Previous Article in Journal
Investigating the Effect of Pipecolic Acid on Specialized Metabolites Involved in Tomato Plant Defense Mechanisms Against Ralstonia solanacearum Wilt Pathogens
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Development of Novel Conductive Inks for Screen-Printed Electrochemical Sensors: Enhancing Rapid and Sensitive Drug Detection

by
Victor Alexandre Ribeiro Leite
,
Sthephane Pereira de Oliveira
,
Larissa Cristina de Souza
,
Léa Júlia de Paula Silva
,
Laís Fonseca Silva
,
Thaís Cristina de Oliveira Cândido
,
Daniela Nunes da Silva
and
Arnaldo César Pereira
*
Natural Sciences Department, Federal University of São João del Rei, 74 Praça Dom Helvécio, São João del Rei 36301-160, MG, Brazil
*
Author to whom correspondence should be addressed.
Analytica 2025, 6(1), 3; https://doi.org/10.3390/analytica6010003
Submission received: 30 October 2024 / Revised: 16 December 2024 / Accepted: 9 January 2025 / Published: 11 January 2025

Abstract

:
The development of screen-printed electrochemical sensors represents a rapidly expanding research field with great potential for applications in the rapid and sensitive determination of drugs in complex matrices. This work presents a review of the state-of-the-art examples of this technology, focusing on its application in real matrices such as water, pharmaceutical formulations, and biological fluids. We discuss the main materials used in developing conductive inks, highlighting their properties and influence on sensor performance. The characterization of materials and sensors is crucial to ensure the reproducibility and reliability of results. Additionally, we address the challenges associated with the application of these sensors in complex matrices, such as interferences from other components and the need for sample pretreatment. Finally, we present future perspectives for developing screen-printed electrochemical sensors, with an emphasis on new technologies and materials that can improve the sensitivity, selectivity, and stability of these devices.

1. Introduction

The pharmaceutical industry has evolved significantly, driven by the growing global demand and the complexity of human diseases. Although historical treatments often lacked scientific rigor, modern medicine offers effective solutions, even for common diseases. Over-the-counter (OTC) medications, such as nonsteroidal anti-inflammatory drugs (NSAIDs), can be harmful when used incorrectly. Self-medication, influenced by various factors, poses significant health risks due to potential overdoses, drug interactions, and adverse effects [1].
The discovery of penicillin in 1928 revolutionized the treatment of bacterial infections. However, the indiscriminate use of antibiotics has led to the emergence of drug-resistant strains, such as methicillin-resistant Staphylococcus aureus (MRSA) and vancomycin-resistant Staphylococcus aureus 1 (VRSA), highlighting the need for responsible use [2]. The World Health Organization (WHO) warns that the inappropriate use of antibiotics is the main factor in developing bacterial resistance, emphasizing the need for control measures to ensure the rational use of these drugs [3,4].
Given the indiscriminate use of drugs, the quantitative analysis of these substances in various matrices (biological fluids, the environment, effluents) becomes essential to monitor exposure, evaluate their therapeutic efficacy, and mitigate the risks to health and the environment [5].
The literature describes various analytical methods, such as chromatography, spectrophotometry, and electroanalytical methods, for the determination of drugs in different matrices. Among them, electroanalytical methods stand out due to their characteristics, which include a rapid response, portability, a low cost, and a lower environmental impact, being frequently used due to their practicality and efficiency [6,7,8,9]. In addition, electrodes offer a high sensitivity, selectivity, and economic viability. This article aims to discuss innovations in printed sensors by applying a screen-printing technique for drug detection. The screen-printing process, the materials used in the composition of conductive inks, and the potential applications will be highlighted.

2. Electrochemical Sensors

Electrodes are crucial elements in electrochemical analyses, allowing the investigation of charge transfer processes at interfaces. The electrochemical cell is composed of a working electrode (WE), a reference electrode (RE), and a counter electrode (CE) immersed in an electrolyte solution, which enables the quantification of analytes through electrochemical measurements [10,11,12,13,14].
The working electrode (WE), where the oxidation or reduction of the analyte occurs, must be composed of an inert material that covers the potential range of the analysis. The reference electrode (RE) maintains a constant and stable potential during measurements. The counter electrode or auxiliary electrode (CE) directs the current generated directly to the working electrode (WE), minimizing current variations during the experiment. The three-electrode system is the most widely used, as it allows analyses in more dilute support solutions and electrolytes with a higher electrical resistance. Figure 1 illustrates a system of an electrochemical cell containing three electrodes (WE, RE, and CE) immersed in an electrolyte solution [14,15,16].
Printed electrodes are devices manufactured using printing techniques, such as screen-printing and 3D printing, enabling various types of applications. Unlike conventional electrodes, these devices present a complete system on a single support, containing the working, auxiliary, and quasi-reference electrodes, as illustrated in Figure 2 [17].
These electrodes were developed as an alternative to conventional electrodes due to their numerous advantages, such as the possibility of miniaturization, a good sensitivity and selectivity, and a low cost [18,19]. They can be manufactured on a large scale and, being disposable, eliminate the need for cleaning, making the analysis process faster and more practical. In addition, they minimize the cross-contamination effects between samples, increasing the reliability of the analytical results [20].
Printed electrodes, manufactured by screen-printing, find wide applications in the electrochemical detection of drugs. The versatility of conductive inks, composed of conductive particles (graphene, carbon nanotubes, carbon black), polymeric binders, and solvents, allows the creation of customized sensors for various molecules of interest. Therefore, these devices offer portable and low-cost solutions for health monitoring, quality control, and clinical analyses [21,22,23,24,25,26,27,28,29,30].

2.1. Printing Electrodes Using the Screen-Printed Technique

Screen-printing has existed for over 2000 years, with rudimentary fabric-printing methods having been developed. In the 18th century, the technique gained popularity in Europe, especially for printing fabrics. Over the years, it has been modernized with the use of synthetic inks and polyester screens, becoming widely used for printing posters, labels, signs, bags, and electronic circuits. Today, this technique continues to evolve and is used in various applications, including the fabrication of printed electrodes and electronic devices [31].
The production of printed electrodes using screen-printing is a widely used technique in fabricating electronic devices, sensors, and biosensors. This method allows for the creation of electrodes with specific electrical properties and the customization of their geometries to meet the needs of different applications. The technique is easily applicable in any laboratory, as it does not require complex equipment, and several stages of the fabrication process can be adjusted using alternative materials for manual production [32].
The fabrication process of these devices involves several steps: substrate preparation, electrode design, conductive material deposition, drying, and curing [33]. Electrode designs are created digitally and optimized for size using software, and are then cut onto an adhesive sheet using a plotter. Conductive inks, often based on carbon materials, nanomaterials, or conductive polymers, are screen-printed onto the substrate using a squeegee to fill the stencil uniformly and achieve the desired thickness [34]. Afterward, these devices undergo a drying process that can occur at room temperature or under controlled conditions, depending on the ink composition and the desired electrode properties. The printed electrodes can also be subjected to additional treatments, such as the application of protective coatings or surface modifications, to increase the sensor’s sensitivity and selectivity [34].
Therefore, screen-printing applied to electrode production has great potential for continued advancement, driven by the increasing demand for more cost-effective, efficient, and customized devices in various sectors of technology and healthcare. Figure 3 illustrates the fabrication process of printed electrodes using the screen-printing technique.

2.2. Conductive Inks for Printing Electrodes

With the advancement of technology and the search for more sustainable solutions, conductive inks have become a promising alternative in various areas, such as printed electronics, sensors, and wearable devices [26]. Their formulations are mainly made up of three main components: conductive particles that transport the charge, a binder responsible for aggregating the particulate material, and a solvent that homogenizes and solubilizes the components (Figure 4) [28].
The most used conductive materials include metal particles, conductive polymers, and carbonaceous materials. Although metal-based conductive inks are more conductive, they are more expensive, less biocompatible, and more unstable to oxidation than carbon-based inks [27,35,36]. Conductive polymers have a good conductivity and mechanical stability, but certain polymers are sensitive to changes in environmental parameters such as temperature, humidity, and pH [37]. Carbon-based inks have excellent properties due to the characteristics of carbonaceous materials, including an excellent electrical and thermal conductivity, chemical stability, a good surface area, a reasonable cost, and easy accessibility [26,27,28,38]. The following topics will discuss the main materials used in developing conductive inks, subdivided into carbonaceous materials, binders, and solvents.

2.2.1. Carbonaceous Materials

Carbonaceous materials have been widely used in the development of conductive inks for printing electrodes. Graphite, carbon nanotubes (CNTs), and graphene have been widely used in the development of inks [39,40]. Graphite is valued for its high purity and low cost and is often used in the production of conductive inks [41]. Verma et al. (2024) produced a conductive ink based on graphite, shellac as a binding agent, and deionized water as a solvent. The proposed ink showed a good homogeneity and viscosity suitable for screen-printing. The devices manufactured with this ink exhibited an excellent electrical conductivity and good adhesion to various types of surfaces. The authors highlighted the sensitivity and reproducibility of the developed sensors, demonstrating the properties that graphite confers to the ink [42].
Carbon nanotubes (CNTs) [43] and graphene [44] are characterized by superior properties compared to other carbonaceous materials [26]. CNTs are cylindrical carbon structures with exceptional electrical and mechanical properties, such as a high electrical conductivity and a large surface area. Their structure, consisting of coiled layers of graphene, can vary from single-walled to multi-walled, which influences their properties [45]. Due to these properties, CNTs have been widely used in developing conductive inks, offering a high conductivity and flexibility. This versatility makes them suitable for fabricating various flexible and wearable electronic devices [45,46]. In this context, Park et al. (2024) used single-walled carbon nanotubes to prepare a highly conductive ink with suitable rheological properties. The ink was composed of a combination of single-walled carbon nanotubes, cellulose, and water. The results of their research showed that the developed ink exhibited a high conductivity due to the alignment of the single-walled carbon nanotubes in the cellulose fibril networks. This alignment was achieved by interfibrillar sliding, which significantly improved the electrical conductivity of the ink [47].
Another material that stands out is graphene, a two-dimensional carbon material with a hexagonal structure that has favorable physical and chemical properties such as a large surface area, an excellent electrical conductivity, a high mechanical strength, thermal conductivity, and ease of surface functionalization [48]. The use of graphene was successfully employed by Kim et al. (2021) in preparing conductive ink. The authors reported that the ink was prepared by the exfoliation process, in which a mixture of graphite and ethyl cellulose (9:1 by weight) was dispersed in terpineol/ethanol (5:5 by volume) and loaded into a fluid dynamic reactor. The intense shear and cavitation forces in the reactor enabled the rapid and efficient production of high-quality graphene ink. This ink was used to fabricate printed sodium ion sensors, demonstrating an excellent detection performance by monitoring the concentration of sodium ions in sweat in real time during exercise. In addition, the ink was found to be highly conductive, and its consistency enabled the production of screen-printed sensors with a high sensitivity, a good reproducibility and selectivity, and a fast response time [27].
Another approach to the production of conductive inks involves the combination of carbonaceous materials, which can result in improvements in analytical performance. In this sense, Poulin et al. (2021) presented the development of a conductive ink based on graphite and carbon black, resulting in a high electrical conductivity. Graphite gives materials flexibility, while carbon black contributes to forming an efficient conductive network. In addition to a high conductivity, the ink demonstrated water resistance and temperature variations, making it ideal for various applications. This combination of properties makes ink a promising alternative for manufacturing flexible electronic devices. However, for these inks to be applied to different surfaces and under different conditions, it is essential to have a component that can provide the necessary cohesion and adhesion for the conductive particles [49].

