Next Article in Journal
Research on Friction Performance of Friction Stir Welding Tools Based on Non-Smooth Structure
Previous Article in Journal
Biomimetic Antifungal Materials: Countering the Challenge of Multidrug-Resistant Fungi
Previous Article in Special Issue
Can Plants Perceive Human Gestures? Using AI to Track Eurythmic Human–Plant Interaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Various Carbon Electrodes on MIP-Based Sensing Proteins Using Poly(Scopoletin): A Case Study of Ferritin

Molecular Biotechnology, Faculty of Science, Turkish-German University, Istanbul 34820, Türkiye
Biomimetics 2024, 9(7), 426; https://doi.org/10.3390/biomimetics9070426
Submission received: 18 May 2024 / Revised: 1 July 2024 / Accepted: 9 July 2024 / Published: 13 July 2024
(This article belongs to the Special Issue Biomimetics in Intelligent Sensor)

Abstract

:
Sensitivity in the sub-nanomolar concentration region is required to determine important protein biomarkers, e.g., ferritin. As a prerequisite for high sensitivity, in this paper, the affinity of the functional monomer to the macromolecular target ferritin in solution was compared with the value for the respective molecularly imprinted polymer (MIP)-based electrodes, and the influence of various surface modifications of the electrode was investigated. The analytical performance of ferritin sensing was investigated using three different carbon electrodes (screen-printed carbon electrodes, single-walled-carbon-nanotube-modified screen-printed carbon electrodes, and glassy carbon electrodes) covered with a scopoletin-based MIP layer. Regardless of the electrode type, the template molecule ferritin was mixed with the functional monomer scopoletin, and electropolymerization was conducted using multistep amperometry. All stages of MIP preparation were followed by evaluating the diffusional permeability of the redox marker ferricyanide/ferrocyanide through the polymer layer by differential pulse voltammetry. The best results were obtained with glassy carbon electrodes. The MIP sensor responded up to 0.5 µM linearly with a Kd of 0.30 µM. Similar results were also obtained in solution upon the interaction of scopoletin and ferritin using fluorescence spectroscopy, resulting in the quenching of the scopoletin signal, with a calculated Kd of 0.81 µM. Moreover, the binding of 1 µM ferritin led to 49.6% suppression, whereas human serum albumin caused 8.6% suppression.

