Next Article in Journal
Correction: Oliveira et al. Effect of Polymer Hydrophobicity in the Performance of Hybrid Gel Gas Sensors for E-Noses. Sensors 2023, 23, 3531
Previous Article in Journal
Differentially Private Client Selection and Resource Allocation in Federated Learning for Medical Applications Using Graph Neural Networks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Innovations in Biosensor Technologies for Healthcare Diagnostics and Therapeutic Drug Monitoring: Applications, Recent Progress, and Future Research Challenges

1
School of Biotechnology, Badr University in Cairo (BUC), Badr City 11829, Egypt
2
Department of Biochemistry, Faculty of Pharmacy, Badr University in Cairo (BUC), Badr City 11829, Egypt
3
Biochemistry and Molecular Biology Department, Faculty of Pharmacy (Boys), Al-Azhar University, Nasr City 11231, Egypt
4
Pharmacology and Toxicology Department, Faculty of Pharmacy, Badr University in Cairo (BUC), Badr City 11829, Egypt
5
Department of Biochemistry, Faculty of Pharmacy, Tanta University, Tanta 31527, Egypt
6
Medical Biophysics, Department of Physics, Faculty of Science, Helwan University, Cairo 11795, Egypt
7
School of Applied Health Sciences, Badr University in Cairo (BUC), Badr City 11829, Egypt
8
Refractories, Ceramics and Building Materials Department, National Research Centre, 33 El Bohouth St., Giza 12622, Egypt
9
Department of Chemical Engineering, Loughborough University, Loughborough LE11 3TU, UK
10
Chemistry Department, Faculty of Science, Benha University, Benha 13511, Egypt
*
Author to whom correspondence should be addressed.
Sensors 2024, 24(16), 5143; https://doi.org/10.3390/s24165143 (registering DOI)
Submission received: 4 July 2024 / Revised: 1 August 2024 / Accepted: 6 August 2024 / Published: 8 August 2024
(This article belongs to the Section Biosensors)

Abstract

:
This comprehensive review delves into the forefront of biosensor technologies and their critical roles in disease biomarker detection and therapeutic drug monitoring. It provides an in-depth analysis of various biosensor types and applications, including enzymatic sensors, immunosensors, and DNA sensors, elucidating their mechanisms and specific healthcare applications. The review highlights recent innovations such as integrating nanotechnology, developing wearable devices, and trends in miniaturisation, showcasing their transformative potential in healthcare. In addition, it addresses significant sensitivity, specificity, reproducibility, and data security challenges, proposing strategic solutions to overcome these obstacles. It is envisaged that it will inform strategic decision-making, drive technological innovation, and enhance global healthcare outcomes by synthesising multidisciplinary insights.

Graphical Abstract

1. Introduction

The healthcare landscape is evolving quickly, propelled by technological innovations, scientific advancements, and paradigmatic shifts, reshaping traditional models, methodologies, and modalities of patient care, disease management, and therapeutic interventions [1,2,3]. Currently, numerous digital biomedical tools are accessible for assessing different facets of human activity and capabilities. The global count of linked wearable gadgets is projected to surge from approximately 300 million in 2016 to more than 1 billion by 2024 [4]. Rapid technology advancement is fuelling the growth of digital hospitals, e-health, and telemedicine. The use of electronic health monitoring in patient diagnosis and treatment is growing dramatically as healthcare staff become scarcer. Wireless ubiquitous monitoring has become a key tool in the biomedical and healthcare industries in recent years [5,6,7].
A biosensor is an advanced analytical device designed with precision to detect subtle changes in complex biological processes, converting these delicate variations into clear electrical signals. Biosensors integrate molecular recognition elements (MREs) with transducers. MREs vary from cells, receptors, and antibodies to synthetic materials like Molecularly Imprinted Polymers (MIPs) and Peptide Nucleic Acids (PNAs). These MREs are categorised into affinity and catalytic groups; the former includes MIPs and antibodies, while the latter features cells and enzymes [8,9,10,11,12]. Enhancing biosensor durability involves using specialised enzymes and microbes. Notably, MIPs offer covalent and non-covalent variants, crucial for biomedical sensors that detect compounds such as herbicides [13,14], beta-estradiol sensors [15,16,17], and chloramphenicol sensors [18,19].
With increasing significance and versatility, biosensors have found various applications, as graphically summarised in Figure 1. Traditional disease management and diagnostics approaches heavily relied on time-consuming clinical examinations, laboratory tests, imaging studies, and invasive procedures in healthcare. These methods often involved resource-intensive protocols and delayed interventions. However, introducing biosensors has transformed diagnostic approaches by enabling swift, precise, and cost-effective detection, identification, and monitoring of diseases in their early stages. This revolution has facilitated prompt interventions, personalised treatment plans, and enhanced patient outcomes [20,21,22,23,24].
In addition to diagnostics, biosensors are revolutionising therapeutic interventions, drug development, and personalised medicine initiatives by facilitating real-time monitoring tailored to individual patient profiles, genetic predispositions, and therapeutic responses. These biosensors enable precise pharmacokinetic/pharmacodynamic analyses, drug efficacy evaluations, and therapeutic monitoring protocols that empower clinicians to optimise drug dosages, minimise adverse effects, and enhance treatment efficacy while minimising therapeutic failures, drug resistance, and medication-related complications [25,26,27,28,29].
Emerging developments like wearable biosensors, point-of-care testing tools, implantable devices, telemedicine applications, and remote monitoring systems are broadening healthcare access, enhancing patient involvement, and reshaping healthcare delivery models to meet evolving societal demands, technological advancements, and healthcare priorities [30,31,32,33].
This review provides an insightful analysis of the progress in the contributions of biosensors in healthcare, emphasising their innovative applications and transformative potential. It also aims to inspire scientific inquiry, stimulate interdisciplinary collaboration, and inform evidence-based practices and policies that address complex healthcare challenges, promote global health equity, and enhance human well-being.

2. Fundamentals of Biosensors: Essential Components, Working Principle and Types

2.1. Essential Components of Biosensors

At the core of biosensors lies the bioreceptor, a pivotal element meticulously crafted to recognise and bind with the target analyte (Figure 2). This component fundamentally dictates the biosensor’s specificity and sensitivity, employing diverse arrays of bioreceptors, each endowed with distinct advantages [34]. Antibodies originating from the immune system exhibit exceptional specificity and can be tailored for various disease indicators, exemplifying their utility in biosensors [35]. Enzymes, acting as biological catalysts, undergo activity changes upon binding to the analyte, furnishing a measurable signal, particularly beneficial in biosensors designed for detecting specific enzymatic reactions [36]. Aptamers, short DNA or RNA sequences, showcase high target affinity and ease of modification, rendering them versatile in diverse biosensor applications [37]. Additionally, cells, whether as a whole or their components like organelles or membranes, offer a dynamic approach to biosensor design, responding to analytes by altering their electrical properties or signalling activity [38]. This intricate interplay of bioreceptors underscores the biosensor functionality’s nuanced and multifaceted nature.
In biosensors, the transducer is crucial as it transforms the biorecognition event into a measurable signal, which is essential for detecting the target substance. The choice of transducer depends on the bioreceptor type and the desired detection approach. Various transducers are commonly employed in biosensor design. Electrochemical transducers, encompassing amperometric, potentiometric, and conductometric devices, measure changes in electric currents, potential, or conductivity, with applications such as amperometric biosensors quantifying analytes through the measurement of current generated during electrochemical reactions [39,40,41]. Optical transducers gauge alterations in light intensity, wavelength, or polarisation upon interaction with the analyte, allowing biosensors to utilise fluorescence to detect specific molecules [42,43]. Thermal transducers measure heat changes associated with the biorecognition event, finding applications in biosensors focused on temperature-sensitive reactions [44]. Piezoelectric transducers sense minute changes in mass or pressure utilised in biosensors, capitalising on alterations in mass due to analyte binding [45]. Mass-sensitive transducers quantify changes in the mass of the bioreceptor induced by analyte binding, providing a sensitive detection method [46,47]. This array of transducers underscores the versatility and adaptability of biosensors across diverse detection modalities.
Electronics emerge as a critical component tasked with amplifying and processing signals generated by the transducer, often characterised by their initial weakness, necessitating enhancement for meaningful interpretation [48,49]. The electronics unit plays a pivotal role in refining the signal, filtering out noise, and converting it into a user-friendly format suitable for display or data analysis. Beyond this, biosensors may incorporate additional components to optimise their functionality for specific applications. Microfluidic channels regulating sample and buffer solution flows enhance precision and efficiency in analyte detection [50]. Integrated microchips reduce the size and improve the performance of biosensors, making them appropriate for portable diagnostics and point-of-care testing [51]. Incorporating biocompatible materials that do not interfere with biorecognition processes is crucial to maintaining the precision and dependability of biosensors in detecting target substances [52].

2.2. Working Principles of Biosensors

Molecular recognition is the fundamental principle driving biosensor functionality, wherein the biorecognition element selectively engages with the target analyte (Figure 3). This specific and precise interaction is the cornerstone of the biosensor’s ability to identify and detect specific substances. A prime example of molecular recognition occurs in immunoassays, marked by the interaction between antigens and antibodies. Within this context, specialised proteins generated by the immune system selectively identify and attach to matching antigens, signalling the existence of the targeted analyte. This high degree of molecular specificity empowers biosensors with exceptional accuracy, enabling them to identify and capture particular molecules amid the complexity of a sample [53,54].
Following molecular recognition, the biosensors employ signal transduction mechanisms to transform the biological interaction into a detectable signal. This process is essential for converting the binding event into a format suitable for quantitative analysis. For instance, biosensors often employ enzymatic reactions to achieve this signal transmission. In these instances, when the analyte binds to the enzyme, it initiates a particular biochemical response that produces an electrochemical indication. This measurable signal provides a quantitative representation of the analyte’s presence, allowing for precise detection. Signal transduction mechanisms thus play a pivotal role in bridging the biological recognition event with a tangible output, facilitating the interpretation of biosensor results in a practical and meaningful way [55,56,57].
The analyte and sensor interactions are crucial in understanding how biosensors operate. When the analyte binds to the biorecognition component, it initiates a series of responses within the sensor [58]. For example, a glucose biosensor where glucose, the analyte, binds to an enzyme specifically designed for glucose recognition. This binding event sets off a responsive mechanism, manifesting as changes in electrical conductivity or the generation of a measurable signal. These responses are instrumental in the accurate detection and quantification of glucose levels [59,60].
A profound comprehension of these interactions between analyte and sensor forms the bedrock for the thoughtful design and optimisation of biosensors across diverse applications. Whether applied to monitor glucose levels in diabetic individuals or to identify specific biomarkers for disease diagnosis, the precision and sensitivity inherent in these interactions underscore the effectiveness and reliability of biosensor technology across a spectrum of fields.

2.3. Non-Specific Adsorption in Biosensors

Non-specific adsorption is the unwanted attachment of molecules to a surface and negatively impacts biosensor performance. It interferes with the desired interactions between molecules and surfaces, leading to inaccurate results and decreased sensitivity. There are two main approaches to reducing non-specific adsorption, namely, passive and active methods [61]. Passive methods involve modifying the surface properties to make it less attractive to non-specific molecules. This can be performed by coating the surface with a blocking agent, such as bovine serum albumin (BSA) or polyethylene glycol (PEG). Blocking agents create a steric barrier that prevents non-specific molecules from coming into close contact with the surface. Active methods involve using external forces to remove non-specific molecules from the surface. This can be performed by applying a shear force, such as with a liquid flow, or by using an electric field. The best approach for reducing non-specific adsorption will depend on the specific application. Passive methods are generally simpler to implement, but they may not be as effective as active methods. Active methods can be more complex, but they can be more effective in removing non-specific molecules [62,63].

2.4. Types of Biosensors

Biosensors play pivotal roles in contemporary diagnostics, highlighting various designs and functionalities. This thorough exploration of various biosensor categories sheds light on their underlying operational principles and presents various applications for each type. Table 1 delineates key applications for each biosensor type, underscoring their versatility and significance in healthcare. The following is an exploration of the diverse types of biosensors.
Enzymatic biosensors, anchored in enzymatic recognition elements, drive reactions that yield measurable signals, ensuring exceptional selectivity. In this process, enzymes facilitate reactions with the target analyte, producing substances that lead to measurable signals. Electrodes, commonly employed as transducers, then convert these signals into electrical responses [64]. A prime example of enzymatic biosensors is found in glucose monitoring, where glucose biosensors, prominently utilising glucose oxidase, are indispensable in managing diabetes. This enzyme catalyses the conversion of glucose, producing hydrogen peroxide, and the ensuing electrochemical signal, proportional to glucose concentration, facilitates real-time monitoring [65]. In a recent paper, Khalife et al. [66] developed an enzymatic biosensor for the quick and non-redox probe detection of vitamin K3. Engineered YaiB was fixed onto a carbon electrode, showing exceptional sensitivity and specificity. The biosensor facilitated rapid measurement in milk samples spiked with the compound, achieving a detection limit of 0.87 µM in buffer and 4.1 µM in milk [66].
Immunosensors utilise antibodies to detect proteins, pathogens, or specific molecules precisely. At their core, immunosensors operate through antibody-antigen interactions, where the existence of the targeted antigen induces binding, generating measurable signals detected through various transduction methods [67]. The swift, sensitive, and specific characteristics of immunosensors find exemplification in applications like pregnancy tests, where antibodies specific to human chorionic gonadotropin (hCG) facilitate its detection, demonstrating their efficacy in clinical diagnostics [68]. In recent research, as illustrated by Chaturvedi et al. [69], an advanced electrochemical immunosensor was developed using a nano-bio-inspired platform of reduced graphene oxide (rGO), polydopamine (pDA), and gold nanoparticles (AuNPs). These components were immobilised with anti-NS1 antibodies to detect the dengue NS1 protein. The sensor exhibits a linear response range from 1 ng mL−1 to 100 µg mL−1, with a sensitivity of 1.78 ng mL−1 mA−1, enabling early diagnosis of dengue virus infections.
DNA biosensors, employing nucleic acids as recognition elements, showcase remarkable precision in identifying specific DNA or RNA sequences. Their operational concept revolves around the pairing of target DNA with a matching probe, causing alterations in electrical or optical characteristics that convert into observable signals [70]. In the realm of genetic diagnostics, DNA biosensors play a pivotal role in detecting genetic mutations linked to diseases, with hybridisation events serving as sensitive indicators for early disease detection [71]. Recent research, as showcased by Ali et al. [72], introduces a label-free electrochemical DNA biosensor using a flexible indium tin oxide electrode to selectively detect Shigella flexneri. The study demonstrates a broad dynamic range, low detection limits, exceptional specificity, and successful deployment in spiked food samples.
Aptasensors, employing aptamers as the recognition elements, utilise single-stranded DNA or RNA molecules, offering remarkable versatility in target detection. Operating akin to immunosensors, aptasensors generate signals through the unique folding of aptamers upon binding to the target, which are then transduced for detection [73]. Their applications span a broad range, finding utility in both environmental monitoring and medical diagnostics due to their adaptability [74]. In recent research, Wu et al. [75] presented an extremely sensitive aptasensor for detecting deoxynivalenol. They utilised gold electrodes coated with polyethyleneimine-functionalised porous graphene oxide modified with gold nanowires. The sensor achieved a detection limit of 2.23 × 10−9 mg/mL through signal amplification using methylene blue-labelled zeolitic imidazolate framework-8.
Electrochemical biosensors, capitalising on changes in electrical properties, provide swift and sensitive measurements. These biosensors operate through biochemical reactions inducing measurable electrical changes, with electrodes facilitating detection through amperometric, potentiometric, or impedimetric methods [76]. Recent studies, such as those by Zhang et al. [77] showcase an electrochemical biosensor designed for highly sensitive detection of the HER2 protein. The biosensor utilises Apt-PNA probes and magnetic nanocomposites, achieving a detection limit of 4.1 femtograms per millilitre (fg·mL−1). The device showed encouraging outcomes in spiked serum samples.
Optical biosensors, engineered to detect variations in light properties, provide essential high sensitivity for real-time monitoring. Their working principle involves alterations in optical properties, such as absorbance or fluorescence, resulting from binding events at the recognition element and producing measurable signals [78,79]. Within this category, Surface Plasmon Resonance (SPR) biosensors play a prominent role and are widely utilised in drug discovery and environmental monitoring [80]. Recent research, exemplified by Yildizhan et al. [81] introduced a highly sensitive bioassay for detecting breast cancer-specific extracellular vesicles (EVs) in blood plasma. The study utilised a fibre-optic surface plasmon resonance (FO-SPR) biosensor, achieving limits of detection (LODs) of 2.1 × 107 particles/mL for SK-BR-3 EVs and 1.1 × 108 particles/mL for MCF7 EVs. This biosensor demonstrated outstanding sensitivity and specificity in clinical applications [81].
Piezoelectric biosensors, crafted to measure mass fluctuations, function by modifying the resonance frequency of a quartz crystal. The mechanism entails binding interactions at the recognition site that cause mass alterations on the crystal surface, resulting in frequency adjustments directly correlated to the analyte’s concentration. Notably, piezoelectric biosensors have proven invaluable for real-time, label-free monitoring of biomolecules, showcasing their exceptional sensitivity and versatility [82]. In their research conducted in 2023, Chen and Shi [83] developed a continuous flow immunoassay system that utilised a piezoelectric quartz crystal biosensor to detect alpha-fetoprotein (AFP) as a marker for tumours. The device demonstrated a quick response time of 12 min, detecting AFP at a rate of 45.9 nanograms per minute and detecting concentrations ranging from 18.8 to 1100 ng/mL. Although the specific details about the limits of detection and quantification were not specified, the biosensor underwent validation using clinical serum samples. The study confirmed that the biosensor did not react with lung cancer markers and strongly agreed with radioimmunoassay (RIA) findings. These findings indicate potential applications in cancer diagnosis and pharmaceutical research [83].
Table 1. Applications of diverse biosensor types in various healthcare scenarios.
Table 1. Applications of diverse biosensor types in various healthcare scenarios.
Biosensor TypeApplicationBiosensor FormulationRef.
Enzymatic BiosensorIdentification of the endocrine-disrupting chemical Bisphenol A (BPA).Carbon paste electrode modified with Pd(II)-loaded PAMAM dendrimer and immobilised tyrosinase enzyme via crosslinking using glutaraldehyde.[84]
Glucose monitoring in diabetes patients.A microneedle composed of polylactic acid (PLA), coated sequentially with gold nanoparticles (AuNPs), glucose oxidase (GOx), overoxidised polypyrrole (OPPy), and Nafion.[85]
ImmunosensorQuick and precise detection of the antibiotic enrofloxacin (EF) in meat.Anti-quinolone antibody on carbon electrodes with difloxacin-amino ferrocene electrochemical probe.[86]
Detecting SARS-CoV-2 in saliva using a combined vertical and lateral flow approach.A container housing immunological reagents attached to magnetic beads alongside a lateral flow device featuring a polyester-based electrode, a magnet, and an absorbent pad.[87]
DNA biosensorDetecting Human papillomavirus-16
(HPV-16)
Glass electrodes coated with indium tin oxide (ITO), modified using graphene oxide and gold nanoparticles coated with silver, and immobilised with specific DNA probes for HPV-16 detection.[88]
Colorimetric AptasensorDetection of Ochratoxin A (OTA).DNA circuit driven by entropy (EDC), catalytic hairpin assembly (CHA), and magnesium ion-assisted DNAzyme catalysis (MNAzyme).[89]
Electrochemical BiosensorHER2 biomarker detection for breast cancerCd2+-aptamer@AMNFs@ZIF-67 nanocomposite[90]
Measuring levels of folic acid in expectant mothersDihydrofolate reductase (DHFR) anchored on a gold electrode modified with carbon-multiwalled carbon nanotubes (c-MWCNT) and titanium dioxide nanoparticles (TiO2NPs)[91]
Optical biosensorDetection of basal cell cancerArrays of one-dimensional photonic crystal (PC) structures linked with two metal–insulator–metal (MIM) plasmonic waveguides, utilising tapering techniques for enhanced alignment.[92]
Tracking biomarkers (MUC-1), utilising fluorescence imaging, and delivering targeted curcuminA nanosystem of mesoporous silica, chitosan, and gold (MCM@CS@Au) targeted by aptamers for MUC-1 positive tumour cells.[93]
Piezoelectric biosensorSelf-monitoring smart vascular grafts.A biocompatible piezo composite with a low filler content of 5% volume and high piezoelectric sensitivity (g33~130 mV m N−1) features structured sodium niobate (NaNbO3) fibres embedded in an elastomeric matrix through dielectrophoresis.[94]