2.2.2. Binders

A binder has the function of agglomerating the particulate material, giving the ink properties such as adhesion, durability, and resistance. The selection of the most suitable binder is contingent upon the specific application and the desired characteristics of the ink, including conductivity, flexibility, and chemical resistance. A range of binder types can be employed, with each exhibiting distinctive properties [27,45].
A literature review revealed that the vast array of existing binders has made possible the development of a diverse range of ink formulations. These include cellulose acetate [50], ethyl cellulose [51], polymeric vinyl acetate [52], polyurethane [53], and others. In a study by Carvalho et al. (2021), a high-quality conductive ink was developed using graphite and cellulose acetate, dispersed in an organic solvent to produce screen-printed sensors. The role of cellulose acetate in forming a polymer matrix, agglomerating the graphite particles, and imparting a gel-like consistency to the ink was pivotal in facilitating the application and formation of a conductive surface after drying. Ultimately, the authors reported the successful development of low-cost sensors that demonstrated an excellent performance, including a wide linear range, a high sensitivity, and a low detection limit. These sensors were shown to be effective in determining levodopa (L-dopa) [54].
The versatility in the choice of binders for conductive ink production is evidenced in the literature, which presents the use of alternative binders such as stained-glass varnish [55], nail polish [56], and chitosan [57] in the production of conductive inks. These materials, in addition to being more commercially accessible, exhibit suitable properties for forming conductive films. The search for binders for conductive inks has increasingly turned to green chemistry. Materials such as shellac stand out for their biocompatible and biodegradable characteristics, making them a promising option. Henrique et al. (2021) demonstrated the potential of shellac in developing ink for printing electrochemical sensors. The authors prepared a conductive ink using shellac powder, graphite, and polyurethane solvent. The resulting sensor was effective in detecting the antibiotic sulfamethoxazole, exhibiting a wide linear range and a low detection limit. Shellac, as a natural bioadhesive polymer, contributed to the simplicity and low cost of the sensor, as well as providing desirable characteristics such as biocompatibility and biodegradability. The results obtained by the authors demonstrate the feasibility of using biodegradable binders in electrochemical sensors, paving the way for the development of more sustainable and efficient technologies [58].

2.2.3. Solvents

Another essential component for the formulation of conductive inks is the solvent. The solvent plays a fundamental role in the production of conductive inks, acting as the medium for making the mixture liquid, facilitating application, and ensuring the uniform dispersion of the components. Without the solvent, the production and application of these inks would become more complex and less efficient [27]. The selection of the solvent is crucial to ensure efficient interactions of the conductive ink components. Various organic solvents, such as acetone [59], cyclohexanone [50], mixtures of toluene and ethanol [51], and dimethylformamide [60], among others, are employed for this purpose. Although organic solvents have been traditionally used, concerns about human health and the environment have led to the exploration of green solvents as a safer and more sustainable alternative [61].
Within the context of green chemistry, which seeks to minimize the environmental impact of chemical processes, the use of water as a solvent stands out. In addition to being abundant and low-cost, water is non-flammable and non-toxic. These attributes make it a promising alternative for producing conductive inks, driving the search for more efficient and sustainable procedures [62]. In this context, Chung et al. (2024) presented a significant advancement by developing a water-based conductive ink composed of gallium, indium, and carbon nanotubes. This formulation demonstrated a high stability and compatibility with various flexible substrates, allowing the fabrication of e-textiles, which are fabrics that incorporate electronic components, such as sensors or circuits, directly into their structure. The produced e-textiles exhibited a high conductivity, flexibility, durability, and biocompatibility, combining a superior mechanical and electrical performance with comfort and safety. These promising results pave the way for new applications of water as an alternative solvent [63].
Different ink compositions can be produced using various types of materials, depending on the desired application. However, the search for more effective and environmentally friendly formulations continues, driven by the need to meet the requirements of increasingly sophisticated applications [26,27,28,63,64].
In addition to the ink composition, the other factors that can influence the electrode printing process are the viscosity and homogeneity of the ink. Viscosity is an essential characteristic of all fluids, as it determines a liquid’s resistance to flow or shear deformation. This resistance arises from cohesive intermolecular forces, which are responsible for the friction between adjacent layers of the fluid that move relative to each other. Several factors, such as the temperature and pressure, influence the viscosity, making it one of the most important properties in the performance of inks [65].
Homogeneity is related to the uniform distribution of the ink components and is an essential parameter in the manufacturing process. This uniformity ensures that the ink maintains a stable conductivity and that the films formed have a consistent thickness and quality. The lack of homogeneity can lead to ink performance failures, such as an irregular conductivity and the formation of defective areas, compromising its application in the printing of flexible sensors. Homogeneity is also crucial to ensure adhesion to the substrate and the durability of the inks after application, especially on complex and flexible surfaces.

3. Chemically Modified Printed Electrodes

The use of electrodes in electrochemical systems is limited by the performance of the electrodes due to several factors, such as the inserted medium, the composition of the electrode, and the potential applied to the surface. Therefore, the selectivity and sensitivity of conventional electrodes are often insufficient for detecting analytes of interest in complex matrices, such as biological and environmental samples. With the emergence of chemically modified electrodes (CMEs), the applications of these devices have grown considerably, becoming a valuable and reliable tool for determining various compounds, from environmental pollutants to disease biomarkers. Therefore, the ability to chemically modify electrodes has paved the way for the development of increasingly sensitive analytical devices for applications in diverse fields, such as environmental monitoring, medical diagnoses, and industrial quality control [66,67,68,69].
CMEs can be defined as electrodes that have active chemical species attached to their surface, to modify the properties of the interface between the electrode and the solution. This modification allows for control of the electrode’s reactivity and selectivity, paving the way for a wide range of applications. These electrodes can be printed; made of glassy carbon, paste, etc.; and have their surface altered by the deposition of a thin layer of conductive or semiconductor material, or by other modifications that enhance the sensor’s performance, depending on its intended use. These modifications confer new electrochemical properties to the electrode, significantly expanding its electroactive area and improving its selectivity, linear response range, catalytic properties, and stability. They can be applied in various fields such as analytical chemistry, biosensors, and electrocatalysis [70,71,72,73].
In recent years, the modification of electrode surfaces has been widely used in the manufacture of highly sensitive devices for determining various compounds. Figure 5 shows the number of articles published between 1974 and 2023 according to the Scopus and Web of Science databases, by using the term “chemically modified electrode” and focusing on the title, article summary, and keywords. The growth observed over the years may be associated with an increase in investments in research in this area and the emergence of several materials with the potential to be used in the modification of electrochemical sensors.

3.1. Materials Applied in Sensor Modifications

Various materials can be used in the chemical modification of screen-printed electrodes, among which the most used are metallic nanoparticles, carbonaceous nanomaterials, polymers, and composites. The high surface area of nanomaterials, their unique electronic properties, and their functionalization capacity allow the creation of electrochemical sensors with a greater sensitivity, selectivity, and stability. Some examples of the most used nanomaterials are carbon nanotubes, graphene, carbon black, gold, and silver nanoparticles (Figure 6). Some of the carbonaceous nanomaterials, such as those mentioned previously, have a high surface area, providing many active sites for interaction with the analyte and increasing the sensitivity of the sensors. Furthermore, the excellent electrical conductivity of these nanomaterials facilitates the transfer of electrons between the electrode and the analyte, improving the efficiency of electrochemical processes. The use of metallic nanoparticles to modify CMEs has also been shown to be advantageous because these nanoparticles, such as gold, silver, and platinum, have a high electrical conductivity, which facilitates electron transfer and increases the sensor sensitivity. Regarding conductive polymers, they present a good versatility, a low market cost, and attractive chemical and physical properties for the development of sensors, in addition to biocompatibility and a high surface area, and are commonly used to immobilize biomolecules in the manufacture of biosensors [74,75,76,77].
Shehata et al. (2020) developed a printed electrochemical sensor (SPE) modified with gold nanoparticles (AuNPs) through electrochemical deposition. This sensor was called GNMSPE and was applied in the electrochemical determination of moxifloxacin hydrochloride (Moxi). The electrochemical performance was compared with that of other sensors (carbon paste electrodes) and modified with other nanomaterials (graphene oxide, multi-walled carbon nanotubes, and silver nanoparticles, among others). The results obtained demonstrated that GNMSPE presented the best electrochemical performance, a high sensitivity, and applicability for determining Moxi in urine samples from human infants, with recoveries between 99.8% and 101.6%, in addition to the advantage of being disposable [78].
In another study conducted by Khosrokhavar et al. (2020), a printed electrochemical sensor modified with a molecularly imprinted polymer (MIP) and graphene nanosheets was developed for the detection of sertraline (STR). To perform the polymerization, precipitation was performed in the presence of sertraline hydrochloride (a template molecule) to synthesize the MIP [79]. The SPE was fabricated in the laboratory and a thin layer of the MIP/graphene suspension was incorporated under the surface of the electrodes. A comparative study was carried out on the sensor modified with the MIP/graphene and with an NIP (non-imprinted polymer)/graphene. The sensor with the MIP/graphene showed a better performance, which can be associated with the high selectivity of the MIP associated with the high conductivity and surface area of graphene nanosheets. Finally, this sensor showed a good sensitivity and was applied in tablet and human serum samples with recoveries between 97.98% and 101.33%. To ensure that the modification process of the screen-printed electrodes was effective, characterization techniques were used, as they can guarantee the quality and performance of these sensors in determining drugs. This step allows the modification process to be optimized, ensuring the reproducibility of the results and the understanding of the molecular recognition mechanisms, which are fundamental for developing sensitive sensors.

3.2. Transition Metal Materials

In addition to the previously mentioned nanomaterials, other materials composed of transition metals and abundant Earth metal oxides have gained prominence in developing electrochemical sensors. Transition metals such as iron (Fe), cobalt (Co), nickel (Ni), copper (Cu), and manganese (Mn), and their oxide-based nanomaterials, exhibit an exceptional set of properties that make them promising candidates for various applications. The oxides of these metals can mimic the catalytic activity of noble metals, making them more accessible and sustainable alternatives for reactions of industrial interest. The high surface area and chemical nature of these metal oxides make them excellent supports for noble metal nanoparticles (gold, silver, platinum, palladium), enhancing their catalytic activity and stability [80,81].
Another relevant characteristic is the abundance of these metals on Earth, which significantly reduces the production costs of the nanomaterials, making them economically viable for large-scale applications. With appropriate surface modifications, metal oxide-based nanomaterials can exhibit biocompatibility, paving the way for applications in biomedicine and biotechnology. Beyond applications in electrochemical devices, these materials offer a wide range of catalytic properties, making them suitable for various energy conversion reactions, such as water electrolysis and CO2 reduction reactions. Maduraiveeran et al. (2019) described in their review the recent progress and main approaches for the synthesis and application of transition metal oxide nanomaterials for various electrochemical devices. Chen et al. (2024) investigated the application of iron-based electrode material composites in electrochemical sensors for real sample detection, demonstrating the main developments around sensor applications. Finally, transition metal oxide-based nanomaterials represent a class of materials with great potential to revolutionize various areas, from catalysis and energy to biomedicine. Their unique properties, combined with their abundance and low cost, make them highly attractive materials for research and development [80,81].

