Next Article in Journal
Antibacterial Activity and Mechanisms of TroHepc2-22, a Derived Peptide of Hepcidin2 from Golden Pompano (Trachinotus ovatus)
Next Article in Special Issue
An All-Green Photo-Electrochemical Biosensor Using Microalgae Immobilized on Eco-Designed Lignin-Based Screen-Printed Electrodes to Detect Sustainable Nanoherbicides
Previous Article in Journal
The Transcription Factor Nrf2 Mediates the Effects of Antrodia camphorata Extract on Neuropathological Changes in a Mouse Model of Parkinson’s Disease
Previous Article in Special Issue
Electrochemical Immunosensor for the Determination of Antibodies against Prostate-Specific Antigen Based on ZnO Nanostructures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Updates on the Biofunctionalization of Gold Nanoparticles for the Rapid and Sensitive Multiplatform Diagnosis of SARS-CoV-2 Virus and Its Proteins: From Computational Models to Validation in Human Samples

by
Rodica Elena Ionescu
Light, Nanomaterials and Nanotechnology (L2n) Laboratory, CNRS EMR 7004, University of Technology of Troyes, 12 Rue Marie Curie, CS 42060, CEDEX, 10004 Troyes, France
Int. J. Mol. Sci. 2023, 24(11), 9249; https://doi.org/10.3390/ijms24119249
Submission received: 30 April 2023 / Revised: 19 May 2023 / Accepted: 22 May 2023 / Published: 25 May 2023
(This article belongs to the Special Issue Molecular Biosensing: From Theory to Point of Care Analytical Device)

Abstract

:
Since the outbreak of the pandemic respiratory virus SARS-CoV-2 (COVID-19), academic communities and governments/private companies have used several detection techniques based on gold nanoparticles (AuNPs). In this emergency context, colloidal AuNPs are highly valuable easy-to-synthesize biocompatible materials that can be used for different functionalization strategies and rapid viral immunodiagnosis. In this review, the latest multidisciplinary developments in the bioconjugation of AuNPs for the detection of SARS-CoV-2 virus and its proteins in (spiked) real samples are discussed for the first time, with reference to the optimal parameters provided by three approaches: one theoretical, via computational prediction, and two experimental, using dry and wet chemistry based on single/multistep protocols. Overall, to achieve high specificity and low detection limits for the target viral biomolecules, optimal running buffers for bioreagent dilutions and nanostructure washes should be validated before conducting optical, electrochemical, and acoustic biosensing investigations. Indeed, there is plenty of room for improvement in using gold nanomaterials as stable platforms for ultrasensitive and simultaneous “in vitro” detection by the untrained public of the whole SARS-CoV-2 virus, its proteins, and specific developed IgA/IgM/IgG antibodies (Ab) in bodily fluids. Hence, the lateral flow assay (LFA) approach is a quick and judicious solution to combating the pandemic. In this context, the author classifies LFAs according to four generations to guide readers in the future development of multifunctional biosensing platforms. Undoubtedly, the LFA kit market will continue to improve, adapting researchers’ multidetection platforms for smartphones with easy-to-analyze results, and establishing user-friendly tools for more effective preventive and medical treatments.

1. Introduction

The contagious respiratory SARS-CoV-2 RNA virus emerged in early 2020 and continues to cause moderate to severe infections in many people [1], where the neutrophil-to-lymphocyte ratio and elevated serum creatinine biomarkers were linked to advanced disease symptoms and mortality [2]. The virus has structural proteins (spike (S), nucleocapsid (N), membrane (M), and envelope (E)) and non-structural proteins (RNA polymerase, 3CL protease) that serve as sensitive biomarkers, which are expressed during the viral infection [3]. These proteins are detected using either time-consuming conventional molecular techniques (e.g., an enzyme-linked immunosorbent assay (ELISA) [4,5], polymerase chain reaction (PCR) [6]) and/or rapid immunochromatographic tests (less than 20 min) known as lateral flow assays, which are easy to handle both for physicians and for the untrained public. To evaluate the presence/absence of a viral infection, different bodily fluids (whole blood, serum, saliva, nasal secretions, urine) are collected and analyzed. Unfortunately, commercial LFAs have lower sensitivity than PCRs and false negative results are often obtained at low viral loads. Therefore, improved detection techniques that are low-cost, selective, and provide reproducible results are urgently needed [7,8,9].
In this context, nanotechnology [10,11] is in high demanded for the rapid detection and treatment of SARS-CoV-2 infection [12]. Among the different nanomaterials [13], inorganic metal nanoparticles, especially noble metals, have made a significant contribution to controlling the spread of COVID-19 and assisting detection, immunodiagnosis, drug development, and photocatalytic viral inactivation. More specifically, colloidal gold nanoparticles (AuNPs) are intensively used in “in vitro” biomedical applications due to their unique physical and optical properties including localized surface plasmon resonance (LSPR) in the visible spectrum, fluorescence resonance energy transfer (FRET), and surface-enhanced Raman scattering (SERS) [14]. Moreover, AuNPs are among the most stable metal nanoparticles; they are highly biocompatible, easy to synthesize, and have numerous biomolecule functionalization abilities [15,16]. They have contributed significantly to the development of varied biosensors with highly improved analytical performance (low detection limits, sensitivity, response times), vaccine formulation [17,18,19,20,21,22], and the large-scale manufacturing of rapid tests [23,24,25].
In the presence of biomolecules, the optical properties of AuNPs change according to their size and shape, leading to different absorption bands that confirm the immobilization strategy step by step [26,27]. To date, the conjugation of AuNPs with viral proteins has typically been carried out by the physical electrostatic interaction at the pH value, varied within a range of ±0.5 around the protein’s isoelectric point [28]. Conversely, the other pH values exhibit a purple or white-gray color, indicating their aggregation and destructive effect on gold particles.
Nowadays, AuNPs are used for the ultrasensitive detection of varied targets [29] such as enzymes [30,31], proteins [32], DNA [33,34], cells [35], viruses [36,37,38,39,40,41], and bacteria [42,43]. Moreover, promising therapeutic approaches have just begun using AuNPs as antiviral agents capable of trapping viruses [44,45,46,47], in anticoagulant therapy [48], in photothermal therapy for breast cancer [49], and in the treatment of inflammatory bowel disease [50]. Preliminary studies using AuNPs as an anticancer agents [51] and medication [52] have also been reported, since bioinspired AuNPs can reach tumors via their enhanced permeability and retention effect [53].
AuNPs have also been used as signal transducers, acting as remarkable optical signal amplifiers, current amplifiers and resonance light-scattering supports in the fabrication of ultrasensitive biosensors, significantly improving the sensitivity of diagnoses. There are examples based on various multilevel optical (reflectometry [54], surface-enhanced Raman scattering (SERS) [55,56], localized surface plasmon resonance (LSPR) [57,58]) electrochemical (non-faradaic electrochemical impedance spectroscopy (n-EIS) [59,60], faradaic EIS [61], square-wave voltammetry (SWV) [62]) and piezoelectric [63]/acoustic [64] sensing platforms for the SARS-CoV-2 virus [65]. Interestingly, the rapid “in-tube” naked-eye detection of SARS-CoV-2 RNA was locally recorded in the presence of AuNPs capped with thiol-modified antisense oligonucleotides specific for the N gene and monitored by UV–visible absorbance spectroscopy, transmission electron microscopy, and hyperspectral microscopy [66]. Recently, the implementation of optical nanosensors based on AuNPs functionalized with oligonucleotides was proposed for the rapid and ultrasensitive detection of SARS-CoV-2 traces (1 ng/mL) on food surfaces and for viral prevention [67]. Additionally, the first attempts to use electrochemical sensing platforms for the internet of healthcare things (IoHT) have been made via the implementation of wearable biosensor patches with internet connection; these enable the rapid recording of viral symptoms and further data transmission to medical services in real time, facilitating the fastest drug treatments of positive individuals [68].
In the context of the COVID-19 pandemic, portable lateral flow assays based on paper strips are in high demand for the naked-eye viral detection obtained over different interconnected membranes (the sample pad, the conjugate pad, the nitrocellulose (NC) membrane, and the absorbent pad) assembled in a sandwich form, when the tested sample moves on the strip via capillary action [69]. LFAs have been intensively optimized in terms of different parameters: the sample drop volume, type of sample pad/conjugation pad/NC flow membrane pore size, blocking conditions, antibody coating concentration, incubation time, running buffers/effects of salts, storage temperature/protective packaging bags, and signal stability. They have been used for the detection of SARS-CoV-2 Ab in whole blood, serum, and plasma. To confirm a SARS-CoV-2 infection, colloidal AuNP-conjugated viral Ab [70] were used for LFA tests. IgA/IgG/IgM Ab [71] were typically used to draw the test line (T-line) and the control line (C-line) to validate the results [72] with the naked eye and with much more sensitive smartphone read-out facilities and high-sensitivity synchrotron X-ray fluorescence imaging [73]. When the sample pad was modified with SARS-CoV-2 receptor binding domain (RBD) Ag from an infected individual, it interacts with the specific AuNP-Ab probe dropped on the conjugate pad, resulting in the formation of immunocomplexes that migrate into the zone of detection, along the NC membrane, with the red T-line becoming visible within about 20 min.
More advanced LFA configurations propose either the use of glycan-based LFA diagnostics, as it can detect both the active and deactivated SARS-CoV-2 virus, with results similar to those obtained with other tests, or the integration of isothermal nucleic acid amplification with LFA, which highlights the possibility of applying genetic testing to affordable point-of-care systems. These LFAs may overcome the current limitations of expensive conventional methods, namely PCR and its related quantitative polymerase chain reaction (qPCR) [74], and the latest generation of digital PCR (dPCR) tests [75].
In the present review, biofunctionalized AuNPs were selected for computational and manual multidetection platforms for the SARS-CoV-2 virus using dry and wet chemistry (Figure 1), either adopting or not adopting buffer washing steps, using the varied biosensing strategies proposed by academic research teams and private markets. In this context, there is still a knowledge gap regarding various computational predictions and patients’ health status, mainly in relation to changes in body fluid composition (viral load, antibody level, antibody types, evolution of blood and urine biomarkers, etc.) in early, middle, and acute SARS-CoV-2 infections. To the author’s knowledge, this review is the first to emphasize the insightful and useful correlations between the theoretical approaches and practicalities of using AuNPs as supports in the construction of multifunctional (portable LFA) platforms for the rapid, sensitive, and selective identification of SARS-CoV-2 in complex biological environments (nasal secretions, saliva, blood, and sera).

2. Synthesis of Gold Nanoparticles for In-Tube and On-Paper Medical Diagnoses

Several bottom-up and top-down syntheses of gold nanoparticles continue to be developed, focusing on controlling their size, shape (nanospheres, nanorods, nanocubes, nanoshells, nananostars) [76], composition, and surface modification, for the detection of SARS-CoV-2 virus. Although the top-down methods (e.g., electron-/ion-beam lithography) provide nanoparticles of various shapes and sizes, with high resolution, good stability, and repeatability, they require expensive instrumentation and longer manufacturing times [77,78]. To overcome these limitations, the nanoparticles used for medical diagnoses are commonly prepared with a bottom-up approach, mainly using gas-phase synthesis (e.g., spray pyrolysis) and liquid-phase formation (e.g., chemical reduction) [79] (Figure 2). To date, the water-dispersed colloidal gold nanoparticles (cAuNPs) are synthesized from the metallic precursor HAuCl4 in the presence of the chemical sodium citrate, with a color change from colorless to red wine [80], This technique is widely used for rapid LFA [81] tests, even though the fluorescence-dye-labeled immunochromatographic [82,83] method exhibits higher sensitivity and selectivity. Elsewhere, positively charged cysteamine-AuNPs have also been prepared for carrying out specific immunoconjugation events [84].
Studies focusing on the direct synthesis and growth of AuNPs on solid wafers, on nanocellulose paper after microwave irradiation, and on PDMS–glass substrates have used sodium citrate/L-lysine [85], sodium citrate/Na3Cl [86], and cetyltrimethylammoniumchloride (CTAB)/ascorbic acid [87] as reducing/capping agent couples. The obtained gold nanostructures were tested as sensitive microfluidic platforms, as SERS substrates for SARS-CoV-2 S protein detection (dried 2 μL, 2 μg/mL, 5 ng/mL), and as ordered arrays of plasmon resonances.