3. Recent Advancements in Biosensor Technology

3.1. Nanotechnology in Biosensors

Incorporating nanotechnology into biosensors marks a revolutionary advancement, significantly impacting their design and performance. This paradigm shift introduces a diverse array of unique nanomaterials and structures that augment sensitivity and expand the horizons of detection limits. Nanoparticles, nanotubes, and nanowires have emerged as pivotal components, playing a transformative role in biosensing applications and contributing to remarkable progress in healthcare [95,96].
The integration of nanomaterials has ushered in a new era of biosensor development, significantly advancing healthcare technologies. Table 2 illustrates the transformative impact of nanotechnology on biosensor sensitivity and detection capabilities. Notably, nanoparticles such as gold nanoparticles have been crucial in enhancing the electrochemical performance of biosensors [97]. This enhancement is largely due to the high surface-to-volume ratios of nanoparticles, which facilitate extensive biomolecule immobilization and interactions. As a result, biosensors incorporating nanomaterials exhibit heightened sensitivity and improved efficiency in binding target molecules. The incorporation of these advanced materials has led to the creation of state-of-the-art biosensors that revolutionize healthcare by enabling precise detection of biomarkers associated with various diseases and providing accurate drug monitoring platforms [98].
Table 2. The utilisation of nano-enhanced biosensors across a spectrum of healthcare applications.
Table 2. The utilisation of nano-enhanced biosensors across a spectrum of healthcare applications.
Biosensor TypeUsed NanoparticleApplicationRef.
Label-free, ultrasensitive electrochemical biosensorGold nanoparticles (AuNPs)Detection of transferrin (Tf), a crucial serum biomarker for atransferrinemia.[99]
Electrochemical biosensorIron oxide nanoparticles capped with L-cysteine.Determination of amino acids using
L-Phenylalanine as the representative analyte.
[100]
Ultrafast and ultrasensitive DNA biosensorNanocubic architecture of MnFe@Pt crystalsFast and precise determination of SARS-CoV-2 RNA via RNA reverse transcription to DNA (rtDNA).[101]
Multiple-signal amplification PEC biosensorCarboxylated graphitic carbon nitride (g-C3N4) paired with avidin-functionalised ruthenium-coated silica nanoparticles (Ru@SiO2).Ultrasensitive detection of kanamycin residues in foods[102]
Electrochemical biosensorMulti-walled carbon nanotubes are functionalised with acid and tungsten wires coated with gold.Rapid and easy detection of Escherichia coli
(E. coli)
[103]
Optical biosensorGold nanoparticles (AuNPs) from green tea leavesBiosensing of CD44 cancer biomarkers with improved sensitivity.[104]
MicroRNA biosensor based on magnetic rod carbon paste electrodesCarbon nanofibers combined with CuBTC-AIA (Copper-based Metal-Organic Framework) and Fe@rGO (Iron-reduced Graphene Oxide).Detection of microRNA 155 for the diagnosis of breast cancer.[105]
Nano immunosensorGold nanoparticles (AuNPs)Detection and discrimination of Zika and Dengue viruses[106]
Electrochemical aptasensorPolypyrrole (PPy)-gold nanoparticles (AuNPs)Detection of cardiac troponin I (cTnI) without labels for diagnosing acute myocardial infarction (AMI).[107]
Fluorescence biosensor without enzymes or labels, operating on a ratiometric principleDNA-silver nanoclusters (DNA-AgNCs)Susceptible detection of tetracycline (TET) for antibiotic detection in food security[108]
Dual-mode nano-biosensor combining ratiometric fluorescence and calorimetryCarbon dots (BCDs) and Manganese dioxide nanosheets (MnO2 NSs)Determination of Staphylococcus aureus
(S. aureus).
[109]
Surface plasmon resonance biosensorTiO2/Au/grapheneA combined TiO2/Au/graphene layer-based surface plasmon resonance (SPR) sensor showed superior sensitivity (210 to 292.86 deg/RIU) for detecting cervical, skin, adrenal gland, breast, and blood cancer cells.[110]

3.2. Integration with Wearable Devices

Integrating biosensors with wearable devices is revolutionising healthcare, bringing about continuous monitoring and personalised insights. Seamlessly woven into everyday attire, wearable biosensors facilitate real-time data acquisition, empowering proactive health management [111]. In a groundbreaking development, Li et al. [112] unveiled a novel wearable device for ongoing surveillance of the highly transmissible severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) and its mutations. This innovative method utilises 2D MXene-graphene composite films in a flexible field-effect transistor (FET) sensor. Embedded on masks and enhanced with Bluetooth connectivity for wireless data transfer, this sensor exhibits remarkable responsiveness to influenza and SARS-CoV-2 viruses in airborne environments and exhaled breath. This innovative wearable biosensor addresses the pressing need for continuous monitoring tools amid the ongoing challenges posed by respiratory viruses, particularly SARS-CoV-2.
In a separate study, Li et al. [113] pioneered a point-of-care (POC) sensing platform for diagnosing early chronic kidney disease (CKD). Employing reduced Graphene Oxide/Polydopamine-Molecularly Imprinted Polymer (rGO/PDA-MIP) with surface-molecular imprinting, the platform detects CKD biomarkers (creatinine, urea and human serum albumin) simultaneously. It integrates a multi-channel electrochemical POC readout system utilising differential pulse voltammetry, achieving an unmatched femtomolar-level limit-of-detection for all biomarkers. Clinical validation in serum and urine demonstrates efficacy, emphasising its cost-effectiveness, user-friendliness, and rapid sample-to-result time. The platform holds promise for resource-limited settings in CKD diagnosis and progression tracking.
Additionally, Yuksel et al. [68] developed an automated electrochemical point-of-care biosensor for detecting human chorionic gonadotropin (hCG) in urine, a critical biomarker for early pregnancy. The biosensor utilised custom gold and Ag/AgCl electrodes, optimised using COMSOL Multiphysics® 5.2a software. DSP self-assembled monolayers were employed for covalently immobilising anti-hCG-beta antibodies, facilitating a sandwich assay with anti-hCG alpha-HRP. The device featured an innovative agitation design to accelerate surface reactions, enabling rapid detection of very low hCG concentrations. This research highlights the biosensor’s potential as a precise and efficient tool for early pregnancy diagnosis in clinical settings.

3.3. Miniaturisation and Portability

The evolution of biosensor technology has witnessed a transformative trend driven by advancements in microfabrication techniques, propelling the miniaturisation of biosensors to the forefront. This paradigm shift, characterised by reduced dimensions and enhanced portability, has profound implications for diagnostics, particularly in point-of-care applications. In this comprehensive exploration, we delve into the intricacies of miniaturisation and portability in biosensors, unveiling their potential through recent studies and innovations [114].
The miniaturisation of biosensors is closely linked to advancements in microfabrication techniques, enabling precise engineering of miniature structures and components. Microelectromechanical systems (MEMS) and microfluidics are pivotal in shaping biosensors at micro and nanoscales. MEMS principles involve the incorporation of mechanical elements, actuators, sensors, and electronics on a microscopic scale, allowing for susceptible and compact biosensor designs. Microfluidics principles focus on manipulating small fluid volumes within microchannels, facilitating controlled sample delivery and reaction kinetics in biosensing applications [115].
Miniaturised biosensors excel in providing rapid analyses with significantly reduced sample volumes, leveraging their inherent advantages of accelerated reaction kinetics for quicker response times. This attribute is vital in point-of-care settings, prioritising timely results. Working with small sample volumes offers significant benefits, particularly when samples are scarce. The compact size of miniaturised biosensors makes them well-suited for point-of-care use, allowing healthcare providers to conduct immediate analyses and make timely decisions and interventions [116].

4. Biosensors Application in Disease Detection

Biosensors are crucial in transforming disease detection and have become essential instruments in healthcare, as seen in Table 3. This section explores the diverse applications of biosensors in diagnosing various diseases, highlighting their versatility and significant impact on early detection and effective disease management. In the current healthcare landscape, where precision and timeliness are crucial, biosensors have emerged as invaluable contributors, providing revolutionary insights that transform diagnostics and enable proactive and personalised patient care. With their innovative capabilities, biosensors serve as essential allies, facilitating the early detection of diseases and enabling timely and targeted interventions. Their influence spans diverse medical domains, showcasing the versatility of biosensors as crucial assets in advancing healthcare outcomes.

4.1. Biosensors in Cancer Detection

Biosensors, utilising molecular recognition principles with antibodies or aptamers, are crucial for cancer diagnosis. They employ diverse sensing mechanisms such as optical (e.g., SPR), electrochemical, and mass-based (e.g., QCM) biosensors, enabling accurate measurement of cancer indicators such as prostate-specific antigen (PSA) or facilitating continuous monitoring of breast cancer biomarkers. These sensors are pivotal in converting specific interactions into detectable signals in diverse contexts, providing essential live data for efficient cancer diagnosis and therapy [117,118].
In a recent research paper, Wan et al. [119] explored innovative methods for detecting advanced breast cancer using biomarkers found in saliva. They developed specialised one-time-use strips that achieved highly sensitive detection levels, detecting HER2 and CA15-3 biomarkers at concentrations as low as 1 femtogram per millilitre. By employing a synchronised double-pulse technique, the study showcased swift testing speeds (under 15 milliseconds) and minimal sample needs (3 microliters of saliva). These advancements suggest a potential revolution in public health practices due to their easy-to-use design.

4.2. Biosensors in Infectious Disease Detection

Biosensors play a crucial role in diagnosing infectious diseases by using specialised recognition components such as antibodies or nucleic acids to interact with molecules associated with pathogens selectively. Immunosensors rely on antibodies binding to pathogen antigens, triggering measurable signals, while nucleic acid sensors detect genetic material through methods like PCR. Whole-cell-based mechanisms employ engineered living cells for signal production. Biosensors, adaptable to emerging threats, were notably used in colourimetric biosensors for visual detection of SARS-CoV-2 RNA during the COVID-19 pandemic [120,121].
In a groundbreaking 2024 study, Khan et al. [122] developed an aptamer-based biosensor for rapid virus detection, significantly advancing large-scale COVID-19 identification. The biosensor formulation used a 60-base DNA aptamer targeting the S protein, immobilised on a gold nanoparticle-modified electrode with methylene blue (MB) as the redox indicator (Figure 4). Computational approaches identified the GLN104-ASP115 pocket as the aptamer’s binding site. The biosensor exhibited a broad linear detection range (10 pM to 6 nM) and a 91.1 pM limit of detection using differential pulse voltammetry. Testing in human saliva and serum demonstrated its capability for direct, single-step detection of viral proteins, highlighting its potential for COVID-19 and other viral agents.

4.3. Biosensors in Diabetes Management

Biosensors have revolutionised diabetes management by providing real-time monitoring, reducing reliance on invasive methods. Continuous glucose monitoring (CGM) systems lead the way in this advancement, seamlessly incorporating biosensor technologies to provide a constant flow of glucose information. This ongoing monitoring enables people to make informed choices regarding insulin doses and lifestyle changes instantly [123].
In an innovative 2024 research led by Li et al. [124], they developed a wearable microneedle sensor designed for continuous glucose monitoring (CGM). The device includes a microneedle electrochemical biosensor with three electrodes and a fully integrated radio-chemical analysis system. By electrodepositing Prussian blue (PB), crosslinking glucose oxidase (GOx), and utilising chitosan to form a 3D network with glutaraldehyde (GA), the sensor demonstrated long-term functionality in diabetic rats. The PB-based CGM system exhibited excellent sensitivity, stability, and resistance to interference, accurately monitoring glucose levels in the interstitial fluid (ISF) in real-time during experiments on diabetic rats. This study underscores the feasibility and promising clinical application of PB-based CGM for managing diabetes effectively. While biosensor advancements in diabetes management are promising, challenges such as accuracy, user adherence, and cost-effectiveness persist. Future directions involve refining sensor accuracy, improving data interpretation algorithms, and addressing issues of affordability and accessibility.

