quantum dots (GQDs) coated with molecularly imprinted polymers (MIPs) for the selective detection of tetracycline (TC) in water. GQDs, as eco-friendly fluorophores, were chemically grafted onto the [...] Read more.
In this paper, we developed a paper-based fluorescent sensor using functional composite materials composed of graphene quantum dots (GQDs) coated with molecularly imprinted polymers (MIPs) for the selective detection of tetracycline (TC) in water. GQDs, as eco-friendly fluorophores, were chemically grafted onto the surface of paper fibers. MIPs, serving as the recognition element, were then wrapped around the GQDs via precipitation polymerization using 3-aminopropyltriethoxysilane (APTES) as the functional monomer. Optimal parameters such as quantum dot concentration, grafting time, and elution time were examined to assess the sensor’s detection performance. The results revealed that the sensor exhibited a linear response to TC concentrations in the range of 1 to 40 µmol/L, with a limit of detection (LOD) of 0.87 µmol/L. When applied to spiked detection in actual water samples, recoveries ranged from 103.3% to 109.4%. Overall, this paper-based fluorescent sensor (MIPs@GQDs@PAD) shows great potential for portable, multi-channel, and rapid detection of TC in water samples in the future. Full article
(This article belongs to the Special Issue Functional Graphene–Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>A schematic illustration of the fabrication process of the MIPs@GQDs@PAD.</p>
Full article ">Figure 2
<p>A schematic illustration of the detection process of the MIPs@GQDs@PAD for TC.</p>
Full article ">Figure 3
<p>Fluorescence spectra of (<b>A</b>) MIPs@GQD emission spectra in the presence and absence of TC; (<b>B</b>) excitation and emission spectra of MIPs@GQDs@PAD.</p>
Full article ">Figure 4
<p>Fluorescence characterization of (<b>A</b>) MIPs@GQDs@PAD after elution and (<b>B</b>) MIPs@GQDs@PAD after readsorption of TC.</p>
Full article ">Figure 5
<p>SEM characterization of (<b>A</b>) bare paper, (<b>B</b>) GQDs@PAD, and (<b>C</b>) MIPs@GQDs@PAD.</p>
Full article ">Figure 6
<p>FT-IR spectra of bare paper, GQDs@PAD, and MIPs@GQDs@PAD before and after elution.</p>
Full article ">Figure 7
<p>Effect of (<b>A</b>) GQD grafting time and (<b>B</b>) amount of GQDs.</p>
Full article ">Figure 8
<p>Effect of (<b>A</b>) pH and (<b>B</b>) amount of functional monomer.</p>
Full article ">Figure 9
<p>Effect of (<b>A</b>) different eluents and (<b>B</b>) number of elution.</p>
Full article ">Figure 10
<p>The (<b>A</b>) fluorescence emission spectra of MIPs@GQDs@PAD under different TC concentrations and the (<b>B</b>) calibration curve and linear equation.</p>
Full article ">Figure 11
<p>The (<b>A</b>) reproducibility and (<b>B</b>) selectivity of MIPs@GQDs@PAD.</p>
Full article ">
19 pages, 3986 KiB  
Article
Molecularly Imprinted Polypyrrole-Modified Screen-Printed Electrode for Dopamine Determination
by Daniele Merli, Alessandra Cutaia, Ines Hallulli, Alessandra Bonanni and Giancarla Alberti
Polymers 2024, 16(17), 2528; https://doi.org/10.3390/polym16172528 - 6 Sep 2024
Viewed by 729
Abstract
This paper introduces a quantitative method for dopamine determination. The method is based on a molecularly imprinted polypyrrole (e-MIP)-modified screen-printed electrode, with differential pulse voltammetry (DPV) as the chosen measurement technique. The dopamine molecules are efficiently entrapped in the polymeric film, creating recognition [...] Read more.