3. Computational Predictions of COVID-19 Infection Using Functionalized Gold Nanoparticles

The development of biosensors based on trial-and-error assays is time consuming when the performances depend on different parameters, including: (i) the stable, fast, reproducible, and reversible interaction of the probe with the target at the microscopic scale, (ii) the size and shape of nanoparticles at the microscopic and mesoscopic scales, and (iii) the optimal experimental conditions for specific targets.
In the present review, three categories of computations using AuNPs for the significant improvement of plasmonic resonances were proposed for the sensitive detection of the COVID-19 virus.
The first category focuses on the probe–target interaction by evaluating its strength, velocity, and configuration, as well as the impact on the response of the biosensor. This type of modeling ensures that the chosen probe family is compatible with effective target detection, but it does not guarantee that the performance of the complete biosensor will be satisfactory.
The second category is based on machine learning, one of the methods of which is deep learning. It allows for convergence on an optimal solution, e.g., a probe to detect a given target or a therapy to treat an infection, based on a large amount of data. This type of model is dependent on the quality of the available data.
Lastly, the third category deals with analytical or parametric models for the optimization of the biosensor configuration (e.g., dimensions, the choice of metals comprising the nanoparticles), from the physical laws characterizing the interaction, the physical properties of materials and interfaces (e.g., the refractive index, permittivity, conduction) and the expected performances (e.g., the wavelength shift, transmitted or reflected intensity). The use of these computations requires the accurate determination of the properties of the materials, which can be difficult in the case of highly heterogeneous media and media of very small sizes. Such properties are often dependent on the experimental conditions.
To summarize, numerous parameters taken into account in the above models aim to improve the resolution of the associated equations, producing more realistic assumptions and neglecting those that only intervene in the second order. Moreover, the selection of the physical law is essential. For example, one study simulated the optical response of different configurations of functionalized AuNPs in the presence of SARS-CoV-2 virions and evaluated the behavior of colloidal solutions containing 105–1013 virions/mL [88]. Interestingly, an equilibration time of 0.3 μs was predicted for peptides (12 derived from ACE2 receptors [89]) to coat the AuNPs (3 nm, 6 nm, and 10 nm) to avoid their aggregation. The resulted peptides-thiol/AuNPs were used for the colorimetric detection of RBD fragments of SARS-CoV-2 S protein. The existence of a strong binding energy between RBD fragments and angiotensin-converting enzyme 2 (ACE2) peptides (66 kcal/mol) was also noted when the peptide-AuNPs were held at a distance of ~2 nm from the spike protein simulated in an aqueous solution [90].
Simulation methods using post-SELEX truncation and mutation were also employed for the evaluation of affinity and the activity characterization of aptamer–target complexes [91]. Moreover, numerical simulations showed that one to three layer(s) of graphene on Ag-Au NPs (15 nm) alloy films should be applied on optical fibers before functionalization with thiol–DNA for transmission measurements in the absence/presence of SARS-CoV-2 RNA, with good sensitivity and resolution [92].
Recent study assumed that 152 AuNPs of 10 nm would partially cover the surface of the SARS-CoV-2 virus and could be used for plasmonic detection in the visible range [93,94]. In contrast, another theoretical study reported an optimal viral extinction detection in the mid-infrared range, since a negligible contribution from the surrounding media was noticed [95]. Furthermore, a fully atomistic molecular dynamic (MD) simulation was applied to AuNPs (2 nm) functionalized with peptides in a physiological solution (150 mM NaCl) as sensitive inhibitory agents against SARS-CoV-2 infection; it proved to have a better affinity for RBD than ACE2 [96].
Finally, the effects of 100 anti-SARS-CoV-2 drug candidates were systematically evaluated using mathematical algorithms [97].

4. Biofunctionalization of Supports and Viral Detection in Spiked Real Samples

4.1. Using Dry Chemistry Functionalization through the Drop-Cast Approach without Rinsing Steps

Drop deposition is commonly used in electrochemical sensing techniques (Figure 3). However, the presence of the “coffee ring” effect can alter the uniform distribution of the nanoparticles that are drop-casted on an electrode. Fortunately, some solutions are proposed to overcome such drawbacks, including the Marangoni effect, anisotropic particles or optimal surfactants [98].
One study reported the use of screen-printed electrodes (SPEs) with platinum (Pt) as a working electrode (WE) for functionalization with carboxyl-modified single-walled carbon nanotubes (SWCNTs-COOH)/peptide thionine (Th)-coated AuNPs/SARS-CoV-2 monoclonal S Ab for square-wave voltammetry (SWV) measurements in the presence of different concentrations of S proteins in PBS and filtered human saliva. When constructing this biosensor, no washing step was performed; therefore, excess bioreagents were not removed from the electrode, possibly affecting the specificity of the immunorecognition events between the S proteins solution and the dried Ab not covalently attached to AuNPs. No measurements using saliva from patients infected with SARS-CoV-2 are reported [99]. SPEs with carbon WE were also modified with multi-walled carbon nanotubes (MWCNTs-COOH) (Table 1) and used for the drop-casting of different bioreagents until complete drying, as follows: chitosan nanoparticles/protein A/D-dimer antibody/bovine serum albumin (BSA)/D-dimer antigen (Ag) present in the diluted plasma of three patients. This rapid and disposable differential pulse voltammetry (DPV) test for D-dimer, a predictive marker of thromboembolic events, was proposed for patients with both thromboembolic and COVID-19 symptoms to improve their medical treatment [100]. It was found that high levels of D-dimer (˃850 μg/L) should not be used for the diagnosis of COVID-19, only for prognosis [100].
Interestingly, pyrolyzed paper was fixed on platinum electrodes patterned on glass slides and served as a porous working electrode for the drop-casting of Tween 20, an irreversible hydrophilic layer against biological fouling/AuNPs/recombinant spike (S) protein. After systematic drying between steps, this biosensor was used for the electrochemical impedance spectroscopy (EIS) detection of SARS-CoV-2 Ab in patient serum diluted with PBS buffer [101].
Table 1. Detection of COVID-19 using electrochemical methods and dried layers of drop-casted biomolecules without rinsing steps (*—target entity).
Table 1. Detection of COVID-19 using electrochemical methods and dried layers of drop-casted biomolecules without rinsing steps (*—target entity).
Drop Casting
SupportMethodAntibodies
(Ab)
Antigens
(Ag)
LODAdvantagesDisadvantagesRef
SPE-PtSWVMonoclonal anti-S AbS protein *200 pM in PBS
500 pM in filtered
saliva
Presence of
SWCNTs
Dried Ab
Short incubation time (5 min)
between
Ab and Ag
No rinsing
[99]
SPE-CDPVD-dimer AbD-dimer *0.6 μg/L
in PBS
Presence of MWCNTs
and
D-dimer, predictive markers of thromboembolic events important for viral infections
Optimized incubation time (15 min)
Dried Ab
No rinsing
[100]
Pyrolized
paper-Pt
EISAb in
serum
diluted with PBS *
S proteinNAPresence of
Tween-20
Dried Ab
No washing steps
[101]

4.2. Using Wet Chemistry for Biofunctionalization in Eppendorf Tubes, including Centrifugation and Rinsing Steps

Colorimetric and SERS measurements are proposed for the rapid (five min) detection of S proteins and SARS-CoV-2 pseudo-virus, using Eppendorf tubes and glass slides as supports for recording optical spectra from a few mL and dried drops of (bio)reagents, respectively (Table 2). For colorimetric assays, aqueous-citrate-coated AuNPs modified with carboxy-PEG-thiol (HS-PEG-COOH) were biofunctionalized with anti-S Ab, developing a pink color visible to the naked eye, which turned blue due to aggregation in the presence of specific antigens, namely, S protein and pseudo virus particles. For the SERS studies, AuNPs modified with 4-aminothiophenol (ATP) were functionalized with anti-S Ab and used for the detection of S protein and pseudo viral particles. In this work, the authors found that AuNPs functionalized with 100 ng/mL of anti-S Ab blocked 100% the viral replication by disrupting the lipid membrane of the SARS-CoV-2 pseudo virus, causing its collapse and preventing it from entering living cells (Figure 4A,B) [102].
Additionally, inactivated SARS-CoV-2 virus spiked in PBS buffer, water rivers, and artificial and human saliva was optically detected in Eppendorf tubes with a smartphone camera and an adapted interface that allowed for the visualization of significant color changes due to the functionalized AuNPs with MUA/SH-PEG/anti-S polyclonal Ab/BSA. Moreover, the contamination of São Carlos water samples spiked with inactivated SARS-CoV-2 virus was optically confirmed using the portable device (Figure 4C) [103].
Colloidal HS-AuNPs conjugated in Eppendorf tubes with anti-S Ab and three concentrations of the SARS-CoV-2 S Ag were dried on gold-coated toroidal planar metasurfaces prepared on silicon wafers. As expected, the AuNPs that dried on the Au films strongly confined the electromagnetic fields, ensuring significant frequency shifts for different antigen concentrations and rinsing steps with PBS. However, these sensitive metastructures have not yet been tested in the presence of real samples (Figure 4D) [104].
A magnetically assisted photoelectrochemical immunosensor (PEC) was constructed to detect the N protein of SARS-CoV-2 using Eppendorf tubes for the magnetic separation and washing steps before the electrochemical (CV, EIS) and optical (photocurrent in the nA range) investigations. Two main synthetic routes were proposed: one for the synthesis of Fe2O3/SiO2/AuNPs (called M-AuNPs) and the other for SiO2/TiO2/CdQDs (called P-QDs). The M-AuNPs were then used for conjugation with anti-N antibody 2D3 (Ab1)/SARS-CoV-2 N protein (Ag), while the P-QSs were used for conjugation with anti-N antibody 3F2 (Ab2). Finally, the biofunctionalized M-AuNPs/Ab1/SARS-CoV2-N and P-QS/Ab2 nanoparticles were mixed. After the PBS washing step, drops of the sandwich-form immune complexes were deposited on ITO electrodes and dried before the photocurrent measurements. As expected, the higher the concentration of the SARS-CoV-2-N protein, the more P-QS particles were combined with M-AuNPs, inducing higher photoelectric response values specific for an increased number of immune complexes (Figure 4E) [105].

4.3. Using Wet Chemistry for the Biofunctionalization of Solid Substrates, including Rinsing Steps

A sandwiched electrochemiluminescence (ECL) immunosensor using two antibodies for the detection of inactivated SARS-CoV-2 virus in PBS in about two hours was reported after two days of a functionalization protocol. The electrodeposition of flower-like AuNPs (60 s) on GCE was followed by the immobilization of Ab1 (polyclonal anti-SARS-CoV-2 Ab in rabbit, 13 μg/mL/SARS-CoV-2 inactivated virus vaccine (Coronavac–Sinovac, 1 mg/mL) and tested using luminol-colloidal AuNPs@Ab2 (the polyclonal anti-SARS-CoV S Ab in goat which identifies the S protein, 1.8 mg/mL). A photomultiplier tube (PMT) and smartphone were used to record the ECL signals related to the formed immunocomplexes, which were validated using saliva as a clinical sample collected from five infected positive and five healthy negative patients (Figure 5) [106].
AuNPs were functionalized with recombinant G protein (G-cAuNPs) and used for the detection of anti-SARS-CoV-2 IgG Ab in the presence of the full-length SARS-CoV-2 spike proteins that were immobilized on homemade eight-well gold interdigitated electrode (IDE) chips. In the case of positive serum samples, the impedance of IDE was altered by the G-cAuNPs, while, for the negative serum samples, the impedance of IDE remained the same since no IgG or G-cAuNPs were bound to the IDE surface [107].
Table 2. Detection of COVID-19 using optical/electrochemical/frequency methods and wet chemistry in tubes with rinsing steps (*—target entity).
Table 2. Detection of COVID-19 using optical/electrochemical/frequency methods and wet chemistry in tubes with rinsing steps (*—target entity).
Tubes/Surfaces
SupportMethodAntibodies
(Ab)
Antigens
(Ag)
LODAdvantagesDisadvantagesRef
Tubes and
glass slide
Colorime-tric
and
SERS
Monoclonal anti-S AbS protein *
1 ng/mL; 4 pg/mL
and
SARS-CoV-2 *
Pseudo-virus
1000
particles/mL:
18 particles/mL
NARapid test
(5 min)
Presence of
100 ng/mL anti-spike Ab
blocked 100% of viral replication
Dried Ag
for SERS
[102]
TubesColorime-tricPolyclonal AbS protein *
and
inactivated
SARS-CoV-2 *
virus in PBS,
water river,
artificial saliva,
human saliva
2.2 PFU/mL
PBS
and
0.28 PFU/mL
human
saliva
and
7; 250; 1000; 5000; 6000 PFU mL
Covalent functionalization of AuNPs with MUA and
thiol-PEG
Optical signal
reading with smartphone
Use of
polyclonal Ab
[103]
Tubes and
toroidal
planar
metasurfaces
Frequency shiftsMonoclonal anti-S AbS protein *
in PBS
(4 fM, 8 fM,
12 fM)
NAUse of colloidal thiol-AuNPsDried Ag on AuNPs
No tests with real samples
[104]
Tubes and
ITO
Photo-
current
Two sources
of anti-N Ab
N protein *
(1 h)
1.8 pg/mLEasy NPs washing by holding
magnet on tubes
Multiple steps for NPs synthesis[105]
GCEECLTwo sources
of polyclonal anti-SARS-CoV-2 Ab
Inactivated SARS-CoV-2 virus *
in PBS
(10 ng–10 μg/mL)
(2 h)
1.93
ng/mL
Use of active probes (luminol)
Optical signal
reading with smartphone
Time-consuming
protocol
(2 days)
[106]
IDEImpedanceSARS-CoV-2 IgG Ab
in
10 × serum *
(4 μL)
S protein
(0.1 mg/mL)
NAEasy monitoring of positive virus with change in frequencyMultistep
preparation
of IDE
supports
[107]

4.4. Using Lateral Flow Wet Chemistry for the Biofunctionalization of Metallic and Flexible Paper-Based Substrates including Rinsing Steps

Flexible nitrocellulose microporous membranes are used in the construction of the immunochromatography tests known as lateral flow assays for the qualitative visualization of immunocomplexes formed in the presence of antigens/antibodies of interest in various bodily fluids (nasal swabs, whole blood, serum, saliva, urine). For general public use or for more advanced medical investigations, LFAs are widely recommended due to their simplicity, single-use nature, cost effectiveness, and portability. These assays mainly consist of a sample pad, a conjugate pad, an NC membrane, and an absorbent pad, which are all packed in a polystyrene cassette for easier handling (Figure 6). When running the test, a sample droplet migrates from the conjugate pad to the absorbent pad due to capillary forces within about 15 min. In this section, LFA generations are discussed.
Conventional LFAs employ a citrate buffer for the stabilization of colloidal gold nanoparticles (cAuNPs), while more recently prepared gap-enhanced Raman polygonal gold nanotags (Au-GNPs) have been proposed as ultrasensitive methods of generating colorimetric and SERS signals. As mentioned above, the target analyte and Ab-labelled AuNPs will form immunocomplexes and bind to the detected antibodies line (the T (test)-line) when a positive signal appears as a red line. However, the main drawback of the LFA approach is the T line’s intensity, which is not always visible to the naked eye and is not always distinct from the background; therefore, false negatives are commonly obtained due to very low target concentrations, generating misleading LFA results. However, for optimal sample testing, the remaining Ab-AuNPs that do not react with the target should bind to the secondary antibodies in the control (C) line and form a red line, which should always appear in the absence or presence of the assessed antigen. Thus, LFAs often suffer from lower sensitivity [108] and require improvements to decrease the risk of viral contamination [109].
In this section, we discuss the latest advances in the fabrication and use of LFAs for the traditional detection of the SARS-CoV-2 virus and its proteins, as well as the first attempts to use LFA detection of the specifically designed synthetic DNA sequence target for the N region of SARS-CoV-2 in real human samples.