4.4. Biosensors in Neurological Disorder Detection

Biosensors have become transformative tools in early neurological disorder diagnosis, particularly in Alzheimer’s and Parkinson’s diseases. These devices, leveraging molecular recognition capabilities, offer unparalleled precision in detecting disease-associated biomarkers, surpassing traditional diagnostic methods. Besides Alzheimer’s, biosensors are crucial in detecting Parkinson’s disease early, facilitating timely identification and personalised treatment initiation. The accuracy and adaptability of biosensors offer a valuable understanding of the intricate realm of neurological conditions, highlighting the importance of early detection and specific interventions [125,126].
In a 2023 study by Altay et al. [127], a low-cost disposable immunosensor was created for the quick and sensitive measurement of the Aβ42 protein linked to Alzheimer’s disease. Though Electrochemical Impedance Spectroscopy (EIS) and Cyclic Voltammetry (CV), the immunosensor exhibited a linear detection span from 1 to 100 picograms per mL with a minimal detectable level of 0.37 picograms per mL, suggesting feasible clinical use for assessing Aβ42 in human blood serum. In a recent 2024 investigation led by Karaboğa et al. [128], an innovative biosensing strategy employed a quartz tuning fork (QTF) for the first time to develop an immunosensor targeting alpha-synuclein protein (SYN alpha), a potential biomarker for Parkinson’s disease. The biosensor, utilising a self-assembled monolayer and gold nanoparticles on QTFs, exhibited a selective detection range from 1 to 500 nanograms per mL and a low detection limit of 0.098 nanograms per mL. Significantly, the QTF-based biosensor effectively captured SYN alpha in cerebrospinal fluid samples, achieving recovery rates between 92% and 104%.

4.5. Biosensors in Cardiovascular Disease Detection

Cardiovascular diseases (CVDs) remain a major global health concern, necessitating sophisticated diagnostic instruments for timely intervention. Biosensors are crucial in quickly identifying cardiovascular conditions, especially myocardial infarction (MI) and heart failure. Representing a shift in cardiovascular disease diagnosis, biosensors offer swift and precise identification of specific biomarkers associated with cardiac events. Their impact is essential for timely interventions, leading to substantial improvements in patient outcomes and relieving the burdens on healthcare systems. The timely and accurate diagnosis of myocardial infarction, a critical aspect of the intervention, is facilitated by biosensors known for their sensitivity and specificity in detecting cardiac biomarkers related to MI [129,130].
In a study by Li et al. [131], they introduced an integrated microfluidic platform (IMS) designed for swift measurement of biomarkers associated with cardiovascular diseases (NT-proBNP, fibrinogen, cTnI, and CRP). Using aptamer-coated interdigitated electrodes and integrated circuits, the system enabled effective plasma filtration, quantification, and detection in just 15 min. This technology shows promise for assessing CVD risk and personalised medicine, achieving reasonable accuracy (>80%) in artificially spiked blood samples. In another study by Hui et al. [132], myocardial infarction (MI) diagnosis is addressed. Existing techniques lack early-stage sensitivity, prompting a focus on blood biomarkers. The study develops an interdigitated electrode sensor for troponin I (cTnI) detection, using aptamer-gold-cTnI-antibody sandwich patterns, achieving a 100 aM limit of detection and superior nonfouling in serum, offering potential for novel MI assays.

4.6. Biosensors in Autoimmune Disorder Detection

Biosensors are essential in diagnosing autoimmune disorders, specifically in conditions such as systemic lupus erythematosus (SLE) and rheumatoid arthritis (RA). Biosensors detect autoantibodies with high sensitivity and specificity, providing a targeted approach for early diagnosis. In autoimmune disorders, biosensors likely employ strategic recognition elements to capture disease-specific autoantibodies, utilising nanomaterials or surface modifications for enhanced sensitivity. The binding of autoantibodies initiates a measurable signal, allowing for quantification through analytical techniques. The ability to detect autoantibodies at low concentrations highlights biosensors’ potential for accurate and early diagnosis [133,134].
Focusing on RA and SLE, biosensors offer a transformative approach to identifying distinct autoantibodies associated with these conditions. Early detection through biosensors enables timely interventions, enhancing symptom control and preventing irreversible tissue damage [135]. In a study by Chen et al. [136], a novel method for rheumatoid arthritis (RA) detection was introduced using bacteriorhodopsin as a photoelectric transducer. The biosensor, employing a PM monolayer-coated electrode and a citrullinated-inter-alpha-trypsin inhibitor heavy chain 3 (ITIH3) 542–556 peptide, showed significantly reduced photocurrents with RA-patient serum compared to healthy controls. Labelling with anti-IgA-AuNP further amplified the distinction, demonstrating a close correlation with commercial assays, offering a sensitive and specific RA diagnosis in a single- or two-step mode. In another study by Jupe et al. [137], a decentralised digital approach utilising a biosensor for systemic lupus erythematosus (SLE) patients was implemented. Machine learning evaluated patient-reported outcomes (PROs), quality of life (QOL), and biometric data to forecast potential disease exacerbations. Digital surveys and smartwatches, acting as biosensors for biometric monitoring, collected data from 550 participants. The study resulted in a 26-feature model accurately distinguishing disease flare risks, facilitating proactive screening for clinical assessment.

4.7. Biosensors in Respiratory Disease Detection

Biosensors are pivotal in the field of respiratory illnesses, providing a revolutionary method for detection and treatment. The mechanism of biosensing involves the precise identification of biomarkers, often present in exhaled breath, through tailored recognition elements. In chronic obstructive pulmonary disease (COPD), biosensors excel in differentiating specific biomarkers associated with the condition, enabling accurate disease identification and personalised treatment strategies. This capability is especially vital in pulmonary fibrosis, where biosensors aid in the early identification of targeted biomarkers, paving the way for timely interventions and enhanced patient outcomes. The non-invasive nature of biosensor-driven diagnostics, combined with their capacity to interpret complex biomarker profiles, establishes them as essential instruments in advancing respiratory healthcare [138,139].
In a groundbreaking study by Qian et al. [140], a microdroplet analysis platform was developed as a biosensor, combining rolling circle amplification (RCA) and surface-enhanced Raman spectroscopy (SERS). This innovative, technology-enabled, accurate detection of miRNA-21 and miRNA-155 in the serum of patients with idiopathic pulmonary fibrosis (IPF) demonstrates high sensitivity and specificity with very low detection limits. The integrated microfluidic system addressed challenges related to SERS reproducibility. This biosensor, when used alongside chest computed tomography, emerged as a supplementary diagnostic tool for assessing IPF risk and monitoring miRNA levels in real-time during chemotherapy, highlighting its potential for versatile miRNA detection across different diseases [140].
Table 3. Examples of key biosensor applications in diverse disease diagnosis.
Table 3. Examples of key biosensor applications in diverse disease diagnosis.
Target DiseaseBiosensorTransducer TypeApplicationLODLinear RangeRef.
Prostate cancerMaackia amurensis (MAA) lectin immobilised on gold-interdigitated microelectrodesElectrochemicalAccurate identification of cancer-linked abnormal glycosylation in prostate-specific antigen (PSA).3.574 pg/mL0.01–100 ng/mL[141]
Gold nanospikes (AuNS) arranged on a quartz crystal microbalance (QCM).ElectrochemicalDetection and assessment of prostate-specific antigen (PSA) for prostate cancer screening24 pg/mL0.1–100 ng/mL[142]
Colon and rectal cancerElectrochemiluminescence immunosensor with Ru@TiO2-MXene as the energy donor and Pd@UiO-66-NH2 as the energy acceptorECL-RET (Electrochemiluminescence Resonance Energy Transfer)Detection of carcinoembryonic antigen (CEA) for diagnosis of colon and rectal cancer2.65 fg/mL1 × 10−5–80 ng/mL[143]
Colorectal cancerAn electrochemical immunosensor employing nucleic acid aptamer recognition and silver depositionElectrochemicalIdentification of epidermal growth factor receptor (EGFR) for colorectal cancer diagnosis0.06 ng/mL1 to 1000 ng/mL[144]
COVID-19 and InfluenzaLiquid-gated graphene field-effect transistors (GFETs) with quadruple architecture and individual functionalisationElectricalRapid and ultraprecise differentiation and sensing of influenza and SARS-CoV-2 surface proteins~50 ag/mL, or 88 zM for COVID-19 and 227 zM for FluN/D[145]
DiabetesNon-invasive tear glucose biosensor based on a photoelectrochemical probe with optical fiberPhotoelectroch-emicalNon-invasive glucose detection.4.1 nM10 nM to 100 μM[146]
Polyaniline/nickel oxide nanohybrid modified graphite sheet (PANI/NiO/GS) biosensorElectrochemicalNon-enzymatic electrochemical detection of methylglyoxal in human saliva.2.64 nM1 to 10 μM[147]
Phenylboronic acid (PBA)-based glucose biosensor with HCNT/PEDOT:PSS dual conductive structureElectrochemicalGlucose detection with exceptional sensitivity, stability, and wide linearity rangeN/D0.20 mM to 2.0 mM[148]
Alzheimer’s diseaseLabel-free impedimetric immunosensor using indium tin oxide polyethylene terephthalate (ITO-PET) electrodes modified with anti-Aβ42 antibodiesElectrochemicalFast, specific, and highly sensitive quantitative assessment of Aβ42 protein for diagnosing Alzheimer’s disease0.37 pg/mL1–100 pg/mL[127]
A dual “turn-on” fluorescence biosensor using aggregation-induced emission fluorogen (AIEgen)-labelled oligonucleotide (TPET-DNA) probes attached to cationic dextran-modified molybdenum disulfide (TPET-DNA@Dex-MoS2).FluorescenceFast, specific, and highly sensitive quantitative analysis of miR-125b for diagnosing Alzheimer’s disease and monitoring in real-time in PC12 cells and brain tissues of a mouse model of Alzheimer’s disease.N/DN/D[149]
Parkinson’s diseaseA biosensor based on a thin-film transistor (TFT) utilising indium gallium zinc oxide (IGZO) thin film technologyElectrochemicalEarly detection of Parkinson’s disease using surface-functionalised IGZO TFT biosensorN/D1 pg mL−1 to 100 ng mL−1[150]
Cardiovascular diseaseAn electrochemical immunosensor utilising nanostructured molybdenum tetraselenide reduced graphene oxide (nMo3Se4-rGO) for detecting cardiac troponin I (cTnI), offering a broader linear range, increased sensitivity, and lower detection limit.ElectrochemicalEarly detection of cardiovascular disease using nMoSe-rGO-based biosensor for cTnI detection1 fg mL−11 fg mL−1–100 ng mL−1[151]
Rheumatoid arthritis (RA)Tilted-fiber Bragg grating (TFBG) biosensor bonded with cyclic citrullinated peptide (CCP) antigens and antibodiesOpticalLaboratory-based detection of rheumatoid arthritis (RA) using cyclic citrullinated peptide (CCP) antibodies and antigens1 ng/mLN/D[152]
Magnetic microbeads functionalised with Neutravidin (NA-MBs) and modified with biotinylated anti-double-stranded DNA (dsDNA) for amperometric detection of anti-dsDNA autoantibodies (IgG, IgA, and IgM AAbs).AmperometricDetection of anti-dsDNA autoantibodies in the sera of rheumatoid arthritis patients0.3 IU mL−11–200 IU mL−1[153]
Chronic obstructive pulmonary disease (COPD)Point-of-care biosensors for the multiplexed detection of interleukin (IL)-6, IL-8, matrix metalloproteinase (MMP)-8, MMP-9, C-reactive protein (CRP), tumour necrosis factor-alpha (TNF-α), and neutrophil elastase (NE)ElectrochemicalIdentification of protein biomarkers in saliva and sputum for the management of chronic obstructive pulmonary disease (COPD).N/DN/D[139]

5. Biosensors in Therapeutic Drug Monitoring

Biosensors are pivotal in monitoring therapeutic drugs, providing swift, sensitive, and precise detection of drug levels in biological samples. This helps in fine-tuning drug doses, ensuring treatment effectiveness, and reducing potential side effects [154,155]. Table 4 reviews some applications of biosensors in therapeutic drug monitoring. By continuously monitoring drug levels in real-time or near-real-time, biosensors enable personalised dosing regimens, enhancing patient care in various clinical settings, including critical care, oncology, psychiatry, and infectious diseases [156]. Additionally, biosensors offer potential cost savings by reducing the need for labour-intensive laboratory techniques and minimising healthcare resource utilisation, thus improving patient outcomes and overall healthcare efficiency [157].
Therapeutic drug monitoring (TDM) involves tracking the plasma concentration of drugs with a narrow therapeutic range. This is achieved through various analytical methods such as HPLC, gas chromatography, and, more recently, biosensors [158,159]. The main problem with drugs with a narrow therapeutic index is that they usually cause toxic effects at the therapeutic dose (an overlap exists between the effectiveness and the toxicity of the drug). For this reason, they require special care during their administration by continuously monitoring and analysing their pharmacokinetic and pharmacodynamic profile [160,161].
Vancomycin, a glycopeptide antibiotic, is a robust antibiotic that faces resistance from an increasing number of bacterial strains, often due to inappropriate administration. Furthermore, vancomycin is associated with a variety of side effects, mainly on the ear and the kidneys, therefore requiring close therapeutic drug monitoring [162,163,164]. Mu et al. [165] designed a fluorescence-based biosensor for rapid, highly specific, and highly sensitive quantitation of vancomycin levels in rabbits (Figure 5). The biosensor utilised a vancomycin-specific fluorescent probe, and a tripeptide (L-Lys-D-Ala-D-Ala) linked it to a dansyl chloride group. This analytical technique was coupled with a microdialysis sampling of the biological fluid to avoid blood sampling. This novel biosensor can effectively replace other analytical methods (HPLC and immunoassay) used to measure and monitor vancomycin in humans since it requires no special sample preparation or complicated instrumentation. It holds the potential to be implemented in analysing the pharmacokinetic profile of vancomycin in clinical settings, thus avoiding unnecessary nephrotoxicity, ototoxicity and patient suitability to vancomycin-resistant strains.
Warfarin, frequently prescribed as an anticoagulant with a narrow therapeutic range, requires careful monitoring to prevent excessive bleeding. It acts as a vitamin K antagonist by inhibiting vitamin K epoxide reductase, thus reducing the levels of active vitamin K and impairing the synthesis of vitamin K-dependent clotting factors (II, VII, IX, and X) [166].
Saeedi et al. fabricated a novel electrochemical biosensor based on ion-sensing capability, creating a biosensing electrode that uses tetradodecylammonium chloride as an ion exchanger [167]. In this study, they designed a portable point-of-care testing device that is exceptionally precise and specific for monitoring warfarin levels in patient blood during treatment. The device achieved a detection limit of 1.4 × 10−5 M in blood samples. These results highlight the potential of the electrochemical-based biosensor in therapeutic drug monitoring protocols to optimise clinical and therapeutic outcomes effectively [167]. Other laborious methods, such as capillary electrophoresis and HPLC, can test warfarin, but they are costly and time-consuming, making routine therapeutic monitoring of warfarin a challenging task [168,169].
In another study, Ozbek et al. [170] constructed a potentiometric biosensor to monitor the level of valproic acid in human blood. The biosensor showed a fast response time (less than 10 s), good stability, and a detection limit of 9.75 × 10−7 mol L−1. The results obtained from the biosensor were also compared with the analytical instrument utilised in the hospital where the research was conducted, and the results were compatible. These results suggest that valproate-sensitive biosensors can replace the conventional, time-consuming, and expensive instruments used in hospitals to monitor patients administering drugs with narrow therapeutic indexes such as valproic acid.
Since chemotherapeutic agents also require close monitoring, a study by Alvau et al. designed an irinotecan-sensitive electrochemical biosensor [171]. Irinotecan’s mode of action in cancer therapy is by inhibiting topoisomerase I enzyme. However, irinotecan can also inhibit acetylcholine esterase, which was exploited to design the irinotecan-sensitive biosensor. Their biosensor was based on using cholinesterase and choline oxidase enzymes. The presence of irinotecan hampers the reaction catalysed by those enzymes, and this decrease in the activity can be correlated with the level of irinotecan in the sample [172]. These findings require further research and analysis to provide an irinotecan monitoring test that can be used to monitor patients administering irinotecan.
Another chemotherapeutic drug that requires better analytical instrumentation in its monitoring is doxorubicin. Doxorubicin, like any other chemotherapeutic agent, possesses off-target toxicity. Therefore, optimising the dose administered to cancer patients and monitoring it is vital in doxorubicin therapy [173]. A doxorubicin-sensitive sensor was designed by using a glassy carbon electrode carrying single-wall carbon nanotubes. The single-wall carbon nanotubes were then loaded with double-stranded DNA. The analysis is based on doxorubicin intercalating into the double-stranded DNA grooves and causing an electrochemical change sensed by the electrode. The biosensor showed promising results regarding the sensitivity and selectivity of doxorubicin. The biosensor had a limit of detection of 0.6 nM [174]. A sensitive and selective biosensor for monitoring doxorubicin is promising to be implemented after further validation studies.
Pharmacokinetic analysis is vital in the healthcare system since it allows healthcare experts to monitor how a candidate drug is absorbed, metabolised and finally excreted from the body. Pharmacokinetics focuses on the level of the drug in the body and eventually linking it to the therapeutic outcome [175]. Nowadays, pharmacokinetic data are generated by measuring the concentration of a candidate drug in blood samples. However, these blood-based data can sometimes reflect poor correlation with the actual drug response and hence alter drug monitoring. This is typically seen in cancer therapy [176].
Seo et al. [177] developed an implantable microelectrode array biosensor to generate real-time tumour tissue-based pharmacokinetic data for a more reliable drug efficacy assessment. There was a significant difference between the tumour tissue-based and the blood-based pharmacokinetics. The tumour tissue can be exposed differently than what is reflected or recognised by analysing the blood samples. This study showed that in the future, this implantable biosensor could replace the traditional needle biopsies that are carried out to study drug levels in the tissue.
Another recent study involved designing a multifunctional biosensor to assess the efficacy of a novel epidermal growth factor receptor tyrosine kinase inhibitor (AC0010). The sensor was utilised to generate electrical impedance data that were processed to reflect tumour cell growth and death and assess the drug’s efficacy. To ensure that the biosensor generates reliable data, results were compared to those obtained from CKK8 assays. In addition, the multifunctional biosensor was also used successfully to monitor the electrophysiological and beating activity of the myocardial cells [178].
Table 4. Biosensors applications in therapeutic drug monitoring (TDM).
Table 4. Biosensors applications in therapeutic drug monitoring (TDM).
Drug MonitoredBiosensorTransducerMatrixLODLinearity RangeRef.
AmoxicillinEnzyme Biosensor with Microfiber Interferometer (MFI) and Fiber Gratings (FBGs) Power VariationOpticalDeionised water, real food, urine samples0.04 nM0.01–100 nM[179]
Tenofovir (TFV)Alkaline phosphatase (ALP) Enzyme with BaTiO3 NanoparticlesElectrochemicalHuman blood serum0.09 nMN/A[180]
KanamycinRu@MOF/Ag+-Dependent DNAzyme ECL BiosensorElectrochemiluminescenceSeawater, Milk13.7 pM30 pM–300 μM[181]
DNA Aptamer-Modified Portable Gold Electrode BiosensorElectrochemicalEnvironmental water samples0.40 μmol/L1–1000 μmol/L[182]
Thionine functionalised graphene and hierarchical nanoporous (HNP) PtCu AptasensorElectrochemicalAnimal-derived food0.42 pg mL−15 × 10−7
5 × 10−2 μg mL−1
[183]
StreptomycinColorimetric Casein hydrolysate peptides-functionalised silver nanoparticles (CHPs@AgNPs)ColourimetricTap water, Dairy whey~98 nM
(tap water)
~56 nM
(dairy whey)
200–650 nM (tap water); 100–700 nM (dairy whey)[184]
FK506Spiky Fe3O4@SiO2@Ag flower magnetic superstructure and hollow Ag@Au superstructure enhanced SERS biosensorMagnetic/PlasmonicBlood of transplant patients0.33 ng/mL0.5–20 ng/mL[185]
VincristineDNA (ds-DNA), polypyrrole (PP), peony-like CuO:Tb3+ nanostructure (P-L CuO:Tb3+ NS)ElectrochemicalPharmaceutical preparations, biological fluids0.21 nM1.0 nM–400.0 μM[186]
Imatinib (IMA)DNA/AuPt/p-L-Met Coated SPE Electrochemical DNA BiosensorElectrochemicalHuman serum, pharmaceutical samples0.18 nM2.33–80 nM[187]
Erlotinib (ERL)ElectrochemicalHuman serum, pharmaceutical samples0.009 nM0.032–1.0 nM
Doxorubicin (DOX)Gold nanoparticles (AuNPs) and DNA tetrahedron (TDN) nanoprobe bifunctional glassy carbon electrodeElectrochemicalHuman serum, cell lysate0.3 nM1.0 nM–50 μM[188]
Daunorubicin (DNR)Upconversion nanoparticles (UCNPs) as energy donor and herring sperm DNA (hsDNA) biosensor.LuminescenceBiological fluids0.60 μg·mL−11–100 μg·mL−1[189]