This paper introduces a quantitative method for dopamine determination. The method is based on a molecularly imprinted polypyrrole (e-MIP)-modified screen-printed electrode, with differential pulse voltammetry (DPV) as the chosen measurement technique. The dopamine molecules are efficiently entrapped in the polymeric film, creating recognition cavities. A comparison with bare and non-imprinted polypyrrole-modified electrodes clearly demonstrates the superior sensitivity, selectivity, and reproducibility of the e-MIP-based one; indeed, a sensitivity of 0.078 µA µM−1, a detection limit (LOD) of 0.8 µM, a linear range between 0.8 and 45 µM and a dynamic range of up to 350 µM are achieved. The method was successfully tested on fortified synthetic and human urine samples to underline its applicability as a screening method for biomedical tests. Full article
Show Figures

Figure 1

Figure 1
<p>A 3-level FCCD layout. (Reproduced with permission from [<a href="#B72-polymers-16-02528" class="html-bibr">72</a>], open-access Creative Common CC licensed 4.0, MDPI).</p>
Full article ">Figure 2
<p>FCCD for the e-MIP film preparation: coefficients plot. Higher values (regardless of the sign) and little black stars suggest a significant influence of the respective parameter or interaction (the possible number of black stars and the relative significance are: * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">Figure 3
<p>FCCD for the e-MIP film preparation: response surface plot.</p>
Full article ">Figure 4
<p>Nyquist plot of the bare electrode (blue dots), e-MIP-modified electrode before template removal (green dots), e-MIP-modified electrode after template removal (grey dots), e-MIP-modified electrode after rebinding with DA 0.02 mM (yellow dots), e-NIP-modified electrode (red dots). Measurements were performed in 5 mM K<sub>4</sub>Fe(CN)<sub>6</sub>/0.1 M KCl solution. Frequency range 100 kHz–10 mHz with a sinusoidal potential modulation of 0.05 V superimposed on a dc potential of 0.2 V for the bare and 0.1 V for the modified electrodes (equilibrium potential of the redox probe).</p>
Full article ">Figure 5
<p>Fractions of species distribution of DA vs. pH in aqueous solution, calculated with the p<span class="html-italic">K</span><sub>a</sub> values reported in [<a href="#B74-polymers-16-02528" class="html-bibr">74</a>,<a href="#B75-polymers-16-02528" class="html-bibr">75</a>].</p>
Full article ">Figure 6
<p>CV profile for DA (2.5 mM) in PBS 0.1 M at pH 7 at different scan rates: (<b>a</b>) bare electrode; (<b>b</b>) e-MIP-modified electrode. The arrows indicate the potential scan direction.</p>
Full article ">Figure 7
<p>DPV voltammograms registered in PBS 0.1 M solution at pH 7 containing DA at different concentrations ranging from 0 µM to 180 µM: (<b>a</b>) bare electrode; (<b>b</b>) e-MIP-modified electrode; (<b>c</b>) e-NIP-modified electrode.</p>
Full article ">Figure 7 Cont.
<p>DPV voltammograms registered in PBS 0.1 M solution at pH 7 containing DA at different concentrations ranging from 0 µM to 180 µM: (<b>a</b>) bare electrode; (<b>b</b>) e-MIP-modified electrode; (<b>c</b>) e-NIP-modified electrode.</p>
Full article ">Figure 8
<p>(<b>a</b>) DA calibration curves obtained with the bare, e-MIP and e-NIP-modified electrodes. Each point is the mean of the current values registered with three electrodes, and the error bars are the standard deviation. (<b>b</b>) Langmuir fitting for the calibration curve of DA with the e-MIP-modified electrode: <span class="html-italic">I</span><sub>pmax</sub> = 10.8 (1) µA; <span class="html-italic">K</span><sub>aff</sub> =8.4 (2) × 10<sup>3</sup> M<sup>−1</sup>; R<sup>2</sup> = 0.994.</p>
Full article ">Figure 9
<p>DPV voltammogram in PBS 0.1 M at pH 7 containing (<b>a</b>) 350 µM AA and (<b>b</b>) 350 µM UA with subsequent additions of DA from 0 µM to 40 µM.</p>
Full article ">Scheme 1
<p>Possible pathway for the electrochemical oxidation of DA in PBS 0.1 M at pH 7.</p>
Full article ">
18 pages, 6059 KiB  
Article
Computational and Experimental Comparison of Molecularly Imprinted Polymers Prepared by Different Functional Monomers—Quantitative Parameters Defined Based on Molecular Dynamics Simulation
by Jing Yuan, Ying Gao, Xinzhuo Tian, Wenhao Su, Yuxin Su, Shengli Niu, Xiangying Meng, Tong Jia, Ronghuan Yin and Jianmin Hu
Molecules 2024, 29(17), 4236; https://doi.org/10.3390/molecules29174236 - 6 Sep 2024
Viewed by 474
Abstract
Background: In recent years, the advancement of computational chemistry has offered new insights into the rational design of molecularly imprinted polymers (MIPs). From this aspect, our study tried to give quantitative parameters for evaluating imprinting efficiency and exploring the formation mechanism of MIPs [...] Read more.