4.4.1. Conventional Protein Antigen/Antibody-Based LFAs: First Generation

Detection of SARS-CoV-2 N protein using nasal swabs

One study reports the use of nasopharyngeal swabs for the detection of SARS-CoV-2 N protein on nitrocellulose (NC) [110]. Before LFA testing, the conjugate pad was modified with dried immunocomplexes made of recombinant anti-SARS-CoV-2 Nucleocapsid antibody (Ab1)-cAuNPs (1:2 ratio). The two fine micro-sprayed lines used as the detection T-line were made of dried Ab1, while the C line was made of dried monoclonal goat anti-mouse IgG (Ab2) (Table 3).

Detection of SARS-CoV-2 RBD Using Nasal/Throat Swabs

The receptor-binding domain of SARS-CoV-2 was selected as the viral antigen, and it was detected in 100 μL PBS buffer and 50 diluted positive samples extracted from the nasopharynx before being added to the LFA strip for a 15 min test. Next, the intensity of the colorimetric signal obtained from the test line was analyzed using a smartphone colorimeter application called ColorGrab. Interestingly, the proposed test showed signal stability over 21 days when stored at 4 °C and no cross-reactivity with the influenza virus. The optimized LFA used a NC membrane modified with anti-RBD antibodies conjugated to cAuNPs for drop-casting on the conjugate pad, an anti-RBD antibody (on T-line) and anti-rabbit IgG antibody (on the C-line) for the detection zone that is visible to the naked eye, and an absorbent pad for the sample flow collection via capillary forces [111].
Recently, the LFA was modified with a trimethylsilyl cellulose barrier added to the CN140 hydrophobic NC membrane before the T-line to slow down the arrival of SARS-CoV-2 S RBD conjugated on cAuNPs. Thus, improvements were recorded in the recognition time between the SARS-CoV-2 S1 IgG and the SARS-CoV-2 S RBD Ag, which continues to accumulate when additional antibodies are present in positive samples. For the C-line, the SARS-CoV-2 S1 IgM antibody was used and validated the LOD of 0.11 ng/mL, 9.1 times lower than the classic LFA. Additionally, eight positive throat swab specimens were confirmed with the new LFA configuration [112].
On the other hand, the GeneTex Company proposes 2 LFA antigenic assays for viral detection in approximately 15 min in bodily fluids: one that screens for neutralizing antibodies to the SARS-CoV-2 S (RBD) that develop on the initial T-line modified with the RBD spike protein, and the second for SARS-CoV-2 virus on the initial T-line modified with the anti-S (RBD) antibody. For positive individuals, using the first kit configuration, the T-line modified with the grafted RBD protein is uncolored in the absence of antibodies from the tested sample, while the C-line modified with the grafted anti-ACE2 antibody is colored red in the presence of displayed human ACE2-AuNPs from the conjugation pad. In the second antigenic configuration, the T-line is colored red-purple in the presence of the SARS-CoV-2 virus due to the immunocomplexes formed with the anti-ARS-CoV2 antibody grafted on the paper. Meanwhile, the C-line is also reddish in color, as displaced Ab-AuNPs from the conjugation pad become attached to the secondary Ab grafted on the paper and promote their agglutination.

Detection of IgA, IgG, and IgM Antibodies in Serum and Nasal Samples

Serological tests for anti-SARS-CoV-2 IgA/IgM/IgG are of great importance to reducing the extent of viral contamination. The presence of IgM [113] and IgA has been noted mostly in asymptomatic and mild infections when IgA plays an important role in mucosal immunity, whereas IgG appears >14 days after the onset of symptoms [114].
The rapid LFA detection (15 min) of IgM antibodies (indicators of the viral acute infection period) in positive SARS-CoV-2 serum samples is proposed. In practical terms, conjugates of anti-IgM human antibody-cAuNPs (30 nm) were dropped on a conjugate pad, while SARS-CoV-2 and goat anti-mouse IgG were used for the T-line and C-line, respectively. The prepared LFA strips exhibited selectivity in the presence of thrombocytopenia syndrome virus (SFTSV) and dengue virus (DFV) [115].
A portable LFA nitrocellulose cassette was developed for the detection of human antibodies in serum samples (50 μL), specifically, the immunoglobulins M and G (early IgM and late IgG) were directed against two recombinant proteins, such as the N protein and the RBD of the SARS-CoV-2 S protein. These proteins were conjugated to 30 nm colloidal Au NPs and deposited on the conjugation pad. Goat anti-human IgG and IgM were used to draw the T-lines, while anti-mouse IgG (raised against both recombinant proteins) was used for the C-lines. It should be noted that, for those individuals infected for less than one week, the T- and C-lines on the NC strips were visible to the naked eye through their reddish color change caused by the presence of IgM and IgG of the N and RBD proteins [116]. Clinical studies confirmed the presence of the anti-IgA antibody in positive infected individuals in both early and late-stage SARS-CoV-2 infection [117]. For example, anti-IgM/IgG/IgA were simultaneously LFA detected (15 min) in positive serum/plasma (10 μL) dispensed onto sample pads that displaced the protein A-cAuNP conjugates dried on the fiberglass conjugation pad. The recombinant SARS-CoV-2 N protein and protein A were immobilized on the NC membrane at the T-line and C-line, respectively. Interestingly, the authors found that the visibility of the T- and C-lines were not affected by the presence of anticoagulants such as TSC, Heparin, and EDTA in the spiked negative, weak positive, and strong positive samples. LFA was also assessed for cross-reactivity with other virus types where the T-line did not show any visible response [118].

Detection of the SARS-CoV-2 Virus Using Glyco-Based LFA: The Influence of Silver Staining

Glycans are often more thermally robust than proteins; they are directly involved in the pathogen adhesion during viral infections and participate in the binding of the angiotensin-converting enzyme 2 (ACE2) receptor during SARS-CoV-2 cell adhesion/entry [119]. An initial glycan-based LFA lacking a T-line but with directly spotted low-volume samples (collected from 50 positive nasal swabs) was developed using an N-acetyl neuraminic-acid-modified polymer to coat AuNPs and capture the active/inactive SARS-CoV-2 virus via sialic acid from protein S’ binding site. For colorimetric signal enhancement, a silver-staining step was introduced, assuming that the silver ions were reduced to insoluble metallic silver in aqueous solutions with precipitation on the surface of AuNPs [120]. Another LFA study proposed the use of anti-S-antibody biofunctionalized AuNPs for the binding of the SARS-CoV-2 virus in the conjugation pad, followed by migration to a test line made of a biocompatible glycocalyx “sugary”-matrix-based biotin polysaccharide conjugated with streptavidin and rabbit antihuman-IgG for the efficient detection of SARS-CoV-2 S/Ab-AuNPs immune complexes [121].

4.4.2. Pressed Protein Antigen/Antibody LFAs in the Center: Second Generation

A study based on LFA principles proposed the use of a manual pressure point on the central zone of the NC paper strip, between the test (T) line and the control (C) line, to enhance signal intensity in the presence of a specific antigen due to a local reduction in the pore size and an increase in the fluid resistance. Therefore, slowing down the flow of the liquid loaded with trace amounts of an antigen such as CRP protein/SARS-CoV-2 N protein resulted in an increase in the anti-CRP antibody(Ab)-AuNPs or kit Ab-AuNPs/Ag conjugation time and in the number of immunocomplexes formed, which significantly improved the colored signal (T-line) and the detection limit. The authors found the optimal pressing condition at 15.69 MPa for CRP detection in PBS/CRP-free diluted human serum and at 23.54 MPa for detection of the SARS-CoV-2 N protein in two commercial COVID kits, namely, the STANDARD™ Q COVID-19 Ag Home Test and Humasis COVID-19 Ag Home [122].