6. Concentration Ranges and S/N Ratios for Key Biomarkers

In addition to detection limits, it is crucial to consider the concentration ranges and signal-to-noise (S/N) ratios in healthy subjects for accurate detection of target molecules. These parameters are vital because they provide a benchmark against which abnormal levels can be identified, enhancing the reliability and sensitivity of biosensors in clinical diagnostics and monitoring. Understanding the typical concentration ranges of target molecules in healthy individuals helps in setting accurate thresholds for detecting deviations that may indicate disease or other health conditions. Similarly, the signal-to-noise (S/N) ratio is essential to distinguish true signals from background noise, ensuring precise measurements [190,191]. Table 5 provides detailed information on the concentration ranges and S/N ratios for various target molecules in healthy subjects, illustrating their clinical relevance and importance for biosensor performance.
Table 5. Concentration ranges and signal-to-noise ratios for some target molecules in healthy subjects. S/N are only for demonstration purposes since it is highly system specific.
Table 5. Concentration ranges and signal-to-noise ratios for some target molecules in healthy subjects. S/N are only for demonstration purposes since it is highly system specific.
Target MoleculeHealthy Subject Concentration RangeS/N RatioClinical RelevanceRef.
Cardiac Troponin I (cTnI)1 pg/mL–100 ng/mL10:1Indicative of myocardial infarction[192]
Anti-dsDNA Autoantibodies1–200 IU/mL10:1Marker for rheumatoid arthritis[193]
Glucose3.9–5.5 mmol/L15:1Monitoring diabetes mellitus[194]
Prostate-Specific Antigen (PSA)0–4 ng/mL15:1Screening for prostate cancer[195]
Cholesterol (Total)<200 mg/dL10:1Assessing cardiovascular health[196]
C-Reactive Protein (CRP)0.8–3.0 mg/L10:1Indicator of inflammation[197]
Hemoglobin A1c (HbA1c)4–5.6%15:1Long-term glucose level monitoring[198]
Thyroid-Stimulating Hormone (TSH)0.4–4.0 mIU/L10:1Assessing thyroid function[199]
These S/N ratios are approximate and are based on typical biosensor performance. They serve to illustrate the importance of distinguishing true signals from background noise, which is critical for the accuracy and reliability of biosensor measurements in clinical diagnostics.

7. Research Challenges for Future Applications

Biosensor technology has undoubtedly undergone remarkable advancements in healthcare. Yet, it is imperative to acknowledge that several formidable research challenges persist on the journey toward its widespread adoption and further advancement. Addressing these multifaceted challenges is essential and pivotal in unlocking the full potential of biosensors to revolutionise patient outcomes and reshape the landscape of healthcare delivery. By confronting these challenges head-on, the healthcare industry can harness the transformative power of biosensors to enhance diagnostic accuracy, improve disease management strategies, and ultimately elevate the standard of patient care to unprecedented heights.

7.1. Sensitivity, Specificity, and Reproducibility

The effectiveness of biosensors in healthcare heavily relies on their sensitivity, specificity, and reproducibility. Achieving high sensitivity ensures the detection of minute biomarker concentrations, which is crucial for early disease detection and monitoring. Specificity is equally vital to accurately identifying the target analytes amidst complex biological samples. Reproducibility, on the other hand, ensures consistent performance across various conditions and settings, minimising variability in results. However, challenges persist in maintaining these parameters due to factors such as sample matrix variations, environmental influences, and user-related variability. Future research endeavours must prioritise refining detection mechanisms, optimising sensor materials, and standardising protocols to enhance the reliability and robustness of biosensor performance [200].

7.2. Integration with Electronic Health Records (EHR) and Data Security

Seamless integration of biosensor-generated data with Electronic Health Records (EHR) holds immense promise for advancing personalised medicine and data-driven healthcare decision-making. However, achieving this integration is fraught with challenges. Interoperability issues between different systems and data formats hinder smooth data exchange, while concerns regarding data privacy and cybersecurity pose significant risks to patient confidentiality. Establishing robust data exchange standards, implementing encryption protocols, and ensuring compliance with stringent privacy regulations are imperative steps towards safeguarding sensitive patient information. By addressing these concerns, healthcare providers can harness the full potential of biosensor data while upholding patient privacy and security [201,202].

7.3. Potential Breakthroughs and Future Directions

Despite the existing research challenges, the future of biosensor technology in healthcare is marked by promising avenues for innovation. Advances in nanotechnology, materials science, and bioinformatics offer unprecedented opportunities for revolutionising biosensor design and functionality [203]. Nanomaterials with unique properties can enhance sensitivity and specificity, while multiplexed biosensor arrays enable simultaneous detection of multiple analytes, paving the way for comprehensive disease profiling [204]. Integrating artificial intelligence (AI) and machine learning algorithms holds immense potential for enhancing data analysis, pattern recognition, and predictive modelling, facilitating early disease detection and personalised treatment strategies. Moreover, the emergence of point-of-care molecular diagnostics, coupled with microfluidics and lab-on-a-chip technologies, promises to democratise healthcare delivery by enabling rapid, cost-effective, and accessible diagnostics, particularly in resource-limited settings [205].

8. Conclusions

In conclusion, biosensor technologies are revolutionising healthcare diagnostics by enabling precise and rapid detection of disease biomarkers and facilitating effective therapeutic drug monitoring. This review has explored the diverse types of biosensors, their mechanisms, and their specific applications in healthcare. The integration of nanotechnology, wearable devices, and miniaturisation trends has demonstrated significant potential to transform patient care. Despite sensitivity, specificity, reproducibility, and data security challenges, strategic solutions and ongoing innovations are paving the way for enhanced global healthcare outcomes. By fostering interdisciplinary collaboration and driving technological advancement, biosensors are poised to play an increasingly pivotal role in modern medicine.