Background: In recent years, the advancement of computational chemistry has offered new insights into the rational design of molecularly imprinted polymers (MIPs). From this aspect, our study tried to give quantitative parameters for evaluating imprinting efficiency and exploring the formation mechanism of MIPs by combining simulation and experiments. Methods: The pre-polymerization system of sulfadimethoxine (SDM) was investigated using a combination of quantum chemical (QC) calculations and molecular dynamics (MD) simulations. MIPs were prepared on the surface of silica gel by a surface-initiated supplemental activator and reducing agent atom transfer radical polymerization (SI-SARA ATRP). Results: The results of the QC calculations showed that carboxylic monomers exhibited higher bonding energies with template molecules than carboxylic ester monomers. MD simulations confirmed the hydrogen bonding sites predicted by QC calculations. Furthermore, it was observed that only two molecules of monomers could bind up to one molecule of SDM, even when the functional monomer ratio was up to 10. Two quantitative parameters, namely, the effective binding number (EBN) and the maximum hydrogen bond number (HBNMax), were defined. Higher values of EBN and HBNMax indicated a higher effective binding efficiency. Hydrogen bond occupancies and RDF analysis were performed to analyze the hydrogen bond formation between the template and the monomer from different perspectives. Furthermore, under the influence of the EBN and collision probability of the template and the monomers, the experimental results show that the optimal molar ratio of template to monomer is 1:3. Conclusions: The method of monomer screening presented in this study can be extended to future investigations of pre-polymerization systems involving different templates and monomers. Full article
Show Figures

Figure 1

Figure 1
<p>Frame simulation diagram in the TFMAA polymerization system (the left side is the complete system, and the right side is the template and monomer that form the double hydrogen bond after local amplification. The pre-polymerization system is composed of SDM:TFMAA:EGDMA = 1:8:40).</p>
Full article ">Figure 2
<p>Analysis of hydrogen bond occupancy generated by different molar ratios of SDM to functional monomers ((<b>a</b>): AA, (<b>b</b>): TFMAA, (<b>c</b>): 4-VBA, (<b>d</b>): MAA, (<b>e</b>): EMA, (<b>f</b>): EHMA, (<b>g</b>): MMA). Since the carboxylic ester monomers have only hydrogen bond acceptors and no hydrogen bond donors, self-polymerization cannot occur. So, the last three figures (<b>e</b>–<b>g</b>) do not have data for the “monomer-monomer” group.</p>
Full article ">Figure 3
<p>Analysis of the lifetime of the hydrogen bond generated by SDM to functional monomers or crosslinkers. (In (<b>a</b>), “solid line” represents template–monomer and “dash line” represents the monomer–monomer; In (<b>b</b>), “solid line” represents template–crosslinker).</p>
Full article ">Figure 4
<p>RDF analysis of TFMAA and SDM binding sites. (<b>a</b>,<b>b</b>): hydrogen bond acceptor of SDM to hydrogen bond donor of TFMAA; (<b>c</b>): hydrogen bond donor of SDM to hydrogen bond acceptor of TFMAA.</p>
Full article ">Figure 5
<p>RDF analysis of EHMA and SDM binding sites. (Hydrogen bond donor of SDM to hydrogen bond acceptor ((<b>a</b>): carbonyl oxygen, (<b>b</b>): ester oxygen) of EHMA).</p>
Full article ">Figure 6
<p>SEM images of SiO<sub>2</sub>@Br (<b>a</b>) and SiO<sub>2</sub>@MIP prepared by AA, TFMAA, and 4-VBA (<b>b</b>–<b>d</b>).</p>
Full article ">Figure 7
<p>Comparison of the adsorption capacities of SMIPs and SNIPs synthesized by carboxylic acid and ester functional monomers. (SDM:monomers = 1:3).</p>
Full article ">Figure 8
<p>Comparison of the adsorption capacities of SMIPs synthesized by AA, TFMAA, and 4-VBA in different molar proportions.</p>
Full article ">Figure 9
<p>“Freeze” the cross-linked structure of the monomer to simulate “holes”. Replacement of SDM with other drug molecules could be used to model selective adsorption. (<b>left</b>: SDM-AA 1:2 complex; <b>right</b>: SDM-TFMAA 1:2 complex).</p>