Graphical Abstract

1. Introduction

Nature needed millions of years for the evolution of proteins and nucleic acids, which exhibit high affinity and selectivity in binding to or chemical conversion of different partners. Nowadays, biomimetic binders and catalysts can be generated within several days using “evolution in the test tube” of non-natural nucleotides or total chemical synthesis of (molecularly imprinted) polymers [1]. According to the International Union of Pure and Applied Chemistry (IUPAC), “biomimetic” refers to a laboratory method designed to replicate a natural chemical process or a compound that mimics the structure or function of a biological material. The lotus effect, observed in water-repellent surfaces, is a prominent example of a biomimetic system.
In recent years, there has been a tremendous interest in substituting biochemical reagents in bioanalysis and separation techniques with fully synthetic organic polymers, specifically molecularly imprinted polymers (MIPs, the so-called artificial antibodies) [1]. In analogy to the binding of antigens and haptens in the antigen-binding sites, the paratope of antibodies, recognition by MIPs is also based on the structural complementarity of the binding pocket and target. MIPs, in particular, have gained attention for their versatile applications in biotechnology. These include their successful use in solid-phase extraction and chromatography, in replacing antibodies in ELISAs, and as sensors for low-molecular-weight substances and biomacromolecules [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21].
The synthesis of MIPs, pioneered by Wulff and Mosbach, involves polymerizing functional monomers (with or without cross-linkers) in the presence of a target analyte, known as the template [22,23]. Removal of the template from the polymer results in the formation of binding cavities with a molecular memory of the template’s size, shape, and functionality (Scheme 1). Unlike enzymes and antibodies, which are built from 20 natural amino acids, MIPs can be prepared using one to six functional monomers, even without a cross-linker. Furthermore, MIPs exhibit higher stability under harsh conditions (e.g., pH, temperature, organic solvents) and are easier to regenerate compared to their biological counterparts [1,24].
A wide range of MIPs have been developed to recognize various targets, especially low-molecular-weight ones, with some already commercially available. However, the development of MIPs for biomacromolecules has been more limited, constituting only around 10% of annually published papers on MIPs [7]. The scarcity of MIPs for high-molecular-weight compounds is largely attributed to stability issues of biomacromolecular templates in the polymerization media used in conventional bulk imprinting processes [7,25].
Another challenge of the traditional bulk imprinting process is the potential entrapment of template molecules within the polymer matrix, hindering their removal and subsequent rebinding. Surface imprinting techniques, including electropolymerization, have partially addressed these limitations [7,26,27,28,29,30]. Electropolymerization enables polymer synthesis under mild conditions with adjustable thickness. Moreover, polymer films can be directly formed on the transducer, eliminating the need for additional immobilization steps. Different types of metal and carbon electrodes have been applied in order to optimize the analytical performance, especially sensitivity [1]. Commonly used electroactive monomers include pyrrole, o-phenylenediamine (o-PD), aniline, scopoletin, p-aminophenylboronic acid, phenol, and thiophene and their derivatives [2,7,14].
The scopoletin molecule offers effective aromatic π–π interactions and hydrogen bonding for the “pre-polymerization complex” with different template molecules. The aromatic system of scopoletin exhibits fluorescence behavior, and quenching resulting from interaction with the respective target analyte has been applied to determine the affinity expressed by the dissociation constant Kd in solution [31]. Polymerization of the natural coumarin derivative scopoletin proceeds at moderate anodic potentials and does not require deoxygenation of the monomer solution. It yields a non-conductive, conformal, hydrophilic film on different electrode materials [7]. These ultra-thin films, in the nanometer range, only partially entrap the protein template, thus enabling the effective removal of the template from the MIP and fast rebinding. For the electrochemical readout, either the direct electron transfer of redox-active proteins, including ferritin, or the modulation of permeability for redox markers by protein binding (“redox gating”) has been applied. The spontaneous adsorption of the protein target at the electrode surface and changes in the polymer structure on changing the background solution will overlay the signal of redox gating for the binding of the target to the MIP [32]. After polymerization, electrostatic interactions are involved in the interaction with the target because the poly(scopoletin) film on Au electrodes has a negative zeta potential [33]. This polarization should suppress the non-specific binding of negatively charged substances, including ferritin (IP 4.52).
This work investigates the analytical performance of ferritin sensing using three different carbon electrodes (screen-printed carbon electrodes, single-walled-carbon-nanotube-modified screen-printed carbon electrodes, and glassy carbon electrodes) covered with a scopoletin-based MIP layer. Ferritin, the central iron storage protein in human iron metabolism, is a biomarker for iron deficiency, cell damage, and inflammation [34]. It has a globular structure with a molecular weight of 450 kDa and a diameter of 12 nm. It also possesses an isoelectric point of 4.52. Various strategies, including immunoassays (enzyme-linked immunosorbent assay [ELISA] and radioimmunoassay) and spectroscopic methods (colorimetry, luminescence measurement, nephelometry, turbidimetry, etc.), have been developed for its analysis [35]. In addition to its function as a biomarker, it is an interesting target for MIP research because its diameter is close to that of colloids and viruses. Thus, the achievements can be transferred to developing imprinted polymers for these bio-particles [36].
Regardless of the electrode type, a simple electrode preparation method, namely random imprinting, was used. As illustrated in the scheme, the template molecule ferritin was mixed with the functional monomer scopoletin, and electropolymerization was conducted using pulse amperometry (Scheme 1) to create films with good adhesion to the electrode surface [7]. The application of pulsed potentials also supports the replenishment of the depleted layer adjacent to the MIP surface by the low diffusion of the macromolecular protein and enhances the incorporation of the template into the polymer. The film thickness can be precisely adjusted by the charge consumed during electropolymerization [7]. Various strategies were used to remove the template molecules from the polymer film. Electrochemistry, mainly differential pulse voltammetry, was used to characterize all stages of MIP preparation. Furthermore, the affinity of the functional monomer to the macromolecular target ferritin in solution was compared with the value for the respective MIP, and the influence of various surface modifications of the electrode was investigated.
Achieving sensitivity in the sub-nanomolar concentration range is essential for the accurate determination of critical protein biomarkers, such as ferritin. A ferritin concentration above 150 µg/L in menstruating women and 200 µg/L in men and non-menstruating women who are otherwise healthy may suggest a risk of iron overload (conditional recommendation based on prior WHO guidelines). In adults with underlying health conditions, a ferritin level exceeding 500 µg/L may also indicate a potential risk of iron overload [37].

2. Materials and Methods

2.1. Chemicals and Reagents

Scopoletin (7-hydroxy-6-methoxycoumarin), ferritin from the equine spleen, human serum albumin (HSA), and HEPES were purchased from Sigma-Aldrich (Steinheim, Germany). Ferricyanide, ferrocyanide, and sodium chloride were supplied by MERCK KGaA (Darmstadt, Germany). All reagents were of analytical grade and used without further purification.

2.2. Preparation of Ferritin-Imprinted Sensors and Measurements

Glassy carbon disk electrodes (3 mm in diameter; CH Instruments, Austin, TX, USA), screen-printed carbon electrodes (4 mm in diameter; SPC from DropSens, Oviedo, Spain), and single-walled-carbon-nanotube-modified screen-printed carbon electrodes (4 mm in diameter; SWSP from DropSens, Oviedo, Spain) were used for MIP synthesis and for voltammetric measurements (PalmSens potentiostat, Utrecht, The Netherlands).
Before electropolymerization, the GCEs were cleaned with 30% nitric acid for 15 min. Afterward, mechanical cleaning was performed with 1.0 μm, 0.3 μm, and 0.05 μm of alumina slurry. The electrodes were rinsed with ethanol and Millipore water by ultrasonication. The screen-printed electrodes were activated in 0.1 M H2SO4 for 2 min at 3 mA using chronopotentiometry.
Based on previous experiences and the suggestion in Malitesta’s recent paper, electropolymerization on the GCEs and screen-printed electrodes was performed in a mixture of 0.5 mM scopoletin (10 mM stock solution in ethanol), 100 mM NaCl (1 M stock in water), and 10 µM ferritin (100 µM stock solution in 10 mM HEPES buffer, pH 7.4) solutions [38,39]. End concentrations in the mixture were obtained by dilution with water. Multistep amperometry was applied, involving 100 cycles, in which each GCE (or SPE) was polarized at 0 V for 1 s and at 0.9 V for 0.1 s. Various strategies, such as electrochemical removal and incubation in alkaline, acidic, or salt solutions, were used to remove the template molecules from the polymer film.
The rebinding and removal of ferritin were followed by changes in DPV responses of 5 mM [Fe(CN)6]3−/4– solution (in 100 mM KCl). This was achieved by applying a potential range of −0.2 to 0.7 V, a potential step of 5 mV, a pulse amplitude of 25 mV, and a pulse duration of 0.7 s. Moreover, cyclic voltammetry was applied to follow the fabrication steps. The potential was swept between −0.2 and 0.7 V, with a scan rate of 50 mV/s (3 scans).
Fluorescence measurements were performed using the M-4/2003 Quanta Master Steady State Spectrofluorometer (Photon Technology International, London, ON, Canada) as follows: The concentration of ferritin (stock solution: 1 mM in 10 mM HEPES buffer, pH 7.4) was varied in 0.5 mM scopoletin solution containing 100 mM NaCl. As ferritin was incrementally added, from 0 to 25 µM, the fluorescence emission of scopoletin (with an excitation wavelength of 430 nm and an emission wavelength of 460 nm) exhibited partial quenching [31]. The spectral alterations corresponding to different ferritin concentrations were analyzed using the One-Site-Specific Binding Model (y = Bmax × x/(Kd + x) to determine the dissociation constant governing the interaction between ferritin and scopoletin (Bmax: maximum response; and Kd: dissociation constant).