Author Contributions

Conceptualization: M.H., M.A.A., A.S.D. and M.M.K.; Methodology: M.H., M.A.A. and I.M.E.; Validation: M.H., S.S.A.M. and M.M.; Formal Analysis: A.S.D., I.M.E. and M.M.; Investigation: M.H. and M.A.A., S.S.A.M.; Data Curation: M.A.A., M.M. and D.B.D.; Writing—Original Draft Preparation: I.M.E., M.M.K., M.H. and M.A.A.; Writing—Review and Editing: A.S.D., S.S.A.M. and A.S.A.; Visualization: M.M. and D.B.D.; Supervision: A.S.D., D.B.D. and A.S.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kim, E.R.; Joe, C.; Mitchell, R.J.; Gu, M.B. Biosensors for Healthcare: Current and Future Perspectives. Trends Biotechnol. 2023, 41, 374–395. [Google Scholar] [CrossRef] [PubMed]
  2. Sadani, K.; Nag, P.; Thian, X.Y.; Mukherji, S. Enzymatic Optical Biosensors for Healthcare Applications. Biosens. Bioelectron. X 2022, 12, 100278. [Google Scholar] [CrossRef]
  3. Zeng, X.; Peng, R.; Fan, Z.; Lin, Y. Self-Powered and Wearable Biosensors for Healthcare. Mater. Today Energy 2022, 23, 100900. [Google Scholar] [CrossRef]
  4. Mohankumar, P.; Ajayan, J.; Yasodharan, R.; Devendran, P.; Sambasivam, R. A Review of Micromachined Sensors for Automotive Applications. Measurement 2019, 140, 305–322. [Google Scholar] [CrossRef]
  5. Huang, H.; Chen, P.-Y.; Hung, C.-H.; Gharpurey, R.; Akinwande, D. A Zero Power Harmonic Transponder Sensor for Ubiquitous Wireless ΜL Liquid-Volume Monitoring. Sci. Rep. 2016, 6, 18795. [Google Scholar] [CrossRef]
  6. Ponmozhi, J.; Frias, C.; Marques, T.; Frazão, O. Smart Sensors/Actuators for Biomedical Applications. Measurement 2012, 45, 1675–1688. [Google Scholar] [CrossRef]
  7. Ng, C.L.; Reaz, M.B.I. Evolution of a Capacitive Electromyography Contactless Biosensor: Design and Modelling Techniques. Measurement 2019, 145, 460–471. [Google Scholar] [CrossRef]
  8. Karunakaran, R.; Keskin, M. Biosensors: Components, Mechanisms, and Applications. In Analytical Techniques in Biosciences: From Basics to Applications; Egbuna, C., Patrick-Iwuanyanwu, K.C., Shah, M.A., Ifemeje, J.C., Rasul, A., Eds.; Academic Press: Cambridge, MA, USA, 2022; pp. 179–190. ISBN 9780128226544. [Google Scholar]
  9. Qin, J.; Wang, W.; Gao, L.; Yao, S.Q. Emerging Biosensing and Transducing Techniques for Potential Applications in Point-of-Care Diagnostics. Chem. Sci. 2022, 13, 2857–2876. [Google Scholar] [CrossRef]
  10. Polat, E.O.; Cetin, M.M.; Tabak, A.F.; Bilget Güven, E.; Uysal, B.Ö.; Arsan, T.; Kabbani, A.; Hamed, H.; Gül, S.B. Transducer Technologies for Biosensors and Their Wearable Applications. Biosensors 2022, 12, 385. [Google Scholar] [CrossRef]
  11. Ramesh, M.; Janani, R.; Deepa, C.; Rajeshkumar, L. Nanotechnology-Enabled Biosensors: A Review of Fundamentals, Design Principles, Materials, and Applications. Biosensors 2022, 13, 40. [Google Scholar] [CrossRef]
  12. UPDIKE, S.J.; HICKS, G.P. The Enzyme Electrode. Nature 1967, 214, 986–988. [Google Scholar] [CrossRef] [PubMed]
  13. Kröger, S.; Turner, A.P.F.; Mosbach, K.; Haupt, K. Imprinted Polymer-Based Sensor System for Herbicides Using Differential-Pulse Voltammetry on Screen-Printed Electrodes. Anal. Chem. 1999, 71, 3698–3702. [Google Scholar] [CrossRef] [PubMed]
  14. Modak, N.; Friebe, V.M. Amperometric Biosensors: Harnessing Photosynthetic Reaction Centers for Herbicide Detection. Curr. Opin. Electrochem. 2023, 42, 101414. [Google Scholar] [CrossRef]
  15. Rachkov, A.; McNiven, S.; El’skaya, A.; Yano, K.; Karube, I. Fluorescence Detection of β-Estradiol Using a Molecularly Imprinted Polymer. Anal. Chim. Acta 2000, 405, 23–29. [Google Scholar] [CrossRef]
  16. Zahraee, S.S.; Alvandi, N.; Ghamari, M.; Esfandiari, N. An Ultra-Sensitive Nano Biosensor for 17β-Estradiol Detection Using Carbon Dots. Nano-Struct. Nano-Objects 2023, 34, 100951. [Google Scholar] [CrossRef]
  17. Garg, R.; Parwani, K.; Bist, R. Development of optical biosensors for diagnosis of amyloid-β peptide and glial fibrillary acidic protein. J. Anal. Comput. 2023, 17, 322–333. [Google Scholar] [CrossRef]
  18. Hong, F.; Xiao, R.; Li, L.; Cai, Z.; Ren, L.; Li, N.; Zhang, F.; Xu, X.; Chen, Y. Win-Win Cooperation: Magnetic Relaxation Switching Biosensor for Chloramphenicol Detection Based on the Pipet Platform and Antigen-Driven Enzyme-Free Dual Signal Amplification Strategy. Sens. Actuators B Chem. 2023, 385, 133686. [Google Scholar] [CrossRef]
  19. McNiven, S.; Kato, M.; Levi, R.; Yano, K.; Karube, I. Chloramphenicol Sensor Based on an in Situ Imprinted Polymer. Anal. Chim. Acta 1998, 365, 69–74. [Google Scholar] [CrossRef]
  20. Ngoepe, M.; Choonara, Y.E.; Tyagi, C.; Tomar, L.K.; Du Toit, L.C.; Kumar, P.; Ndesendo, V.M.K.; Pillay, V. Integration of Biosensors and Drug Delivery Technologies for Early Detection and Chronic Management of Illness. Sensors 2013, 13, 7680–7713. [Google Scholar] [CrossRef]
  21. Metkar, S.K.; Girigoswami, K. Diagnostic Biosensors in Medicine—A Review. Biocatal. Agric. Biotechnol. 2019, 17, 271–283. [Google Scholar] [CrossRef]
  22. Shoaib, A.; Darraj, A.; Khan, M.E.; Azmi, L.; Alalwan, A.; Alamri, O.; Tabish, M.; Khan, A.U. A Nanotechnology-Based Approach to Biosensor Application in Current Diabetes Management Practices. Nanomaterials 2023, 13, 867. [Google Scholar] [CrossRef]
  23. Pandey, R.; Lu, Y.; McConnell, E.M.; Osman, E.; Scott, A.; Gu, J.; Hoare, T.; Soleymani, L.; Li, Y. Electrochemical DNAzyme-Based Biosensors for Disease Diagnosis. Biosens. Bioelectron. 2023, 224, 114983. [Google Scholar] [CrossRef] [PubMed]
  24. Zamiri, G.; Babadi, A.A.; Chaudhary, V.; Numan, A.; Khalid, M.; Walvekar, R.; Khosla, A. Infection Management of Virus-Diagnosing Biosensors Based on Mxenes: An Overview. J. Electrochem. Soc. 2023, 170, 37501. [Google Scholar] [CrossRef]
  25. Bhatia, D.; Paul, S.; Acharjee, T.; Ramachairy, S.S. Biosensors and Their Widespread Impact on Human Health. Sens. Int. 2024, 5, 100257. [Google Scholar] [CrossRef]
  26. Sarkar, P. Revolutionizing Drug Transport: Unleashing Futuristic Biosensors with Arduino Programming. Eur. Chem. Bull. 2023, 12, 2126–2139. [Google Scholar]
  27. Lu, T.; Ji, S.; Jin, W.; Yang, Q.; Luo, Q.; Ren, T.-L. Biocompatible and Long-Term Monitoring Strategies of Wearable, Ingestible and Implantable Biosensors: Reform the Next Generation Healthcare. Sensors 2023, 23, 2991. [Google Scholar] [CrossRef] [PubMed]
  28. Zhang, S. Current Development on Wearable Biosensors towards Biomedical Applications. Front. Bioeng. Biotechnol. 2023, 11, 1264337. [Google Scholar] [CrossRef]
  29. Sengupta, J.; Hussain, C.M. CNT and Graphene-Based Transistor Biosensors for Cancer Detection: A Review. Biomolecules 2023, 13, 1024. [Google Scholar] [CrossRef]
  30. Zhao, V.X.T.; Wong, T.I.; Zheng, X.T.; Tan, Y.N.; Zhou, X. Colorimetric Biosensors for Point-of-Care Virus Detections. Mater. Sci. Energy Technol. 2020, 3, 237–249. [Google Scholar] [CrossRef]
  31. Parihar, A.; Ranjan, P.; Sanghi, S.K.; Srivastava, A.K.; Khan, R. Point-of-Care Biosensor-Based Diagnosis of COVID-19 Holds Promise to Combat Current and Future Pandemics. ACS Appl. Bio Mater. 2020, 3, 7326–7343. [Google Scholar] [CrossRef]
  32. Kalasin, S.; Surareungchai, W. Challenges of Emerging Wearable Sensors for Remote Monitoring toward Telemedicine Healthcare. Anal. Chem. 2023, 95, 1773–1784. [Google Scholar] [CrossRef]
  33. Polshettiwar, S.A.; Deshmukh, C.D.; Wani, M.S.; Baheti, A.M.; Bompilwar, E.; Choudhari, S.; Jambhekar, D.; Tagalpallewar, A. Recent Trends on Biosensors in Healthcare and Pharmaceuticals: An Overview. Int. J. Pharm. Investig. 2021, 11, 131–136. [Google Scholar] [CrossRef]
  34. Huang, X.; Zhu, Y.; Kianfar, E. Nano Biosensors: Properties, Applications and Electrochemical Techniques. J. Mater. Res. Technol. 2021, 12, 1649–1672. [Google Scholar] [CrossRef]
  35. Shamsuddin, S.H.; Gibson, T.D.; Tomlinson, D.C.; McPherson, M.J.; Jayne, D.G.; Millner, P.A. Reagentless Affimer-and Antibody-Based Impedimetric Biosensors for CEA-Detection Using a Novel Non-Conducting Polymer. Biosens. Bioelectron. 2021, 178, 113013. [Google Scholar] [CrossRef]
  36. Kucherenko, I.S.; Soldatkin, O.O.; Kucherenko, D.Y.; Soldatkina, O.V.; Dzyadevych, S.V. Advances in Nanomaterial Application in Enzyme-Based Electrochemical Biosensors: A Review. Nanoscale Adv. 2019, 1, 4560–4577. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, Q.; Wang, J.; Huang, Y.; Du, Y.; Zhang, Y.; Cui, Y.; Kong, D. Development of the DNA-Based Biosensors for High Performance in Detection of Molecular Biomarkers: More Rapid, Sensitive, and Universal. Biosens. Bioelectron. 2022, 197, 113739. [Google Scholar] [CrossRef] [PubMed]
  38. Gupta, N.; Renugopalakrishnan, V.; Liepmann, D.; Paulmurugan, R.; Malhotra, B.D. Cell-Based Biosensors: Recent Trends, Challenges and Future Perspectives. Biosens. Bioelectron. 2019, 141, 111435. [Google Scholar] [CrossRef]
  39. Singh, A.; Sharma, A.; Ahmed, A.; Sundramoorthy, A.K.; Furukawa, H.; Arya, S.; Khosla, A. Recent Advances in Electrochemical Biosensors: Applications, Challenges, and Future Scope. Biosensors 2021, 11, 336. [Google Scholar] [CrossRef]
  40. Cesewski, E.; Johnson, B.N. Electrochemical Biosensors for Pathogen Detection. Biosens. Bioelectron. 2020, 159, 112214. [Google Scholar] [CrossRef]
  41. Zeng, H.; Xie, Y.; Liu, T.; Chu, Z.; Dempsey, E.; Jin, W. Conductive Polymer Nanocomposites: Recent Advances in the Construction of Electrochemical Biosensors. Sens. Diagn. 2024, 3, 165–180. [Google Scholar] [CrossRef]
  42. Taha, B.A.; Al-Jubouri, Q.; Chahal, S.; Al Mashhadany, Y.; Rustagi, S.; Chaudhary, V.; Arsad, N. State-of-the-Art Telemodule-Enabled Intelligent Optical Nano-Biosensors for Proficient SARS-CoV-2 Monitoring. Microchem. J. 2024, 197, 109774. [Google Scholar] [CrossRef]
  43. Naresh, V.; Lee, N. A Review on Biosensors and Recent Development of Nanostructured Materials-Enabled Biosensors. Sensors 2021, 21, 1109. [Google Scholar] [CrossRef] [PubMed]
  44. Acar, T.; Nur Çimen, K.; Özalp, E.; Ilıca, Ö.; Ahlatcıoğlu Özerol, E. Recent Advances in Biosensors Based on Conducting Polymers for Biomedical Applications. ChemistrySelect 2023, 8, e202300819. [Google Scholar] [CrossRef]
  45. Huang, S.; Gao, Y.; Hu, Y.; Shen, F.; Jin, Z.; Cho, Y. Recent Development of Piezoelectric Biosensors for Physiological Signal Detection and Machine Learning Assisted Cardiovascular Disease Diagnosis. RSC Adv. 2023, 13, 29174–29194. [Google Scholar] [CrossRef] [PubMed]
  46. Zhou, C.X.; Huang, L.Q.; Li, S.F.Y. Microgravimetric DNA sensor based on quartz crystal microbalance: Comparison of oligonucleotide immobilization methods and the application in genetic diagnosis. Biosens. Bioelectron. 2001, 16, 85–95. [Google Scholar] [CrossRef] [PubMed]
  47. Alawajji, R.A.; Alsudani, Z.A.N.; Biris, A.S.; Kannarpady, G.K. Biosensor Design for the Detection of Circulating Tumor Cells Using the Quartz Crystal Resonator Technique. Biosensors 2023, 13, 433. [Google Scholar] [CrossRef] [PubMed]
  48. Khan, H.; Shah, M.R.; Barek, J.; Malik, M.I. Cancer Biomarkers and Their Biosensors: A Comprehensive Review. TrAC Trends Anal. Chem. 2023, 158, 116813. [Google Scholar] [CrossRef]
  49. Singh, A.K.; Mittal, S.; Das, M.; Saharia, A.; Tiwari, M. Optical Biosensors: A Decade in Review. Alexandria Eng. J. 2023, 67, 673–691. [Google Scholar] [CrossRef]
  50. Liu, Y.; Shen, H.; Yang, X.; Kang, S.; Cai, L.; Tian, T.; Su, R.; Yang, C.; Zhu, Z. Recent Progress in Microfluidic Biosensors with Different Driving Forces. TrAC Trends Anal. Chem. 2023, 158, 116894. [Google Scholar] [CrossRef]
  51. Yang, L.; Bai, R.; Xie, B.; Zhuang, N.; Lv, Z.; Chen, M.; Dong, W.; Zhou, J.; Jiang, M. A Biosensor Based on Oriented Immobilization of an Engineered L-Glutamate Oxidase on a Screen-Printed Microchip for Detection of l-Glutamate in Fermentation Processes. Food Chem. 2023, 405, 134792. [Google Scholar] [CrossRef]
  52. Liu, G.; Lv, Z.; Batool, S.; Li, M.; Zhao, P.; Guo, L.; Wang, Y.; Zhou, Y.; Han, S. Biocompatible Material-Based Flexible Biosensors: From Materials Design to Wearable/Implantable Devices and Integrated Sensing Systems. Small 2023, 19, 2207879. [Google Scholar] [CrossRef] [PubMed]
  53. Shanbhag, M.M.; Manasa, G.; Mascarenhas, R.J.; Mondal, K.; Shetti, N.P. Fundamentals of Bio-Electrochemical Sensing. Chem. Eng. J. Adv. 2023, 16, 100516. [Google Scholar] [CrossRef]
  54. Gao, Z.F.; Zhu, H.; Li, Y.; Yang, X.; Ren, X.; Wu, D.; Ma, H.; Wei, Q.; Xia, F.; Ju, H. Revolutionizing Biosensing with Superwettability: Designs, Mechanisms, and Applications. Nano Today 2023, 53, 102008. [Google Scholar] [CrossRef]
  55. Ventura-Aguilar, R.I.; Bautista-Baños, S.; Mendoza-Acevedo, S.; Bosquez-Molina, E. Nanomaterials for Designing Biosensors to Detect Fungi and Bacteria Related to Food Safety of Agricultural Products. Postharvest Biol. Technol. 2023, 195, 112116. [Google Scholar] [CrossRef]
  56. Wei, L.-N.; Luo, L.; Wang, B.-Z.; Lei, H.-T.; Guan, T.; Shen, Y.-D.; Wang, H.; Xu, Z.-L. Biosensors for Detection of Paralytic Shellfish Toxins: Recognition Elements and Transduction Technologies. Trends Food Sci. Technol. 2023, 133, 205–218. [Google Scholar] [CrossRef]
  57. Xiang, X.; Song, M.; Xu, X.; Lu, J.; Chen, Y.; Chen, S.; He, Y.; Shang, Y. Microfluidic Biosensor Integrated with Signal Transduction and Enhancement Mechanism for Ultrasensitive Noncompetitive Assay of Multiple Mycotoxins. Anal. Chem. 2023, 95, 7993–8001. [Google Scholar] [CrossRef] [PubMed]
  58. Liu, Y.; Yang, L.; Cui, Y. Transdermal Amperometric Biosensors for Continuous Glucose Monitoring in Diabetes. Talanta 2023, 253, 124033. [Google Scholar] [CrossRef]
  59. Johnston, L.; Wang, G.; Hu, K.; Qian, C.; Liu, G. Advances in Biosensors for Continuous Glucose Monitoring towards Wearables. Front. Bioeng. Biotechnol. 2021, 9, 733810. [Google Scholar] [CrossRef] [PubMed]
  60. Lee, I.; Probst, D.; Klonoff, D.; Sode, K. Continuous Glucose Monitoring Systems-Current Status and Future Perspectives of the Flagship Technologies in Biosensor Research. Biosens. Bioelectron. 2021, 181, 113054. [Google Scholar] [CrossRef] [PubMed]
  61. Lichtenberg, J.Y.; Ling, Y.; Kim, S. Non-Specific Adsorption Reduction Methods in Biosensing. Sensors 2019, 19, 2488. [Google Scholar] [CrossRef]
  62. Lv, Y.; Zhao, M.; Fan, J.; Wu, R.; Xu, Y.; Li, J.; Li, N.; Shen, H.; Guo, F.; Li, L.S. Self-Assembly Strategy to Reduce Non-Specific Adsorption for the Development of High Sensitivity Quantitative Immunoassay. Anal. Chim. Acta 2022, 1229, 340367. [Google Scholar] [CrossRef] [PubMed]
  63. Karrat, A.; Amine, A. Innovative Approaches to Suppress Non-Specific Adsorption in Molecularly Imprinted Polymers for Sensing Applications. Biosens. Bioelectron. 2024, 250, 116053. [Google Scholar] [CrossRef] [PubMed]
  64. Daurai, B.; Ramchiary, S.S.; Gogoi, M. Enzymatic Biosensors for Healthcare Applications. In Enzyme-Based Biosensors: Recent Advances and Applications in Healthcare; Springer: Berlin/Heidelberg, Germany, 2024; pp. 1–29. [Google Scholar]
  65. Bi, R.; Ma, X.; Miao, K.; Ma, P.; Wang, Q. Enzymatic Biosensor Based on Dendritic Gold Nanostructure and Enzyme Precipitation Coating for Glucose Sensing and Detection. Enzyme Microb. Technol. 2023, 162, 110132. [Google Scholar] [CrossRef] [PubMed]
  66. Khalife, M.; Stankovic, D.; Stankovic, V.; Danicka, J.; Rizzotto, F.; Costache, V.; Schwok, A.S.; Gaudu, P.; Vidic, J. Electrochemical Biosensor Based on NAD(P)H-Dependent Quinone Reductase for Rapid and Efficient Detection of Vitamin K3. Food Chem. 2024, 433, 137316. [Google Scholar] [CrossRef]
  67. Pittman, T.W.; Zhang, X.; Punyadeera, C.; Henry, C.S. Electrochemical Immunosensor for the Quantification of Galectin-3 in Saliva. Sensors Actuators B Chem. 2024, 400, 134811. [Google Scholar] [CrossRef]
  68. Yuksel, M.; Luo, W.; McCloy, B.; Mills, J.; Kayaharman, M.; Yeow, J.T.W. A Precise and Rapid Early Pregnancy Test: Development of a Novel and Fully Automated Electrochemical Point-of-Care Biosensor for Human Urine Samples. Talanta 2023, 254, 124156. [Google Scholar] [CrossRef] [PubMed]
  69. Chaturvedi, M.; Patel, M.; Mondal, D.P.; Srivastava, A.K.; Dwivedi, N.; Dhand, C. Bio-Inspired Graphene Nanocomposite Enabled Electrochemical Immunosensor for Detection and Quantification of NS1 Protein of Dengue Virus. Electrochim. Acta 2024, 475, 143630. [Google Scholar] [CrossRef]
  70. Ghasemi, L.; Jahani, S.; Ghazizadeh, M.; Foroughi, M.M. A Novel and Ultrasensitive Electrochemical DNA Biosensor for Pralatrexate Detection. Anal. Methods 2023, 15, 631–638. [Google Scholar] [CrossRef] [PubMed]
  71. Lomae, A.; Preechakasedkit, P.; Hanpanich, O.; Ozer, T.; Henry, C.S.; Maruyama, A.; Pasomsub, E.; Phuphuakrat, A.; Rengpipat, S.; Vilaivan, T. Label Free Electrochemical DNA Biosensor for COVID-19 Diagnosis. Talanta 2023, 253, 123992. [Google Scholar] [CrossRef]
  72. Ali, M.R.; Bacchu, M.S.; Das, S.; Akter, S.; Rahman, M.M.; Saad Aly, M.A.; Khan, M.Z.H. Label Free Flexible Electrochemical DNA Biosensor for Selective Detection of Shigella Flexneri in Real Food Samples. Talanta 2023, 253, 123909. [Google Scholar] [CrossRef]
  73. Guo, X.; Wang, M. Recent Progress in Optical and Electrochemical Aptasensor Technologies for Detection of Aflatoxin B1. Crit. Rev. Food Sci. Nutr. 2023, 1–19. [Google Scholar] [CrossRef] [PubMed]
  74. Novakovic, Z.; Khalife, M.; Costache, V.; Camacho, M.J.; Cardoso, S.; Martins, V.; Gadjanski, I.; Radovic, M.; Vidic, J. Rapid Detection and Identification of Vancomycin-Sensitive Bacteria Using an Electrochemical Apta-Sensor. ACS Omega 2024, 9, 2841–2849. [Google Scholar] [CrossRef]