Full article ">Figure 10
<p>Adsorption amounts and weak interaction energies of the two SMIPs on seven sulfonamides. (<b>left</b>: SDM-AA; <b>right</b>: SDM-TFMAA).</p>
Full article ">Figure 11
<p>The synthetic route for SMIPs, taking TFMAA as a functional monomer for example.</p>
Full article ">
17 pages, 4566 KiB  
Article
A Highly Efficient Fluorescent Turn-Off Nanosensor for Quantitative Detection of Teicoplanin Antibiotic from Humans, Food, and Water Based on the Electron Transfer between Imprinted Quantum Dots and the Five-Membered Cyclic Boronate Esters
by Yansong Zhang, Daojin Li and Xiping Tian
Molecules 2024, 29(17), 4115; https://doi.org/10.3390/molecules29174115 - 30 Aug 2024
Viewed by 367
Abstract
Teicoplanin has been banned in the veterinary field due to the drug resistance of antibiotics. However, teicoplanin residue from the antibiotic abuse of humans and animals poses a threat to people’s health. Therefore, it is necessary to develop an efficient way for the [...] Read more.
Teicoplanin has been banned in the veterinary field due to the drug resistance of antibiotics. However, teicoplanin residue from the antibiotic abuse of humans and animals poses a threat to people’s health. Therefore, it is necessary to develop an efficient way for the highly accurate and reliable detection of teicoplanin from humans, food, and water. In this study, novel imprinted quantum dots of teicoplanin were prepared based on boronate affinity-based precisely controlled surface imprinting. The imprinting factor (IF) for teicoplanin was evaluated and reached a high value of 6.51. The results showed excellent sensitivity and selectivity towards teicoplanin. The relative fluorescence intensity was inversely proportional to the concentration of teicoplanin, in the range of 1.0–17 μM. And its limit of detection (LOD) was obtained as 0.714 μM. The fluorescence quenching process was mainly controlled by a static quenching mechanism via the non-radiative electron-transfer process between QDs and the five-membered cyclic boronate esters. The recoveries for the spiked urine, milk, and water samples ranged from 95.33 to 104.17%, 91.83 to 97.33, and 94.22 to 106.67%, respectively. Full article
(This article belongs to the Section Analytical Chemistry)
Show Figures

Figure 1

Figure 1
<p>Structure of teicoplanin.</p>
Full article ">Figure 2
<p>Synthesis routes of APBA-functionalized CdTe QDs (<b>A</b>) and the formation mechanism of Tei-imprinted magnetic nanoparticles by boronate affinity-based precisely controlled surface imprinting (<b>B</b>).</p>
Full article ">Figure 3
<p>Transmission electron microscopy characterization (TEM) of imprinted APBA-functionalized CdTe QDs.</p>
Full article ">Figure 4
<p>XPS (<b>A</b>) and XRD (<b>B</b>) spectra of QDs@APBA@MIPs.</p>
Full article ">Figure 5
<p>(<b>A</b>) The dependence of imprinting effect on polymerization time for preparing imprinted APBA-functionalized CdTe QDs and non-imprinted APBA-functionalized CdTe QDs, [Tei] = 1.7 × 10<sup>−5</sup> M; (<b>B</b>) fluorescence emission spectra of QDs@APBA@NIPs (a), QDs@APBA@MIPs (after washing, b), QDs@APBA@MIPs (before washing, c).</p>
Full article ">Figure 6
<p>Comparison of adsorption performance of QDs@APBA@MIPs and QDs@APBA@NIPs.</p>
Full article ">Figure 7
<p>Fluorescence emission spectra of the QDs@APBA@MIPs (<b>A</b>), the QDs@APBA@NIPs (<b>B</b>) with various concentrations of Tei and Stern–Volmer plots (<b>C</b>) for the QDs@APBA@MIPs, and the QDs@APBA@NIPs with the target Tei. The concentrations were 0.1 mg/mL; [Tei] = 0, 1, 2, 3, 4, 5, 8, 11, 14, 17 (1 × 10<sup>−6</sup> mol/L) (1–10).</p>
Full article ">Figure 8
<p>Fluorescence decay curves for QDs@APBA@MIPs.</p>
Full article ">Figure 9
<p>Selectivity of QDs@APBA@MIPs toward Tei using boronate affinity-based precisely controlled surface imprinting.</p>
Full article ">Figure 10
<p>Reproducibility (<b>A</b>) and chemical stability of QDs@APBA@MIPs (<b>B</b>).</p>
Full article ">
27 pages, 5737 KiB  
Review
Electrochemical Sensors for Antibiotic Detection: A Focused Review with a Brief Overview of Commercial Technologies