3. Results and Discussion

3.1. Screen-Printed Electrodes

3.1.1. Screen-Printed Carbon Electrodes

Over time, sensor technology using screen-printed electrodes has found its potential in various applications. It is known for its capacity to enable fast, accurate, and portable analyses at low cost [40,41]. Figure 1 presents the formation of an MIP layer on an SPC electrode by the electropolymerization of scopoletin in the presence of ferritin through multistep amperometry, involving 100 cycles, in which the SPC electrode was polarized at 0 V for 1 s and at 0.9 V for 1 s.
The predominant method used for the characterization of electrochemical MIP sensors evaluates the diffusional permeability of the polymer layer to a redox marker, such as ferricyanide/ferrocyanide, using cyclic voltammetry, differential pulse voltammetry, square wave voltammetry, or electrochemical impedance spectroscopy (as depicted in Figure 2) [1]. As demonstrated in Figure 2, a non-conducting film was formed after electropolymerization (red curve).
A pivotal stage in the fabrication of sensitive and selective MIP sensors is the removal step. Preferably, mild procedures are used to avoid the destruction of the polymeric film. Several approaches have been applied to remove the template ferritin from an MIP-covered SPC electrode. An increase in ionic strength by incubation in 3 M KCl had no effect, while electrochemical treatment in HEPES buffer caused partial removal of template molecules. Nevertheless, the rebinding of ferritin was not achieved. Milder acidic conditions, such as incubation in 50 mM glycine–HCl buffer at pH 2.2, or harsher acidic conditions, such as incubation in 0.1 M H2SO4, also had a similar effect. In our previous work involving an MIP-covered gold SPR chip, the template ferritin was successfully removed from 1 mM NaOH solution, enabling the rebinding of ferritin [38]. However, the use of three different concentrations of NaOH solution (0.5 mM, 1 mM, and 5 mM) failed to achieve template removal.

3.1.2. Single-Walled-Carbon-Nanotube-Modified Screen-Printed Carbon Electrodes

Carbon nanotubes are frequently used in bio(mimetic) sensors due to their ability to increase the electrochemical active surface area and improve electrochemical properties [42,43,44]. In this study, SWSP electrodes were used for MIP sensor preparation. The multistep amperometry method described earlier was used. Following electropolymerization, a non-conducting film was formed (Figure 3). Partial removal of template molecules was achieved in 50 mM glycine–HCl buffer using cyclic voltammetry within the potential range of −0.2 V to 1.0 V (50 scans), whereas incubation in the same buffer alone had no effect. Despite these efforts, successful rebinding was not achieved. Additionally, various other template removal procedures were attempted, with no favorable outcomes. Consequently, further investigation was conducted using a glassy carbon electrode.