  75. Wu, J.; He, B.; Wang, Y.; Zhao, R.; Zhang, Y.; Bai, C.; Wei, M.; Jin, H.; Ren, W.; Suo, Z.; et al. ZIF-8 Labelled a New Electrochemical Aptasensor Based on PEI-PrGO/AuNWs for DON Detection. Talanta 2024, 267, 125257. [Google Scholar] [CrossRef]
  76. Liu, C.; Guan, C.; Li, Y.; Li, Z.; Wang, Y.; Han, G. Advances in Electrochemical Biosensors for the Detection of Common Oral Diseases. Crit. Rev. Anal. Chem. 2024, 1–21. [Google Scholar] [CrossRef]
  77. Zhang, Y.; Wang, J.; Liu, M.; Ni, Y.; Yue, Y.; He, D.; Liu, R. Magnetically Induced Self-Assembly Electrochemical Biosensor with Ultra-Low Detection Limit and Extended Measuring Range for Sensitive Detection of HER2 Protein. Bioelectrochemistry 2024, 155, 108592. [Google Scholar] [CrossRef]
  78. Hemdan, M.; Ali, M.A.; Amin, A.S. Eco-Friendly Optical Sensor for Precise Detection of Gold Ions in Diverse Matrices through the Integration of β-2-Hydroxybenzyl-3-Methoxy-2-Hydroxyazastyrene in a PVC Membrane. Anal. Bioanal. Chem. 2024, 416, 3835–3846. [Google Scholar] [CrossRef] [PubMed]
  79. Alshehri, R.F.; Hemdan, M.; Babalghith, A.O.; Amin, A.S.; Darwish, E.R. An Innovative Approach in Titanium Determination Based on Incorporating 2-Amino-4-((4-Nitrophenyl) Diazenyl) Pyridine-3-Ol in a PVC Membrane. RSC Adv. 2024, 14, 712–724. [Google Scholar] [CrossRef]
  80. Uniyal, A.; Srivastava, G.; Pal, A.; Taya, S.; Muduli, A. Recent Advances in Optical Biosensors for Sensing Applications: A Review. Plasmonics 2023, 18, 735–750. [Google Scholar] [CrossRef]
  81. Yildizhan, Y.; Driessens, K.; Tsao, H.S.; Boiy, R.; Thomas, D.; Geukens, N.; Hendrix, A.; Lammertyn, J.; Spasic, D. Detection of Breast Cancer-Specific Extracellular Vesicles with Fiber-Optic SPR Biosensor. Int. J. Mol. Sci. 2023, 24, 3764. [Google Scholar] [CrossRef] [PubMed]
  82. Xu, Q.; Gao, X.; Zhao, S.; Liu, Y.; Zhang, D.; Zhou, K.; Khanbareh, H.; Chen, W.; Zhang, Y.; Bowen, C. Construction of Bio-piezoelectric Platforms: From Structures and Synthesis to Applications. Adv. Mater. 2021, 33, 2008452. [Google Scholar] [CrossRef]
  83. Chen, Y.; Shi, H. Rapid and Label-Free Analysis of Antigen–Antibody Dynamic Binding of Tumor Markers Using Piezoelectric Quartz Crystal Biosensor. Biosensors 2023, 13, 917. [Google Scholar] [CrossRef]
  84. Ergezer, E.E.; Bodur, O.C.; Özkan, E.H.; Sarı, N.; Arslan, F. Development of a Novel Biosensor for the Ultra-Sensitive Detection of the Endocrine Disruptor Bisphenol A and Detection of Bisphenol A Leakage from Some Plastic Containers. J. Iran. Chem. Soc. 2024, 21, 767–779. [Google Scholar] [CrossRef]
  85. Zhang, B.L.; Yang, Y.; Zhao, Z.Q.; Guo, X.D. A Gold Nanoparticles Deposited Polymer Microneedle Enzymatic Biosensor for Glucose Sensing. Electrochim. Acta 2020, 358, 136917. [Google Scholar] [CrossRef]
  86. Aymard, C.; Kanso, H.; Serrano, M.J.; Pagán, R.; Noguer, T.; Istamboulie, G. Development of a New Dual Electrochemical Immunosensor for a Rapid and Sensitive Detection of Enrofloxacin in Meat Samples. Food Chem. 2022, 370, 131016. [Google Scholar] [CrossRef] [PubMed]
  87. Fabiani, L.; Fiore, L.; Fillo, S.; D’Amore, N.; De Santis, R.; Lista, F.; Arduini, F. Smartphone-Assisted Paper-Based Electrochemical Immunosensor for SARS-CoV-2 Detection in Saliva. Bioelectrochemistry 2024, 156, 108619. [Google Scholar] [CrossRef]
  88. Pareek, S.; Jain, U.; Bharadwaj, M.; Saxena, K.; Roy, S.; Chauhan, N. An Ultrasensitive Electrochemical DNA Biosensor for Monitoring Human Papillomavirus-16 (HPV-16) Using Graphene Oxide/Ag/Au Nano-Biohybrids. Anal. Biochem. 2023, 663, 115015. [Google Scholar] [CrossRef] [PubMed]
  89. Li, D.; Xia, H.; Sun, Y.; Liu, W.; Liu, W.; Yu, J.; Jing, G.; Zhang, J.; Li, W. Colorimetric Aptasensor for the Sensitive Detection of Ochratoxin A Based on a Triple Cascade Amplification Strategy. Anal. Chim. Acta 2023, 1237, 340616. [Google Scholar] [CrossRef]
  90. Zhang, Y.; Li, N.; Xu, Y.; Liu, X.; Ma, Y.; Huang, Z.; Luo, H.; Hou, C.; Huo, D. A Novel Electrochemical Biosensor Based on AMNFs@ ZIF-67 Nano Composite Material for Ultrasensitive Detection of HER2. Bioelectrochemistry 2023, 150, 108362. [Google Scholar] [CrossRef] [PubMed]
  91. Batra, B.; Yadav, S.; Kalra, V.; Sharma, M.; Rana, J.S. An Electrochemical Biosensor for the Determination of Folic Acid in Pregnant Women Based on DHFR/c-MWCNTs/TiO2NPs Modified Gold Electrode. Sens. Int. 2023, 4, 100235. [Google Scholar] [CrossRef]
  92. Khani, S.; Hayati, M. Optical Biosensors Using Plasmonic and Photonic Crystal Band-Gap Structures for the Detection of Basal Cell Cancer. Sci. Rep. 2022, 12, 5246. [Google Scholar] [CrossRef]
  93. Esmaeili, Y.; Khavani, M.; Bigham, A.; Sanati, A.; Bidram, E.; Shariati, L.; Zarrabi, A.; Jolfaie, N.A.; Rafienia, M. Mesoporous Silica@chitosan@gold Nanoparticles as “on/off” Optical Biosensor and PH-Sensitive Theranostic Platform against Cancer. Int. J. Biol. Macromol. 2022, 202, 241–255. [Google Scholar] [CrossRef] [PubMed]
  94. D’Ambrogio, G.; Zahhaf, O.; Le, M.-Q.; Bordet, M.; Lermusiaux, P.; Della Schiava, N.; Liang, R.; Cottinet, P.-J.; Capsal, J.-F. Piezoelectric Biosensor for Smart Cardiovascular Grafts Based on NaNbO3 Fibers/PDMS Structured Composite. Mater. Des. 2022, 223, 111195. [Google Scholar] [CrossRef]
  95. Mahmoud, A.E.D.; Fawzy, M. Nanosensors and Nanobiosensors for Monitoring the Environmental Pollutants. Waste Recycl. Technol. Nanomater. Manuf. 2021, 229–246. [Google Scholar] [CrossRef]
  96. Tang, X.; Zhu, Y.; Guan, W.; Zhou, W.; Wei, P. Advances in Nanosensors for Cardiovascular Disease Detection. Life Sci. 2022, 305, 120733. [Google Scholar] [CrossRef] [PubMed]
  97. Vu, Q.K.; Tran, Q.H.; Vu, N.P.; Anh, T.-L.; Le Dang, T.T.; Matteo, T.; Nguyen, T.H.H. A Label-Free Electrochemical Biosensor Based on Screen-Printed Electrodes Modified with Gold Nanoparticles for Quick Detection of Bacterial Pathogens. Mater. Today Commun. 2021, 26, 101726. [Google Scholar] [CrossRef]
  98. Singh, P.; Yadava, R.D.S. Nanosensors for Health Care. In Nanosensors for Smart Cities; Elsevier: Amsterdam, The Netherlands, 2020; pp. 433–450. [Google Scholar]
  99. Rabbani, G.; Khan, M.E.; Khan, A.U.; Ali, S.K.; Zamzami, M.A.; Ahmad, A.; Bashiri, A.H.; Zakri, W. Label-Free and Ultrasensitive Electrochemical Transferrin Detection Biosensor Based on a Glassy Carbon Electrode and Gold Nanoparticles. Int. J. Biol. Macromol. 2024, 256, 128312. [Google Scholar] [CrossRef] [PubMed]
  100. Sarkar, T.; Brahma, D.; Gupta, A.N. L-Cysteine Capped Fe3O4 Nanoparticles-Functionalized Conductive Carbon Yarn as a Flexible Electrochemical Biosensor for Detection of L-Phenylalanine. Microchem. J. 2024, 196, 109541. [Google Scholar] [CrossRef]
  101. Zhu, M.; Liu, Y.; Wang, M.; Liu, T.; Chu, Z.; Jin, W. Facile Construction of Nanocubic Mn3[Fe(CN)6]2@Pt Based Electrochemical DNA Sensors for Ultrafast Precise Determination of SARS-CoV-2. Bioelectrochemistry 2024, 156, 108598. [Google Scholar] [CrossRef] [PubMed]
  102. Yao, J. A Multiple Signal Amplification Photoelectrochemical Biosensor Based on Biotin-Avidin System for Kanamycin Sensing in Fish and Milk via Synergism of g-C3N4 and Ru@SiO2. Anal. Chim. Acta 2024, 1288, 342141. [Google Scholar] [CrossRef]
  103. Ertaş, T.; Dinç, B.; Üstünsoy, R.; Eraslan, H.; Ergenç, A.F.; Bektaş, M. Novel Electrochemical Biosensor for Escherichia Coli Using Gold-Coated Tungsten Wires and Antibody Functionalized Short Multiwalled Carbon Nanotubes. Instrum. Sci. Technol. 2024, 52, 109–124. [Google Scholar] [CrossRef]
  104. Ashikbayeva, Z.; Bekmurzayeva, A.; Myrkhiyeva, Z.; Assylbekova, N.; Atabaev, T.S.; Tosi, D. Green-Synthesized Gold Nanoparticle-Based Optical Fiber Ball Resonator Biosensor for Cancer Biomarker Detection. Opt. Laser Technol. 2023, 161, 109136. [Google Scholar] [CrossRef]
  105. Sadrabadi, E.A.; Benvidi, A.; Azimzadeh, M.; Asgharnejad, L.; Dezfuli, A.S.; Khashayar, P. Novel Electrochemical Biosensor for Breast Cancer Detection, Based on a Nanocomposite of Carbon Nanofiber, Metal–Organic Framework, and Magnetic Graphene Oxide. Bioelectrochemistry 2024, 155, 108558. [Google Scholar] [CrossRef] [PubMed]
  106. Cuapa-González, M.A.; Santos-López, G.; Orduña-Díaz, A.; Martínez-Gutiérrez, H.; Rojas-López, M. Detection of Zika Virus by the Development of a Colloidal Gold Nanoparticle-Based Immunosensor. Anal. Lett. 2024, 1–20. [Google Scholar] [CrossRef]
  107. Eshlaghi, S.N.; Syedmoradi, L.; Amini, A.; Omidfar, K. A Label-Free Electrochemical Aptasensor Based on Screen Printed Carbon Electrodes With Gold Nanoparticles-Polypyrrole Composite for Detection of Cardiac Troponin I. IEEE Sens. J. 2023, 23, 3439–3445. [Google Scholar] [CrossRef]
  108. Yang, S.; Li, C.; Yu, Y.; Zhan, H.; Zhai, J.; Liu, R.; Chen, W.; Zou, Y.; Xu, K. An Enzyme-Free and Label-Free Ratiometric Fluorescence Signal Amplification Biosensor Based on DNA-Silver Nanoclusters and Catalytic Hairpin Assembly for Tetracycline Detection. Sens. Actuators B Chem. 2024, 404, 135216. [Google Scholar] [CrossRef]
  109. Gao, X.; Zhang, H.; Liu, L.; Jia, M.; Li, X.; Li, J. Nano-Biosensor Based on Manganese Dioxide Nanosheets and Carbon Dots for Dual-Mode Determination of Staphylococcus Aureus. Food Chem. 2024, 432, 137144. [Google Scholar] [CrossRef] [PubMed]
  110. Mostufa, S.; Akib, T.B.; Rana, M.M.; Islam, M.R. Highly Sensitive TiO2/Au/Graphene Layer-Based Surface Plasmon Resonance Biosensor for Cancer Detection. Biosensors 2022, 12, 603. [Google Scholar] [CrossRef] [PubMed]
  111. Ye, S.; Feng, S.; Huang, L.; Bian, S. Recent Progress in Wearable Biosensors: From Healthcare Monitoring to Sports Analytics. Biosensors 2020, 10, 205. [Google Scholar] [CrossRef] [PubMed]
  112. Li, Y.; Peng, Z.; Li, J.; Wei, C.; Liu, S.; Hao, W.; Cheng, H.; Burton, C.; Wang, Y.; Huang, Y. Wearable MXene-Graphene Sensing of Influenza and SARS-CoV-2 Virus in Air and Breath: From Lab to Clinic. Adv. Mater. Technol. 2024, 9, 2201787. [Google Scholar] [CrossRef]
  113. Li, Y.; Luo, L.; Kong, Y.; George, S.; Li, Y.; Guo, X.; Li, X.; Yeatman, E.; Davenport, A.; Li, Y. A Point-of-Care Sensing Platform for Multiplexed Detection of Chronic Kidney Disease Biomarkers Using Molecularly Imprinted Polymers. Adv. Funct. Mater. 2024, 34, 2316865. [Google Scholar] [CrossRef]
  114. Kulkarni, M.B.; Ayachit, N.H.; Aminabhavi, T.M. A Short Review on Miniaturized Biosensors for the Detection of Nucleic Acid Biomarkers. Biosensors 2023, 13, 412. [Google Scholar] [CrossRef] [PubMed]
  115. Alzahrani, K.E.; Assaifan, A.K.; Al-Gawati, M.; Alswieleh, A.M.; Albrithen, H.; Alodhayb, A. Microelectromechanical System-based Biosensor for Label-free Detection of Human Cytomegalovirus. IET Nanobiotechnol. 2023, 17, 32–39. [Google Scholar] [CrossRef] [PubMed]
  116. Kulkarni, M.B.; Ayachit, N.H.; Aminabhavi, T.M. Biosensors and Microfluidic Biosensors: From Fabrication to Application. Biosensors 2022, 12, 543. [Google Scholar] [CrossRef] [PubMed]
  117. Kaur, B.; Kumar, S.; Kaushik, B.K. Recent Advancements in Optical Biosensors for Cancer Detection. Biosens. Bioelectron. 2022, 197, 113805. [Google Scholar] [CrossRef] [PubMed]
  118. Hasan, M.R.; Ahommed, M.S.; Daizy, M.; Bacchu, M.S.; Ali, M.R.; Al-Mamun, M.R.; Aly, M.A.S.; Khan, M.Z.H.; Hossain, S.I. Recent Development in Electrochemical Biosensors for Cancer Biomarkers Detection. Biosens. Bioelectron. X 2021, 8, 100075. [Google Scholar] [CrossRef]
  119. Wan, H.-H.; Zhu, H.; Chiang, C.-C.; Li, J.-S.; Ren, F.; Tsai, C.-T.; Liao, Y.-T.; Neal, D.; Esquivel-Upshaw, J.F.; Pearton, S.J. High Sensitivity Saliva-Based Biosensor in Detection of Breast Cancer Biomarkers: HER2 and CA15-3. J. Vac. Sci. Technol. B 2024, 42, 023202. [Google Scholar] [CrossRef] [PubMed]
  120. Menon, S.; Mathew, M.R.; Sam, S.; Keerthi, K.; Kumar, K.G. Recent Advances and Challenges in Electrochemical Biosensors for Emerging and Re-Emerging Infectious Diseases. J. Electroanal. Chem. 2020, 878, 114596. [Google Scholar] [CrossRef] [PubMed]
  121. Wang, C.; Liu, M.; Wang, Z.; Li, S.; Deng, Y.; He, N. Point-of-Care Diagnostics for Infectious Diseases: From Methods to Devices. Nano Today 2021, 37, 101092. [Google Scholar] [CrossRef]
  122. Khan, R.; Deshpande, A.S.; Proteasa, G.; Andreescu, S. Aptamer-Based Electrochemical Biosensor with S Protein Binding Affinity for COVID-19 Detection: Integrating Computational Design with Experimental Validation of S Protein Binding Affinity. Sens. Actuators B Chem. 2024, 399, 134775. [Google Scholar] [CrossRef]
  123. Pullano, S.A.; Greco, M.; Bianco, M.G.; Foti, D.; Brunetti, A.; Fiorillo, A.S. Glucose Biosensors in Clinical Practice: Principles, Limits and Perspectives of Currently Used Devices. Theranostics 2022, 12, 493. [Google Scholar] [CrossRef]
  124. Li, L.; Zhou, Y.; Sun, C.; Zhou, Z.; Zhang, J.; Xu, Y.; Xiao, X.; Deng, H.; Zhong, Y.; Li, G.; et al. Fully Integrated Wearable Microneedle Biosensing Platform for Wide-Range and Real-Time Continuous Glucose Monitoring. Acta Biomater. 2024, 175, 199–213. [Google Scholar] [CrossRef] [PubMed]
  125. Mobed, A.; Razavi, S.; Ahmadalipour, A.; Shakouri, S.K.; Koohkan, G. Biosensors in Parkinson’s Disease. Clin. Chim. Acta 2021, 518, 51–58. [Google Scholar] [CrossRef] [PubMed]
  126. Ameri, M.; Shabaninejad, Z.; Movahedpour, A.; Sahebkar, A.; Mohammadi, S.; Hosseindoost, S.; Ebrahimi, M.S.; Savardashtaki, A.; Karimipour, M.; Mirzaei, H. Biosensors for Detection of Tau Protein as an Alzheimer’s Disease Marker. Int. J. Biol. Macromol. 2020, 162, 1100–1108. [Google Scholar] [CrossRef] [PubMed]
  127. Altay, D.N.; Yagar, H.; Ozcan, H.M. A New ITO-Based Aβ42 Biosensor for Early Detection of Alzheimer’s Disease. Bioelectrochemistry 2023, 153, 108501. [Google Scholar] [CrossRef] [PubMed]
  128. Karaboğa, M.N.S.; Ünal, M.A.; Arı, F.; Sezgintürk, M.K.; Özkan, S.A. An Innovative Method for the Detection of Alpha Synuclein, a Potential Biomarker of Parkinson’s Disease: Quartz Tuning Fork-Based Mass Sensitive Immunosensor Design. Phys. Chem. Chem. Phys. 2024, 26, 5106–5114. [Google Scholar] [CrossRef] [PubMed]
  129. Tang, L.; Yang, J.; Wang, Y.; Deng, R. Recent Advances in Cardiovascular Disease Biosensors and Monitoring Technologies. ACS Sens. 2023, 8, 956–973. [Google Scholar] [CrossRef] [PubMed]
  130. Ouyang, M.; Tu, D.; Tong, L.; Sarwar, M.; Bhimaraj, A.; Li, C.; Cote, G.L.; Di Carlo, D. A Review of Biosensor Technologies for Blood Biomarkers toward Monitoring Cardiovascular Diseases at the Point-of-Care. Biosens. Bioelectron. 2021, 171, 112621. [Google Scholar] [CrossRef] [PubMed]
  131. Li, P.-R.; Kiran Boilla, S.; Wang, C.-H.; Lin, P.-C.; Kuo, C.-N.; Tsai, T.-H.; Lee, G.-B. A Self-Driven, Microfluidic, Integrated-Circuit Biosensing Chip for Detecting Four Cardiovascular Disease Biomarkers. Biosens. Bioelectron. 2024, 249, 115931. [Google Scholar] [CrossRef] [PubMed]
  132. Hui, H.; Gopinath, S.C.B.; Ismail, Z.H.; Chen, Y.; Pandian, K.; Velusamy, P. Cardiovascular Biomarker Troponin I Biosensor: Aptamer-gold-antibody Hybrid on a Metal Oxide Surface. Biotechnol. Appl. Biochem. 2023, 70, 581–591. [Google Scholar] [CrossRef]
  133. Imas, J.J.; Ruiz Zamarreno, C.; Zubiate, P.; Sanchez-Martín, L.; Campión, J.; Matías, I.R. Optical Biosensors for the Detection of Rheumatoid Arthritis (RA) Biomarkers: A Comprehensive Review. Sensors 2020, 20, 6289. [Google Scholar] [CrossRef]
  134. Mobed, A.; Dolati, S.; Shakouri, S.K.; Eftekharsadat, B.; Izadseresht, B. Recent Advances in Biosensors for Detection of Osteoarthritis and Rheumatoid Arthritis Biomarkers. Sens. Actuators A Phys. 2021, 331, 112975. [Google Scholar] [CrossRef]
  135. Golfinopoulou, R.; Kintzios, S. Biosensing for Autoimmune Chronic Disease—A Review. Chemosensors 2023, 11, 366. [Google Scholar] [CrossRef]
  136. Chen, H.-M.; Tsai, Y.-H.; Hsu, C.-Y.; Wang, Y.-Y.; Hsieh, C.-E.; Chen, J.-H.; Chang, Y.-S.; Lin, C.-Y. Peptide-Coated Bacteriorhodopsin-Based Photoelectric Biosensor for Detecting Rheumatoid Arthritis. Biosensors 2023, 13, 929. [Google Scholar] [CrossRef] [PubMed]
  137. Jupe, E.R.; Lushington, G.H.; Purushothaman, M.; Pautasso, F.; Armstrong, G.; Sorathia, A.; Crawley, J.; Nadipelli, V.R.; Rubin, B.; Newhardt, R.; et al. Tracking of Systemic Lupus Erythematosus (SLE) Longitudinally Using Biosensor and Patient-Reported Data: A Report on the Fully Decentralized Mobile Study to Measure and Predict Lupus Disease Activity Using Digital Signals—The OASIS Study. BioTech 2023, 12, 62. [Google Scholar] [CrossRef] [PubMed]
  138. Ribeiro, B.V.; Cordeiro, T.A.R.; E Freitas, G.R.O.; Ferreira, L.F.; Franco, D.L. Biosensors for the Detection of Respiratory Viruses: A Review. Talanta Open 2020, 2, 100007. [Google Scholar] [CrossRef] [PubMed]
  139. Dong, T.; Santos, S.; Yang, Z.; Yang, S.; Kirkhus, N.E. Sputum and Salivary Protein Biomarkers and Point-of-Care Biosensors for the Management of COPD. Analyst 2020, 145, 1583–1604. [Google Scholar] [CrossRef] [PubMed]
  140. Qian, J.; Zhao, L.; Huang, Y.; Zhao, C.; Liu, H.; Liu, X.; Cheng, Z.; Yu, F. A Microdroplet SERS-RCA Biosensor with Enhanced Specificity and Reproducibility for Profiling Dual MiRNAs in Idiopathic Pulmonary Fibrosis Diagnosis and Monitoring. Chem. Eng. J. 2024, 482, 149012. [Google Scholar] [CrossRef]
  141. Rahman, S.F.A.; Arshad, M.K.M.; Gopinath, S.C.B.; Fathil, M.F.M.; Sarry, F.; Ibau, C.; Elmazria, O.; Hage-Ali, S. Interdigitated Impedimetric-Based Maackia Amurensis Lectin Biosensor for Prostate Cancer Biomarker. Microchim. Acta 2024, 191, 118. [Google Scholar] [CrossRef] [PubMed]
  142. Sonklin, T.; Tongpeng, S.; Munthala, D.; Suksaweang, S.; Janphuang, P.; Bharti, A.; Mathur, A.; Avasthi, D.K.; Jiansirisomboon, S.; Pojprapai, S. Development of Gold Nanospikes-Modified Quartz Crystal Microbalance Biosensor for Prostate Specific Antigen Detection. Surf. Interfaces 2024, 45, 103877. [Google Scholar] [CrossRef]