4. Characterizations of Screen-Printed Electrodes

The characterization of electrodes and the materials that constitute the inks is a crucial step in developing electrochemical sensors. Through them, it is possible to obtain crucial information about electrochemical characteristics (sensor resistivity) and structural and morphological characteristics. Several studies exist in the literature that include characterization techniques to correlate the properties of the materials with the electrochemical performance of the sensor, in addition to validating the results obtained [82,83]. The most used techniques are cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), scanning electron microscopy (SEM), transmission electron microscopy (TEM), Fourier-transform infrared spectroscopy (FTIR), Raman spectroscopy, and X-ray diffraction (XRD).

4.1. Cyclic Voltammetry and Electrochemical Impedance Spectroscopy

Electrochemical characterizations are generally carried out using cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) techniques. CV is an electrochemical technique that involves applying a cyclically varying potential to an electrode and measuring the resulting current. EIS, on the other hand, applies a small-amplitude alternating potential perturbation to the electrochemical system and measures the resulting alternating current response. Both techniques can provide information about the reaction mechanism, the material resistivity, the charge transfer at the electrode/electrolyte interface, the kinetics of electrochemical processes, and a material’s ability to act as an electrode in an electrochemical device. These techniques are widely used to optimize the performance of electrochemical devices [82,83,84,85,86,87,88].
In the literature, Ehsan et al. (2021) employed cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) to characterize the modification of SPEs with various biomolecules. Their results indicated that the immobilization of proteins such as protein A and antibodies can substantially impact electron transfer at the electrode/electrolyte interface. By analyzing the changes in the peak current and charge transfer resistance (Rct) values, they were able to assess the immobilization efficiency and the influence of different biomolecules on the electrochemical properties of the sensor [88].

4.2. Scanning Electron Microscopy (SEM)

SEM is an essential tool for characterizing materials used in electrochemical sensors. This technique allows high-resolution images of the surface of materials and electrodes to be obtained, revealing details about the morphology, such as the particle size, roughness, and shape. This information is crucial to understanding how the surface structure influences the interaction of the electrode with chemical species and, consequently, the sensor performance [89,90,91,92,93].
Saenchoopa et al. (2021), in addition to using CV to compare the performance of the sensors after modification with each of the compounds employed, used the SEM technique to analyze the morphology of the SPE after modification with silver/hydroxypropylmethylcellulose/chitosan/urease nanowires. The results indicated that the sensor surface became more homogeneous due to the action of the hydroxypropylmethylcellulose, chitosan, and urease solution, filling the gaps between the silver nanowires. Using these characterizations, it was possible to verify that the modification promoted an increase in the electroactive surface area, facilitating electron transfer and optimizing the sensor’s performance [93].

4.3. Transmission Electron Microscopy

Transmission electron microscopy (TEM) is essential for developing printed electrochemical sensors. By providing high-resolution images, TEM allows the detailed characterization of the microstructure and the composition of these materials. This technique enables a correlation between the morphological characteristics (grain size, porosity) and the electrochemical properties, elucidating detection mechanisms and optimizing sensor performance. Impurity identification and elemental quantification, also performed by TEM, ensure the reproducibility and reliability of the devices.
In this context, Liao et al. (2019) developed a conductive ink based on graphite, acrylic resin, and water. Through TEM, it was possible to confirm the presence of graphene with few layers in the ink. In addition, the authors were able to correlate the dispersibility of the material in the ink with its conductivity. Wang et al. (2023) also applied TEM to characterize the conductive ink developed based on carbon black and graphite (1:2), polymer resin, and water. Through the analyses, the authors were able to identify a structure similar to graphene, indicating the presence of transparent few-layer graphene [94,95].

4.4. Fourier-Transform Infrared Spectroscopy (FTIR)

FTIR is a widely used technique for the structural characterization of conductive inks and electrochemical sensors, which allows detailed information to be obtained about the chemical bonds, functional groups, and molecular structure of the analyzed compounds [94,95,96,97,98,99]. This technique has been widely used to characterize the biofunctionalization process of SPEs.
In a study conducted by Inam et al. (2022), this technique was applied to confirm the successful immobilization of each component of the aptasensor BSA-Aptamer-N(3-dimethylaminopropyl) N-ethylcarbodiimide/N-hydroxysulfosuccinimide sodium salt-11-mercaptoundecanoic acid-Ag. By analyzing the characteristic bands of each molecule, the authors showed the formation of bonds between the different components, demonstrating the effectiveness of the biofunctionalization protocol. The presence of bands corresponding to the functional groups of each molecule corroborated the formation of the proposed structure for the aptasensor. In addition to this technique, CV and EIS were employed to evaluate the effectiveness of the fabrication steps of the developed sensor. Together, these techniques contributed to providing valuable information about the aptasensor fabrication process [100].

4.5. Raman Spectroscopy

Raman spectroscopy has also been proven to be indispensable for the detailed characterization of electrochemical sensors. By interacting monochromatic light with matter, this technique provides a unique molecular vibrational profile, including detailed information about the molecular structure, interfacial interactions, and the dynamics of electrochemical processes and significantly contributing to the development of more efficient, selective, and robust sensors, with applications in several areas [101,102,103,104,105].
By applying Raman spectroscopy, Gomes et al. (2021) investigated the effects of different modifications on the surface of SPEs, showing the formation of polymer networks of polypyrrole and their composites with carbon nanotubes with conjugated C=C bonds, so that the technique allowed the different materials to be compared and the influence of the presence of polymeric materials in the film structure to be identified in devices with a molecularly printed poly(Py) matrix (MPPy) and with a non-imprinted poly(Py) matrix (NPPy) [105].

4.6. X-Ray Diffraction (DRX)

XRD is a fundamental analytical technique in the development of printed electrochemical sensors. When an X-ray beam is incident on a material, the interaction with the crystal structure generates a characteristic diffraction pattern, which allows the identification of the crystalline phases present, the calculation of the lattice parameters, and the evaluation of the degree of crystallinity. In the context of sensors, XRD is used to determine the chemical composition of materials, evaluate the purity of the substances used, analyze the crystal structure of modified electrodes, and identify the presence of secondary phases that may affect the sensor performance. In addition, XRD can be used to monitor the formation of new phases during electrode surface modification processes, providing valuable information for optimizing the manufacturing of printed electrochemical sensors.
Pradella-Filho et al. (2020) developed an ink based on graphite and stained-glass varnish, and applied XRD to the graphite powder, ink, and varnish to better understand the crystallinity of the materials. The results obtained suggest the presence of the materials in the formation of the conductive carbon ink [106].
The characterization of SPEs, especially those that are chemically modified, is an extremely important step to ensure reliability and accuracy in drug detection. By subjecting the sensors and ink components to the techniques mentioned above, it is possible to obtain fundamental information for understanding the mechanisms of interaction between the sensor and the drug, optimizing the analysis conditions, and validating the analytical method (Figure 7). In this way, the characterization of the sensors allows more accurate and reproducible results to be obtained in the determination of drugs in real samples, such as biological fluids, river water, and pharmaceutical samples, among others.

5. Applications of SPEs in the Determination of Drugs

The increasing interdependence between health and the environment, driven by the pursuit of a higher quality of life and environmental preservation, has increased the demand for precise analytical technologies. Electrochemical sensors, capable of detecting pollutants with a high sensitivity, stand out in this context. The search for new drugs and the need for faster and more accurate analytical methods drive the development of innovative sensors for detecting and quantifying drugs in various matrices [20].
In their work, Hoppe (2024) used electrosynthesis to create a molecularly imprinted polymer on a screen-printed carbon electrochemical sensor, developed for the quantitative determination of paracetamol in water samples [107]. The developed electrode demonstrated an excellent performance in the detection of paracetamol, presenting a linear detection range of 5.0 × 10−8 mol L−1, indicating a high sensor sensitivity [107].
Modifying the electrode surface with synthetic materials such as molecularly imprinted polymers (MIPs) aids in constructing an electrode with an enhanced selective performance. In this context, Kim (2023) developed an electrochemical sensor modified with a molecularly imprinted polymer quantifying cortisol. Synthetic cortisol, also known as hydrocortisone, is a medication prescribed for the treatment of infections, allergies, and autoimmune diseases. The developed sensor exhibited good mechanical resistance after successive measurements, contributing to the production of MIP-modified sensors. The performance in cortisol detection was positive, presenting a detection limit of 0.036 nM [108]. The primary application of this sensor lies in the non-invasive monitoring of cortisol levels, enabling continuous real-time detection. This can be beneficial in fields such as sports medicine, medical diagnoses, and the management of stress-related conditions.
Beyond their application in water and bodily fluid samples, these sensors are also employed in food samples for the detection of drugs administered for the prevention and treatment of diseases in food-producing animals destined for human consumption [109].
Souza et al. (2022) presented the development of a printed electrochemical sensor for the detection of ciprofloxacin (CIP) in honey and milk samples. This drug is used to treat bacterial infections such as American foulbrood, caused by the bacterium Paenibacillus larvae. This disease affects bee larvae, leading to the death of the colony if not treated adequately. It is also used as a preventive and therapeutic measure in dairy cattle for the treatment of mastitis, among others. If ciprofloxacin is not managed correctly, it can contaminate bee products (honey, wax, etc.) and milk produced by dairy cattle with drug residues [104]. In this context, the researchers developed a sensor using paper as a support material, making it sustainable and easily accessible. The device was fabricated using a low-cost conductive ink made from graphite powder and nail polish. In addition, the device underwent a polishing process that exposed the active sites of the graphite, thus improving its sensitivity. The device demonstrated a good precision and reproducibility in quantifying the antibiotic, with detection limits suitable for food samples such as honey and milk. The method stood out for its simplicity and low cost, making it suitable for rapid analyses in food-production environments. The developed sensor showed a good performance with a wide linear range of 9.90–220 µmol L −1 and a detection limit of 4.96 µmol L −1 [109].
Table 1 shows other studies where printed electrodes were applied to the analytical determination of drugs.
The advances highlighted in this review underscore the significant potential of screen-printed electrochemical sensors for drug detection across diverse matrices. The versatility of screen-printing enables the fabrication of miniaturized, disposable, and cost-effective devices, making them ideal for point-of-care applications, drug production monitoring, and environmental quality control. The ability to modify conductive inks with various materials and optimize experimental conditions expands the range of detectable drugs. However, further research is needed to assess these sensors’ long-term stability and delve deeper into potential interferences from other substances present in real-world samples.

6. Perspectives and Challenges

Developing new conductive inks for screen-printed electrochemical sensors represents a promising field of research with the potential to revolutionize the area of drug detection. However, there are still some challenges to be overcome. Complex matrices such as water, biological fluids, and pharmaceutical formulations can impair analyses, so it is necessary to develop conductive inks that present a high selectivity, minimizing interference from other components. This problem can be solved by inserting new components into the ink and modifying the electrode surface. Ensuring the electrochemical and mechanical stability of conductive inks over time and in different environmental conditions is also a challenge. To this end, the choice of the ink component materials and the support on which this ink will be applied must be carefully chosen, giving preference to those materials that present a greater stability.
Ensuring the reproducibility of the manufactured sensors, both in terms of performance and manufacturing, is also a crucial factor in guaranteeing efficient analyses. Exploring new conductive materials (nanomaterials and conductive polymers) and employing surface modification techniques enable an improved sensor sensitivity and selectivity, in addition to expanding the application possibilities. Another relevant factor that can be explored in developing printed sensors is the use of machine learning algorithms to improve the interpretation of data obtained by the sensors and develop more robust analysis methods.