3.2. Glassy Carbon Electrodes

Glassy carbon electrodes have been widely used in electrochemical biosensing, including MIP-based sensors, owing to their good chemical stability, electrical conductivity, biocompatibility, and cost-effectiveness [45,46]. Furthermore, the integration of modifiers and nanomaterials has been observed to enhance the performance of these sensors.
The multistep amperometry method described earlier was used. depicted in Figure 4, the MIP preparation process can be monitored by observing the reduction in the signal intensity of redox markers. In contrast to screen-printed electrodes, ferritin molecules can be effectively removed using 5 mM NaOH solution (blue). Furthermore, ferritin molecules can be rebound, reaching equilibration within an hour (green).
Rebinding studies were conducted in 10 mM HEPES buffer (pH 7.4) with varying ferritin concentrations. The change in the signal of the redox marker was investigated using differential pulse voltammetry. The ferritin–MIP electrode linearly suppressed the signal of the redox marker up to 0.5 µM with an R2 of 0.97277 (Figure 5A,B), consistent with our previous findings and those of several previously published results (Table 1) [38,47,48]. The limit of detection (LOD) was calculated using the equation 3.3 × SD/m (SD: standard deviation of the blank and the slope of the calibration graph) as 0.097 µM.
Using the One-Site-Specific Binding Model, the dissociation constant (Kd) was calculated to be 0.30 µM. Similar results were also obtained in solution upon the interaction of scopoletin and ferritin using fluorescence spectroscopy (resulting in the quenching of the scopoletin signal), with a calculated Kd of 0.81 µM.
Since the cutoff value of ferritin in the serum of healthy persons is at 3 × 10−11 M, the sensitivity of the MIP sensor in this study is not sufficient for the lower ferritin range in serum. The sensor presented by Cai et al. claims to have the required concentration [49] However, the reported measuring range in the lower attomolar concentration range is highly questionable. In contrast, the measuring range and Kd value in this study are in accordance with other publications and compatible with thermodynamic data [24].
Electrosynthesized MIPs typically exhibit Kd values in the range of μM to nM, i.e., the change in free enthalpy is between −30 and −60 KJoule. A few papers claim even lower limits of detection, in the picomolar range, or even lower measuring ranges over more than four decades of target concentration (Table 2). An explanation for the “ultra-sensitivity” may be that the MIP acts as an ion channel sensor with a semi-logarithmic concentration dependence over several orders of magnitude [50].
In addition, the cross-reactivity of the MIP sensor was studied using human serum albumin (HSA). The binding of 1 µM ferritin led to 49.6% suppression, whereas HSA caused 8.6% suppression. This result demonstrates the low cross-reactivity of the MIP toward a non-target protein and is in accordance with reports in the literature. MIPs for macromolecular targets typically exhibit cross-reactivities above 2% versus the most abundant proteins in serum (HSA, IgG, and hemoglobin). While “real samples,” like human serum, contain the relevant protein biomarker, e.g., for cancer, in the sub-nanomolar range, these potentially interfering proteins are in the millimolar concentration range. Therefore, the MIP should be saturated by an excess of the interfering proteins, and dilution should not solve this problem. However, lateral-flow devices with many separation plates could govern the separation effect of the MIP.

4. Conclusions

The general concept of increasing the active surface of electrochemical sensors by using electrodes with high roughness or by integrating nanoparticles, nanotubes, and graphene has been widely studied and transferred to MIP sensors. The comparison of the three different MIP-covered carbon electrodes underlines the findings described in the literature that despite identical MIP synthesis, each surface has a specific performance in template removal. For poly(scopoletin)-based MIPs for the huge globular protein ferritin, an increase in the electroactive surface prevents the preparation of effective MIP layers either by blocking template removal or by rebinding. However, the MIP formed on a flat glassy carbon electrode has an analytical performance, i.e., measuring range (up to 0.5 µM and Kd 0.30 µM) comparable to that of the MIP on the gold surface of an SPR chip.