4.4.3. SERS Optics for Protein Antigen/Antibody LFAs: Third Generation with Dual- and Tri-Mode Detection Signals

Recently, four studies [123,124,125,126] have proposed the use of Raman molecules in the functionalization of AuNPs as enhancers of colorimetric and SERS signals on one/two T-line LFA strips for single anti-SARS-CoV-2 N/S antibodies or simultaneous IgM/IgG detections (Table 3). Thus, the first study reports the use of gap-enhanced Raman polygonal gold nanotags (Au-GNPs), which are biofunctionalized with recombinant COVID-19 antigens (S-protein/N-protein), for the instantaneous detection of IgM/IgG in positive blood/serum samples. The homemade LFA strips had two T-lines (IgM/IgG) that were analyzed using rapid naked-eye colorimetric evaluation and advanced microscopic surface-enhanced Raman spectroscopy (SERS) measurements; they were more accurate than colorimetric data and had 100-fold the sensitivity. For the preparation of the Au-GNPs, the adsorption of 4-nitrobenzenethiol (4-NBT) on the Au’s core surface over 30 min was an obligate spacer molecule for the formation of a 1 mm gap between the core and the shell. Then, the C line was coated with polyclonal goat anti-chicken IgY Ab to capture the AuNPs [123] (Figure 7A).
In a second study, the simultaneous detection of anti-SARS-CoV-2 IgM/IgG in 19 positive clinical samples was reported using a naked eye colorimetric visualization and a portable Raman system for SERS signals on an LFA strip. To improve the SERS signal, the authors prepared nanoparticles labelled with dual layers of Raman dye (5,5-dithiobis-(2-nitrobenzoic acid—DTNB), namely, SiO2 (200 nm)/AuNPs (3 nm) and DTNBs (1), followed by coating with AgNPs(3 nm) and DTNBs (2). The obtained nanocomposites were biofunctionalized with a SARS-CoV-2 S protein that conjugated with anti-SARS-CoV-2 IgM/IgG from positive samples in the conjugation pad, which moved on the NC membrane to the IgM T-line, the IgG-T line, and the C-line, respectively [124] (Figure 7B).
Table 3. Detection of COVID-19 using optical/electrochemical/frequency methods and four generations of LFA tests (*—target entity).
Table 3. Detection of COVID-19 using optical/electrochemical/frequency methods and four generations of LFA tests (*—target entity).
First LFA
Paper
Support
MethodBioreagents
on AuNPs
Antigen
(Ag)
Bioreagents on
T-Line/C-Line
AdvantagesDisadvantagesRef
NCColorimetricAnti-N
20 ng/mL
(minimal concentration)/cAuNPs (1:2)
N-protein *
(2.14 ng/mL)
on nasal swabs
Ab1
2 mg/mL
and
Ab2
2 mg/mL
Low concentrations
and volumes of Ab
for conjugation
on AuNPs
High optimal concentration of Ab on
T/C-lines
Local
optimization
[110]
NC
(12 μ pore size)
ColorimetricAnti- RBD
(1 mg/mL)/
cAuNPs
(~15 nm, pH 9)
RBD *
on nasal swabs
in PBS
and
spiked PBS
(LOD 1 ng/mL)
Ab1
1 mg/mL
and
Ab2
1 mg/mL
100 μL sample
T-line readings
with smartphone
Signal stability for
21 days
Local
optimization
[111]
NC
CN140
ColorimetricRBD
(1 mg/mL in PBS):cAuNPs
(~20 nm, pH 9)
RBD * in PBS
LOD 0.11 ng/mL
RBD * in
throat swabs
S-IgG (500 μL/Ml
in PBS)
on T-spot
and
IgM (1 mg/mL in PBS) on C-spot
Presence of trimethylsilyl cellulose barrier and upper extra sample pad layer before T-line to slow down the flow
Improved LOD
Target detected on
T-spot with no line
Lack of robustness of local optimization:
Spacing between barrier/T spot/C spot
Homogeneity of spot size
[112]
NC +
1 mg/mL BSA
ColorimetricIgM human
(1.5 μg)-cAuNPs
(30 nm)
SARS-CoV-2 Ig M *
in serum samples
(20 μL/80 μL running buffer PBS/BSA/Triton)
SARS-CoV-2
(1 mg/mL) on T-line
and goat anti-mouse IgG (2 mg/mL)
on C-line
Good selectivity in the presence of
other viruses
10x test/sample
very low sample
volume
10–20 μL)
Laborious strip optimization[115]
NCColorimetricN (70–80 μg/mL) /Au NPs (30 nm)
and
RBD (60–70 μg/mL)/AuNPs
(30 nm) pH 9
Simultaneous
IgM * and IgG * in serum samples
On T-line:
goat anti-human IgG
(1 mg/mL)
and IgM
(1.2 mg/mL)
On C-line:
anti-mouse IgG 2 mg/mL for N and
1.8 mg/mL for RBD
Early detection of SARS-CoV-2 infection (<1 week)
Visible T-line
No calibration curve
No LOD
[116]
NCColorimetric Protein A-cAuNP
(pH 8.5–9)
Simultaneous
IgM */IgG */IgA *
in
serum/plasma
N protein (0.9 μg/cm) on T-line
and
Protein A
(0.3 μg/cm) on C-line
No influence from
anticoagulants on the color of T-/C-lines
No cross-reactivity with other viruses
No calibration curve
No LOD
[118]
NC
Immunopore
Colorimetric Glycan-AuNPs (35 nm) Inactivated SARS-CoV-2 *
collected from
nasal swabs
Spot of SARS-CoV-2 S protein-bearing lentivirus
and
Lectine on C-line
Sample spot instead
of T-line
Very low sample
volume (2 μL)
Improved color
visibility due to silver staining step
Need for polymeric coating to reduce nonspecific interaction[120]
NCColorimetric
and
enhanced signal in the
presence of HRP/AEC
reagents
Anti-S Ab/AuNPs
(10 nm)
SARS-CoV-2 S *
in saliva
Visible to naked eye for: HEP at 3.13 μg/mL in 25 μL,
and on HS at 3.13 μg/mL in 50 μL
On T-line:
GAG with
streptavidin
(1 mg/mL)
and
on C-line:
Anti IgG
(1 mg/mL)
Signal stability over
47 days at RT
Computation data
LOD (0.78 μg/mL in
25 μL) 4-fold lower compared with unamplified results
Laborious strip optimization [121]
Second LFA
NC
10 μm pore size (CNPF-SN12)
Colorimetric Anti-CRP
(2 mg/mL)/
AuNPs
(20 nm)
CRP protein * in
PBS and serum
and
COVID-19 Ag kit
2-fold increased signal
Anti-CRP Ab
(1 mg/mL)
on T-line
and
goat anti-mouse
IgG Ab
on C-line
Pressure zone between T-line and C-line induces flow delay and signal is enhancedNo tests with samples from SARS-CoV-2 infected
patients
[122]
Third LFA
NC
8 μm pore size
(CN140)
Colorimetric
and
SERS
S-protein/
N-protéiné on
Au-GNPs
(10 μg/
1.1 × 10−3 g GNPs
(55 nm)
+ 4-NBT
Simultaneous
IgM */IgG *
in positive blood/serum samples
Mouse anti-human IgM (1 mg/mL)/
mouse anti-human IgG (1 mg/mL)
on two T-lines
and
polyclonal goat anti-chicken IgY antibody (0.5 mg/mL)
on C-line
Two T lines
4-NBT (5 μg/mL)
on AuGNPs
100 times more
sensitive than
colorimetric result
LODs:
1 ng/mL for IgM and
0.1 ng/mL for IgG
Instability of
4-NBT on AuNPs
No data on storage of
Au-GNPs /buffer/tempe-rature
[123]
NC
~10 μm pore size (CN140)
Colorimetric
and
SERS
SARS-CoV-2 S/
nanocomposites
(30 μg of
S protein/ mL of
SERS tags) + DTNB
simultaneous
IgM */IgG *
in 103×, 104×, 105×, 106× diluted serum and 1% whole blood samples
after 25 min
immunoreaction
Goat anti-human IgM (0.5 mg/mL)/
goat anti-human IgG (0.6 mg/mL)
on two T-lines
and
SARS-CoV-2 S Ab (0.5 mg/mL) on
C-line
DTNB
Raman dye
SERS signal at 1328 cm–1 for IgM andIgG
800 times more sensitive that LFA based on AuNPs
LODSERS 1 pg/mL
No SERS data on individual components of blood/serum samples
for patients
with different
pathologies
[124]
NCColorimetric
and
SERS
SARS-CoV-2 N
(10 μg/mL)/
trimetallic
hybrid MNPs + MBA
anti-SARS-CoV-2 N
Ab * in
positive serum
samples
Mouse anti-SARS-CoV-2 N Ab
(1 mg/mL) on
T-line
and
rabbit anti-human Ab
(1 mg/mL) on C-line
LOD 10−8 mg/mL
MBA
Raman dye
SERS signal at 1075 cm−1, linear within the range of 10−10 to 10−6 mg /mL Ab *
LOD 0.08 pg/mL
No data on the stability of hybrid MNPs with MBA over time and influence of washing, RT and 4 °C in SERS signal evolution[127]
NC
CN140
Colorimetric
and
Photothermal
and
SERS
SARS-CoV-2 S protein
(10 μg/mL)/
bimetallic Au–Ag HNSs + MBA
Neutralizing
SARS-CoV-2
Ab * in
spiked serum samples 10x diluted with PBS
(20 to 1500 ng/mL)
and blood
ACE2 protein
on T-line
and
SARS-CoV-2 S
protein Ab
on C-line
MBA
Raman dye
LOD 160 ng/mL
LOD 20 ng/mL
LOD 20 ng/mL
Missing the SERS signal information used for the calibration curve
No data on the stability of NPs with MBA over time and influence of washing, RT and 4 °C in SERS signal evolution
[125]
Fourth LFA
NC
M170
ColorimetricThiol–DNA probe
(4 and 8 μM)/
AuNPs
SARS-CoV-2 N gene *SARS-CoV-2 N gene complementary on T-line
and
complementary DNA probe
on C-line
3 sandwich nLFA models
50 bp length of the target influence LOD 5 pM in SSC buffer
Laborious explanation
Manually micro-
pipetting lines
No calibration curves for the three proposed nLFA
models
[126]
NC
HF120
ColorimetricThiol–DNA detection probe/AuNPs
(~13 nm)
(7 μL)
SARS-CoV-2 RNA *Biotin–DNA capture probe (0.5 μL of 25 μM) on
T-line
and
biotin–DNA
control probe
(0.5 μL of 25 μM) on
C-line
-Two sugar barriers (10 %) before (2.5 μL) and after (1 μL) the
T-line to slow down the flow and improve the test sensitivity by 5-fold
No data on storage and signal stability
No data
using positive
samples
[128]
NCColorimetric
and
digital camera
and
SERS
(laser 532 nm)
Two antisense oligonucleotides
ASO (25 μM) labeled with FAM (30 μM) and biotin (30 μM)
Optimization:
ASO1-biotin
ASO2-FAM
Inactivated
SARS-CoV-2 N gene *
(RNA/cDNA) in
10x diluted swab
sample
and
67250; 3362; 168; 8; 0.42; 0.02; 0.001
copies/μL
Anti-FAM Ab +
Av-AuNPs
on T-line
and
anti-Av Ab
on C-line
30 min test after
collection of swab
Cys-AuNPs as
signal enhancer of
T-line color for concentrations below
8 copies/μL
LOD of 0.02 copies/μL
SERS signal at 1617 cm−1 strongly
increased
Validated in 30 positive nasal/nasopharyngeal samples
Naked-eye color evaluation not possible until
8 copies/μL
No calibration curve based on SERS signals
Several parameters require optimizations by end users
[129]
NCColorimetric
and
smartphone
Anti-FITC Ab/AuNPSARS-CoV-2 N gene detection in cDNA
and
SARS-CoV-2
RNA *
from swabs
tested positive with real-time
qRT-PCR
(5.6 × 106 to 3.9 × 103 copies/mL)
Streptavidin
on T-line
and
anti-goat IgG Ab on C-line
5 μL of LAMP-amplified product based on biotin-dUTP/FITC-LF + 20 μL running
buffer
LOD 3.9 × 103
RNA copies/mL
Laborious protocol
Temperature ˃60 °C for active polymerase enzyme and not applicable for kit market
[130]
NC ColorimetricAnti-FAM Ab/AuNPs RBD * within
the S gene of
SARS-CoV-2,
(synthetic DNA,
16.7 nM)
SA on T-line
and
anti-rabbit Ab
on C-line
Biotin primer for easy linking to SA-NC
FAM probe
Costly
protocol
No data with real positive samples
[131]
In a third study, colorimetric and SERS dual-LFA (dLFA) were also successfully performed for the quantification of anti-SARS-CoV-2 N Ab in positive serum samples (from 9 unvaccinated and 98 vaccinated volunteers). This study used trimetallic hybrid (MNPs) nanomaterials consisting of magnetic nanoparticles (Fe2O4NPs) that were electrostatically coated with silver/4-mercaptobenzoic acid (MBA-ERS probe) and AuNPs and biofunctionalized with the SARS-CoV-2 N protein. The resulting N-protein/MNPs were incubated/mixed with positive samples containing nucleocapsid antibodies using the Eppendorf wet chemistry approach; this was followed by magnetic separation and purification (p). Finally, pN-Ab/MNPs were dropped on an LFA strip; they then moved to a T-line made of rabbit anti-human IgG antibodies (the color changed to red), while excess immunocomplexes became bound to the C-line, which was made of mouse anti-N-protein antibody (color also changed to red), thus validating the LFA test [127] (Figure 7C).
For the more sensitive and accurate quantification of SARS-CoV-2-neutralizing antibodies, a three-mode competitive LFIA (tLFA) based on colorimetric, photothermal, and SERS detections was developed using hybrid bimetallic nanoparticles based on gold–silver alloy hollow nanoshells (Au/AgNPs) modified with 4-mercaptobenzoic acid (MBA) for the enhancement of SERS signals. This tLFA, which detects neutralizing antibodies against SARS-CoV-2, evaluated the ability of Ab to inhibit the binding between the SARS-CoV-2 S protein RBD and the angiotensin-converting enzyme 2 (ACE2). Therefore, no color was noticed on the T-line sprayed with the ACE2 protein, whereas a red color was visible on the C-line sprayed with SARS-CoV-2 S protein IgG for 79 out of 98 vaccinated individuals [125] (Figure 7D).

4.4.4. Nucleic Acid LFAs: Fourth Generation

An initial nucleic sandwich assay (nLFA) [126] for the detection of synthetic target DNA (50 bases long) specific for the N gene of SARS-CoV-2 in a saline-sodium citrate (SSC) running buffer was reported. The nLFA used two concentrations of thiol–DNA probes (4 and 8 μM) for immobilization on cAuNPs, while streptavidin-modified biotin-oligonucleotides were micro-pipetted directly onto the T- and C-lines (Table 3).
One nLFA prototype proposed the addition of two sugar barriers (drops) before and after the T-line, resulting in the five-fold improved sensitivity of SARS-CoV-2 RNA (44 bp) detection due to an increased reaction time with the 5’thiol-detecting DNA probe cAuNP and further with the biotin–DNA capture probe (20 bp) on the T-line [128].
Aggregates of cysteamine-AuNPs/streptavidin-AuNPs were used as nano-amplifiers of colorimetric/SERS LFA signals formed in the presence of extracted target SARS-CoV-2 RNA from 30 positive nasal/nasopharyngeal swab samples (30 min test) and two specific 6-carboxyfluorescein (FAM)/biotin-labeled antisense DNA oligonucleotide probes (20 bp). The anti-FAM antibody and anti-streptavidin antibody were used for the T-line and C-line, respectively [129].
Furthermore, the loop-mediated isothermal amplification (LAMP) technique for specific nucleic acid amplification was combined with the fLFA for SARS-CoV-2 N gene detection in 82 cDNA and clinical swab-extracted RNA samples. The nLFA was analyzed with a smartphone interface; a streptavidin (SA) printed T-line appeared with a red color only when the LAMP-amplified product based on biotin-dUTP/FITC-LF was detected by the AuNP anti-FITC antibody. Unfortunately, the implementation of LAMP for commercial kits is not yet possible due to the higher temperatures (˃60 °C) required to keep the polymerase enzyme active [130]. When using rolling circle amplification (RCA) coupled with nLFA, it made possible the colorimetric detection of synthetic RBD target DNA (16.7 nM) amplified by a biotin primer to a padlock probe specific for RBD (8.33 nM) modified with a FAM probe (0.1 μM), which bound to anti-FAM Ab/AuNPs and was captured by SA on the T-line [131].

5. Conclusions and Perspectives

The SARS-CoV-2 virus is a persistent global disease, so prevention and treatment schemes are essential. Even if significant efforts have been made to ensure the rapid detection of viral loads, further improvements are expected, in particular for single-use disposable tests that are accessible to non-medical personnel (e.g., pregnant women [132]). In this review, the author discusses about the advances in the biofunctionalization of gold nanoparticles with different viral proteins against SARS-CoV-2 for the direct screening of positive human samples. At present, only a few studies have used nanomaterials for the biosensing of SARS-CoV-2. In the first part of the review, the author focuses on the applications of AuNPs that have been used for computational predictions against viral infections; meanwhile, in the second part, the author summarizes the use of dry chemistry on solid supports with no rinsing steps and wet “in-tube” chemistry/paper flexible supports with rinsing steps. Finally, AuNPs have been implicated in the development of lateral flow assays [133], producing low-cost tools that are sensitive, specific, user friendly, and less time consuming. In the near future, biofunctionalized AuNPs will be deeply concerned with the development of modern sensing strategies, focusing on the ultrasensitive rapid assays based on SERS-ELISA in plates [134], multiple detection capabilities in LFA-SEM [135], and the simultaneous detection of IgA/IgM/IgG antibodies [136] against SARS-CoV-2, enabling the more effective monitoring of the evolution of the pandemic [137]. Efforts will also focus on the construction of high-throughput quantitative PCR tests for the immediate tracking of SARS-CoV-2 variants [138], as well as the full integration of LFA with biocompatible nanomaterials [139] for better discrimination between false negative results due to low viral loads and false positive results due to a mutant strain in the tested sample. Moreover, the next generation of LFA tests will certainly serve as routine diagnostic tools for vaccine quality control and the monitoring of active antibody titers developed in the vaccinated population, enabling the direct public coordination and understanding of vaccine dose(s). In addition, prevention conditions must be adopted as a permanent priority, with a particular focus on the development of commercial disinfectants based on biocompatible nanopolymers to facilitate smart transport for human activities [140].
In conclusion, remarkable progress has been made during the COVID-19 pandemic, and the world must remain vigilant [141] and prepare for possible future pandemics. To face this uncertainty, nanotechnological [142,143,144] miniaturized portable devices with multidetection [145] and high-throughput multimodal microtiter plate immunoassays [146] will play a central role in enabling the early detection of targeted viral loads that will save precious time during the isolation of patients and the delivery of medical treatment. Moreover, the preparation of AuNPs using green chemistry, which is less toxic to the environment and to humans, may contribute to an increase in the antiviral properties of metal nanoparticles for the future formulation of viral vaccines [147,148].

Funding

This research was funded by the Agence Nationale de la Recherche, grant number ANR-20-COVI-0080.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The author thanks ANR Flash COVID-19 for funding the AcOstoVie project.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations

AA—amino acids; ACE2—angiotensin-converting enzyme 2; Au-GNPs—polygonal gold nanotags; Au-Ag HNSs—gold–silver alloy hollow nanoshells; AEC—3-amino-9-ethylcarbazole; Av—streptavidin; ASO—unique target regions of the N gene (ASO1 starting with 421 and ending with 440; ASO2 starting with 443 and ending with 462); DTNB—5,5-dithiobis-2-nitrobenzoic acid; Cys—cysteamine; FAM—6-carboxyfluorescein; GAG—glycosaminoglycans; HEP—heparin; HS—heparin sulfate (HS) proteoglycans; HRP—peroxidase; MBA—4-mercaptobenzoic acid; 4-NBT—4-nitrobenzenethiol; PEG—thiol-PEG-carboxyl; SSC—saline sodium citrate; SA-NC—streptavidin-modified nitrocellulose.