by Margaux Frigoli, Mikolaj P. Krupa, Geert Hooyberghs, Joseph W. Lowdon, Thomas J. Cleij, Hanne Diliën, Kasper Eersels and Bart van Grinsven
Sensors 2024, 24(17), 5576; https://doi.org/10.3390/s24175576 - 28 Aug 2024
Viewed by 1362
Abstract
Antimicrobial resistance (AMR) poses a significant threat to global health, powered by pathogens that become increasingly proficient at withstanding antibiotic treatments. This review introduces the factors contributing to antimicrobial resistance (AMR), highlighting the presence of antibiotics in different environmental and biological matrices as [...] Read more.
Antimicrobial resistance (AMR) poses a significant threat to global health, powered by pathogens that become increasingly proficient at withstanding antibiotic treatments. This review introduces the factors contributing to antimicrobial resistance (AMR), highlighting the presence of antibiotics in different environmental and biological matrices as a significant contributor to the resistance. It emphasizes the urgent need for robust and effective detection methods to identify these substances and mitigate their impact on AMR. Traditional techniques, such as liquid chromatography-mass spectrometry (LC-MS) and immunoassays, are discussed alongside their limitations. The review underscores the emerging role of biosensors as promising alternatives for antibiotic detection, with a particular focus on electrochemical biosensors. Therefore, the manuscript extensively explores the principles and various types of electrochemical biosensors, elucidating their advantages, including high sensitivity, rapid response, and potential for point-of-care applications. Moreover, the manuscript investigates recent advances in materials used to fabricate electrochemical platforms for antibiotic detection, such as aptamers and molecularly imprinted polymers, highlighting their role in enhancing sensor performance and selectivity. This review culminates with an evaluation and summary of commercially available and spin-off sensors for antibiotic detection, emphasizing their versatility and portability. By explaining the landscape, role, and future outlook of electrochemical biosensors in antibiotic detection, this review provides insights into the ongoing efforts to combat the escalating threat of AMR effectively. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of (<b>a</b>) a generic biosensor device showing a sample containing the target interacting with the bioreceptor. After the recognition, the transducer is responsible for generating a signal. (<b>b</b>) examples of electrochemical transduction signals. Copyright Sensors and Actuators Reports. 2022 [<a href="#B71-sensors-24-05576" class="html-bibr">71</a>].</p>
Full article ">Figure 2
<p>Schematic illustration of the working principle of electrochemically based aptamer sensors. Copyright Toxics 2023 [<a href="#B91-sensors-24-05576" class="html-bibr">91</a>].</p>
Full article ">Figure 3
<p>Portable aptamer-based sensor for kanamycin developed by Bao et al. Copyright Chinese Jour. Of Chem., 2023 [<a href="#B93-sensors-24-05576" class="html-bibr">93</a>].</p>
Full article ">Figure 4
<p>A modified ssDNA aptamer sensor for the selective detection of tetracycline designed by Malecka-Baturo et al. Copyright Int. J. Mol. Sci. 2022 [<a href="#B94-sensors-24-05576" class="html-bibr">94</a>]. (<b>A</b>) Scheme of the aptasensor preparation, and (<b>B</b>) representation of the steps needed to use the aptasensor.</p>
Full article ">Figure 5
<p>Schematic representation of the extraction and rebinding process on an antibiotic (cefquinome) after electropolymerization on the surface of an electrode and consequent electrochemical analysis. Copyright Sensors and Act. B: Chemical, 2019 [<a href="#B108-sensors-24-05576" class="html-bibr">108</a>].</p>
Full article ">Figure 6
<p>The dual recognition sensor for amoxicillin. Copyright Analytica Chimica Acta, 2020 [<a href="#B110-sensors-24-05576" class="html-bibr">110</a>].</p>
Full article ">Figure 7
<p>Overview of the norfloxacin sensor developed by Thi Vu et al. [<a href="#B111-sensors-24-05576" class="html-bibr">111</a>]. Copyright ACS Omega, 2023.</p>