Funding

This study was supported by Turkish—German University Scientific Research Projects Commission under the grant no: 2024BF04.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The author would like to especially thank Frieder W. Scheller for fruitful discussions.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Yarman, A.; Scheller, F.W. How Reliable Is the Electrochemical Readout of MIP Sensors? Sensors 2020, 20, 2677. [Google Scholar] [CrossRef] [PubMed]
  2. Sharma, P.S.; Pietrzyk-Le, A.; D’Souza, F.; Kutner, W. Electrochemically Synthesized Polymers in Molecular Imprinting for Chemical Sensing. Anal. Bioanal. Chem. 2012, 402, 3177–3204. [Google Scholar] [CrossRef] [PubMed]
  3. Chianella, I.; Guerreiro, A.; Moczko, E.; Caygill, J.S.; Piletska, E.V.; De Vargas Sansalvador, I.M.P.; Whitcombe, M.J.; Piletsky, S.A. Direct Replacement of Antibodies with Molecularly Imprinted Polymer Nanoparticles in ELISA—Development of a Novel Assay for Vancomycin. Anal. Chem. 2013, 85, 8462–8468. [Google Scholar] [CrossRef] [PubMed]
  4. Ozcelikay, G.; Kurbanoglu, S.; Yarman, A.; Scheller, F.W.; Ozkan, S.A. Au-Pt Nanoparticles Based Molecularly Imprinted Nanosensor for Electrochemical Detection of the Lipopeptide Antibiotic Drug Daptomycin. Sens. Actuators B Chem. 2020, 320, 128285. [Google Scholar] [CrossRef]
  5. Chunta, S.; Suedee, R.; Boonsriwong, W.; Lieberzeit, P.A. Biomimetic Sensors Targeting Oxidized-Low-Density Lipoprotein with Molecularly Imprinted Polymers. Anal. Chim. Acta 2020, 1116, 27–35. [Google Scholar] [CrossRef] [PubMed]
  6. Schirhagl, R.; Podlipna, D.; Lieberzeit, P.A.; Dickert, F.L. Comparing Biomimetic and Biological Receptors for Insulin Sensing. Chem. Commun. 2010, 46, 3128–3130. [Google Scholar] [CrossRef]
  7. Erdossy, J.; Horváth, V.; Yarman, A.; Scheller, F.W.; Gyurcsányi, R.E. Electrosynthesized Molecularly Imprinted Polymers for Protein Recognition. TrAC—Trends Anal. Chem. 2016, 79, 179–190. [Google Scholar] [CrossRef]
  8. Latif, U.; Seifner, A.; Dickert, F.L. Selective Detection of Erythrocytes with QCMs—ABO Blood Group Typing. Sensors 2023, 23, 7533. [Google Scholar] [CrossRef]
  9. Erol, K.; Hasabnis, G.; Altintas, Z. A Novel NanoMIP–SPR Sensor for the Point-of-Care Diagnosis of Breast Cancer. Micromachines 2023, 14, 1086. [Google Scholar] [CrossRef]
  10. Saylan, Y.; Kılıç, S.; Denizli, A. Biosensing Applications of Molecularly Imprinted-Polymer-Based Nanomaterials. Processes 2024, 12, 177. [Google Scholar] [CrossRef]
  11. Sarvutiene, J.; Prentice, U.; Ramanavicius, S.; Ramanavicius, A. Molecular Imprinting Technology for Biomedical Applications. Biotechnol. Adv. 2024, 71, 108318. [Google Scholar] [CrossRef] [PubMed]
  12. Ramanaviciene, A.; Ramanavicius, A. Molecularly Imprinted Polypyrrole-Based Synthetic Receptor for Direct Detection of Bovine Leukemia Virus Glycoproteins. Biosens. Bioelectron. 2004, 20, 1076–1082. [Google Scholar] [CrossRef]
  13. Tse Sum Bui, B.; Mier, A.; Haupt, K. Molecularly Imprinted Polymers as Synthetic Antibodies for Protein Recognition: The Next Generation. Small 2023, 19, 2206453. [Google Scholar] [CrossRef] [PubMed]
  14. Mazzotta, E.; Di Giulio, T.; Malitesta, C. Electrochemical Sensing of Macromolecules Based on Molecularly Imprinted Polymers: Challenges, Successful Strategies, and Opportunities. Anal. Bioanal. Chem. 2022, 414, 5165–5200. [Google Scholar] [CrossRef]
  15. Di Giulio, T.; Mazzotta, E.; Malitesta, C. Molecularly Imprinted Polyscopoletin for the Electrochemical Detection of the Chronic Disease Marker Lysozyme. Biosensors 2021, 11, 3. [Google Scholar] [CrossRef] [PubMed]
  16. Turco, A.; Corvaglia, S.; Pompa, P.P.; Malitesta, C. An Innovative and Simple All Electrochemical Approach to Functionalize Electrodes with a Carbon Nanotubes/Polypyrrole Molecularly Imprinted Nanocomposite and Its Application for Sulfamethoxazole Analysis. J. Colloid. Interface Sci. 2021, 599, 676–685. [Google Scholar] [CrossRef]
  17. Bossi, A.M.; Maniglio, D. BioMIPs: Molecularly Imprinted Silk Fibroin Nanoparticles to Recognize the Iron Regulating Hormone Hepcidin. Microchim. Acta 2022, 189, 66. [Google Scholar] [CrossRef]
  18. Uzun, L.; Turner, A.P.F. Molecularly-Imprinted Polymer Sensors: Realising Their Potential. Biosens. Bioelectron. 2016, 76, 131–144. [Google Scholar] [CrossRef]
  19. Zhang, X.; Yarman, A.; Kovács, N.; Bognár, Z.; Gyurcsányi, R.E.; Bier, F.F.; Scheller, F.W. Specific Features of Epitope-MIPs and Whole-Protein MIPs as Illustrated for AFP and RBD of SARS-CoV-2. Microchim. Acta 2024, 191, 242. [Google Scholar] [CrossRef]