References

  1. Kamble, P.; Daulatabad, V.; Patil, R.; John, N.A.; John, J. Omicron variant in COVID-19 current pandemic: A reason for apprehension. Horm. Mol. Biol. Clin. Investig. 2023, 44, 89–96. [Google Scholar] [CrossRef]
  2. Tufa, A.; Gebremariam, T.H.; Manyazewal, T.; Manyazewal, T.; Asrat, Y.; Getinet, T.; Hundie, T.G.; Webb, D.-L.; Hellström, P.M.; Genet, S. Limited value of neutrophil-to-lymphocyte ratio and serum creatinine as point-of-care biomarkers of disease severity and infection mortality in patients hospitalized with COVID-19. PLoS ONE 2022, 17, e0275391. [Google Scholar] [CrossRef] [PubMed]
  3. Yadav, R.; Chaudhary, J.K.; Jain, N.; Chaudhary, P.K.; Khanra, S.; Dhamija, P.; Sharma, A.; Kumar, A.; Handu, S. Role of structural and non-structural proteins and therapeutic targets of SARS-CoV-2 for COVID-19. Cells 2021, 10, 821. [Google Scholar] [CrossRef] [PubMed]
  4. Hofmann, N.; Grossegesse, M.; Neumann, M.; Schaade, L.; Nitsche, A. Evaluation of a commercial ELISA as alternative to plaque reduction neutralization test to detect neutralizing antibodies against SARS-CoV-2. Sci. Rep. 2022, 12, 3549. [Google Scholar] [CrossRef] [PubMed]
  5. Ma, L.; Abugalyon, Y.; Li, X.J. Multicolorimetric ELISA biosensors on a paper/polymer hybrid analytical device for visualpoint-of-care detection of infection diseases. Anal. Bioanal. Chem. 2021, 413, 4655–4663. [Google Scholar] [CrossRef]
  6. Perivolaropoulos, C.; Vlacha, V. A reduction of the number of assays and turnaround time by optimizing polymerase chain reaction (PCR) pooled testing for SARS-CoV-2. J. Med. Virol. 2021, 93, 4508–4515. [Google Scholar] [CrossRef] [PubMed]
  7. Seo, S.E.; Ryu, E.; Kim, J.; Shin, C.J.; Kwon, O.S. Fluorophore-encapsulated nanobeads for on-site, rapid, and sensitive lateral flow assay. Sens. Actuators B Chem. 2023, 381, 133364. [Google Scholar] [CrossRef]
  8. Jaisankar, A.; Krishnan, S.; Ramalingam, M. Design and development of nanoscale aptasensors for viral diagnostics. Lett. Appl. NanoBioScience 2023, 12, 34. [Google Scholar]
  9. Hemida, M.G. The next-generation coronavirus diagnosis techniques with particular emphasis on the SARS-CoV-2. J. Med. Virol. 2021, 93, 4219–4241. [Google Scholar] [CrossRef]
  10. Cardeso, V.M.d.O.; Moreira, B.J.; Comparetti, E.J.; Sampaio, I.; Ferreira, L.M.B.; Lins, P.M.P.; Zucolotto, V. Is nanotechnology helping in the fight against COVID-19. Front. Nanotechnol. 2020, 2, 588915. [Google Scholar] [CrossRef]
  11. Moabelo, K.L.; Martin, D.R.; Fadaka, A.O.; Sibuyi, N.R.S.; Meyer, M.; Madiehe, A.M. Nanotechnology-based strategies for effective and rapid detection of SARS-CoV-2. Materials 2021, 14, 7851. [Google Scholar] [CrossRef]
  12. Somu, P.; Mohanty, S.; Chakraborty, S.; Paul, S. Application of nanoscale materials and nanotechnology against viral infection: A special focus on coronaviruses. Adv. Exp. Med. Biol. 2021, 1352, 173–193. [Google Scholar] [PubMed]
  13. Srivastava, M.; Srivastava, N.; Mishra, P.K.; Malhotra, B.D. Prospects of nanomaterials-enabled biosensors for COVID-19 detection. Sci. Total Environ. 2021, 754, 142363. [Google Scholar] [CrossRef]
  14. Anh, N.H.; Doan, M.Q.; Dinh, N.X.; Huy, T.Q.; Tri, D.Q.; Ngoc Loan, L.T.; Van Hao, B.; Le, A.-T. Gold nanoparticle-based optical nanosensors for food and health safety monitoring: Recent advances and future perspectives. RSC Adv. 2022, 12, 10950–10988. [Google Scholar] [CrossRef] [PubMed]
  15. Jazaverin, M.H.; Amani, H.; Pourfatollah, A.A.; Pazoki-Toroudi, H.; Sedighimoghaddam, B. Various methods of gold nanoparticles (GNP) conjugation to antibodies. Sens. Bio-Sens. Res. 2016, 9, 17–22. [Google Scholar]
  16. Stine, K.J.; Jefferson, K.; Shulga, O.V. Nanoporous gold for enzyme immobilization. Methods Mol. Biol. 2017, 1504, 37–60. [Google Scholar]
  17. Mateu Ferrando, R.; Lay, L.; Polito, L. Gold nanoparticle-based platforms for vaccine development. Drug Discov. Today Technol. 2020, 38, 57–67. [Google Scholar] [CrossRef]
  18. Okay, S.; Özcan, O.O.; Karahan, M. Nanoparticles-based delivery platforms for mRNA vaccine development. AIMS Biophys. 2020, 7, 323–338. [Google Scholar] [CrossRef]
  19. Sengupta, A.; Azharuddin, M.; Al-Otaibi, N.; Hinkula, J. Efficacy and immune response elicited by gold nanoparticle-based nanovaccines against infectious diseases. Vaccines 2022, 10, 505. [Google Scholar] [CrossRef]
  20. Jafari, A.; Danesh Pouya, F.; Niknam, Z.; Abdollahpour-Alitappeh, M.; Rezaei-Tavirani, M.; Rasmi, Y. Curent advances and challenges in COVID-19 vaccine development: From conventional vaccines to next-generation vaccine platforms. Molec. Biol. Rep. 2022, 49, 4943–4957. [Google Scholar] [CrossRef]
  21. Moshref Javadi, M.; Taghdisi Hosseinzadeh, M.; Soleimani, N.; Rommasi, F. Evaluating the immunogenity of gold nanoparticles conjugated RBD with Freud’s adjuvant as a potential vaccine against SARS-CoV-2. Microb. Pathog. 2022, 170, 105687. [Google Scholar] [CrossRef]
  22. Zagorski, K.; Pandey, K.; Rajaiah, R.; Olwenyi, O.A.; Bade, A.N.; Acharya, A.; Johnston, M.; Filliaux, S.; Lyubchenko, Y.L.; Byrareddy, S.N. Modular nanoarray vaccine for SARS-CoV-2. Nanomed. Nanotechnol. Biol. Med. 2022, 46, 102604. [Google Scholar] [CrossRef]
  23. Mallakpou, S.; Azadi, E.; Hussain, C.M. The latest strategies in the fight against the COVID-19 pandemic: The role of metal and metal oxide nanoparticles. New J. Chem. 2021, 45, 6167–6179. [Google Scholar] [CrossRef]
  24. Biddram, E.; Esmaelli, Y.; Amini, A.; Sartorious, R.; Tay, F.R.; Shariati, L.; Makvandi, P. Nanobased platforms for diagnosis and treatment of COVID-19: From benchtop to bedside. ACS Biomater. Sci. Eng. 2021, 7, 2150–2176. [Google Scholar] [CrossRef] [PubMed]
  25. Guarise, C.; Pasquato, L.; De Filippis, V.; Scrimin, P. Gold nanoparticles-based protease assay. Proc. Natl. Acad. Sci. USA 2006, 103, 3978–3982. [Google Scholar] [CrossRef] [PubMed]
  26. Vidotti, M.; Carvalhal, R.F.; Mendes, R.K.; Ferreira, D.; Kubota, L.T. Biosensors based on gold nanostructures. J. Braz. Chem. Soc. 2011, 22, 3–20. [Google Scholar] [CrossRef]
  27. Dreaden, E.C.; Alkilany, A.M.; Huang, X.; Murphy, C.J.; El-Sayed, M.A. The golden age: Gold nanoparticles for biomedicine. Chem. Soc. Rev. 2012, 41, 2740–2779. [Google Scholar] [CrossRef]
  28. Tokmakov, A.A.; Kurotani, A.; Sato, K.-I. Protein pI and intracellular localization. Front. Mol. Biosci. 2021, 8, 775736. [Google Scholar] [CrossRef]
  29. Hori, K.; Yoshimoto, S.; Yoshino, T.; Zako, T.; Hirao, G.; Fujita, S.; Nakamura, C.; Yamagishi, A.; Kamiya, N. Recent advances in research on biointerfaces: From cell surfaces to artificial interfaces. J. Biosci. Bioeng. 2022, 133, 195–207. [Google Scholar] [CrossRef]
  30. Hutter, E.; Maysinger, D. Gold-nanoparticles-based biosensors for detection of enzyme activity. Trends Pharmacol. Sci. 2013, 34, 497–507. [Google Scholar] [CrossRef]
  31. Ramsey, A.V.; Bischoff, A.J.; Francis, M.B. Enzyme activated gold nanoparticles for versatile site-selective bioconjugation. J. Am. Chem. Soc. 2021, 143, 7342–7350. [Google Scholar] [CrossRef]
  32. Mohammadi, M.; Antoine, D.; Vitt, M.; Dickie, J.M.; Jyoti, S.S.; Wall, J.G.; Johnson, P.A.; Wawrousek, K.E. A fast, ultrasensitive SERS immunoassay to detect SARS-CoV-2 in saliva. Anal. Chim. Acta 2022, 1229, 340290. [Google Scholar] [CrossRef] [PubMed]
  33. Xu, Y.; Chen, B.; He, M.; Hu, B. A homogeneous nucleic acid assay for simultaneous detection of SARS-CoV-2 and influenza A (H3N2) by single-particle inductively coupled plasma mass spectrometry. Anal. Chim. Acta 2021, 1186, 339134. [Google Scholar] [CrossRef]
  34. Alafeef, M.; Dighe, K.; Moitra, P.; Pan, D. Rapid, ultrasensitive, and quantitative detection of SARS-CoV-2 using antisense oligonucleotides directed electrochemical biosensor chip. ACS Nano 2020, 14, 17028–17045. [Google Scholar] [CrossRef] [PubMed]
  35. Elwakkad, A.; Gamal el Din, A.A.; Saleh, H.A.; Ibrahim, N.E.; Hebishy, M.A.; Mourad, H.H.; El-Kassaby, M.; Abou-Seif, H.S.; Elgattan, G.M. Gold nanoparticles combined baker’s yeast as a successful approach for breast cancer treatment. J. Genet. Eng. Biotechnol. 2023, 21, 27. [Google Scholar] [CrossRef] [PubMed]
  36. Bao, Q.; Li, G.; Yang, Z.; Liu, J.; Wang, H.; Pang, G.; Guo, Q.; Wei, J.; Cheng, W.; Lin, L. Electrochemical biosensorbased on antibody-modified Au nanoparticles for rapid and senstive analysis of influenza A virus. Ionics 2023, 29, 2021–2029. [Google Scholar] [CrossRef]
  37. Sinawang, P.D.; Rai, V.; Ionescu, R.E.; Marks, R.S. Electrochemical lateral flow immunosensor for detection and quantification of dengue NS1 protein. Biosens. Bioelectron. 2016, 77, 400–408. [Google Scholar] [CrossRef]
  38. Karki, M.; Rajak, K.K.; Singh, P.; Fayaz, A.; Kumar, A.; Bhatt, M.; Rai, V.; Einstein, C.; Yadav, A.K.; Singh, R.P. Optimization of lateral flow assay for Canine morbilivirus detection and the application of the strip as sample substitute. J. Immunol. Methods 2023, 514, 113438. [Google Scholar] [CrossRef]
  39. Shyam, K.U.; Kong, K.-H.; Oh, M.-J.; Kim, T.; Kim, C.-S.; Kim, W.-S. Development of a lateral flow immunochromatographic assay (LFIA) for the detection of hirame novirhabdovirus (HIRRV) in olive flounder (Paralichthys olivaceus). Aquaculture 2023, 568, 739341. [Google Scholar] [CrossRef]
  40. Martinez-Liu, C.; Machain-Williams, C.; Martinez-Acuna, N.; Lozano-Sepulveda, S.; Galan-Huerta, K.; Arellanos-Soto, D.; Meléndez-villanueva, M.; Avalos-Nolazco, D.; Pérez-Ibarra, K.; Galindo-Rodriguez, S.; et al. Development of a rapid gold nanoparticle-based lateral flow immunoassay for the detection of Dengue virus. Biosensors 2022, 12, 495. [Google Scholar] [CrossRef]
  41. Najafabad, M.B.; Rastin, S.J.; Taghvaei, F.; Khiyavi, A.A. A review on applications of gold nanoparticles-based biosensors for pathogen detection. Adv. Nat. Sci. Nanosci. Nanotechnol. 2022, 13, 033002. [Google Scholar] [CrossRef]
  42. Das, A.; Terry, L.R.; Guo, H. A surface-enhanced Raman spectroscopy based smart Petri dish for sensitive and rapid detection of bacterial contamination in shrimp. Food Chem. 2023, 2, 100222. [Google Scholar] [CrossRef]
  43. Deb, M.; Hunter, R.; Taha, M.; Abdelbary, H.; Anis, H. Rapid detection of bacteria using gold nanoparticles in SERS with three different capping agents: Thioglucose, polyvinylpyrrolidone, and citrate. Spectrochim. Acta Mol. Biomol. Spectrosc. 2022, 280, 121533. [Google Scholar] [CrossRef]
  44. Peplow, M. Nanotechnology offers alternative ways to fight COVID-19 pandemic with antivirals. Nat. Biotechnol. 2021, 39, 1172–1174. [Google Scholar] [CrossRef]
  45. Ramaiah, G.B.; Tegegne, A.; Melese, B. Functionality of nanomaterials and its technological aspects-used in preventing, diagnosing and treating COVID-19. Mater. Today Proc. 2021, 47, 2337–2344. [Google Scholar] [CrossRef]