Full article ">Figure 8
<p>Working principle of the ConA arsanilic acid developed by You et al. Copyright J. Electroanalytical Chem., 2024 [<a href="#B135-sensors-24-05576" class="html-bibr">135</a>].</p>
Full article ">
29 pages, 1862 KiB  
Review
Molecularly Imprinted Microspheres in Active Compound Separation from Natural Product
by Husna Muharram Ahadi, Firghi Muhammad Fardhan, Driyanti Rahayu, Rimadani Pratiwi and Aliya Nur Hasanah
Molecules 2024, 29(17), 4043; https://doi.org/10.3390/molecules29174043 - 26 Aug 2024
Viewed by 432
Abstract
Molecularly Imprinted Microspheres (MIMs) or Microsphere Molecularly Imprinted Polymers represent an innovative design for the selective extraction of active compounds from natural products, showcasing effectiveness and cost-efficiency. MIMs, crosslinked polymers with specific binding sites for template molecules, overcome irregularities observed in traditional Molecularly [...] Read more.
Molecularly Imprinted Microspheres (MIMs) or Microsphere Molecularly Imprinted Polymers represent an innovative design for the selective extraction of active compounds from natural products, showcasing effectiveness and cost-efficiency. MIMs, crosslinked polymers with specific binding sites for template molecules, overcome irregularities observed in traditional Molecularly Imprinted Polymers (MIPs). Their adaptability to the shape and size of target molecules allows for the capture of compounds from complex mixtures. This review article delves into exploring the potential practical applications of MIMs, particularly in the extraction of active compounds from natural products. Additionally, it provides insights into the broader development of MIM technology for the purification of active compounds. The synthesis of MIMs encompasses various methods, including precipitation polymerization, suspension polymerization, Pickering emulsion polymerization, and Controlled/Living Radical Precipitation Polymerization. These methods enable the formation of MIPs with controlled particle sizes suitable for diverse analytical applications. Control over the template-to-monomer ratio, solvent type, reaction temperature, and polymerization time is crucial to ensure the successful synthesis of MIPs effective in isolating active compounds from natural products. MIMs have been utilized to isolate various active compounds from natural products, such as aristolochic acids from Aristolochia manshuriensis and flavonoids from Rhododendron species, among others. Based on the review, suspension polymerization deposition, which is one of the techniques used in creating MIPs, can be classified under the MIM method. This is due to its ability to produce polymers that are more homogeneous and exhibit better selectivity compared to traditional MIP techniques. Additionally, this method can achieve recovery rates ranging from 94.91% to 113.53% and purities between 86.3% and 122%. The suspension polymerization process is relatively straightforward, allowing for the effective control of viscosity and temperature. Moreover, it is cost-effective as it utilizes water as the solvent. Full article
Show Figures

Figure 1

Figure 1
<p>Principle of MIP.</p>
Full article ">Figure 2
<p>General flow of MIP synthesis.</p>
Full article ">Figure 3
<p>Illustration of the ATRPP mechanism.</p>
Full article ">Figure 4
<p>Process of suspension polymerization.</p>
Full article ">
13 pages, 3648 KiB  
Article
Enhancing Selectivity with Molecularly Imprinted Polymers via Non-Thermal Dielectric Barrier Discharge Plasma
by Samira Amiri Khoshkar Vandani, Qianwei Liu, Yuki Lam and Hai-Feng Ji
Polymers 2024, 16(16), 2380; https://doi.org/10.3390/polym16162380 - 22 Aug 2024
Viewed by 487
Abstract
Molecularly imprinted polymers (MIPs) are synthetic polymers that mimic the functions of antibodies. Though MIPs are promising tools in various areas, achieving high selectivity in MIPs can be difficult. To improve selectivity, various approaches have been implemented; however, the role of polymerization methods [...] Read more.