7. Conclusions

This review provides information on developing conductive inks to obtain screen-printed electrochemical sensors applied in drug detection. Various compositions of conductive inks are discussed, and their main properties, such as their conductivity, malleability, ease of modification, chemical stability, and high surface area, are highlighted. These sensors have been extensively modified with various materials capable of expanding the properties of SPEs, giving them a greater sensitivity, selectivity, and stability, as well as low detection and quantification limits. In the long term, screen-printed sensors are expected to revolutionize drug monitoring in various matrices, making analyses more reliable, simple, and accessible. Finally, this review presents an overview of the current scenario, focusing on developing new conductive inks for fabricating SPEs focused on drug detection.

Author Contributions

V.A.R.L.; S.P.d.O.; L.C.d.S. and L.J.d.P.S. contributed to the writing of the text. T.C.d.O.C. and D.N.d.S. contributed to the revised text. A.C.P. contributed to the general writing and proofreading of the manuscript. All authors the contributed to the spelling and a semantic review of the text. All authors have read and agreed to the published version of this manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data are contained within this article.

Acknowledgments

We acknowledge Minas Gerais State Research Support Foundation (FAPEMIG), National Council for Scientific and Technological Development (CNPq), Coordination for the Improvement of Higher Education Personnel (CAPES), National Institute of Alternative Technologies for Detection, Toxicological Assessment and Removal of Emerging and Radioactive Contaminants (INCT-DATREM), and Federal University of São João del Rei (UFSJ).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Santos, A.D.F.O.; de Alencar, A.C.S.; Cruz, E.E.S.; Duarte, F.T.; Constantino, G.K.G.; Duarte, J.N.M.; Pereira, K.V.; Viana, L.G.F.C.; Bomfim, R.S.S.; Rodrigues, V.R.A.; et al. Medication Intoxication by NSAIDs: An Integrative Revie. Braz. J. Health Rev. 2024, 7, 71066. [Google Scholar] [CrossRef]
  2. da Silva, R.V.; Pinheiro-Machado, R.; de Sousa, A.C.A. Penicillin: Penicillin: From discovery to patenting the large-scale production process. Hist. Ciênc. Saúde-Manguinhos 2024, 31, 202402. [Google Scholar] [CrossRef]
  3. Brito, A.M.G.; Versiani, A.P.; Dias, M.R.S.; Piris, A.P. Indiscriminate use of antibiotics: An integrative review. Bionorte 2022, 11, 219–225. [Google Scholar] [CrossRef]
  4. de Souza, F.S.; de Lima, R.Q.; Figueiredo, E.F.G. Intoxication by non-steroid anti-inflammatory drugs: Risks of use. Braz. J. Health Rev. 2021, 4, 14873–14891. [Google Scholar] [CrossRef]
  5. Xavier, M.S.; Castro, H.N.; de Souza, L.G.D.; de Oliveira, Y.S.L.; Tafuri, N.F.; Amâncio, N.F.G. Self-medication and health risk: A literature review. Braz. J. Health Rev. 2021, 4, 225–240. [Google Scholar] [CrossRef]
  6. Ates, H.C.; Roberts, J.A.; Lipman, J.; Cass, A.E.; Urban, G.A.; Dincer, C. Therapeutic monitoring of non-local medications. Trends Biotechnol. 2020, 38, 1262–1277. [Google Scholar] [CrossRef]
  7. Karimi-Maleh, H.; Karimi, F.; Alizadeh, M.; Sanati, A.L. Electrochemical sensors, a bright future in the fabrication of portable kits in analytical systems. Chem. Rec. 2020, 20, 682–692. [Google Scholar] [CrossRef]
  8. de Oliveira, M.; Frihling, B.E.F.; Velasques, J.; Magalhães Filho, F.J.C.; Cavalheri, P.S.; Migliolo, L. Pharmaceutical residues, and xenobiotic contaminants: Occurrence, analytical techniques and sustainable alternatives for wastewater treatment. Total Environ. Sci. 2020, 705, 135568. [Google Scholar] [CrossRef]
  9. Akca, Z.; Özok, H.İ.; Yardim, Y.; Şentürk, Z. Electroanalytical investigation and voltammetric quantification of antiviral drug favipiravir in the pharmaceutical formulation and urine sample using a glassy carbon electrode in anionic surfactant media. Turk. J. Chem. 2022, 46, 869–880. [Google Scholar] [CrossRef]
  10. Sussuchi, E.M.; Monteiro, M.D.S.; dos Santos Júnior, J.C. Electrochemistry: Modified electrodes and their potentialities. Rev. Virtual Quím. 2020, 12, 1145–1160. [Google Scholar]
  11. Heard, D.M.; Lennox, A.J.J. Electrode Materials in Modern Organic Electrochemistry. Angew. Chem. Int. Ed. 2020, 59, 18866–18884. [Google Scholar] [CrossRef] [PubMed]
  12. de Freitas, J.M. Development of Simple, Portable, and Low-Cost Electroanalytical Methods for Forensic Analysis Applications. Ph.D. Thesis, Federal University of Uberlândia, Uberlândia, Brazil, 2019. [Google Scholar]
  13. Asadian, E.; Ghalkhani, M.; Shahrokhian, S. Electrochemical Sensing Based on Carbon Nanoparticles: A Review. Sens. Actuators B Chem. 2019, 293, 183. [Google Scholar] [CrossRef]
  14. Heikenfeld, J.; Jajack, A.; Rogers, J.; Gutruf, P.; Tian, L.; Pan, T.; Li, R.; Khine, M.; Kim, J.; Wang, J.; et al. Wearable sensors: Modalities, challenges, and prospects. Lab Chip 2018, 18, 217–248. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, H.; Zhao, H.; Khan, M.A.; Zou, W.; Xu, J.; Zhang, L.; Zhang, J. Recent progress in advanced electrode materials, separators and electrolytes for lithium batteries. J. Mater. Chem. A 2018, 6, 20564–20620. [Google Scholar] [CrossRef]
  16. Jian, X.; Liu, S.; Gao, Y.; Tian, W.; Jiang, Z.; Xiao, X.; Tang, H.; Yin, L. Carbon-Based Electrode Materials for Supercapacitor: Progress, Challenges and Prospective Solutions. J. Electr. Eng. 2016, 4, 75–87. [Google Scholar] [CrossRef]
  17. Pereira, D.F. Development of Paper-Based Printed Electrodes from Graphite Ink Modified with Magnetite for the Determination of Vitamins. Master’s Thesis, Federal University of Santa Catarina, Florianópolis, Brazil, 2022. [Google Scholar]
  18. Pérez-Fernández, B.; Costa-García, A.; Muñiz, A.d.l.E. Electrochemical (Bio)Sensors for Pesticides Detection Using Screen-Printed Electrodes. Biosensors 2020, 10, 32. [Google Scholar] [CrossRef]
  19. Ferrari, A.G.-M.; Rowley-Neale, S.J.; Banks, C.E. Screen-printed electrodes: Transitioning the laboratory in-to-the field. Talanta Open 2021, 3, 100032. [Google Scholar] [CrossRef]
  20. Malvessi, G.S.; Hoppe, T.D.; Zapp, E.; Brondani, D. Lab-made screen-printed electrode with conductive ink based on carbon nanomaterials for simultaneous electrochemical analysis of bisphenol A, catechol, and 4-nitrophenol. Measurement 2024, 231, 114601. [Google Scholar] [CrossRef]
  21. Dinu, A.; Apetrei, C. Development of Polypyrrole Modified Screen-Printed Carbon Electrode Based Sensors for Determination of L-Tyrosine in Pharmaceutical Products. Int. J. Mol. Sci. 2021, 22, 7528. [Google Scholar] [CrossRef]
  22. Demir, E.; Silah, H. Development of a New Analytical Method for Determination of Veterinary Drug Oxyclozanide by Electrochemical Sensor and Its Application to Pharmaceutical Formulation. Chemosensors 2020, 8, 25. [Google Scholar] [CrossRef]
  23. Hrioua, A.; Loudiki, A.; Farahi, A.; Bakasse, M.; Lahrich, S.; Saqrane, S.; El Mhammedi, M.A. Recent advances in electrochemical sensors for amoxicillin detection in biological and environmental samples. Bioelectrochemistry 2021, 137, 107687. [Google Scholar] [CrossRef]
  24. dos Santos, A.M. Development of Electrochemical Sensors Based on Composites for the Simultaneous Determination of Drugs. Ph.D. Thesis, Federal University of São Carlos, São Carlos, Brazil, 2019. [Google Scholar]
  25. Mincu, N.-B.; Lazar, V.; Stan, D.; Mihailescu, C.M.; Iosub, R.; Mateescu, A.L. Screen-Printed Electrodes (SPE) for In Vitro Diagnostic Purpose. Diagnostics 2020, 10, 517. [Google Scholar] [CrossRef]
  26. Qin, Y.; Ouyang, X.; Lv, Y.; Liu, W.; Liu, Q.; Wang, S. A Review of Carbon-Based Conductive Inks and Their Printing Technologies for Integrated Circuits. Coatings 2023, 13, 1769. [Google Scholar] [CrossRef]
  27. Camargo, J.R.; Orzari, L.O.; Araújo, D.A.G.; de Oliveira, P.R.; Kalinke, C.; Rocha, D.P.; dos Santos, A.L.; Takeuchi, R.M.; Munoz, R.A.A.; Bonacin, J.A.; et al. Development of conductive inks for electrochemical sensors and biosensors. Microchem. J. 2021, 164, 105998. [Google Scholar] [CrossRef]
  28. Pouria, K.; Gehan, C.J.; Chao, W.; Mehran, T. High conductivity and stability intercalated carbonaceous conductors. Carbon Trends 2022, 9, 100229. [Google Scholar]
  29. Maddipatla, D.; Narakathu, B.B.; Atashbar, M. Recent Progress in Manufacturing Techniques of Printed and Flexible Sensors: A Review. Biosensors 2020, 10, 199. [Google Scholar] [CrossRef] [PubMed]
  30. Hassani, F.A.; Shi, Q.; Wen, F.; He, T.; Haroun, A.; Yang, Y.; Feng, Y.; Lee, C. Smart materials for smart healthcare– moving from sensors and actuators to self-sustained nanoenergy nanosystems. Smart Mater. Med. 2020, 1, 92–124. [Google Scholar]
  31. Barhoum, A.; Hamimed, S.; Slimi, H.; Othmani, A.; Abdel-Haleem, F.M.; Bechelany, M. Modern designs of electrochemical sensor platforms for environmental analyses: Principles, nanofabrication opportunities, and challenges. Trends Environ. Anal. Chem. 2023, 38, 00199. [Google Scholar] [CrossRef]
  32. Bounegru, A.V.; Apetrei, C. Carbonaceous Nanomaterials Employed in the Development of Electrochemical Sensors Based on Screen-Printing Technique—A Review. Catalysts 2020, 10, 680. [Google Scholar] [CrossRef]
  33. Musa, A.M.; Kiely, J.; Luxton, R.; Honeychurch, K.C. Recent progress in screen-printed electrochemical sensors and biosensors for the detection of estrogens. TrAC Trends Anal. Chem. 2021, 139, 116254. [Google Scholar] [CrossRef]
  34. Buffon, E. Disposable, Molecularly Imprinted Electrochemical Sensors Based on Reduced Graphene Oxide and Metal Nanoparticles for Phenolic Acid Analysis in Brazilian Fruit Waste. Ph.D. Thesis, São Paulo State University, Araraquara, Brazil, 2023. [Google Scholar]
  35. Mendez-Rossal, H.R.; Wallner, G.M. Printability and Properties of Conductive Inks on Primer-Coated Surfaces. Int. J. Polym. Sci. 2019, 3874181, 8. [Google Scholar] [CrossRef]
  36. Pajor-Świerzy, A.; Pawłowski, R.; Warszyński, P.; Szczepanowicz, K. The conductive properties of ink coating based on Ni–Ag core–shell nanoparticles with the bimodal size distribution. J. Mater. Sci. Mater. Electron. 2020, 31, 12991–12999. [Google Scholar] [CrossRef]
  37. Megha, G.; Kulwant, S.; Nitu, B. Conductive polymers: A multipurpose material for protecting coating. Prog. Org. Coat. 2024, 187, 108083. [Google Scholar]
  38. Gholami, F.; Tomas, M.; Gholami, Z.; Mirzaei, S.; Vakili, M. Surface Characterization of Carbonaceous Materials Using Inverse Gas Chromatography: A Review. Electrochem 2020, 1, 367–387. [Google Scholar] [CrossRef]
  39. Li, X.; Fan, B.; Wang, Z.; Guan, G. Carbonaceous matrixes-based free-standing electrode materials for energy storage. Carbon Trends 2024, 16, 100397. [Google Scholar] [CrossRef]
  40. Patel, T.; Huang, J.; Krukiewicz, K. Multifunctional organic monolayer-based coatings for implantable biosensors and bioelectronic devices: Review and perspectives. Biosens. Bioelectron. X 2023, 14, 100349. [Google Scholar] [CrossRef]
  41. Dokur, E.; Uruc, S.; Kurteli, R.; Gorduk, O.; Sahin, Y. Investigation of flexible electrochemical sensor properties of graphite conductive ink modified textile products and their applications in electrical circuits. Mater. Today Commun. 2023, 37, 106920. [Google Scholar] [CrossRef]