  46. Sarkar, J.; Das, S.; Aich, S.; Bhattacharyya, P.; Acharya, K. Antiviral potential of nanoparticles for the treatment of Coronavirus infections. J. Trace Elem. Med. Biol. 2022, 72, 126977. [Google Scholar] [CrossRef] [PubMed]
  47. Chakravarty, M.; Vora, A. Nanotechnology-based antiviral therapeutics. Drug Deliv. Transl. Res. 2021, 11, 748–787. [Google Scholar] [CrossRef]
  48. Barabadi, H.; Noqani, H.; Ashouri, E.; Prasad, A.; Jounaki, K.; Mobaraki, K.; Mohanta, Y.K.; Mostafavi, E. Nanobiotechnological approaches in anticoagulant therapy: The role of bioengineered silver and gold nanomaterials. Talanta 2023, 256, 124279. [Google Scholar] [CrossRef]
  49. Dheyab, M.A.; Aziz, A.A.; Khaniabadi, P.M.; Jameel, M.S.; Oladzadabbasadi, N.; Rahman, A.A.; Braim, F.S.; Mehrdel, B. Gold nanoparticles-based photothermal therapy for breast cancer. Photodiagn. Photodyn. Ther. 2023, 42, 103312. [Google Scholar] [CrossRef]
  50. Li, M.; Liu, J.; Shi, L.; Zhou, C.; Zou, M.; Fu, D.; Yuan, Y.; Yao, C.; Zhang, L.; Qin, S.; et al. Gold nanoparticles-embedded ceria with enhanced antioxidant activities for treating inflammatory bowel disease. Bioact. Mater. 2023, 25, 95–106. [Google Scholar] [CrossRef]
  51. Baker, A.; Wahid, I.; Hassan Baig, M.; Alotaibi, S.S.; Khalid, M.; Uddin, I.; Dong, J.-J.; Khan, M.S. Silk coccon-derived protein bioinspired gold nanoparticles as a formidable anticacer agent. J. Biomed. Nanotechnol. 2021, 17, 615–626. [Google Scholar] [CrossRef] [PubMed]
  52. Kim, H.S.; Lee, D.Y. Photothermal therapy with gold nanoparticles as an anticancer medication. J. Pharm. Investig. 2017, 47, 19–26. [Google Scholar] [CrossRef]
  53. Alle, M.; Sharma, G.; Lee, S.-H.; Kim, J.-C. Next-generation engineered nanogold for multimodal cancer therapy and imaging: A clinical perspectives. J. Nanobiotechnol. 2022, 20, 222. [Google Scholar] [CrossRef]
  54. Diyana Jamaluddin, N.; Ibrahim, N.; Yuziana Mohd Yusof, N.; Ta Goh, C.; Ling Tan, L. Optical reflectometric measurement of SARS-CoV-2 (COVID-19) RNA based on cationic cysteamine-capped gold nanoparticles. Opt. Laser Technol. 2023, 157, 108763. [Google Scholar] [CrossRef]
  55. Luo, X.; Yue, W.; Zhang, S.; Liu, H.; Chen, Z.; Qiao, L.; Wu, C.; Li, P.; He, Y. SARS-CoV-2 proteins monitored by long-range surface plasmon field-enhanced Raman scattering with hybrid bowtie nanoperture arrays and nanocavities. Lab Chip 2022, 23, 388–399. [Google Scholar] [CrossRef]
  56. Luo, Y.; Zhou, M.; Fan, C.; Song, Y.; Wang, L.; Xu, T.; Zhang, X. Active enrichment of nanoparticles for ultra-trace point-of-care COVID-19 detection. Anal. Chem. 2023, 95, 5316–5322. [Google Scholar] [CrossRef] [PubMed]
  57. Behrouzi, K.; Lin, L. Gold nanoparticle based plasmonic sensing for the detection of SARS-CoV-2 nucleocapsid proteins. Biosens. Biolectron. 2022, 195, 113669. [Google Scholar] [CrossRef]
  58. Oliveira, B.B.; Ferreira, D.; Fernandes, A.R.; Baptista, P.V. Engineering gold nanoparticles for molecular diagnosis and biosensing. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2023, 15, e1836. [Google Scholar]
  59. Zeng, J.; Duarte, P.A.; Ma, Y.; Savchenko, O.; Shoute, L.; Khaniani, Y.; Babiuk, S.; Zhuo, R.; Abdelrasoul, G.N.; Charlton, C.; et al. An impedimetric biosensor for COVID-19 serology test and modification of sensor performance via diaelectrophoresis force. Biosens. Bioelectron. 2022, 213, 114476. [Google Scholar] [CrossRef] [PubMed]
  60. Khaniani, Y.; Ma, Y.; Ghadiri, M.; Zeng, J.; Wishart, D.; Babiuk, S.; Charlton, C.; Kanji, J.N.; Chen, J. A gold nanoparticle-protein G electrochemical affinity biosensor for the detection of SARS-CoV-2 antibodies: A surface modification approach. Sci. Rep. 2022, 12, 12850. [Google Scholar] [CrossRef] [PubMed]
  61. Brazaca, L.C.; Imamura, A.H.; Gomes, N.O.; Almeida, M.B.; Scheidt, D.T.; Raymundo-Pereira, P.A.; Oliveira, O.N., Jr.; Janegitz, B.C.; Machado, S.A.S.; Carriho, E. Electrochemical immunosensors using electrodeposited gold nanostructures for detecting the S proteins from SARS-CoV and SARS-CoV-2. Anal. Bioanal. Chem. 2022, 414, 5507–5517. [Google Scholar] [CrossRef]
  62. Zeng, R.; Qiu, M.; Wan, Q.; Huang, Z.; Liu, X.; Tang, D.; Lnopp, D. Smartphone-based electrochemical immunoassay for point-of-care detection of SARS-CoV-2 nucleocapsid protein. Anal. Chem. 2022, 94, 15155–15161. [Google Scholar] [CrossRef]
  63. Mandal, D.; Indaleeb, M.M.; Younan, A.; Banerjee, S. Piezoelectric point-of-care biosensor for the detection of SARS-CoV-2 (COVID-19) antibodies. Sens. Bio-Sens. Res. 2022, 37, 100510. [Google Scholar] [CrossRef]
  64. Peng, Y.-C.; Cheng, C.-H.; Yatsud, H.; Liu, S.-H.; Liu, S.-J.; Kogai, T.; Kuo, C.-Y.; Wang, R.Y.L. Novel rapid test to detect anti-SARS-CoV-2 N protein IgG based on shear horizontal surface acoustic wave (SH-SAW). Diagnostics 2021, 11, 1838. [Google Scholar] [CrossRef]
  65. Iravani, S. Nano- and biosensors for the detection of SARS-CoV-2: Challenges and opportunities. Adv. Mater. 2020, 1, 3092–3103. [Google Scholar] [CrossRef]
  66. Moitra, P.; Alafeef, M.; Dighe, K.; Frieman, M.B.; Pan, D. Selective naked-eye detection of SARS-CoV-2 mediated by N gene targeted antisense oligonucleotide capped plasmonic nanoparticles. ACS Nano 2020, 14, 7617–7627. [Google Scholar] [CrossRef] [PubMed]
  67. Tessaro, L.; Aquino, A.; Panzenhagen, P.; Ochioni, A.C.; Mutz, Y.S.; Raymundo-Pereira, P.A.; Vieira, I.R.S.; Belem, N.K.R.; Conte-Junior, C.A. Development and application of an SPR nanobiosensor based on AuNPs for the detection of SARS-CoV-2 on food surfaces. Biosensors 2022, 12, 1101. [Google Scholar] [CrossRef] [PubMed]
  68. Paladhi, A.G.; Manohar, M.; Pal, K.; Vallinayagam, S.; Packirisamy, A.S.B.; Bashreer, V.A.; Sai Nandhini, R.; Ukhurebor, K. Novel electrochemical biosensor key significance of smart intelligence IoMT & IoHT) of COVID-19 virus control management. Process Biochem. 2022, 122, 105–109. [Google Scholar] [PubMed]
  69. Lee, A.S.; Kim, S.M.; Kim, K.R.; Park, C.; Lee, D.-G.; Heo, H.R.; Cha, H.J.; Kim, C.S. A colorimetric lateral flow immunoassay based on oriented antibody immobilization for sensitive detection of SARS-CoV-2. Sens. Actuators B Chem. 2023, 379, 133245. [Google Scholar] [CrossRef] [PubMed]
  70. Fox, T.; Geppert, J.; Dinnes, J.; Scandrett, K.; Bigio, J.; Sulis, G.; Hettiarachchi, D.; Mathangasinghe, Y.; Weeratunga, P.; Wickramasinghe, D.; et al. Antibody tests for identification of current and past infection with SARS-CoV-2. Cochrane Database Syst. Rev. 2022, 11, CD013652. [Google Scholar]
  71. Shaw, A.M.; Hyde, C.; Merrick, B.; James-Pemberton, P.; Squires, B.K.; Olkhov, R.V.; Batra, R.; Patel, A.; Bisnauthsing, K.; Nebbia, G.; et al. Real-world evaluation of a novel technology for quantitative simultaneous antibody detection against multiple SARS-CoV-2 antigens in a cohort of patients presenting with COVID-19 syndrome. Analyst 2020, 145, 5638–5646. [Google Scholar] [CrossRef] [PubMed]
  72. Guglielmi, G. Rapid coronavirus tests: A guide for the perplexed. Nature 2021, 590, 202–205. [Google Scholar] [CrossRef] [PubMed]
  73. Koller, G.; Morrel, A.P.; Galao, R.P.; Pickering, S.; MacMahon, E.; Johnson, J.; Ignatyev, K.; Neil, S.J.D.; Elsharkawy, S.; Fleck, R.; et al. More than the eye can see: Shedding new light on SARS-CoV-2 lateral flow device-based immunoassays. ACS Appl. Mater. Interfaces 2021, 13, 25694–25700. [Google Scholar] [CrossRef]
  74. Higuchi, R.; Fockler, C.; Dollinger, G.; Watson, R. Kinetic PCR analysis: Real-time monitoring of DNA amplification reactions. Bio/Technology 1993, 11, 1026–1030. [Google Scholar] [CrossRef] [PubMed]
  75. Nyaruaba, R.; Mwaliko, C.; Dobnik, D.; Neuzil, P.; Amoth, P.; Mwau, M.; Yu, J.; Yang, H.; Wei, H. Digital PCR applications in the SARS-CoV-2/COVID-19 era: A roadmap for future outbreaks. Clin. Microbiol. Rev. 2022, 35, e00168-21. [Google Scholar] [CrossRef] [PubMed]
  76. Hassan, H.; Sharma, P.; Hassan, M.R.; Singh, S.; Thakur, D.; Narang, J. Gold nanomaterials—the golden approach from synthesis to applications. Mater. Sci. Energy Technol. 2022, 5, 375–390. [Google Scholar] [CrossRef]
  77. Fu, X.; Cai, J.; Zhang, X.; Li, W.-D.; Ge, H.; Hu, Y. Top-down fabrication of shape-controlled, monodisperse nanoparticles for biomedical applications. Adv. Drug Deliv. Rev. 2018, 132, 169–187. [Google Scholar] [CrossRef]
  78. Aryal, S.; Park, H.; Leary, J.F.; Key, J. Top-down fabrication-based nano/microparticles for molecular imaging and drug delivery. Int. J. Nanomed. 2019, 14, 6631–6644. [Google Scholar] [CrossRef]
  79. Kumar, S.; Bhushan, P. Bhattacharya, Fabrication of nanostructures with bottom-up approach and their utility in diagnosis, therapeutics, and others. In Environmental, Chemical and Medical Sensors; Energy, Environment, and Sustainability; Springer: Singapore, 2018; pp. 167–198. [Google Scholar]
  80. Ji, X.; Song, X.; Li, J.; Bai, Y.; Yang, W.; Peng, X. Size control of gold nanocrystals in citrate reduction: The third role of citrate. J. Am. Chem. Soc. 2007, 129, 13939–13948. [Google Scholar] [CrossRef]
  81. Borse, V.; Konwar, A.N. Synthesis and characterization of gold nanoparticles as a sensing tool for the lateral flow immunoassay development. Sens. Inter. 2020, 1, 100051. [Google Scholar] [CrossRef]
  82. Feng, M.; Chen, J.; Xun, J.; Dai, R.; Zhao, W.; Lu, H.; Xu, J.; Chen, L.; Sui, G. Development of a sensitive immunochromatographic method using lanthanide fluorescent microsphere for rapid serodiagnosis of COVID-19. ACS Sens. 2020, 5, 2331–2337. [Google Scholar] [CrossRef] [PubMed]
  83. Xie, Q.Y.; Wu, Y.H.; Xiong, Q.R.; Xu, H.Y.; Xiong, Y.H.; Liu, K.; Jin, Y.; Lai, W.H. Advantages of fluorescent microspheres compared with colloidal gold as a label in immunochromatographic lateral flow assays. Biosens. Bioelectron. 2014, 54, 262–265. [Google Scholar] [CrossRef]
  84. Thangavelu, R.M.; Kadirvel, N.; Balasubramaniam, P.; Viswanathan, R. Ultrasensitive nano-gold labelled, duplex lateral flow immunochromatographic assay for early detection of sugarcane mosaic viruses. Sci. Rep. 2022, 12, 4144. [Google Scholar] [CrossRef] [PubMed]
  85. Prabhabar, A.; Bansal, I.; Jaiswar, A.; Roy, N.; Verma, D. A simple cost-effective microfluidic platform for rapid synthesis of diverse metal nanoparticles: A novel approach towards fighting SARS-CoV-2. Mater. Today Proc. 2023, 80, 1852–1857. [Google Scholar]
  86. Marques, A.C.; Pinheiro, T.; Morais, M.; Martins, C.; Andrade, A.F.; Martins, R.; Sales, M.G.F.; Fortunto, E. Bottom-up microwave-assisted seed-mediated syhthesis of gold nanopartciles onto nanocellulose to boos stability and high performane for SERS applications. Appl. Surf. Sci. 2021, 561, 150060. [Google Scholar] [CrossRef]