Molecularly imprinted polymers (MIPs) are synthetic polymers that mimic the functions of antibodies. Though MIPs are promising tools in various areas, achieving high selectivity in MIPs can be difficult. To improve selectivity, various approaches have been implemented; however, the role of polymerization methods or synthetic techniques in enhancing the selectivity of MIPs has not been studied and remains a crucial area for further research. MIPs are typically prepared from free radical reactions. Recently, we found that Dielectric Barrier Discharge (DBD) plasma can be used to initiate the polymerization of vinyl monomers. The DBD plasma method allows the monomers to associate with the template molecules and initiate polymerization with minimal disruption to the positioning of the monomers. We hypothesize that this could be a preferred method to prepare MIPs over the traditional radical reaction that may cause a disturbance of the pre-associated monomers on the templates for the polymerization. Chicken egg white serum albumin (CESA) was used as the template protein for the MIPs. Our results show that in all test conditions, approximately twofold improvement in selectivity was achieved, which is the primary performance metric for MIPs. This enhancement was evident across all categories, including MIPs prepared from various monomer combinations. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Scheme of a DBD plasma setup. (<b>b</b>) The DBD electrode was made of a 38 mm × 64 mm copper plate covered with a 1 mm-thick glass strip. The discharge gap for plasma was 5 mm. Plasma was observed on a glass slide between the two insulated electrodes.</p>
Full article ">Figure 2
<p>FTIR spectra of MIPs synthesized from the traditional radical reactions and the DBD plasma using the combination of six monomers before and after washing to remove CESA.</p>
Full article ">Figure 3
<p>(<b>a</b>) UV-Vis spectrum of CESA upon binding with an MIP synthesized using the combination of six monomers and radical polymerization methodology. (<b>b</b>) UV-Vis spectrum of BSA upon binding with an MIP synthesized using the combination of six monomers in <a href="#polymers-16-02380-t003" class="html-table">Table 3</a> and radical polymerization methodology. (<b>c</b>) UV-Vis spectrum of CESA upon binding with an MIP synthesized using the combination of six monomers and plasma polymerization methodology. (<b>d</b>) UV-Vis spectrum of BSA upon binding with an MIP synthesized using the combination of six monomers in <a href="#polymers-16-02380-t003" class="html-table">Table 3</a> and plasma polymerization methodology.</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>) UV-Vis spectrum of CESA upon binding with an MIP synthesized using the combination of six monomers and radical polymerization methodology. (<b>b</b>) UV-Vis spectrum of BSA upon binding with an MIP synthesized using the combination of six monomers in <a href="#polymers-16-02380-t003" class="html-table">Table 3</a> and radical polymerization methodology. (<b>c</b>) UV-Vis spectrum of CESA upon binding with an MIP synthesized using the combination of six monomers and plasma polymerization methodology. (<b>d</b>) UV-Vis spectrum of BSA upon binding with an MIP synthesized using the combination of six monomers in <a href="#polymers-16-02380-t003" class="html-table">Table 3</a> and plasma polymerization methodology.</p>
Full article ">Figure 4
<p>Comparison of selectivity among A to G monomer combinations listed in <a href="#polymers-16-02380-t003" class="html-table">Table 3</a>.</p>
Full article ">
22 pages, 4474 KiB  
Review
Sensors Based on Molecularly Imprinted Polymers in the Field of Cancer Biomarker Detection: A Review
by Camila Quezada, S. Shiva Samhitha, Alexis Salas, Adrián Ges, Luis F. Barraza, María Carmen Blanco-López, Francisco Solís-Pomar, Eduardo Pérez-Tijerina, Carlos Medina and Manuel Meléndrez
Nanomaterials 2024, 14(16), 1361; https://doi.org/10.3390/nano14161361 - 19 Aug 2024
Viewed by 755
Abstract
Biomarkers play a pivotal role in the screening, diagnosis, prevention, and post-treatment follow-up of various malignant tumors. In certain instances, identifying these markers necessitates prior treatment due to the complex nature of the tumor microenvironment. Consequently, advancing techniques that exhibit selectivity, specificity, and [...] Read more.