  20. Li, Y.; Luo, L.; Kong, Y.; Li, Y.; Wang, Q.; Wang, M.; Li, Y.; Davenport, A.; Li, B. Recent Advances in Molecularly Imprinted Polymer-Based Electrochemical Sensors. Biosens. Bioelectron. 2024, 249, 116018. [Google Scholar] [CrossRef]
  21. Bartold, K.; Iskierko, Z.; Sharma, P.S.; Lin, H.-Y.; Kutner, W. Idiopathic Pulmonary Fibrosis (IPF): Diagnostic Routes Using Novel Biomarkers. Biomed. J. 2024, 100729. [Google Scholar] [CrossRef] [PubMed]
  22. Wulff, G.; Sarhan, A. Macromolecular Colloquium. Angew. Chem. Int. Ed. Engl. 1972, 11, 334–342. [Google Scholar] [CrossRef]
  23. Arshady, R.; Mosbach, K. Synthesis of Substrate-Selective Polymers by Host-Guest Polymerizatioa. Die Makromol.Chem. Macromol. Chem. Phys. 1981, 692, 687–692. [Google Scholar] [CrossRef]
  24. Yarman, A.; Kurbanoglu, S.; Zebger, I.; Scheller, F.W. Simple and Robust: The Claims of Protein Sensing by Molecularly Imprinted Polymers. Sens. Actuators B Chem. 2021, 330, 129369. [Google Scholar] [CrossRef]
  25. Turner, N.W.; Jeans, C.W.; Brain, K.R.; Allender, C.J.; Hlady, V.; Britt, D.W. From 3D to 2D: A Review of the Molecular Imprinting of Proteins. Biotechnol. Prog. 2006, 22, 1474–1489. [Google Scholar] [CrossRef]
  26. Pluhar, B.; Ziener, U.; Mizaikoff, B. Surface Imprinting of Pepsin via Miniemulsion Polymerization. J. Mater. Chem. B 2013, 1, 5489–5495. [Google Scholar] [CrossRef] [PubMed]
  27. Van Ho Phan, N.; Sussitz, H.F.; Ladenhauf, E.; Pum, D.; Lieberzeit, P.A. Combined Layer/Particle Approaches in Surface Molecular Imprinting of Proteins: Signal Enhancement and Competition. Sensors 2018, 18, 180. [Google Scholar] [CrossRef] [PubMed]
  28. Lin, H.Y.; Hsu, C.Y.; Thomas, J.L.; Wang, S.E.; Chen, H.C.; Chou, T.C. The Microcontact Imprinting of Protein: The Effect of Cross-Linking Monomers for Lysozyme, Ribonuclease A and Myoglobin. Biosens. Bioelectron. 2006, 22, 534–543. [Google Scholar] [CrossRef]
  29. Hayden, O.; Lieberzeit, P.A.; Blaas, D.; Dickert, F.L. Artificial Antibodies for Bioanalyte Detection—Sensing Viruses and Proteins. Adv. Funct. Mater. 2006, 16, 1269–1278. [Google Scholar] [CrossRef]
  30. Ayankojo, A.G.; Boroznjak, R.; Reut, J.; Öpik, A.; Syritski, V. Molecularly Imprinted Polymer Based Electrochemical Sensor for Quantitative Detection of SARS-CoV-2 Spike Protein. Sens. Actuators B Chem. 2022, 353, 131160. [Google Scholar] [CrossRef]
  31. Bosserdt, M.; Gajovic-Eichelman, N.; Scheller, F.W. Modulation of Direct Electron Transfer of Cytochrome c by Use of a Molecularly Imprinted Thin Film. Anal. Bioanal. Chem. 2013, 405, 6437–6444. [Google Scholar] [CrossRef]
  32. Zhang, X.; Yarman, A.; Erdossy, J.; Katz, S.; Zebger, I.; Jetzschmann, K.J.; Altintas, Z.; Wollenberger, U.; Gyurcsányi, R.E.; Scheller, F.W. Electrosynthesized MIPs for Transferrin: Plastibodies or Nano-Filters? Biosens. Bioelectron. 2018, 105, 29–35. [Google Scholar] [CrossRef]
  33. Dechtrirat, D. Combination of Self-Assembled Monolayers (SAMs) and Molecularly Imprinted Polymers (MIPs) in Biomimetic Sensors; University of Potsdam: Potsdam, Germany, 2013. [Google Scholar]
  34. Knovich, M.A.; Storey, J.A.; Coffman, L.G.; Torti, S.V.; Torti, F.M. Ferritin for the Clinician. Blood Rev. 2009, 23, 95–104. [Google Scholar] [CrossRef]
  35. Garcia-Casal, M.N.; Peña-Rosas, J.P.; Urrechaga, E.; Escanero, J.F.; Huo, J.; Martinez, R.X.; Lopez-Perez, L. Performance and Comparability of Laboratory Methods for Measuring Ferritin Concentrations in Human Serum or Plasma: A Systematic Review and Meta-Analysis. PLoS ONE 2018, 13, e0196576. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, X.; Yarman, A.; Bagheri, M.; El-Sherbiny, I.M.; Hassan, R.Y.A.; Kurbanoglu, S.; Waffo, A.F.T.; Zebger, I.; Karabulut, T.C.; Bier, F.F.; et al. Imprinted Polymers on the Route to Plastibodies for Biomacromolecules (MIPs), Viruses (VIPs), and Cells (CIPs). In Trends in Biosensing Research: Advances, Challenges and Applications; Springer: Berlin/Heidelberg, Germany, 2023; pp. 107–148. [Google Scholar]
  37. WHO. Use of Ferritin Concentrations to Assess Iron Status in Individuals and Populations. Available online: https://www.who.int/publications/i/item/9789240000124 (accessed on 10 May 2024).
  38. Bosserdt, M.; Erdossy, J.; Lautner, G.; Witt, J.; Köhler, K.; Gajovic-Eichelmann, N.; Yarman, A.; Wittstock, G.; Scheller, F.W.; Gyurcsányi, R.E. Microelectrospotting as a New Method for Electrosynthesis of Surface-Imprinted Polymer Microarrays for Protein Recognition. Biosens. Bioelectron. 2015, 73, 123–129. [Google Scholar] [CrossRef]
  39. Goldoni, R.; Thomaz, D.V.; Di Giulio, T.; Malitesta, C.; Mazzotta, E. An Insight into Polyscopoletin Electrosynthesis by a Quality-by-Design Approach. J. Mater. Sci. 2022, 57, 12161–12175. [Google Scholar] [CrossRef] [PubMed]