  87. Vinnacombe-Wilson, G.A.; Conti, Y.; Jonas, S.J.; Weiss, P.S.; Mihi, A.; Scarabelli, L. Surface lattice plasmon resonaces by direct in situ substrate growth of gold nanoparticles in ordered arrays. Adv. Matter. 2022, 34, 2205330. [Google Scholar] [CrossRef]
  88. Minopoli, A.; Scardapane, E.; Acunzo, A.; Della Ventura, B.; Veotta, R. Antonio Minpopoli Analysis of the optical response of a SARS-CoV-2-directed colorimetric immunosensor. AIP Adv. 2021, 11, 065319. [Google Scholar] [CrossRef]
  89. Lan, J.; Ge, J.; Yu, J.; Shan, S.; Zhou, H.; Fan, S.; Zhang, Q.; Shi, X.; Wang, Q.; Zhang, L.; et al. Structure of the SARS-CoV-2 antibodiesreceptor-binding domain bound to the ACE2 receptor. Nature 2020, 581, 215–220. [Google Scholar] [CrossRef]
  90. Vermma, N.; Badhe, Y.; Gupta, R.; Maparu, A.K.; Rai, B. Peptide mediated colorimetric detection of SARS-CoV-2 using gold nanoparticles: A molecular dynamics simulation study. J. Mol. Model. 2022, 28, 202. [Google Scholar] [CrossRef]
  91. Sun, D.; Sun, M.; Zhang, J.; Lin, X.; Zhang, Y.; Lin, F.; Zhang, P.; Yang, C.; Song, J. Computational tools for aptamer identification and optimization. TrAC Trends Anal. Chem. 2022, 157, 116767. [Google Scholar] [CrossRef]
  92. Saad, Y.; Gazzah, M.H.; Mougin, K.; Selmi, M.; Belmabrouk, H. Sensitive detection of SARS-CoV-2 using a novel plasmonic fiber optic biosensor design. Plasmonics 2022, 17, 1489–1500. [Google Scholar] [CrossRef] [PubMed]
  93. Baidya, S.; Hassan, A.M. SARS-CoV-2 detection using colorimetric plasmonic sensors: A proof-of-concept computational study. IEEE Trans. Nanobiosci. 2023, 22, 71–77. [Google Scholar] [CrossRef]
  94. Baidya, S.; Hassan, A.M. A computational study of COVID-19 detection using colorimetric plasmonic sensors. In Proceedings of the 2021 IEEE International Symposium on Antennas and Propagation and North American Radio Science Meeting, APS/URSI 2021—Proceedings, Singapore, 4–10 December 2021; pp. 1731–1732. [Google Scholar]
  95. Valagiannopoulos, C.; Sihvola, A. Maximal interaction of electromagnetic radiation with corona virions. Phys. Rev. B 2021, 103, 014114. [Google Scholar] [CrossRef]
  96. Mehranfar, A.; Izadyar, M. Theoretical design of functionalized gold nanoparticles as antiviral agents against severe acute respiratoty syndrome coronavirus 2 (SARS-CoV-2). J. Phys. Chem. Lett. 2020, 11, 10284–10289. [Google Scholar] [CrossRef]
  97. De, P.; Kumar, V.; Kar, S.; Roy, K.; Leszczynski, J. Repurposing FDA approved drugs as possible anti-SARS-CoV-2 medications using ligand-based computational approaches: Sum of ranking difference-based model selection. Struc. Chem. 2022, 33, 1741–1753. [Google Scholar] [CrossRef]
  98. Kaliyaraj Selva Kumar, A.; Zhang, Y.; Li, D.; Compton, R.G. A mini-review: How reliable is the drop casting technique? Electrochem. Commun. 2020, 121, 106867. [Google Scholar] [CrossRef]
  99. Farsaeivahid, N.; Grenier, C.; Nazarian, S.; Wang, M.L. A rapid label-free disposable electrochemical salivary point-of-care sensor for SARS-CoV-2 detection and quatification. Sensors 2023, 23, 433. [Google Scholar] [CrossRef] [PubMed]
  100. Tortolini, C.; Gigli, V.; Angeloni, A.; Galantini, L.; Tasca, F.; Antiochia, R. Disposable voltammetric immunosensor for D-dimer detection as early biomarker of thromboembolic disease and of COVID-19 prognosis. Biosensors 2023, 13, 43. [Google Scholar] [CrossRef]
  101. Nicoliche, C.Y.N.; Pascon, A.M.; Bezerra, I.R.S.; de Castro, A.C.H.; Martos, G.R.; Bettini, J.; Alves, A.W.; Santhiago, M.; Lima, R.S. In situ nanocoating on porous pyrolyzed paper enables antibiofouling and sensitive electrochemical analyses in biological fluids. ACS Appl. Mater. 2022, 14, 2522–2533. [Google Scholar] [CrossRef] [PubMed]
  102. Pramanik, A.; Gao, Y.; Patibandla, S.; Mitra, D.; McCandless, M.G.; Fassero, L.A.; Gates, K.; Tandon, R.; Chandra Ray, P. The rapid diagnosis and effective inhibition of coronavirus using spike antibody attached gold nanoparticles. Nanoscale Adv. 2021, 3, 1588–1596. [Google Scholar] [CrossRef]
  103. Materon, E.M.; Gomez, F.R.; Almeidia, M.B.; Azevedo, R.B.; Goncalves, D. Colorimetric detection of SARS-CoV-2 using plasmonic biosensors and smartphones. ACS Appl. Mater. Interfaces 2022, 14, 44538–54527. [Google Scholar] [CrossRef]
  104. Ahmadivand, A.; Gerislioglu, B.; Ramezani, Z.; Kaushik, A.; Manickam, P.; Ghoreishi, S.A. Functionalized terahertz plasmonic metasensors: Femtomolar-level detection of SARS-CoV-2 spike proteins. Biosens. Bioelectron. 2021, 177, 112971. [Google Scholar] [CrossRef]
  105. Guo, A.; Pei, F.; Hu, W.; Xia, M.; Mu, X.; Tong, Z.; Wang, F.; Liu, B. CdTe QDs-sensitized TiO2 nanocomposite for magnetic-assisted photoelectrochemical immunoassay of SARS-CoV-2 nucleocapsid protein. Bioelectrochemistry 2023, 150, 108358. [Google Scholar] [CrossRef] [PubMed]
  106. Hosseini, M.; Sobhanie, E.; Salehnia, F.; Xu, G.; Rabbani, H.; Sheikholeslami, M.N.; Firoozbakhtian, A.; Sadeghi, N.; Farajollah, M.H.; Reza Ganjali, M.; et al. Development of sandwich electrochemi-luminescence immunosensor for COVID-19 diagnosis by SARS-CoV-2 spike protein detection based on Au@BSA-luminol nanocomposites. Bioelectrochemistry 2022, 147, 108161. [Google Scholar] [CrossRef] [PubMed]
  107. Ye, X.; Jiang, T.; Ma, Y.; TO, D.; Wang, S.; Chen, J. A portable, low-cost and high-throughput electrochemical impedance spectroscopy device for point-of-care biomarker detection. Biosens. Bioelectron. X 2023, 13, 100301. [Google Scholar] [CrossRef]
  108. Khlebtsov, B.N.; Tumskys, R.S.; Burov, A.M.; Pylaev, T.E.; Khlebtsov, N.G. Quantifying the numbers of gold nanoparticles in the test zone of lateral flow immunoassay strips. ACS Appl. Nano Mater. 2019, 2, 5020–5028. [Google Scholar] [CrossRef]
  109. Foncea, P.; Mondschein, S.; Olivares, M. Replacing quarantine of COVID-19 contacts with periodic testing is also effective in mitigating the risk of transmission. Sci. Rep. 2022, 12, 3620. [Google Scholar] [CrossRef]
  110. Singh, S.; Singha, T.; Maheshwari, R.; Yadav, P.; Kumar, A.; Bhoj, H.; Sharma, B.; Verma, A.; Gupta, A.; Singh, H.; et al. Development of colloidal gold nanoparticle based lateral-flow assay for rapid detection of SARS-CoV-2 showing enhanced sensitivity and specificity. J. Appl. Biol. Biotechnol. 2023, 11, 253–258. [Google Scholar] [CrossRef]
  111. Prakashan, D.; Shrikrishna, N.S.; Byakodi, M.; Nagamani, K.; Gandhi, S. Gold nanoparticle conjugate-based lateral flow immunoassay (LFIA) for rapid detection of RBD antigen of SARS-CoV-2 in clinical samples using a smartphone-based application. J. Med. Virol. 2023, 95, e28416. [Google Scholar] [CrossRef]
  112. Srithong, P.; Chaiyo, S.; Pasomsub, E.; Rengpipat, S.; Chailapakul, O.; Praphairaksit, N. A novel delayed lateral flow immunoassay for enhanced detection of SARS-CoV-2 spike NC. Microchim. Acta 2022, 189, 386. [Google Scholar] [CrossRef]
  113. Higgins, V.; Fabros, A.; Wang, X.Y.; Bhandari, M.; Daghfai, D.J.; Kulasingam, V. Anti-SARS-CoV-2 IgM improves clinical sensitivity early in disease course. Clin. Biochem. 2021, 90, 1–7. [Google Scholar] [CrossRef] [PubMed]
  114. Nicol, T.; Lefeuvre, C.; Serri, O.; Pivert, A.; Joubaud, F.; Dubée, V.; Kouatchet, A.; Ducancelle, A.; Lunel-Fabiani, F.; Le Guillou-Guilemette, H. Assessment of SARS-CoV-2 serological tests for the diagnosis of COVID-19 through the evaluation of three immunoassays: Two automated immunoassays (Euroimmun and Abbott) and one rapid lateral flow immunoassay (NG Biotech). J. Clin. Virol. 2020, 129, 104511. [Google Scholar] [CrossRef] [PubMed]
  115. Huang, C.; Wen, T.; Shi, F.-J.; Zeng, X.-Y.; Jiao, Y.-J. Rapid detection of IgM antibodies against the SARS-CoV-2 virus via colloidal gold nanoparticles based lateral-flow assay. ACS Omega 2020, 5, 12550–12556. [Google Scholar] [CrossRef] [PubMed]
  116. Salem, R.; Elshamy, A.M.; Kamel, N.; Younes, S.; Marie, O.M.; Waly, F.R.; El-Kholy, A.A.; Elmenofy, W. A gold nanoparticles-based lateral flow assay utilizing baculovirus expressed recombinant nucelocapsid and receptor binding domain proteins for serodetection of IgG and IgM against SARS-CoV-2. Biotechnol. Lett. 2022, 4, 1507–1517. [Google Scholar] [CrossRef]
  117. Guo, L.; Ren, L.; Yang, S.; Xiao, M.; Chang, D.; Yang, F.; Dela Cruz, C.S.; Wang, Y.; Wu, C.; Xiao, Y.; et al. Profiling early humoral response to diagnose novel coronavirus disease (COVID-19). Clin. Infect. Dis. 2020, 71, 778–785. [Google Scholar] [CrossRef]
  118. Ngo, H.T.; Van Tran, S.; Nguyen, H.D.; Truong, P.Q. Humoral immune response in COVID-19 patients and novel design of lateral flow assay strip for simultaneous rapid detection of IgA/IgM/IgG antibodies against the SARS-CoV-2 virus. J. Appl. Biol. Biotechnol. 2023, 11, 102–113. [Google Scholar] [CrossRef]
  119. Mehdipour, A.R.; Hummer, G. Dual nature of human ACE2 glycosylation in binding to SARS-CoV-2 spike. Proc. Natl. Acad. Sci. USA 2021, 118, e2100425118. [Google Scholar] [CrossRef]
  120. Baker, A.N.; Richards, S.-J.; Pandey, S.; Guy, C.S.; Ahmad, A.; Hasan, M.; Biggs, C.I.; Georgiou, P.G.; Zwetsloot, A.J.; Straube, A.; et al. Glycan-based flow-through device for the detection of SARS-CoV-2. ACS Sens. 2021, 6, 3696–3705. [Google Scholar] [CrossRef]
  121. Kim, S.H.; Kearns, F.L.; Rosenfeld, M.A.; Casalino, L.; Papanikolas, M.J.; Simmerling, C.; Amaro, R.E.; Freeman, R. GlycoGrip: Cell surface-inspired universal sensor for betacoronaviruses. ACS Cent. Sci. 2022, 8, 22–42. [Google Scholar] [CrossRef]
  122. Park, S.B.; Shin, J.H. Pressed lateral flow assay strips for flow delay-induced signal enhancement in lateral flow assay strips. Biochip J. 2022, 16, 480–489. [Google Scholar] [CrossRef]
  123. Chen, S.; Meng, L.; Wang, L.; Huang, X.; Ali, S.; Chen, X.; Yu, M.; Yi, M.; Li, L.; Chen, X.; et al. SERS-based lateral flow immunoassay for sensitive and simultaneous detection of anti-SARS-CoV-2 IgM and IgG antibodies by using gap-enhanced Raman nanotags. Sens. Actuators B Chem. 2021, 348, 130706. [Google Scholar] [CrossRef]
  124. Liu, H.; Dai, E.; Xiao, R.; Zhou, Z.; Zhang, M.; Bai, Z.; Shao, Y.; Qi, K.; Tu, J.; Wang, C.; et al. Development of a SERS-based lateral flow immunoassay for rapid and ultra-sensitive detection of anti-SARS-CoV-2 IgM/IgG in clinical samples. Sens. Actuators B Chem. 2021, 329, 129196. [Google Scholar] [CrossRef] [PubMed]
  125. Zhao, T.; Liang, P.; Ren, J.; Zhu, J.; Yang, X.; Bian, H.; Li, J.; Cui, X.; Fu, C.; Xing, J.; et al. Gold-silver alloy hollow nanoshells-based lateral flow immunoassay for colorimetric, photothermal and SERS tri-mode detection of SARS-CoV-2 neutralizing antibody. Anal. Chim. Acta 2023, 1255, 341102. [Google Scholar] [CrossRef]