  42. Verma, D.; Dubey, N.; Yadav, A.K.; Saraya, A.; Sharma, R.; Solanki, P.R. Disposable paper-based screen-printed electrochemical immunoplatform for dual detection of esophageal cancer biomarkers in patients’ serum samples. Mater. Adv. 2024, 5, 2153–2168. [Google Scholar] [CrossRef]
  43. Marchianò, V.; Tricase, A.; Macchia, E.; Bollella, P.; Torsi, L. Self-powered wearable biosensor based on stencil-printed carbon nanotube electrodes for ethanol detection in sweat. Anal. Bioanal. Chem. 2024, 416, 5303–50316. [Google Scholar] [CrossRef]
  44. Kim, D.S.; Jeong, J.M.; Park, H.J.; Kim, Y.K.; Lee, K.G.; Choi, B.G. Highly Concentrated, Conductive, Defect-free Graphene Ink for Screen-Printed Sensor Application. Nano-Micro Lett. 2021, 13, 87. [Google Scholar] [CrossRef] [PubMed]
  45. Hairi, N.I.I.M.; Ralib, A.A.M.; Nordin, A.N.; Yunos, M.F.A.M.; Ming, L.L.; Tung, L.H.; Samsudin, Z. Recent advance in using eco-friendly carbon-based conductive ink for printed strain sensor: A review. Clean. Mater. 2024, 12, 100248. [Google Scholar] [CrossRef]
  46. Szroeder, P.; Ziółkowski, P.; Mosińska, L.; Trykowski, G. Boosting electrochemical performance of single-walled carbon nanotube three-dimensional architectures through appropriate selection of organic dispersant. Diam. Relat. Mater. 2024, 148, 111440. [Google Scholar] [CrossRef]
  47. Park, S.; Song, Y.; Ryu, B.; Song, Y.-B.; Lee, H.; Kim, Y.; Lim, J.; Lee, D.; Yoon, H.; Lee, C.; et al. Highly Conductive Ink Based on Self-Aligned Single-Walled Carbon Nanotubes through Inter-Fiber Sliding in Cellulose Fibril Networks. Adv. Sci. 2024, 11, 2402854. [Google Scholar] [CrossRef] [PubMed]
  48. Lee, J.-H.; Park, S.-J.; Choi, J.-W. Electrical Property of Graphene and Its Application to Electrochemical Biosensing. Nanomaterials 2019, 9, 297. [Google Scholar] [CrossRef] [PubMed]
  49. Poulin, A.; Aeby, X.; Siqueira, G.; Nyströn, G. Versatile carbon-loaded shellac ink for disposable printed electronics. Sci. Rep. 2021, 11, 23784. [Google Scholar] [CrossRef] [PubMed]
  50. Zappi, D.; Varani, G.; Cozzoni, E.; Iatsunskyi, I.; Laschi, S.; Giardi, M.T. Innovative Eco-Friendly Conductive Ink Based on Carbonized Lignin for the Production of Flexible and Stretchable Bio-Sensors. Nanomaterials 2021, 11, 3428. [Google Scholar] [CrossRef]
  51. Gemeiner, P.; Sarakhman, O.; Hatala, M.; Ház, A.; Roupcová, P.; Mackuľak, T.; Barek, J.; Švorc, Ľ. A new generation of fully-printed electrochemical sensors based on biochar/ethylcellulose-modified carbon electrodes: Fabrication, characterization and practical applications. Electrochim. Acta 2024, 487, 144161. [Google Scholar] [CrossRef]
  52. Zhang, Y.; Cen, Q.; Xu, X.; Li, W.; Zhao, Y.; Li, W.; Liu, Q.; Chen, M.; Guo, N.; Wu, W.; et al. The effect of PVAc in silver ink for adhesion and conductivity of conductive pattern. J. Mater. Res. Technol. 2022, 18, 4277–4284. [Google Scholar] [CrossRef]
  53. van Hazendonk, L.S.; Vonk, C.F.; van Grondelle, W.; Vonk, N.H.; Friedrich, H. Towards a predictive understanding of direct ink writing of graphene-based inks. Appl. Mater. Today 2024, 36, 102014. [Google Scholar] [CrossRef]
  54. Carvalho, J.H.S.; Gogola, J.L.; Bergamini, M.F.; Marcolino-Junior, L.H.; Janegitz, B.C. Disposable and low-cost lab-made screen-printed electrodes for voltammetric determination of L-dopa. Sens. Actuators Rep. 2021, 3, 100056. [Google Scholar] [CrossRef]
  55. Carvalho, J.H.S.; Stefano, J.S.; Brazaca, L.C.; Janegitz, B.C. New conductive ink based on carbon nanotubes and glass varnish for the construction of a disposable electrochemical sensor. J. Electroanal. Chem. 2023, 937, 117428. [Google Scholar] [CrossRef]
  56. Andreotti, I.A.A.; Orzari, L.O.; Camargo, J.R.; Faria, R.C.; Marcolino-Junior, L.H.; Bergamini, M.F.; Gatti, A.; Janegitz, B.C. Disposable and flexible electrochemical sensor made by recyclable material and low cost conductive ink. J. Electroanal. Chem. 2019, 840, 109–116. [Google Scholar] [CrossRef]
  57. Shitanda, I.; Oda, K.; Loew, N.; Watanabe, H.; Itagaki, M.; Tsujimura, S.; Zebda, A. Chitosan-based enzyme ink for screen-printed bioanodes. RSC Adv. 2021, 11, 20550–20556. [Google Scholar] [CrossRef] [PubMed]
  58. Henrique, J.M.; Camargo, J.R.; de Oliveira, G.G.; Stefano, J.S.; Janegitz, B.C. Disposable electrochemical sensor based on shellac and graphite for sulfamethoxazole detection. Microchem. J. 2021, 170, 106701. [Google Scholar] [CrossRef]
  59. de Lima, L.F.; Ferreira, A.F.; Maciel, C.C.; Ferreira, M.; de Araujo, W.R. Disposable and low-cost electrochemical sensor based on the colorless nail polish and graphite composite material for tartrazine detection. Talanta 2021, 227, 122200. [Google Scholar] [CrossRef]
  60. Raja, I.S.; Kim, C.; Kang, M.S.; Joung, Y.K.; Lee, J.H.; Han, D.W. Studies on cytocompatibility of human dermal fibroblasts on carbon nanofiber nanoparticle-containing bioprinted constructs. Discov. Nano 2024, 19, 149. [Google Scholar] [CrossRef]
  61. Corzo, D.; Rosas-Villalva, D.; Amruth, C.; Tostado-Bláquez, G.; Alexandre, E.B.; Hernandez, L.H.; Han, J.; Xu, H.; Babics, M.; De Wolf, S.; et al. High-performing organic electronics using terpene green solvents from renewable feedstocks. Nat. Energy 2023, 8, 62–73. [Google Scholar] [CrossRef]
  62. Zhou, F.; Hearne, Z.; Li, C.J. Water—The greenest solvent overall. Curr. Opin. Green Sustain. Chem. 2019, 18, 118–123. [Google Scholar] [CrossRef]
  63. Chung, K.Y.; Xu, B.; Tan, D.; Yang, Q.; Li, Z.; Fu, H. Naturally Crosslinked Biocompatible Carbonaceous Liquid Metal Aqueous Ink Printing Wearable Electronics for Multi-Sensing and Energy Harvesting. Nano-Micro Lett. 2024, 16, 149. [Google Scholar] [CrossRef]
  64. Htwe, Y.Z.N.; Abdullah, M.K.; Mariatti, M. Optimization of graphene conductive ink using solvent exchange techniques for flexible electronics applications. Synth. Met. 2021, 274, 116719. [Google Scholar] [CrossRef]
  65. Oliveira, V.C.; Calábria, L.K.; de Rezende, A.A.A. Electrochemical sensors for pesticide detection in environmental samples: A literature review. Online perspectives 2022, 12, 12–21. [Google Scholar]
  66. Alkire, R.C.; Kolb, D.M.; Lipkowski, J.; Ross, P.N. Chemically Modified Electrodes. In Advances in Electrochemical Science and Engineering, 1st ed.; Wiley-VCH: Weinheim, Germany, 2009; Volume 11, pp. 1–50. ISBN 9783527627059. [Google Scholar]
  67. Yoon, H.; Nah, J.; Kim, H.; Ko, S.; Sharifuzzaman, M.; Barman, S.C.; Xuan, X.; Kim, J.; Park, J.Y. A chemically modified laser-induced porous graphene based flexible and ultrasensitive electrochemical biosensor for sweat glucose detection. Sens. Actuators B Chem. 2020, 311, 127866. [Google Scholar] [CrossRef]
  68. Rocha, D.P.; Squissato, A.L.; da Silva, S.M.; Richter, E.M.; Munoz, R.A. Improved electrochemical detection of metals in biological samples using 3D-printed electrode: Chemical/electrochemical treatment exposes carbon-black conductive sites. Electrochim. Acta 2020, 335, 135688. [Google Scholar] [CrossRef]
  69. Zhao, G.; Wang, X.; Liu, G.; Cao, Y.; Liu, N.; Thuy, N.T.D.; Zhang, L.; Yu, M. A flexible and disposable electrochemical sensor for the evaluation of arsenic levels: A new and efficient method for the batch fabrication of chemically modified electrodes. Anal. Chim. Acta 2022, 1194, 339413. [Google Scholar] [CrossRef] [PubMed]
  70. Cardenas, A.; Frontana, C. Evaluation of a carbon ink chemically modified electrode incorporating a copper-neocuproine complex for the quantification of antioxidants. Sens. Actuators B Chem. 2020, 313, 128070. [Google Scholar] [CrossRef]
  71. Giancarla, A.; Zanoni, C.; Merli, D.; Magnaghi, L.R.; Biesuz, R. A new cysteamine-copper chemically modified screen-printed gold electrode for glyphosate determination. Talanta 2024, 269, 125436. [Google Scholar] [CrossRef]
  72. Torres-Rojas, F.; Muñoz, D.; Canales, C.P.; Hevia, S.A.; Leyton, F.; Veloso, N.; Isaacs, M.; Vargas, I.T. Synergistic effect of electrotrophic perchlorate reducing microorganisms and chemically modified electrodes for enhancing bioelectrochemical perchlorate removal. Environ. Res. 2023, 233, 116442. [Google Scholar] [CrossRef] [PubMed]
  73. Jahani, P.M.; Mohammadi, S.Z.; Khodabakhshzadeh, A.; Cha, J.H.; Asl, M.S.; Shokouhimehr, M.; Zhang, K.; Le, Q.V.; Peng, W. Simultaneous detection of morphine and diclofenac using screen-printed electrode modified with graphene nanoribbon. Int. J. Electrochem. Sci. 2020, 15, 9037–9048. [Google Scholar] [CrossRef]
  74. Abdel-Raoof, A.M.; Abdel-Monem, A.H.; Almrasy, A.A.; Mohamed, T.F.; Nasr, Z.A.; Mohamed, G.F. Optimization of highly sensitive screen-printed electrodes modified with cerium (IV) oxide nanoparticles for the electrochemical determination of oxymetazoline hydrochloride using response surface methodology. J. Electrochem. Soc. 2020, 167, 047502. [Google Scholar] [CrossRef]
  75. Silva, D.N.; Cândido, T.C.O.; Pereira, A.C. Simple and disposable device based on gold nanoparticles modified screen-printed carbon electrode for detection of ciprofloxacin. Journal of Solid State Electrochemistry 2024, 1–14. [Google Scholar] [CrossRef]
  76. Dinu, A.; Apetrei, C. Voltammetric Determination of Phenylalanine Using Chemically Modified Screen-Printed Based Sensors. Chemosensors 2020, 8, 113. [Google Scholar] [CrossRef]
  77. Dăscălescu, D.; Apetrei, C. Voltammetric Determination of Levodopa Using Mesoporous Carbon—Modified Screen-Printed Carbon Sensors. Sensors 2021, 21, 6301. [Google Scholar] [CrossRef]
  78. Shehata, M.; Fekry, A.M.; Walcarius, A. Moxifloxacin Hydrochloride Electrochemical Detection at Gold Nanoparticles Modified Screen-Printed Electrode. Sensors 2020, 20, 2797. [Google Scholar] [CrossRef] [PubMed]
  79. Khosrokhavar, R.; Motaharian, A.; Hosseini, M.R.M.; Mohammadsadegh, S. Screen-printed carbon electrode (SPCE) modified by molecularly imprinted polymer (MIP) nanoparticles and graphene nanosheets for determination of sertraline antidepressant drug. Microchem. J. 2020, 159, 105348. [Google Scholar] [CrossRef]
  80. Maduraiveerana, G.; Sasidharana, M.; Jinb, W. Earth-abundant transition metal and metal oxide nanomaterials: Synthesis and electrochemical applications. Prog. Mater. Sci. 2019, 106, 100574. [Google Scholar] [CrossRef]
  81. Chen, W.-H.; Maheshwaran, S.; Park, Y.-K.; Ong, H.C. Iron-based electrode material composites for electrochemical sensor application in the environment: A review. Sci. Total Environ. 2024, 953, 176128. [Google Scholar] [CrossRef]
  82. Skoog, D.A.; West, D.M.; Holler, F.J.; Crouch, S. Fundamentals of Analytical Chemistry, 9th ed.; Cengage Learning: São Paulo, Brazil, 2015; pp. 602–640. ISBN 9788522121373. [Google Scholar]
  83. Lazanas, A.C.; Prodromidis, M.I. Electrochemical impedance spectroscopy—A tutorial. ACS Meas. Sci. Au 2023, 3, 162–193. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, S.; Zhang, J.; Gharbi, O.; Vivier, V.; Gao, M.; Orazem, M.E. Electrochemical impedance spectroscopy. Nat. Rev. Methods Primers 2021, 1, 41. [Google Scholar] [CrossRef]
  85. Gharbi, O.; Tran, M.T.; Tribollet, B.; Turmine, M.; Vivier, V. Revisiting cyclic voltammetry and electrochemical impedance spectroscopy analysis for capacitance measurements. Electrochim. Acta 2020, 343, 136109. [Google Scholar] [CrossRef]