  40. Paimard, G.; Ghasali, E.; Baeza, M. Screen-Printed Electrodes: Fabrication, Modification, and Biosensing Applications. Chemosensors 2023, 11, 113. [Google Scholar] [CrossRef]
  41. Couto, R.A.S.; Lima, J.L.F.C.; Quinaz, M.B. Recent Developments, Characteristics and Potential Applications of Screen-Printed Electrodes in Pharmaceutical and Biological Analysis. Talanta 2016, 146, 801–814. [Google Scholar] [CrossRef]
  42. Mansuriya, B.D.; Altintas, Z. Carbon Nanomaterials for Sensing Applications. In Fundamentals of Sensor Technology; Elsevier: Amsterdam, The Netherlands, 2023; pp. 367–400. [Google Scholar]
  43. Ferrier, D.C.; Honeychurch, K.C. Carbon Nanotube (CNT)-Based Biosensors. Biosensors 2021, 11, 486. [Google Scholar] [CrossRef]
  44. Pandey, R.R.; Chusuei, C.C. Carbon Nanotubes, Graphene, and Carbon Dots as Electrochemical Biosensing Composites. Molecules 2021, 26, 6674. [Google Scholar] [CrossRef]
  45. Preda, D.; David, I.G.; Popa, D.-E.; Buleandra, M.; Radu, G.L. Recent Trends in the Development of Carbon-Based Electrodes Modified with Molecularly Imprinted Polymers for Antibiotic Electroanalysis. Chemosensors 2022, 10, 243. [Google Scholar] [CrossRef]
  46. Bou-Maroun, E. Carbon Electrode Modified with Molecularly Imprinted Polymers for the Development of Electrochemical Sensor: Application to Pharmacy, Food Safety, Environmental Monitoring, and Biomedical Analysis. Chemosensors 2023, 11, 548. [Google Scholar] [CrossRef]
  47. Li, T.; Deng, Z.; Bu, J.; Liu, H.; Yang, Y.; Zhong, S. Quantum Dot Based Molecularly Imprinted Polymer Test Strips for Fluorescence Detection of Ferritin. Sens. Actuators B Chem. 2022, 358, 131548. [Google Scholar] [CrossRef]
  48. Stojanovic, Z.; Erdőssy, J.; Keltai, K.; Scheller, F.W.; Gyurcsányi, R.E. Electrosynthesized Molecularly Imprinted Polyscopoletin Nanofilms for Human Serum Albumin Detection. Anal. Chim. Acta 2017, 977, 1–9. [Google Scholar] [CrossRef] [PubMed]
  49. Cai, D.; Ren, L.; Zhao, H.; Xu, C.; Zhang, L.; Yu, Y.; Wang, H.; Lan, Y.; Roberts, M.F.; Chuang, J.H.; et al. A Molecular-Imprint Nanosensor for Ultrasensitive Detection of Proteins. Nat. Nanotechnol. 2010, 5, 597–601. [Google Scholar] [CrossRef] [PubMed]
  50. Yoshimi, Y.; Ohdaira, R.; Iiyama, C.; Sakai, K. ‘Gate Effect’ of Thin Layer of Molecularly-Imprinted Poly(Methacrylic Acid-Co-Ethyleneglycol Dimethacrylate). Sens. Actuators B Chem. 2001, 73, 49–53. [Google Scholar] [CrossRef]
  51. Stephen, A.N.; Dennison, S.R.; Holden, M.A.; Reddy, S.M. Rapid Sub-Nanomolar Protein Determination in Serum Using Electropolymerized Molecularly Imprinted Polymers (E-MIPs). Analyst 2023, 148, 5476–5485. [Google Scholar] [CrossRef] [PubMed]
  52. Tretjakov, A.; Syritski, V.; Reut, J.; Boroznjak, R.; Volobujeva, O.; Öpik, A. Surface Molecularly Imprinted Polydopamine Films for Recognition of Immunoglobulin G. Microchim. Acta 2013, 180, 1433–1442. [Google Scholar] [CrossRef]
  53. Zuo, J.; Zhao, X.; Ju, X.; Qiu, S.; Hu, W.; Fan, T.; Zhang, J. A New Molecularly Imprinted Polymer (MIP)-based Electrochemical Sensor for Monitoring Cardiac Troponin I (CTnI) in the Serum. Electroanalysis 2016, 28, 2044–2049. [Google Scholar] [CrossRef]
  54. Hassine, A.B.; Raouafi, N.; Moreira, F.T.C. Novel Electrochemical Molecularly Imprinted Polymer-Based Biosensor for Tau Protein Detection. Chemosensors 2021, 9, 238. [Google Scholar] [CrossRef]
  55. Rebelo, T.S.C.R.; Pereira, C.M.; Sales, M.G.F.; Noronha, J.P.; Silva, F. Protein Imprinted Material Electrochemical Sensor for Determination of Annexin A3 in Biological Samples. Electrochim. Acta 2016, 190, 887–893. [Google Scholar] [CrossRef]
  56. Silva, B.V.M.; Rodríguez, B.A.G.; Sales, G.F.; Sotomayor, M.D.P.T.; Dutra, R.F. An Ultrasensitive Human Cardiac Troponin T Graphene Screen-Printed Electrode Based on Electropolymerized-Molecularly Imprinted Conducting Polymer. Biosens. Bioelectron. 2016, 77, 978–985. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Schematic representation of the preparation and characterization of the ferritin–MIP sensor. GCE: glassy carbon electrode; SPE: screen-printed electrode (Blue curve: Without ferritin and Red Curve: With ferritin).
Scheme 1. Schematic representation of the preparation and characterization of the ferritin–MIP sensor. GCE: glassy carbon electrode; SPE: screen-printed electrode (Blue curve: Without ferritin and Red Curve: With ferritin).
Biomimetics 09 00426 sch001
Figure 1. MIP film formation showing current–time curves for 0.5 mM scopoletin solution containing 10 µM ferritin (0 V for 1 s, 0.9 V for 0.1 s, 100 cycles).
Figure 1. MIP film formation showing current–time curves for 0.5 mM scopoletin solution containing 10 µM ferritin (0 V for 1 s, 0.9 V for 0.1 s, 100 cycles).