  126. Dighe, K.; Moitra, P.; Alafeef, M.; Gunaseelan, N.; Pan, D. A rapid RNA extraction-free lateral flow assay for molecular point-of-care detection of SARS-CoV-2 augmented by chemical probes. Biosens. Bioelectron. 2022, 200, 113900. [Google Scholar] [CrossRef] [PubMed]
  127. Li, J.; Liang, P.; Zhao, T.; Guo, G.; Zhu, J.; Wen, C.; Zeng, J. Colorimetric and Raman dual-mode lateral flow immunoassay detection of SARS-CoV-2 N protein antibody based on Ag nanoparticles with ultrathin Au shell assembled onto Fe3O4 nanoparticles. Anal. Bioanal. Chem. 2023, 415, 545–554. [Google Scholar] [CrossRef]
  128. Tang, R.; Alam, N.; Li, M.; Xie, M.; Ni, Y. Dissolvable sugar barriers to enhance the sensitivity of nitrocellulose lateral flow assay for COVID-19 nucleic acid. Carbohydr. Polym. 2021, 268, 118259. [Google Scholar] [CrossRef] [PubMed]
  129. Çam Derin, D.; Gültekin, E.; Içen Taskin, I.; Yakupogullan, Y. Development of nucleic acid based lateral flow assays for SARS-CoV-2 detection. J. Biosci. Bioeng. 2023, 135, 87–92. [Google Scholar] [CrossRef]
  130. Agarwal, S.; Warmt, C.; Henkel, J.; Schrick, L.; Nitschne, A.; Bier, F.F. Lateral flow-based nucleic acid detection of SARS-CoV-2 using enzymatic incorporation of biotin-labeled dUTP for POCT use. Anal. Bioanal. Chem. 2022, 414, 3177–3186. [Google Scholar] [CrossRef]
  131. Lee, H.N.; Lee, J.; Kang, Y.K.; Lee, J.H.; Yang, S.; Chung, H.J. A lateral flow assay for nucleic acid detection based on rolling circle amplification using capture ligand-modified oligonucleotides. Biochip J. 2022, 16, 441–450. [Google Scholar] [CrossRef]
  132. Fabre, M.; Ruiz-Martinez, S.; Monserrat Cantera, M.E.; Cortizo Garrido, A.; Beunza Fabra, Z.; Peran, M.; Benito, R.; Mateo, P.; Paules, C.; Oros, D. SARS-CoV-2 immunochromatographic IgM/IgG rapid test in pregnancy: A false friend? Anal. Clin. Biochem. 2021, 58, 149–152. [Google Scholar] [CrossRef]
  133. Kim, S.; Lee, J.-H. Current advances in paper-based biosensor technologies for rapid COVID-19 diagnosis. Biochip J. 2022, 16, 376–396. [Google Scholar] [CrossRef] [PubMed]
  134. Yu, Q.; Trinh, H.D.; Lee, Y.; Kang, T.; Chen, L.; Yoon, S.; Choo, J. SERS-ELISA using silica-encapsulated Au core-satellite nanotags for sensitive detection of SARS-CoV-2. Sens. Actuators B Chem. 2023, 382, 133521. [Google Scholar] [CrossRef] [PubMed]
  135. Kawasaki, H.; Suzuki, H.; Maekawa, M.; Hariyama, T. Combination of the Nanosuit method and gold/platinum particle-based lateral flow assay for quantitative and highly sensitive diagnosis using a desktop scanning electron microscope. J. Pharm. Biomed. Anal. 2021, 196, 113924. [Google Scholar] [CrossRef] [PubMed]
  136. Chen, M.; Qin, R.; Jiang, M.; Yang, Z.; Wen, W.; Li, J. Clinical applications of detecting IgG, IgM or IgA antibody for the diagnosis of COVID-19: A meta-analysis and systematic review. Int. J. Infect. Dis. 2021, 104, 415–422. [Google Scholar] [CrossRef]
  137. Nuccetelli, M.; Pieri, M.; Gisone, F.; Sarubbi, S.; Ciotti, M.; Andreoni, M.; Bernardini, S. Evaluation of a new simultaneous anti-SARS-CoV-2 IgA, IgM and IgG screening automated assay based on native inactivated virus. Int. Immunopharmacol. 2021, 92, 107330. [Google Scholar] [CrossRef]
  138. Malla, B.; Thakali, O.; Shrestha, S.; Segawa, T.; Kitajima, M.; Haramoto, E. Application of a high-throughput quantitative PCR system for simultaneous monitoring of SARS-CoV-2 variants and other viruses in wastewater. Sci. Total Environ. 2022, 853, 158659. [Google Scholar] [CrossRef]
  139. Sena-Torralba, A.; Alvarez-Diduk, R.; Parolo, C.; Piper, A.; Merkoci, A. Toward next generation lateral flow assays: Integration of nanomaterials. Chem. Rev. 2022, 122, 14881–14910. [Google Scholar] [CrossRef]
  140. Kamat, S.; Kumari, M.; Jayabaskaran, C. Nano-engineered tools in the diagnosis, therapeutics, prevention, and mitigation of SARS-CoV-2. J. Control Release 2021, 338, 813–836. [Google Scholar] [CrossRef]
  141. Tyagi, P.K.; Tyagi, S.; Mishra, M.; Dashora, K. Prevention, diagnosis and treatment of COVID-19: A nanotechnological perspective. Curr. Nanosci. 2021, 17, 418–422. [Google Scholar] [CrossRef]
  142. Ramakrishnan, S.G.; Robert, B.; Salim, A.; Ananthan, P.; Sivaramakrishnan, M.; Subramaniam, S.; Natesan, S.; Suresh, R.; Rajeshkumar, G.; Maran, J.P.; et al. Nanotechnology based solutions to combat zootic viruses with special attention to SARS, MERS, and COVID-19: Detection, protection and medication. Microb. Pathog. 2021, 159, 1051133. [Google Scholar] [CrossRef]
  143. Ionescu, R.E.; Marks, R.S.; Gheber, L.A. Nanolithography using protease etching of protein surfaces. Nano Lett. 2003, 3, 1639–1642. [Google Scholar] [CrossRef]
  144. Ionescu, R.E.; Marks, R.S.; Gheber, L.A. Manufacturing of nanochannels with controlled dimensions using protease nanolithography. Nano Lett. 2005, 5, 821–827. [Google Scholar] [CrossRef] [PubMed]
  145. Taha, B.A.; Al Mashhadany, Y.; Al-Jubouri, Q.; Rashid, A.R.B.A.; Luo, Y.; Chen, Z.; Rustagi, S.; Chaudhary, V.; Arsad, N. Next-generation nanophotonic-enabled biosensors for intelligent diagnosis of SARS-CoV-2 variants. Sci. Total Environ. 2023, 880, 163333. [Google Scholar] [CrossRef] [PubMed]
  146. Shan, Y.; Wang, B.; Huang, H.; Yan, K.; Li, W.; Wang, S.; Liu, F. Portable high-throughput multimodal immunoassay platform for rapid on-site COVID-19 diagnosis. Anal. Chim. Acta 2023, 1238, 340634. [Google Scholar] [CrossRef]
  147. Fatima, M.; Sadaf Zaidi, N.U.S.; Amraiz, D.; Afzal, F. In vitro antiviral activity of Cinnamomum cassia and its nanoparticles against H7N3 influenza a virus. J. Microbiol. Biotechnol. 2016, 26, 151–159. [Google Scholar] [CrossRef] [PubMed]
  148. Ahmadi, S.; Rabiee, N.; Fatahi, Y.; Hooshmad, S.E.; Bagherzadeh, M.; Rabiee, M.; Jajarmi, V.; Dinarvand, R.; Habibzadeh, S.; Saeb, M.R.; et al. Green chemistry and coronavirus. Sustain. Chem. Pharm. 2021, 21, 100415. [Google Scholar] [CrossRef]
Figure 1. Detection of the SARS-CoV-2 virus and its proteins (S,N) using biofunctionalized gold nanoparticles (AuNPs) for computational prediction and dry/wet chemistry (LFA—lateral flow assay with test (T)-line and control (C)-line).
Figure 1. Detection of the SARS-CoV-2 virus and its proteins (S,N) using biofunctionalized gold nanoparticles (AuNPs) for computational prediction and dry/wet chemistry (LFA—lateral flow assay with test (T)-line and control (C)-line).
Ijms 24 09249 g001
Figure 2. Synthesis of AuNPs using bottom-up methods applicable to the detection of the SARS-CoV-2 virus and its proteins.
Figure 2. Synthesis of AuNPs using bottom-up methods applicable to the detection of the SARS-CoV-2 virus and its proteins.
Ijms 24 09249 g002
Figure 3. Electrochemical detection of SARS-CoV-2 S proteins and D-dimer with drop-casted biofunctionalized AuNPs using dry chemistry on screen-printed electrodes (A,B) and Pt/glass slides (C). SWV—square-wave voltammetry [99]; DPV—differential pulse voltammetry [100]; EIS—Faradic electrochemical impedance spectroscopy [101].
Figure 3. Electrochemical detection of SARS-CoV-2 S proteins and D-dimer with drop-casted biofunctionalized AuNPs using dry chemistry on screen-printed electrodes (A,B) and Pt/glass slides (C). SWV—square-wave voltammetry [99]; DPV—differential pulse voltammetry [100]; EIS—Faradic electrochemical impedance spectroscopy [101].
Ijms 24 09249 g003
Figure 4. Optical and electrochemical detection of SARS-CoV-2 S and N-proteins using biofunctionalized gold nanoparticles (AuNPs) and wet chemistry “in tube” with (A,B), [102], (C), [103], (D), [104], and (E), [105]; SERS—surface-enhanced Raman spectroscopy; CV—cyclic voltammetry; EIS—electrochemical impedance spectroscopy.
Figure 4. Optical and electrochemical detection of SARS-CoV-2 S and N-proteins using biofunctionalized gold nanoparticles (AuNPs) and wet chemistry “in tube” with (A,B), [102], (C), [103], (D), [104], and (E), [105]; SERS—surface-enhanced Raman spectroscopy; CV—cyclic voltammetry; EIS—electrochemical impedance spectroscopy.
Ijms 24 09249 g004
Figure 5. Optical detection of SARS-CoV-2 S proteins using biofunctionalized AuNPs and wet chemistry on solid electrodes with rinsing steps. ECL—electrochemiluminescence.
Figure 5. Optical detection of SARS-CoV-2 S proteins using biofunctionalized AuNPs and wet chemistry on solid electrodes with rinsing steps. ECL—electrochemiluminescence.
Ijms 24 09249 g005
Figure 6. Different configurations of LFA cassettes for screening for the presence of SARS-CoV-2 in nasal swabs/serum/plasma/saliva samples (B) and the production of its specific IgA/IgM/IgG Ab (A) using biofunctionalized gold nanoparticles (AuNPs) and wet chemistry “on-paper” support electrodes with rinsing steps for colorimetric, SERS, and photothermal rapid testing.
Figure 6. Different configurations of LFA cassettes for screening for the presence of SARS-CoV-2 in nasal swabs/serum/plasma/saliva samples (B) and the production of its specific IgA/IgM/IgG Ab (A) using biofunctionalized gold nanoparticles (AuNPs) and wet chemistry “on-paper” support electrodes with rinsing steps for colorimetric, SERS, and photothermal rapid testing.
Ijms 24 09249 g006
Figure 7. LFAs for simultaneous and individual SARS-CoV-2 IgM and IgG detection in human sera using different coatings on AuNPs and dyes as Raman enhancers: (A) one layer of NBT—4-nitrobenzenethiol [123]: (B) two layers of DTMB—5,5-dithiobis-(2-nitrobenzoic acid [124]; (C,D) one layer of MBA—4-mercaptobenzoic acid [125,127].
Figure 7. LFAs for simultaneous and individual SARS-CoV-2 IgM and IgG detection in human sera using different coatings on AuNPs and dyes as Raman enhancers: (A) one layer of NBT—4-nitrobenzenethiol [123]: (B) two layers of DTMB—5,5-dithiobis-(2-nitrobenzoic acid [124]; (C,D) one layer of MBA—4-mercaptobenzoic acid [125,127].
Ijms 24 09249 g007
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

Ionescu, R.E. Updates on the Biofunctionalization of Gold Nanoparticles for the Rapid and Sensitive Multiplatform Diagnosis of SARS-CoV-2 Virus and Its Proteins: From Computational Models to Validation in Human Samples. Int. J. Mol. Sci. 2023, 24, 9249. https://doi.org/10.3390/ijms24119249

AMA Style

Ionescu RE. Updates on the Biofunctionalization of Gold Nanoparticles for the Rapid and Sensitive Multiplatform Diagnosis of SARS-CoV-2 Virus and Its Proteins: From Computational Models to Validation in Human Samples. International Journal of Molecular Sciences. 2023; 24(11):9249. https://doi.org/10.3390/ijms24119249

Chicago/Turabian Style

Ionescu, Rodica Elena. 2023. "Updates on the Biofunctionalization of Gold Nanoparticles for the Rapid and Sensitive Multiplatform Diagnosis of SARS-CoV-2 Virus and Its Proteins: From Computational Models to Validation in Human Samples" International Journal of Molecular Sciences 24, no. 11: 9249. https://doi.org/10.3390/ijms24119249

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