  86. Kozak, J.; Tyszczuk-Rotko, K. Screen-printed gold electrode for ultrasensitive voltammetric determination of the antipsychotic drug thioridazine. Measurement 2023, 217, 113107. [Google Scholar] [CrossRef]
  87. Kozak, J.; Tyszczuk-Rotko, K.; Wójciak, M.; Sowa, I.; Rotko, M. First screen-printed sensor (electrochemically activated screen-printed boron-doped diamond electrode) for quantitative determination of rifampicin by adsorptive stripping voltammetry. Materials 2021, 14, 4231. [Google Scholar] [CrossRef] [PubMed]
  88. Ehsan, M.A.; Khan, S.A.; Rehman, A. Screen-printed graphene/carbon electrodes on paper substrates as impedance sensors for detection of coronavirus in nasopharyngeal fluid samples. Diagnostics 2021, 11, 1030. [Google Scholar] [CrossRef] [PubMed]
  89. Akhtar, K.; Khan, S.A.; Khan, S.B.; Asiri, A.M. Scanning Electron Microscopy: Principle and Applications in Nanomaterials Characterization, 1st ed.; Springer: Cham, Switzerland, 2018; pp. 113–145. ISBN 978-3-319-92955-2. [Google Scholar]
  90. Hasanpour, M.; Pardakhty, A.; Tajik, S. The development of disposable electrochemical sensor based on MoSe2-rGO nanocomposite modified screen-printed carbon electrode for amitriptyline determination in the presence of carbamazepine, application in biological and water samples. Chemosphere 2022, 308, 136336. [Google Scholar] [CrossRef] [PubMed]
  91. Kondori, T.; Tajik, S.; Akbarzadeh-T, N.; Beitollahi, H.; Graiff, C. Screen-printed electrode modified with co-NPs, as an electrochemical sensor for simultaneous determination of doxorubicin and dasatinib. J. Iran. Chem. Soc. 2022, 19, 4423–4434. [Google Scholar] [CrossRef]
  92. Hassan, S.S.; Kamel, A.H.; Fathy, M.A. A novel screen-printed potentiometric electrode with carbon nanotubes/polyaniline transducer and molecularly imprinted polymer for the determination of nalbuphine in pharmaceuticals and biological fluids. Anal. Chim. Acta 2022, 1227, 340239. [Google Scholar] [CrossRef] [PubMed]
  93. Saenchoopa, A.; Klangphukhiew, S.; Somsub, R.; Talodthaisong, C.; Patramanon, R.; Daduang, J.; Daduang, S.; Kulchat, S. A disposable electrochemical biosensor based on screen-printed carbon electrodes modified with silver nanowires/hpmc/chitosan/urease for the detection of mercury (ii) in water. Biosensors 2021, 11, 351. [Google Scholar] [CrossRef]
  94. Liao, Y.; Zhang, R.; Wang, H.; Ye, S.; Zhou, Y.; Ma, T.; Zhu, J.; Pfefferleb, L.D.; Qian, J. Highly conductive carbon-based aqueous inks toward electroluminescent devices, printed capacitive sensors and flexible wearable electronics. RSC Adv. 2019, 9, 15184–15189. [Google Scholar] [CrossRef]
  95. Wang, D.; Zhang, W.; Wang, J.; Li, X.; Liu, Y. A high-performance, all-solid-state Na+ selective sensor printed with eco-friendly conductive ink. RSC Adv. 2023, 13, 16610–16618. [Google Scholar] [CrossRef] [PubMed]
  96. Sharma, V.; Singh, P.; Kumar, A.; Gupta, N. Electrochemical detection of dopamine by using nickel supported carbon nanofibers modified screen-printed electrode. Diam. Relat. Mater. 2023, 133, 109677. [Google Scholar] [CrossRef]
  97. Fahelelbom, K.M.; Saleh, A.; Al-Tabakha, M.M.; Ashames, A.A. Recent applications of quantitative analytical FTIR spectroscopy in pharmaceutical, biomedical, and clinical fields: A brief review. Rev. Anal. Chem. 2022, 41, 21–33. [Google Scholar] [CrossRef]
  98. Mattioli, I.A.; Cervini, P.; Cavalheiro, É.T.G. Screen-printed disposable electrodes using graphite-polyurethane composites modified with magnetite and chitosan-coated magnetite nanoparticles for voltammetric epinephrine sensing: A comparative study. Microchim. Acta 2020, 187, 318. [Google Scholar] [CrossRef] [PubMed]
  99. Montiel, N.F.; Parrilla, M.; Beltrán, V.; Nuyts, G.; Van Durme, F.; De Wael, K. The opportunity of 6-monoacetylmorphine to selectively detect heroin at pre-anodized screen-printed electrodes. Talanta 2021, 226, 122005. [Google Scholar] [CrossRef]
  100. Inam, A.K.M.S.; Angeli, M.A.C.; Douaki, A.; Shkodra, B.; Lugli, P.; Petti, L. An Aptasensor based on a flexible screen-printed silver electrode for the rapid detection of chlorpyrifos. Sensors 2022, 22, 2754. [Google Scholar] [CrossRef] [PubMed]
  101. Sala, O. Fundamentals of Raman and Infrared Spectroscopy, 2nd ed.; UNESP: São Paulo, Brazil, 2011; pp. 30–65. ISBN 9788571398689. [Google Scholar]
  102. Romay, L.; Nuñez-Marinero, P.; Perales-Rondon, J.V.; Heras, A.; del Campo, F.J.; Colina, A. New Screen-Printed Electrodes for Raman Spectroelectrochemistry. Determination of p-Aminosalicylic Acid. Anal. Chim. Acta 2024, 1325, 343095. [Google Scholar] [CrossRef] [PubMed]
  103. Castaño-Guerrero, Y.; Romaguera-Barcelay, Y.; Moreira, F.T.C.; Brito, W.R.; Fortunato, E.; Sales, M.G.F. Poly(Thionine)-Modified Screen-Printed Electrodes for CA 19-9 Detection and Its Properties in Raman Spectroscopy. Chemosensors 2022, 10, 92. [Google Scholar] [CrossRef]
  104. Matvieiev, O.; Šelešovská, R.; Vojs, M.; Marton, M.; Michniak, P.; Hrdlička, V.; Hatala, M.; Janíková, L.; Chýlková, J.; Skopalová, J.; et al. Novel Screen-Printed Sensor with Chemically Deposited Boron-Doped Diamond Electrode: Preparation, Characterization, and Application. Biosensors 2022, 12, 241. [Google Scholar] [CrossRef] [PubMed]
  105. Gomes, R.S.; Gomez-Rodríguez, B.A.; Fernandes, R.; Sales, M.G.F.; Moreira, F.T.C.; Dutra, R.F. Plastic Antibody of Polypyrrole/Multiwall Carbon Nanotubes on Screen-Printed Electrodes for Cystatin C Detection. Biosensors 2021, 11, 175. [Google Scholar] [CrossRef]
  106. Pradela-Filhoa, L.A.; Andreotti, I.A.A.; Carvalho, J.H.S.; Araujoa, D.A.G.; Orzari, L.O.; Gatti, A.; Takeuchia, R.M.; Santos, A.L.; Janegitz, B.C. Glass varnish-based carbon conductive ink: A new way to produce disposable electrochemical sensors. Sens. Actuators B Chem. 2020, 305, 127433. [Google Scholar] [CrossRef]
  107. Hoppe, T.D. Electrochemically Synthesized Molecularly Imprinted Polymer on Screen-Printed Carbon Electrode for Electrochemical Determination of Paracetamol in Water. Master’s Thesis, Federal University of Santa Catarina, Blumenau, Brazil, 2024. [Google Scholar]
  108. Kim, M.; Park, D.; Park, J.; Park, J. Bio-Inspired Molecularly Imprinted Polymer Electrochemical Sensor for Cortisol Detection Based on O-Phenylenediamine Optimization. Biomimetics 2023, 8, 282. [Google Scholar] [CrossRef]
  109. de Souza, C.C.; Alves, G.F.; Lisboa, T.P.; Matos, M.A.C.; Matos, R.C. Low-cost paper-based electrochemical sensor for the detection of ciprofloxacin in honey and milk samples. J. Food Compos. Anal. 2022, 112, 104700. [Google Scholar] [CrossRef]
  110. Guo, M.; Chen, Y.; Mo, X.; Sun, K.; Du, Y.; Hu, F. Construction of a microfluidic electrochemical sensor based on screen printing electrode for the detection of fluoxetine. Microchem. J. 2024, 204, 111103. [Google Scholar] [CrossRef]
  111. Silva, F.W.L.; Name, L.L.; Tiba, D.Y.; Braz, B.F.; Santelli, R.E.; Canevari, T.C.; Cincotto, F.H. High sensitivity, low-cost, and disposability: A novel screen-printed electrode developed for direct electrochemical detection of the antibiotic ceftriaxone. Talanta 2024, 266, 125075. [Google Scholar] [CrossRef] [PubMed]
  112. Rodríguez, J.; Castañeda, G.; Lizcano, I. Electrochemical sensor for leukemia drug imatinib determination in urine by adsorptive striping square wave voltammetry using modified screen-printed electrodes. Electrochim. Acta 2024, 269, 668–675. [Google Scholar] [CrossRef]
  113. Qiu, S.; Wu, C.; Akter, S.; Hong, S.; Liu, H.; Mahmud, S. Copper oxide–modified screen-printed carbon electrodes for the detection of mycobacterium tuberculosis infection treating drug: Isoniazid. Ionics 2024, 30, 8631–8645. [Google Scholar] [CrossRef]
  114. Mehmandoust, M.; Soylak, M.; Erk, N. Innovative molecularly imprinted electrochemical sensor for the nanomolar detection of Tenofovir as an anti-HIV drug. Talanta 2023, 253, 123991. [Google Scholar] [CrossRef] [PubMed]
  115. Tajik, S.; Beitollahi, H.; Nejad, F.G. Novel and simple electrochemical sensing platform based on polypyrrole nanotubes/ZIF-67 nanocomposite/screen printed graphite electrode for sensitive determination of metronidazole. Electrochem. Commun. 2024, 169, 107824. [Google Scholar] [CrossRef]
  116. Naghian, E.; Marzi Khosrowshahi, E.; Sohouli, E.; Ahmadi, F.; Rahimi-Nasrabadi, M.; Safarifard, V. A new electrochemical sensor for the detection of fentanyl lethal drug by a screen-printed carbon electrode modified with the open-ended channels of Zn(ii)-MOF. New J. Chem. 2020, 44, 9271–9277. [Google Scholar] [CrossRef]
  117. Materón, E.M.; Wong, A.; Freitas, T.A.; Faria, R.C.; Oliveira, O.N. A sensitive electrochemical detection of metronidazole in synthetic serum and urine samples using low-cost screen-printed electrodes modified with reduced graphene oxide and C60. J. Pharm. Anal. 2021, 11, 646–652. [Google Scholar] [CrossRef] [PubMed]
  118. Stelmaszczyk, P.; Białkowska, K.; Sekuła, K.; Stanaszek, R.; Wietecha-Posłuszny, R. Screen-printed electrode-based sensor for rapid ketamine determination: Optimization and on-site application for seized drugs analysis. Monatshefte Chem.—Chem. Mon. 2024, 155, 881–888. [Google Scholar] [CrossRef]
  119. Zinatloo-Ajabshir, S.; Mahmoudi-Moghaddam, H.; Amiri, M.; Akbari Javar, H. A green and simple procedure to synthesize dysprosium cerate plate-like nanostructures and their application in the electrochemical sensing of mesalazine. J. Mater. Sci. Mater. Electron. 2024, 35, 500. [Google Scholar] [CrossRef]
  120. Ali, M.R.; Bacchu, M.S.; Al-Mamun, M.R.; Ahommed, M.S.; Saad Aly, M.A.; Khan, M.Z.H. N-Hydroxysuccinimide crosslinked graphene oxide–gold nanoflower modified SPE electrode for sensitive detection of chloramphenicol antibiotic. RSC Adv. 2021, 11, 15565–15572. [Google Scholar] [CrossRef] [PubMed]
  121. Silva, L.R.G.; dos Santos, G.F.S.; Vasconcellos, M.L.S.; Ferreira, R.Q. Development of electroanalytical procedure for monitoring of metamizole in organic fertilizers (human urine and struvite) associated with portable equipment. J. Environ. Manag. 2020, 266, 110587. [Google Scholar] [CrossRef] [PubMed]
  122. Seguro, I.; Rebelo, P.; Pacheco, J.G.; Delerue-Matos, C. Electropolymerized, Molecularly Imprinted Polymer on a Screen-Printed Electrode—A Simple, Fast, and Disposable Voltammetric Sensor for Trazodone. Sensors 2022, 22, 2819. [Google Scholar] [CrossRef]
  123. Wang, T.; Jafar, N.N.A.; Al-Rihaymee, A.M.A.; Alhameedi, D.Y.; Rasen, F.A.; Hashim, F.S.; Hussein, T.K.; Ramadan, M.F.; Alasedi, K.K.; Suliman, M.; et al. Highly efficient electrocatalytic oxidation of levodopa as a Parkinson therapeutic drug based on modified screen-printed electrode. Heliyon 2024, 10, 34689. [Google Scholar] [CrossRef] [PubMed]
  124. Martoni, L.V.L. Development of Printed Photoelectrochemical Sensors Modified with Carbon Nanospheres for Quantification of Emerging Pollutants in Water Supply Samples. Ph.D. Thesis, Institute of Chemistry of São Carlos, University of São Paulo, São Carlos, Brazil, 2023. [Google Scholar]
  125. Kozak, J.; Tyszczuk-Rotko, K.; Gorylewski, D. A nanoporous screen-printed carbon sensor for environmental and clinical monitoring of the antibiotic ciprofloxacin. Measurement 2023, 222, 113626. [Google Scholar] [CrossRef]
  126. Jahani, P.M.; Tajik, S.; Dourandish, Z. Electrochemical Sensor Based on Ce-MOF Modified Screen-Printed Electrode for Metronidazole Determination. Chem. Methodol. 2024, 8, 123–132. [Google Scholar]
  127. Tajik, S.; Dourandish, Z.; Nejad, F.G.; Aghaei Afshar, A.; Beitollahi, H. Voltammetric Determination of Isoniazid in the Presence of Acetaminophen Utilizing MoS2-Nanosheet-Modified Screen-Printed Electrode. Micromachines 2022, 13, 369. [Google Scholar] [CrossRef] [PubMed]
  128. Tyszczuk-Rotko, K.; Kozak, J.; Węzińska, A. Electrochemically Activated Screen-Printed Carbon Electrode for Determination of Ibuprofen. Appl. Sci. 2021, 11, 9908. [Google Scholar] [CrossRef]
  129. Kozak, J.; Tyszczuk-Rotko, K.; Wójciak, M.; Sowa, I. Electrochemically Activated Screen-Printed Carbon Sensor Modified with Anionic Surfactant (aSPCE/SDS) for Simultaneous Determination of Paracetamol, Diclofenac and Tramadol. Materials 2021, 14, 3581. [Google Scholar] [CrossRef] [PubMed]
  130. Sopaj, F.; Loshaj, F.; Contini, A.; Mehmeti, E.; Veseli, A. Preparation of an efficient and selective voltammetric sensor based on screen printed carbon ink electrode modified with TiO2 nanoparticles for Azithromycin quantification. Results Chem. 2023, 6, 101123. [Google Scholar] [CrossRef]
  131. Sanli, S. Single-drop electrochemical immunosensor with 3D-printed magnetic attachment for onsite smartphone detection of amoxicillin in raw milk. Food Chem. 2024, 437, 137823. [Google Scholar] [CrossRef] [PubMed]
  132. Silva, L.R.G.; Rodrigues, J.G.A.; Vasconcellos, M.L.S.; Ribeiro, E.S.; D’Elia, E.; de, Q. Ferreira, R. Portable and simple electroanalytical procedure for simultaneous detection of dipyrone and norfloxacin with disposable commercial electrodes in water and organic fertilizers. Ionics 2022, 28, 4833–4841. [Google Scholar] [CrossRef]
  133. Pacheco, J.C.; Brito, J.A.R.; dos Santos, C.C.; Bezerra, C.W.B.; Damos, F.S.; Luz, R.C.S. A simple method for determination of chloroquine based on electrogenerated chemiluminescence. J. Braz. Chem. Soc. 2023, 34, 1410–1419. [Google Scholar] [CrossRef]
  134. Zanoni, C.; Rovida, R.; Magnaghi, L.R.; Biesuz, R.; Alberti, G. Voltammetric Detection of Irbesartan by Molecularly Imprinted Polymer (MIP)-Modified Screen-Printed Electrodes. Chemosensors 2022, 10, 517. [Google Scholar] [CrossRef]
  135. Haššo, M.; Kekeľáková, A.; Hanko, M.; Švorc, Ľ. Powerful Analytical Platform for Diazepam Determination in Pharmaceuticals and Alcoholic Drinks Based on Batch Injection Analysis Coupled with Adsorptive Stripping Voltammetry. J. Electrochem. Soc. 2024, 171, 047517. [Google Scholar] [CrossRef]
  136. Alfadhel, M.; Alrobaian, M.; Arida, H. Fabrication of New Potentiometric Microsensor for Metformin Based on Modified Screen-Printed Microchip. Int. J. Electrochem. Sci. 2021, 16, 210660. [Google Scholar] [CrossRef]
  137. Garima; Sachdev, A.; Matai, I. An electrochemical sensor based on cobalt oxyhydroxide nanoflakes/reduced graphene oxide nanocomposite for detection of illicit drug-clonazepam. J. Electroanal. Chem. 2022, 919, 116537. [Google Scholar] [CrossRef]
  138. Wang, K.; Yang, L.; Huang, H.; Lv, N.; Liu, J.; Liu, Y. Nanochannel Array on Electrochemically Polarized Screen-Printed Carbon Electrode for Rapid and Sensitive Electrochemical Determination of Clozapine in Human Whole Blood. Molecules 2022, 27, 2739. [Google Scholar] [CrossRef]
  139. Beitollahi, H.; Tajik, S.; Aflatoonian, M.R.; Nejad, F.G.; Zhang, K.; Asl, M.S.; Van Le, Q.; Cha, J.H.; Shokouhimehr, M.; Peng, W. A Novel Screen-Printed Electrode Modified by graphene Nanocomposite for Detecting Clozapine. Int. J. Electrochem. Sci. 2020, 15, 9271–9281. [Google Scholar] [CrossRef]
Figure 1. Representation of a conventional electrochemical cell, composed of three electrodes: working, reference, and auxiliary. The images are copyright material owned by the authors.
Figure 1. Representation of a conventional electrochemical cell, composed of three electrodes: working, reference, and auxiliary. The images are copyright material owned by the authors.
Analytica 06 00003 g001
Figure 2. Representation of a printed electrode, indicating the auxiliary, working, and reference electrodes. The images are copyright material owned by the authors.
Figure 2. Representation of a printed electrode, indicating the auxiliary, working, and reference electrodes. The images are copyright material owned by the authors.
Analytica 06 00003 g002
Figure 3. Schematic representation of the process of printing electrodes by screen-printing. The images are copyright material owned by the authors.
Figure 3. Schematic representation of the process of printing electrodes by screen-printing. The images are copyright material owned by the authors.
Analytica 06 00003 g003
Figure 4. Schematic representation of the components of a carbon-based conductive ink. The images are copyright material owned by the authors.
Figure 4. Schematic representation of the components of a carbon-based conductive ink. The images are copyright material owned by the authors.
Analytica 06 00003 g004
Figure 5. Searches related to article publications between 1974 and 2023, with the term “chemically modified electrode”, carried out in the databases (A) Scopus, available at https://bit.ly/2BLC3Jo (accessed on 11 September 2024), and (B) Web of Science, available at https://bit.ly/3dIngwu (accessed 11 September 2024).
Figure 5. Searches related to article publications between 1974 and 2023, with the term “chemically modified electrode”, carried out in the databases (A) Scopus, available at https://bit.ly/2BLC3Jo (accessed on 11 September 2024), and (B) Web of Science, available at https://bit.ly/3dIngwu (accessed 11 September 2024).
Analytica 06 00003 g005
Figure 6. Schematic representation of different materials applied to modify the surface of screen-printed electrodes. The images are copyright material owned by the authors.
Figure 6. Schematic representation of different materials applied to modify the surface of screen-printed electrodes. The images are copyright material owned by the authors.
Analytica 06 00003 g006
Figure 7. Representation of information obtained by the CV, EIS, SEM, TEM, FTIR, XDR, and Raman techniques. The images are copyright material owned by the authors.
Figure 7. Representation of information obtained by the CV, EIS, SEM, TEM, FTIR, XDR, and Raman techniques. The images are copyright material owned by the authors.
Analytica 06 00003 g007
Table 1. Printed electrochemical sensors for drug determination.
Table 1. Printed electrochemical sensors for drug determination.
Sensor ConfigurationDetection TechniqueAnalyteLinear RangeDetection LimitAnalyte/SampleRef.
COFTpTt@c-MWCNTs/SPCEDPVFluoxetine62.5–2000 ng mL−124.3 ng mL−1Spiked rat serum[110]
SPE/SiO2/ZrO2/Cdot-NSWVCeftriaxone0.0078–40.02 μmol L−10.2 nmol L−1Urine and tap water[111]
SPCE/MWCNT-COOHSWVImatinib50–912 nmol L−17.0 nmol L−1Urine[112]
CuO/SPCEDPVIsoniazid4.0–200 μmol L−18.39 μmol L−1 Pharmaceutical formulations[113]
MIP-Pt@g-C3N4/F-MWCNT/SPEDPVTenofovir 0.005–0.69 μmol L−10.0030 μmol L−1Human plasma, urine, and water[114]
PPyNTs/ZIF-67/SPGEDPVMetronidazole0.01–500 μmol L−10.004 μmol L−1Pharmaceutical formulations and urine samples[115]
Zn(II)-MOF/SPCEDPVFentanyl 1–100 μmol L−1 0.3 μmol L−1Urine and plasma[116]
C60-rGO-NF/SPESWVMetronidazole 0.25-34 µmol L−10.21 µmol L−1Serum and urine[117]
SPEDPVKetamine 50–500 µmol L−115.0 µmol L−1Pharmaceutical formulations and seized drugs[118]
PL-Dy2Ce2O7/SPEDPVMesalazine0.02–145 µmol L−10.008 µmol L−1Pharmaceutical formulations, human blood serum, and urine[119]
SPE/rGO–NHS–AuNFsDPVChloramphenicol0.05–100 µmol L⁻¹1.0 nmol L⁻¹Blood serum, poultry feed, milk, eggs, honey, and powdered milk[120]
SPCESWVMetamizole 0.35–50 µmol L−10.09 µmol L−1Human urine[121]
MIP/SPCEDPVTrazodone5–80 µmol L−11.6 µmol L−1Human serum and water[122]
ZnO/Co3O4NC/SPEDPVLevodopa 0.001–800.0 µmol L−10.81 nmol L−1Levodopa pills and urine[123]
SPE-CSSSWVAcetaminophen,
hydroquinone
1.5–14.0 µmol L−1
3.0–23.0 µmol L−1
0.85 µmol L−1
2.7 µmol L−1
Tap water and tablets[124]
NPSPCESWVCiprofloxacin0.0005–0.03 µmol L−16.3 × 10−5 µmol L−1Sewage and bodily fluids[125]
Ce-BTC MOF/SPGEDPVMetronidazole0.05–400.0 µmol L⁻¹0.02 µmol L−1Pharmaceutical formulations and urine[126]
MoS2 NSs-SPEDPVIsoniazid 0.035–390.0 µmol L⁻¹10.0 n mol L−1Pharmaceutical formulations and urine[127]
aSPCEDPVIbuprofen0.50–20.0 µmol L−1 and
20.0–500.0 µmol L−1
0.059 µmol L−1Pharmaceutical formulations[128]
aSPCE/SDSDPAdSVParacetamol,
diclofenac,
tramadol
0.0149 µmol L⁻¹
0.00021 µmol L⁻¹
0.00171 µmol L⁻¹
14.87 nmol L−1
0.21 nmol L−1
1.71 nmol L−1
River water, serum samples, and pharmaceuticals formulations[129]
SPCIE-TiO2 NPsCVAzithromycin 0.05–50 µmol L⁻¹0.93 µmol L⁻¹Urine and water[130]
SPE/MNP-TA/Ab AmxSWVAmoxicillin 0.50–100 µmol L⁻¹0.44 µmol L⁻¹Milk[131]
SPE-BDDSWVDipyrone,
norfloxacin
2.0–250 µmol L⁻¹
2.0–62.5 µmol L⁻¹
0.30 µmol L⁻¹
0.40 µmol L⁻¹
Water and organic fertilizers[132]
CB/SPELSVChloroquine 0.5–500 µmol L−10.5 µmol L−1Milk[133]
MIP/SPCsSWVIrbesartan 20–220 nmol L⁻¹0.012 µmol L⁻¹Water[134]
SPCEBIA-SWAdSVDiazepam 5–40 µmol L⁻¹2.0 µmol L⁻¹Pharmaceuticals and spiked alcoholic drinks[135]
SPEsPOTMetformin 10−5–1.0 mol L−14 µmol L⁻¹Pharmaceutical formulations[136]
CoOOH-rGO/SPCEDVPClonazepam0.1–350 µmol L⁻¹5.6 µmol L⁻¹Drink samples[137]
VMSF/pSPCEDPVClozapine0.050–20 μmol L⁻¹28 nmol L⁻¹Human blood[138]
Go/Fe3O4/SiO2/SPEDPVClozapine0.1–700.0 μmol L⁻¹0.03 μmol L⁻¹Tablets and urine[139]
DPV: differential pulse voltammetry; SWV: square wave voltammetry; MIP: molecularly imprinted polymer; CV: cyclic voltammetry; COFTpTt: covalent organic framework material; c-MWCNTs: carboxylated multi-walled carbon nanotubes; Cdot-Ns: nitrogen-doped carbon quantum dots; MWCNT-COOH: multi-walled carbon nanotubes modified with carboxyl groups; MOF: metal–organic framework; NF: Nafion; C60: fullerene; rGO: graphene oxide; SDS: sodium dodecyl sulfate; DPAdSV: differential pulse adsorptive stripping voltammetry; Dy: dysprosium; Ln2Ce2O7: lanthanide cerate; MOF: metal–organic framework; SPCIE: screen-printed carbon ink electrode; TiO2 NPs: tetragonal (rutile) nanostructures of TiO2 nanoparticles; MNP: magnetic nanoparticles; LSV: linear sweep voltammetry; ePADs: electrochemical paper-based analytical devices; BIA: batch injection analysis; SWAdSV: square-wave adsorptive stripping voltammetry; SPCE: screen-printed carbon electrode; BDD: boron-doped diamond; SPEs: screen-printed electrodes; SPCs: graphite working electrode of screen-printed cells; SPE/MNP-TA/Ab Amx: screen-printed electrode, magnetic nanoparticle amino-functionalized iron oxide anti-amoxillim antibodies; CB: carbon black; POT: potentiometry; Dp: dipyrone; NF: norfloxacin; PA: paracetamol; TR: tramadol; DF: diclofenac; SWCNT: single-walled carbon nanotube; CSS: carbon spherical shells; NPSPCE: nanoporous screen-printed; PRO: propranolol; TIM: timolol; AML: amlodipine; TRI: triamterene; SPE: screen-printed electrode; SCPE: screen-printed carbon electrode; VMSF: mesoporous silica nanomembrane film.
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