Biomimetics 09 00426 g001
Figure 2. DPVs of SPC electrodes in 5 mM ferricyanide/ferrocyanide solution (in 100 mM KCl): bare electrode (black), after electropolymerization (red), and after electrochemical removal from HEPES buffer (2 × 120; blue).
Figure 2. DPVs of SPC electrodes in 5 mM ferricyanide/ferrocyanide solution (in 100 mM KCl): bare electrode (black), after electropolymerization (red), and after electrochemical removal from HEPES buffer (2 × 120; blue).
Biomimetics 09 00426 g002
Figure 3. DPVs of SWSP electrodes in 5 mM ferricyanide/ferrocyanide solution (in 100 mM KCl): bare electrode (black), after electropolymerization (red), after overnight incubation in glycine–HCl buffer (50 mM, pH 2.2) (blue), after electrochemical removal from glycine–HCl buffer (50 mM, pH 2.2) using cyclic voltammetry within the potential range of −0.2 V to 1.0 V (50 scans) (green), and after 1 µM ferritin binding for 1 h (purple).
Figure 3. DPVs of SWSP electrodes in 5 mM ferricyanide/ferrocyanide solution (in 100 mM KCl): bare electrode (black), after electropolymerization (red), after overnight incubation in glycine–HCl buffer (50 mM, pH 2.2) (blue), after electrochemical removal from glycine–HCl buffer (50 mM, pH 2.2) using cyclic voltammetry within the potential range of −0.2 V to 1.0 V (50 scans) (green), and after 1 µM ferritin binding for 1 h (purple).
Biomimetics 09 00426 g003
Figure 4. CVs of GCEs in 5 mM ferricyanide/ferrocyanide solution (in 100 mM KCl): bare electrode (black), after electropolymerization (red), after overnight incubation in 5 mM NaOH (blue), and after 0.25 µM ferritin binding for 1 h (green).
Figure 4. CVs of GCEs in 5 mM ferricyanide/ferrocyanide solution (in 100 mM KCl): bare electrode (black), after electropolymerization (red), after overnight incubation in 5 mM NaOH (blue), and after 0.25 µM ferritin binding for 1 h (green).
Biomimetics 09 00426 g004
Figure 5. (A,B) Concentration dependence of relative current decrease for the binding of ferritin to MIP–GCE based on differential pulse voltammetry; (C,D) fluorescence titration of scopoletin (0.5 mM) with ferritin (1 mM).
Figure 5. (A,B) Concentration dependence of relative current decrease for the binding of ferritin to MIP–GCE based on differential pulse voltammetry; (C,D) fluorescence titration of scopoletin (0.5 mM) with ferritin (1 mM).
Biomimetics 09 00426 g005
Table 1. Examples of MIPs for ferritin sensing.
Table 1. Examples of MIPs for ferritin sensing.
MaterialsMethodLinearity RangeReference
Quantum dot/ferritin-imprinted nanoparticle-modified test stripsFluorescence spectroscopy1.0–6.0 µM[47]
Ferritin-imprinted poly(phenol) film on CNT-array-modified titanium-coated glass substrateEIS2 × 10−17–1 × 10−12 M[49]
Ferritin-imprinted poly(scopoletin) film on an Au electrodeCV0.9–8 µM[48]
Ferritin-imprinted poly(scopoletin) film on an Au chipSPRup to 0.5 µM[38]
Ferritin-imprinted poly(scopoletin) film on a GCEDPVup to 0.5 µMThis work
SPR: surface plasmon resonance; DPV: differential pulse voltammetry; GCE: glassy carbon electrode.
Table 2. Electrosynthesized MIPs for proteins with redox markers.
Table 2. Electrosynthesized MIPs for proteins with redox markers.
TemplateMonomerElectrodeLinearity RangeKdReference
TransferrinScopoletinAu0.1–1 µM0.56 µM[32]
HemoglobinNHMASPE0.1–1 nM0.86 ± 0.11 nM[51]
LysozymeScopoletinAu150 nM–20 µM
(MIP 2)
-[15]
IgGDopamineSAM-modified QCM chip-296 nM[52]
Cardiac troponin Io-APGCE0.05–5.0 nM-[53]
p-Tau-441AMPSPE2.18 pM–2.18 nM-[54]
Annexin A3Caffeic acidSPE2.8 pM–5.5 nM-[55]
Troponin TPyrroleSPE0.26–2.6 pM0.7 pM[56]
NHMA: N-hydroxymethylacrylamide; SPE: screen-printed electrode; QCM: quartz crystal microbalance; SAM: self-assembled monolayer; MIP 2: molecularly imprinted polymer (two different methods have been applied for MIP synthesis); GCE: glassy carbon electrode; o-PD: o-phenylenediamine; o-AP: o-aminophenol; AMP: 3-aminophenol.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yarman, A. Effect of Various Carbon Electrodes on MIP-Based Sensing Proteins Using Poly(Scopoletin): A Case Study of Ferritin. Biomimetics 2024, 9, 426. https://doi.org/10.3390/biomimetics9070426

AMA Style

Yarman A. Effect of Various Carbon Electrodes on MIP-Based Sensing Proteins Using Poly(Scopoletin): A Case Study of Ferritin. Biomimetics. 2024; 9(7):426. https://doi.org/10.3390/biomimetics9070426

Chicago/Turabian Style

Yarman, Aysu. 2024. "Effect of Various Carbon Electrodes on MIP-Based Sensing Proteins Using Poly(Scopoletin): A Case Study of Ferritin" Biomimetics 9, no. 7: 426. https://doi.org/10.3390/biomimetics9070426

Article Metrics

Back to TopTop