  143. Cheng, Y.; Li, Y.; Ren, H.; Wen, B.; Liang, W.; Zhang, S.; Cong, B.; Jiang, M.; Hong, C. Electrochemiluminescent Biosensor Based on ECL-RET between Ru@TiO2-MXene and Pd@UiO-66-NH2 for the Detection of Carcinoembryonic Antigens. Sens. Actuators B Chem. 2024, 405, 135381. [Google Scholar] [CrossRef]
  144. Tang, S.; Xu, Q.; Liu, M.; Zhu, Y.; Zhang, G.; Tang, X. Highly Sensitive Electrochemical Immunosensor Based on SiO2 Nanospheres for Detection of EGFR as Colorectal Cancer Biomarker. Alexandria Eng. J. 2024, 89, 53–59. [Google Scholar] [CrossRef]
  145. Kumar, N.; Towers, D.; Myers, S.; Galvin, C.; Kireev, D.; Ellington, A.D.; Akinwande, D. Graphene Field Effect Biosensor for Concurrent and Specific Detection of SARS-CoV-2 and Influenza. ACS Nano 2023, 17, 18629–18640. [Google Scholar] [CrossRef] [PubMed]
  146. Wen, X.; Yang, X.; Ge, Z.; Ma, H.; Wang, R.; Tian, F.; Teng, P.; Gao, S.; Li, K.; Zhang, B.; et al. Self-Powered Optical Fiber Biosensor Integrated with Enzymes for Non-Invasive Glucose Sensing. Biosens. Bioelectron. 2024, 253, 116191. [Google Scholar] [CrossRef] [PubMed]
  147. Vasanth, S.; Amir, H.; Ponpandian, N.; Viswanathan, C. Non-Enzymatic Electrochemical Detection of Methylglyoxal in Saliva Using a Polyaniline/Nickel Oxide Nanohybrid Biosensor: A Noninvasive Approach for Diabetes Diagnosis. Biosens. Bioelectron. X 2024, 17, 100444. [Google Scholar] [CrossRef]
  148. Wang, X.; Huo, H.; Xu, C.; Lin, H.; Wang, Q.; Yang, J.; Vogel, F.; Wang, X.; Lin, Z.; Cao, L.; et al. A Sensitive Non-Enzymatic Dual-Conductive Biosensor for Continuous Glucose Monitoring. Anal. Chim. Acta 2023, 1279, 341845. [Google Scholar] [CrossRef] [PubMed]
  149. Zhang, Q.; Yin, B.; Huang, Y.; Gu, Y.; Yan, J.; Chen, J.; Li, C.; Zhang, Y.; Wong, S.H.D.; Yang, M. A Dual “Turn-on” Biosensor Based on AIE Effect and FRET for in Situ Detection of MiR-125b Biomarker in Early Alzheimer’s Disease. Biosens. Bioelectron. 2023, 230, 115270. [Google Scholar] [CrossRef] [PubMed]
  150. Li, T.; Xu, T.; Yao, Z.; Ding, Y.; Liu, G.; Shan, F. Highly Sensitive Biosensor Based on IGZO Thin-Film Transistors for Detection of Parkinson’s Disease. Appl. Phys. Lett. 2023, 122, 243701. [Google Scholar] [CrossRef]
  151. Chauhan, D.; Pooja; Nirbhaya, V.; Srivastava, C.M.; Chandra, R.; Kumar, S. Nanostructured Transition Metal Chalcogenide Embedded on Reduced Graphene Oxide Based Highly Efficient Biosensor for Cardiovascular Disease Detection. Microchem. J. 2020, 155, 104697. [Google Scholar] [CrossRef]
  152. Wen, H.-Y.; Chiang, C.-C.; Chen, R.-Y.; Ni, W.-Z.; Weng, Y.-Q.; Yeh, Y.-T.; Hsu, H.-C. Immunosensing for Early Detection of Rheumatoid Arthritis Biomarkers: Anti-Cyclic Citrullinated Peptide Antibodies Based on Tilted-Fiber Bragg Grating Biosensor. Bioengineering 2023, 10, 261. [Google Scholar] [CrossRef]
  153. Arévalo, B.; Serafín, V.; Sánchez-Paniagua, M.; Montero-Calle, A.; Barderas, R.; López-Ruíz, B.; Campuzano, S.; Yáñez-Sedeño, P.; Pingarrón, J.M. Fast and Sensitive Diagnosis of Autoimmune Disorders through Amperometric Biosensing of Serum Anti-DsDNA Autoantibodies. Biosens. Bioelectron. 2020, 160, 112233. [Google Scholar] [CrossRef]
  154. Liang, W.S.; Beaulieu-Jones, B. Emerging Therapeutic Drug Monitoring Technologies: Considerations and Opportunities in Precision Medicine. Front. Pharmacol. 2024, 15, 1348112. [Google Scholar] [CrossRef] [PubMed]
  155. Garzón, V.; Bustos, R.-H.G.; Pinacho, D. Personalized Medicine for Antibiotics: The Role of Nanobiosensors in Therapeutic Drug Monitoring. J. Pers. Med. 2020, 10, 147. [Google Scholar] [CrossRef] [PubMed]
  156. Mohankumar, P.; Ajayan, J.; Mohanraj, T.; Yasodharan, R. Recent Developments in Biosensors for Healthcare and Biomedical Applications: A Review. Measurement 2021, 167, 108293. [Google Scholar] [CrossRef]
  157. Song, L.; Chen, J.; Xu, B.B.; Huang, Y. Flexible Plasmonic Biosensors for Healthcare Monitoring: Progress and Prospects. ACS Nano 2021, 15, 18822–18847. [Google Scholar] [CrossRef] [PubMed]
  158. Garzón, V.; Pinacho, D.G.; Bustos, R.H.; Garzón, G.; Bustamante, S. Optical Biosensors for Therapeutic Drug Monitoring. Biosensors 2019, 9, 132. [Google Scholar] [CrossRef] [PubMed]
  159. Wang, Q.; Li, S.; Chen, J.; Yang, L.; Qiu, Y.; Du, Q.; Wang, C.; Teng, M.; Wang, T.; Dong, Y. A Novel Strategy for Therapeutic Drug Monitoring: Application of Biosensors to Quantify Antimicrobials in Biological Matrices. J. Antimicrob. Chemother. 2023, 78, 2612–2629. [Google Scholar] [CrossRef] [PubMed]
  160. Habet, S. Narrow Therapeutic Index Drugs: Clinical Pharmacology Perspective. J. Pharm. Pharmacol. 2021, 73, 1285–1291. [Google Scholar] [CrossRef]
  161. Maiti, R.; Patel, B.; Patel, N.; Patel, M.; Patel, A.; Dhanesha, N. Antibody Drug Conjugates as Targeted Cancer Therapy: Past Development, Present Challenges and Future Opportunities. Arch. Pharm. Res. 2023, 46, 361–388. [Google Scholar] [CrossRef] [PubMed]
  162. Ahmed, M.O.; Baptiste, K.E. Vancomycin-Resistant Enterococci: A Review of Antimicrobial Resistance Mechanisms and Perspectives of Human and Animal Health. Microb. Drug Resist. 2018, 24, 590–606. [Google Scholar] [CrossRef]
  163. Stogios, P.J.; Savchenko, A. Molecular Mechanisms of Vancomycin Resistance. Protein Sci. 2020, 29, 654–669. [Google Scholar] [CrossRef]
  164. Martin, J.H.; Norris, R.; Barras, M.; Roberts, J.; Morris, R.; Doogue, M.; Jones, G.R.D. Therapeutic Monitoring of Vancomycin in Adult Patients: A Consensus Review of the American Society of Health-System Pharmacists, the Infectious Diseases Society of America, and the Society Of Infectious Diseases Pharmacists. Clin. Biochem. Rev. 2010, 31, 21–24. [Google Scholar] [PubMed]
  165. Mu, F.; Zhou, X.; Fan, F.; Chen, Z.; Shi, G. A Fluorescence Biosensor for Therapeutic Drug Monitoring of Vancomycin Using in Vivo Microdialysis. Anal. Chim. Acta 2021, 1151, 338250. [Google Scholar] [CrossRef]
  166. Gage, B.F.; Fihn, S.D.; White, R.H. Management and Dosing of Warfarin Therapy. Am. J. Med. 2000, 109, 481–488. [Google Scholar] [CrossRef] [PubMed]
  167. Saeedi, I.; Ahmadi, S.; Thompson, M.; Hashemi, P.; Ramezani, Z. Electrochemical Sensor for the Direct Determination of Warfarin in Blood. Chemosensors 2022, 10, 44. [Google Scholar] [CrossRef]
  168. Nooraee Nia, N.; Hadjmohammadi, M.R. Nanofluid of Magnetic-Activated Charcoal and Hydrophobic Deep Eutectic Solvent: Application in Dispersive Magnetic Solid-Phase Extraction for the Determination and Preconcentration of Warfarin in Biological Samples by High-Performance Liquid Chromatograp. Biomed. Chromatogr. 2021, 35, e5113. [Google Scholar] [CrossRef] [PubMed]
  169. Nowak, P.; Olechowska, P.; Mitoraj, M.; Woźniakiewicz, M.; Kościelniak, P. Determination of Acid Dissociation Constants of Warfarin and Hydroxywarfarins by Capillary Electrophoresis. J. Pharm. Biomed. Anal. 2015, 112, 89–97. [Google Scholar] [CrossRef] [PubMed]
  170. Özbek, O.; Isildak, Ö.; Isildak, I. A Potentiometric Biosensor for the Determination of Valproic Acid: Human Blood–Based Study of an Anti–Epileptic Drug. Biochem. Eng. J. 2021, 176, 108181. [Google Scholar] [CrossRef]
  171. Alvau, M.D.; Tartaggia, S.; Meneghello, A.; Casetta, B.; Calia, G.; Serra, P.A.; Polo, F.; Toffoli, G. Enzyme-Based Electrochemical Biosensor for Therapeutic Drug Monitoring of Anticancer Drug Irinotecan. Anal. Chem. 2018, 90, 6012–6019. [Google Scholar] [CrossRef] [PubMed]
  172. Xu, Y.; Villalona-Calero, M.A. Irinotecan: Mechanisms of Tumor Resistance and Novel Strategies for Modulating Its Activity. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2002, 13, 1841–1851. [Google Scholar] [CrossRef]
  173. Carvalho, C.; Santos, R.X.; Cardoso, S.; Correia, S.; Oliveira, P.J.; Santos, M.S.; Moreira, P.I. Doxorubicin: The Good, the Bad and the Ugly Effect. Curr. Med. Chem. 2009, 16, 3267–3285. [Google Scholar] [CrossRef]
  174. Hassani Moghadam, F.; Taher, M.A.; Karimi-Maleh, H. Doxorubicin Anticancer Drug Monitoring by Ds-DNA-Based Electrochemical Biosensor in Clinical Samples. Micromachines 2021, 12, 808. [Google Scholar] [CrossRef] [PubMed]
  175. Dunvald, A.-C.D.; Iversen, D.B.; Svendsen, A.L.O.; Agergaard, K.; Kuhlmann, I.B.; Mortensen, C.; Andersen, N.E.; Järvinen, E.; Stage, T.B. Tutorial: Statistical Analysis and Reporting of Clinical Pharmacokinetic Studies. Clin. Transl. Sci. 2022, 15, 1856–1866. [Google Scholar] [CrossRef] [PubMed]
  176. Peng, P.-C.; Hong, R.-L.; Tsai, T.; Chen, C.-T. Co-Encapsulation of Chlorin E6 and Chemotherapeutic Drugs in a PEGylated Liposome Enhance the Efficacy of Tumor Treatment: Pharmacokinetics and Therapeutic Efficacy. Pharmaceutics 2019, 11, 617. [Google Scholar] [CrossRef] [PubMed]
  177. Seo, J.-W.; Fu, K.; Correa, S.; Eisenstein, M.; Appel, E.A.; Soh, H.T. Real-Time Monitoring of Drug Pharmacokinetics within Tumor Tissue in Live Animals. Sci. Adv. 2022, 8, eabk2901. [Google Scholar] [CrossRef]
  178. Jiang, D.; Wei, X.; Zhu, Y.; Qiu, Y.; Liu, X.; Kong, L.; Li, F.; Liu, J.; Zhuang, L.; Wan, H.; et al. Evaluating the Efficacy and Cardiotoxicity of EGFR-TKI AC0010 with a Novel Multifunctional Biosensor. Microsyst. Nanoeng. 2023, 9, 57. [Google Scholar] [CrossRef] [PubMed]
  179. Sun, D.; Hou, Z.; Yan, H.; Yang, Y.; Wang, G.; Wu, J.; Ma, J. An All-Fiber System Biosensor for Trace β-Lactam Antibiotics Detection Enhanced by Functionalized Microfiber and Fiber Bragg Grating. J. Colloid Interface Sci. 2024, 658, 903–912. [Google Scholar] [CrossRef] [PubMed]
  180. Draz, M.U.; Haq, M.Z.U.; Hayat, A.; Ajab, H. An ALP Enzyme-Based Electrochemical Biosensor Coated with Signal-Amplifying BaTiO 3 Nanoparticles for the Detection of an Antiviral Drug in Human Blood Serum. Nanoscale Adv. 2024, 6, 534–547. [Google Scholar] [CrossRef] [PubMed]
  181. Zheng, L.; Li, Q.; Deng, X.; Guo, Q.; Liu, D.; Nie, G. A Novel Electrochemiluminescence Biosensor Based on Ru(Bpy)32+-Functionalized MOF Composites and Cycle Amplification Technology of DNAzyme Walker for Ultrasensitive Detection of Kanamycin. J. Colloid Interface Sci. 2024, 659, 859–867. [Google Scholar] [CrossRef] [PubMed]
  182. Bao, Y.; Cai, K.; Liu, Y.; Li, B. Aptamer-Based Antibiotic Electrochemical Detection Platform Using Portable Plastic Gold Electrode and Multichannel Chip. Chin. J. Chem. 2024, 42, 171–176. [Google Scholar] [CrossRef]
  183. Qin, X.; Yin, Y.; Yu, H.; Guo, W.; Pei, M. A Novel Signal Amplification Strategy of an Electrochemical Aptasensor for Kanamycin, Based on Thionine Functionalized Graphene and Hierarchical Nanoporous PtCu. Biosens. Bioelectron. 2016, 77, 752–758. [Google Scholar] [CrossRef]
  184. Kim, D.-Y.; Sharma, S.K.; Rasool, K.; Koduru, J.R.; Syed, A.; Ghodake, G. Development of Novel Peptide-Modified Silver Nanoparticle-Based Rapid Biosensors for Detecting Aminoglycoside Antibiotics. J. Agric. Food Chem. 2023, 71, 12883–12898. [Google Scholar] [CrossRef] [PubMed]
  185. Zheng, X.; Ye, J.; Chen, W.; Wang, X.; Li, J.; Su, F.; Ding, C.; Huang, Y. Ultrasensitive Sandwich-Type SERS-Biosensor-Based Dual Plasmonic Superstructure for Detection of Tacrolimus in Patients. ACS Sens. 2022, 7, 3126–3134. [Google Scholar] [CrossRef] [PubMed]
  186. Abbasi, M.; Alsaikhan, F.; Obaid, R.F.; Jahani, S.; Biroudian, S.; Oveisee, M.; Arab, M.R.; Aramesh-Boroujeni, Z.; Foroughi, M.M. Development of the DNA-Based Voltammetric Biosensor for Detection of Vincristine as Anticancer Drug. Front. Chem. 2023, 10, 1060706. [Google Scholar] [CrossRef]
  187. Eskiköy Bayraktepe, D.; Yıldız, C.; Yazan, Z. The Development of Electrochemical DNA Biosensor Based on Poly-l-Methionine and Bimetallic AuPt Nanoparticles Coating: Picomolar Detection of Imatinib and Erlotinib. Talanta 2023, 257, 124361. [Google Scholar] [CrossRef] [PubMed]
  188. Zhao, Y.; Hao, Y.; Cui, M.; Li, N.; Sun, B.; Wang, Y.; Zhao, H.; Zhang, C. An Electrochemical Biosensor Based on DNA Tetrahedron Nanoprobe for Sensitive and Selective Detection of Doxorubicin. Bioelectrochemistry 2024, 157, 108652. [Google Scholar] [CrossRef]
  189. He, W.; Chen, Z.; Yu, C.; Shen, Y.; Wu, D.; Liu, N.; Zhang, X.; Wu, F.; Chen, J.; Zhang, T. Unlabelled LRET Biosensor Based on Double-Stranded DNA for the Detection of Anthraquinone Anticancer Drugs. Microchim. Acta 2024, 191, 15. [Google Scholar] [CrossRef]
  190. Mollarasouli, F.; Kurbanoglu, S.; Ozkan, S.A. The Role of Electrochemical Immunosensors in Clinical Analysis. Biosensors 2019, 9, 86. [Google Scholar] [CrossRef]
  191. Barreiros dos Santos, M.; Rodriguez-Lorenzo, L.; Queirós, R.; Espiña, B. Fundamentals of Biosensors and Detection Methods. In Microfluidics and Biosensors in Cancer Research: Applications in Cancer Modeling and Theranostics; Springer: Berlin/Heidelberg, Germany, 2022; pp. 3–29. [Google Scholar]
  192. Wang, B.; Wang, C.; Li, Y.; Liu, X.; Wu, D.; Wei, Q. Electrochemiluminescence Biosensor for Cardiac Troponin I with Signal Amplification Based on a MoS2@Cu2O–Ag-Modified Electrode and Ce: ZnO-NGQDs. Analyst 2022, 147, 4768–4776. [Google Scholar] [CrossRef]
  193. Sohrabian, A.; Parodis, I.; Carlströmer-Berthén, N.; Frodlund, M.; Jönsen, A.; Zickert, A.; Sjöwall, C.; Bengtsson, A.A.; Gunnarsson, I.; Rönnelid, J. Increased Levels of Anti-DsDNA Antibodies in Immune Complexes before Treatment with Belimumab Associate with Clinical Response in Patients with Systemic Lupus Erythematosus. Arthritis Res. Ther. 2019, 21, 259. [Google Scholar] [CrossRef]
  194. Kai, T.; Hirayama, S.; Soda, S.; Fuwa, F.; Nakagawa, S.; Ueno, T.; Hori, A.; Miida, T. Higher Concentration of 25-Hydroxycholesterol in Treatment-Naïve Patients with Type 2 Diabetes Compared to Healthy Individuals. J. Clin. Lipidol. 2023, 17, 384–391. [Google Scholar] [CrossRef]
  195. Uruc, S.; Dokur, E.; Gorduk, O.; Sahin, Y. Disposable and Ultrasensitive Label-Free Gold Nanoparticle Patterned Poly (3, 4-Ethylenedioxythiophene-Co-3-Methylthiophene) Electrode for Electrochemical Immunosensing of Prostate-Specific Antigen. New J. Chem. 2024, 48, 10415–10426. [Google Scholar] [CrossRef]
  196. Bae, H.-J.; Kim, S.-W.; Kim, I.-S. Comparison of Low-Density Lipoprotein Cholesterol Estimation Methods in Individuals with Insulin Resistance: A Cross-Sectional Study. Clin. Chim. Acta 2023, 547, 117393. [Google Scholar] [CrossRef] [PubMed]
  197. Meng, F.; Zhang, L.; Huo, W.; Lian, J.; Jesorka, A.; Shi, X.; Gao, Y. Dynamic Range Expansion of the C-Reactive Protein Quantification with a Tandem Giant Magnetoresistance Biosensor. ACS Omega 2021, 6, 12923–12930. [Google Scholar] [CrossRef] [PubMed]
  198. Thiruppathi, M.; Lee, J.-F.; Chen, C.C.; Ho, J.A. A Disposable Electrochemical Sensor Designed to Estimate Glycated Hemoglobin (HbA1c) Level in Whole Blood. Sens. Actuators B Chem. 2021, 329, 129119. [Google Scholar] [CrossRef]
  199. Karami, P.; Gholamin, D.; Johari-Ahar, M. Electrochemical Immunoassay for One-Pot Detection of Thyroxin (T4) and Thyroid-Stimulating Hormone (TSH) Using Magnetic and Janus Nanoparticles. Anal. Bioanal. Chem. 2023, 415, 4741–4751. [Google Scholar] [CrossRef] [PubMed]
  200. Yoon, J.; Cho, H.-Y.; Shin, M.; Choi, H.K.; Lee, T.; Choi, J.-W. Flexible Electrochemical Biosensors for Healthcare Monitoring. J. Mater. Chem. B 2020, 8, 7303–7318. [Google Scholar] [CrossRef]
  201. Cernian, A.; Tiganoaia, B.; Sacala, I.; Pavel, A.; Iftemi, A. PatientDataChain: A Blockchain-Based Approach to Integrate Personal Health Records. Sensors 2020, 20, 6538. [Google Scholar] [CrossRef]
  202. Ebada, A.I.; Abdelrazek, S.; Elhenawy, I. Applying Cloud Based Machine Learning on Biosensors Streaming Data for Health Status Prediction. In Proceedings of the 2020 11th International Conference on Information, Intelligence, Systems and Applications, Piraeus, Greece, 15–17 July 2020; pp. 1–8. [Google Scholar]
  203. Foroutan, B.; Najafabadi, A.R.A. Capabilities of Bioinformatics Tools for Optimizing Physicochemical Features of Proteins Used in Nano Biosensors: A Short Overview of the Tools Related to Bioinformatics. Biochem. Biophys. Rep. 2021, 27, 101094. [Google Scholar] [CrossRef] [PubMed]
  204. Preetam, S.; Dash, L.; Sarangi, S.S.; Sahoo, M.M.; Pradhan, A.K. Application of Nanobiosensor in Health Care Sector. Bio-Nano Interface Appl. Food Healthc. Sustain. 2022, 251–270. [Google Scholar] [CrossRef]
  205. Jin, X.; Liu, C.; Xu, T.; Su, L.; Zhang, X. Artificial Intelligence Biosensors: Challenges and Prospects. Biosens. Bioelectron. 2020, 165, 112412. [Google Scholar] [CrossRef]
Figure 1. Various uses of biosensors in healthcare, biotechnology, environmental monitoring, and food safety.
Figure 1. Various uses of biosensors in healthcare, biotechnology, environmental monitoring, and food safety.
Sensors 24 05143 g001
Figure 2. Key components of biosensors.
Figure 2. Key components of biosensors.
Sensors 24 05143 g002
Figure 3. The operational principles and sensing mechanism employed by a biosensor.
Figure 3. The operational principles and sensing mechanism employed by a biosensor.
Sensors 24 05143 g003
Figure 4. Schematic representation of the aptamer-based biosensor formulation and detection mechanism for COVID-19.
Figure 4. Schematic representation of the aptamer-based biosensor formulation and detection mechanism for COVID-19.
Sensors 24 05143 g004
Figure 5. Fluorescence-based biosensor for rapid and sensitive quantitation of vancomycin levels in rabbits.
Figure 5. Fluorescence-based biosensor for rapid and sensitive quantitation of vancomycin levels in rabbits.
Sensors 24 05143 g005
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