Biomarkers play a pivotal role in the screening, diagnosis, prevention, and post-treatment follow-up of various malignant tumors. In certain instances, identifying these markers necessitates prior treatment due to the complex nature of the tumor microenvironment. Consequently, advancing techniques that exhibit selectivity, specificity, and enable streamlined analysis hold significant importance. Molecularly imprinted polymers (MIPs) are considered synthetic antibodies because they possess the property of molecular recognition with high selectivity and sensitivity. In recent years, there has been a notable surge in the investigation of these materials, primarily driven by their remarkable adaptability in terms of tailoring them for specific target molecules and integrating them into diverse analytical technologies. This review presents a comprehensive analysis of molecular imprinting techniques, highlighting their application in developing sensors and analytical methods for cancer detection, diagnosis, and monitoring. Therefore, MIPs offer great potential in oncology and show promise for improving the accuracy of cancer screening and diagnosis procedures. Full article
(This article belongs to the Section Biology and Medicines)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Template molecules. (<b>B</b>) Reagents used in the molecular printing process. (<b>C</b>) General scheme of molecular imprinting polymers, from (<b>a</b>–<b>c</b>), the pre-polymerization, polymerization, and template extraction complex, is presented. (<b>D</b>–<b>H</b>) Scheme of the most common molecular printing methods.</p>
Full article ">Figure 2
<p>Solid-phase synthesis template molecule immobilization by different functional groups [<a href="#B68-nanomaterials-14-01361" class="html-bibr">68</a>].</p>
Full article ">Figure 3
<p>Sensors made with molecular imprinting polymers for the detection of lung cancer. (<b>A</b>) Proteomic analysis [<a href="#B90-nanomaterials-14-01361" class="html-bibr">90</a>]. (<b>B</b>) Electrochemical sensor for epitope identification [<a href="#B91-nanomaterials-14-01361" class="html-bibr">91</a>]. (<b>C</b>) Electrochemical sensor for the identification of endogenous substances [<a href="#B92-nanomaterials-14-01361" class="html-bibr">92</a>]. (<b>D</b>) Chemoresistive sensor for the determination of hexanal [<a href="#B93-nanomaterials-14-01361" class="html-bibr">93</a>].</p>
Full article ">Figure 4
<p>Schematics of sensors designed for the detection of tumor markers for the reproductive system using MIPs as identification material. (<b>A</b>) Sensitive and specific plasmonic biosensor for the detection of PSA [<a href="#B118-nanomaterials-14-01361" class="html-bibr">118</a>]. (<b>B</b>) Electrochemical sensor aimed at the quantitative analysis of PSA [<a href="#B120-nanomaterials-14-01361" class="html-bibr">120</a>]. (<b>C</b>) Electrochemical sensor for the identification of PSA [<a href="#B121-nanomaterials-14-01361" class="html-bibr">121</a>].</p>
Full article ">
Back to TopTop