Leite, V.A.R.; Oliveira, S.P.d.; Souza, L.C.d.; Silva, L.J.d.P.; Silva, L.F.; Cândido, T.C.d.O.; Silva, D.N.d.; Pereira, A.C. Development of Novel Conductive Inks for Screen-Printed Electrochemical Sensors: Enhancing Rapid and Sensitive Drug Detection. Analytica 2025, 6, 3. https://doi.org/10.3390/analytica6010003

AMA Style

Leite VAR, Oliveira SPd, Souza LCd, Silva LJdP, Silva LF, Cândido TCdO, Silva DNd, Pereira AC. Development of Novel Conductive Inks for Screen-Printed Electrochemical Sensors: Enhancing Rapid and Sensitive Drug Detection. Analytica. 2025; 6(1):3. https://doi.org/10.3390/analytica6010003

Chicago/Turabian Style

Leite, Victor Alexandre Ribeiro, Sthephane Pereira de Oliveira, Larissa Cristina de Souza, Léa Júlia de Paula Silva, Laís Fonseca Silva, Thaís Cristina de Oliveira Cândido, Daniela Nunes da Silva, and Arnaldo César Pereira. 2025. "Development of Novel Conductive Inks for Screen-Printed Electrochemical Sensors: Enhancing Rapid and Sensitive Drug Detection" Analytica 6, no. 1: 3. https://doi.org/10.3390/analytica6010003

APA Style

Leite, V. A. R., Oliveira, S. P. d., Souza, L. C. d., Silva, L. J. d. P., Silva, L. F., Cândido, T. C. d. O., Silva, D. N. d., & Pereira, A. C. (2025). Development of Novel Conductive Inks for Screen-Printed Electrochemical Sensors: Enhancing Rapid and Sensitive Drug Detection. Analytica, 6(1), 3. https://doi.org/10.3390/analytica6010003

Article Metrics

Back to TopTop