Hemdan, M.; Ali, M.A.; Doghish, A.S.; Mageed, S.S.A.; Elazab, I.M.; Khalil, M.M.; Mabrouk, M.; Das, D.B.; Amin, A.S. Innovations in Biosensor Technologies for Healthcare Diagnostics and Therapeutic Drug Monitoring: Applications, Recent Progress, and Future Research Challenges. Sensors 2024, 24, 5143. https://doi.org/10.3390/s24165143

AMA Style

Hemdan M, Ali MA, Doghish AS, Mageed SSA, Elazab IM, Khalil MM, Mabrouk M, Das DB, Amin AS. Innovations in Biosensor Technologies for Healthcare Diagnostics and Therapeutic Drug Monitoring: Applications, Recent Progress, and Future Research Challenges. Sensors. 2024; 24(16):5143. https://doi.org/10.3390/s24165143

Chicago/Turabian Style

Hemdan, Mohamed, Mohamed A. Ali, Ahmed S. Doghish, Sherif S. Abdel Mageed, Ibrahim M. Elazab, Magdy M. Khalil, Mostafa Mabrouk, Diganta B. Das, and Alaa S. Amin. 2024. "Innovations in Biosensor Technologies for Healthcare Diagnostics and Therapeutic Drug Monitoring: Applications, Recent Progress, and Future Research Challenges" Sensors 24, no. 16: 5143. https://doi.org/10.3390/s24165143

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop