Next Article in Journal
From Immune Dysregulations to Therapeutic Perspectives in Myelodysplastic Syndromes: A Review
Next Article in Special Issue
Valid and Reliable Assessment of Upper Respiratory Tract Specimen Collection Skills during the COVID-19 Pandemic
Previous Article in Journal
Cerebrospinal Fluid Levels of AFP and hCG: Validation of the Analytical Method and Application in the Diagnosis of Central Nervous System Germ Cell Tumors
Previous Article in Special Issue
Development of an Enzyme-Linked Immunosorbent Assay (ELISA) for Accurate and Prompt Coronavirus Disease 2019 (COVID-19) Diagnosis Using the Rational Selection of Serological Biomarkers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tracking SARS-CoV-2: Novel Trends and Diagnostic Strategies

1
Centro de Investigación Biomédica, Facultad de Ciencias de la Salud Eugenio Espejo, Universidad UTE, Quito 170147, Ecuador
2
Immunology and Virology Laboratory, Department of Life Science and Agriculture, Universidad de las Fuerzas Armadas, Quito 171103, Ecuador
3
Grupo de Investigación en Biotecnología Aplicada a Biomedicina (BIOMED), Universidad de Las Américas, Quito 170125, Ecuador
4
One Health Research Group, Faculty of Medicine, Universidad de Las Américas (UDLA), Quito 170125, Ecuador
5
Institute for Bioengineering, School of Engineering, University of Edinburgh, Edinburgh EH8 9LE, UK
6
Centre for Synthetic and Systems Biology (SynthSys), University of Edinburgh, Edinburgh EH8 9LE, UK
7
Experimental Medicine Research Unit, Facultad de Medicina, Universidad Nacional Autónoma de México, Mexico City 4510, Mexico
8
Centro de Investigación Genética y Genómica, Facultad de Ciencias de la Salud Eugenio Espejo, Universidad UTE, Quito 170147, Ecuador
9
Escuela de Medicina, Colegio de Ciencias de la Salud Quito, Universidad San Francisco de Quito USFQ, Quito 170901, Ecuador
*
Author to whom correspondence should be addressed.
Diagnostics 2021, 11(11), 1981; https://doi.org/10.3390/diagnostics11111981
Submission received: 14 July 2021 / Revised: 18 September 2021 / Accepted: 23 September 2021 / Published: 26 October 2021

Abstract

:
The COVID-19 pandemic has had an enormous impact on economies and health systems globally, therefore a top priority is the development of increasingly better diagnostic and surveillance alternatives to slow down the spread of the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). In order to establish massive testing and contact tracing policies, it is crucial to have a clear view of the diagnostic options available and their principal advantages and drawbacks. Although classical molecular methods such as RT-qPCR are broadly used, diagnostic alternatives based on technologies such as LAMP, antigen, serological testing, or the application of novel technologies such as CRISPR-Cas for diagnostics, are also discussed. The present review also discusses the most important automation strategies employed to increase testing capability. Several serological-based diagnostic kits are presented, as well as novel nanotechnology-based diagnostic methods. In summary, this review provides a clear diagnostic landscape of the most relevant tools to track COVID-19.

1. Introduction

In late 2019, multiple cases of atypical pneumonia caused by a then unknown pathogen were reported in Wuhan, province of Hubei in China. The disease named COVID-19 quickly spread globally and, after considerable research efforts, the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) was identified as the pathogen causing the disease [1]. Coronaviruses are a large group of viruses that can infect a wide range of vertebrate hosts [2]. Diseases caused by each of those viruses show different clinical manifestations, usually affecting the respiratory tract in humans and digestive disorders in other animals [3]. In the last 20 years, three highly pathogenic beta coronaviruses have emerged from zoonotic events and have caused human disease [4]. In February 2003, the severe acute respiratory syndrome-related coronavirus 1 (SARS-CoV-1) was first reported in Asia. According to the World Health Organization (WHO), a total of 8,098 people worldwide became infected and 774 died (global case–fatality ratio (CFR) of 11%) [5] across two dozen countries in North America, South America, Europe, and Asia [6]. Since September 2012, WHO has been notified of 2562 laboratory-confirmed cases of infection with Middle East respiratory syndrome-related coronavirus (MERS-CoV), with a global CFR of 34.4% since 2012 [7]. Finally, the current disease caused by SARS-CoV-2, to date, has resulted in 217,914,903 confirmed cases and 4,523,984 deaths in 223 affected countries, areas, and territories [8]. There have been many attempts to estimate the global CFR of COVID-19 [9,10,11,12,13,14], but it has proven challenging, mainly because different regions of the world are experiencing different stages of the pandemic [15] affecting those estimates.
SARS-CoV-2 belongs to the Coronaviridae family, subfamily Orthocoronavirinae, genus Betacoronavirus, subgenus Sarbecovirus, and the species severe acute respiratory syndrome-related coronavirus [16]. Coronaviruses are enveloped, positive-sense, single-stranded RNA viruses that are distributed broadly among humans, other mammalian, and avian hosts [17]. The genome of SARS CoV-2 (NCBI Reference Sequence: NC_045512.2) [18] is very similar to the genomes of other known SARS-CoV and SARS-related coronaviruses, including the one that caused the SARS epidemic in 2003 (SARS CoV, NCBI Reference sequence: NC_004718.3) [19] and other related coronaviruses (e.g., MERS-CoV) [20]. Therefore, studies of protein structure and function from those coronaviruses have been important for understanding the function and mechanisms of SARS-CoV-2 proteins [21]. In addition, recent structural and functional studies on SARS-CoV-2 proteins have provided critical knowledge of SARS-CoV-2 biology [17].
The SARS-CoV-2 genome is about 30 kb (Zhou et al., 2020), and it encodes at least 29 proteins, including 16 non-structural proteins (NSP), 4 structural proteins:(S), envelope (E), membrane (M), and nucleocapsid (N), and 9 accessory proteins (Figure 1). Some structural and non-structural proteins (S, E, M, N, 3CL protease, papain-like protease, RNA polymerase [22], and helicase) have been suggested as viable antiviral drug and diagnostic targets [23,24,25]. Although coronaviruses accessory proteins are frequently considered as non-critical for viral replication in vitro, some have an important role in virus-host interactions in vivo [26] and in vitro [27]. Different studies have shown that SARS-CoV-2 is a pathogen with a high efficiency in invading the cells of its host [28,29]. Currently, there is enough evidence supporting the preponderant role of angiotensin-converting enzyme 2 (ACE2) as a key factor allowing the binding and entry of SARS-CoV-2 to host cells [30,31,32].

2. Diagnosis

Due to the pandemic nature of COVID-19, testing has become an indispensable tool for monitoring the diagnosing and spread of the disease [33,34]. Multiple approaches and strategies for detecting SARS-CoV-2 have been developed. In this context, a clear understanding of the principles and differences between these strategies, their impact on the results, their correct interpretation, and the best conditions for their adequate use, is crucial [34]. In general, the major approaches for COVID-19 testing are: (1) detection of viral structural components (RNA and proteins) and (2) detection of molecules produced by the host immune system in response to SARS-CoV-2 infection.
RNA is a key structural component of SARS-CoV-2 and, given the previous development of several technologies allowing for the efficient isolation, amplification, and detection of specific nucleic acid sequences, the viral genome (RNA) is the most targeted molecule used for the detection of SARS-CoV-2. (Figure 1) [35,36]. This is the principle of widely used detection methods such as RT-qPCR, RT-LAMP, and next generation sequencing based assays, among others. A second group of methods have been developed targeting SARS-CoV-2 proteins in samples. These methods include viral antigen detection using proteomics-based tests and the nanosensor-based detection of viral proteins.
On the other hand, the concentration of the different types of antibodies (IgM, IgG, IgA) produced by the host immune system in response to SARS-CoV-2 can be used as a tool for surveillance to better understand how much of the population has been infected with SARS-CoV-2, as well as how the virus is spreading through the population over time [37] (Figure 2). The detection of antibodies is the common principle of all serological tests.
It is important to highlight that although serological tests are very useful from an epidemiological perspective, given that antibodies can take several days or even weeks to be produced and may stay in your blood for several weeks or more after recovery, serological tests should not be used to diagnose COVID-19 and are not approved for the diagnosis of acute, active SARS-CoV-2 infection [38,39].

3. Nucleic Acid Amplification-Based Tests (NAAT) for Detection of SARS-CoV-2

Since the SARS-CoV-2 genomic sequence was made public in early January 2020, several primers, probes, and detection strategies have been developed and optimized worldwide (Table 1). Despite the increasing availability of assays, most strategies are based on the same basic principle: the accurate and sensitive identification of SARS-CoV-2 RNA sequences. As is further discussed in this section, the major differences of all those assays are in the method used to measure/detect the presence of viral genetic material.

3.1. Reverse Transcription-Polymerase Chain Reaction (RT-PCR)

RT-PCR is a widely used method in molecular biology and is considered the gold standard for the detection of SARS-CoV-2 [106] and other respiratory viruses [107,108,109]. Early detection of infection caused by SARS-CoV-2 by RT-PCR relies on the efficient detection of viral genomic sequences in clinical samples. Based on the established practices for other respiratory infections such as influenza, the nasopharyngeal swab has been widely adopted as the preferred sampling technique for SARS-CoV-2 along with other techniques such as saliva, oropharyngeal, nasal mid-turbinate, and nasal swabs [110]. However, some studies have also reported using blood, sputum, feces, and urine samples for SARS-CoV-2 detection [111,112].
Most of the described protocols for SARS-CoV-2 detection require the isolation of viral RNA from biological samples as a first step. RNA isolation must be performed in a careful and quick manner to avoid degradation caused by nucleases released from cells and present in the environment [113]. First, the sample is mixed with a lysis buffer to release the genetic material and preserve RNA integrity. The following steps are designed to eliminate DNA, protein, and other cell components in order to isolate the RNA by serial washes [113]. After RNA isolation, the next step for RT-PCR is the conversion of viral genome (RNA) into DNA. This process requires the specific binding of DNA primers to viral RNA sequences and the activity of an RNA-dependent DNA polymerase (reverse transcriptase) to generate a short complementary DNA copy (cDNA) of the viral RNA. In the real-time RT-PCR, the polymerase amplifies the viral cDNA, and the synthesis is monitored in real-time by using a fluorescent dye or a sequence-specific DNA probe labeled with a fluorescent molecule and a quencher molecule [114]. To date, most molecular diagnostic tests developed have targeted different SARS-CoV-2 genomic regions, including the ORF1b or ORF8 regions and the nucleocapsid (N), spike (S) protein, RNA-dependent RNA polymerase (RdRP), or envelope (E) genes, while the most-used genes employed as targets in diagnostic tests are N gene, ORF1b, ORF1a and the E gene [115].
Since the development of PCR in the 1980s, many assays used for detecting nucleic acid have been created, and a growing number of research and diagnostic laboratories around the world perform PCR routinely. The wide adoption of this method has been, and still is, critical because it allowed for the rapid implementation of RT-qPCR-based diagnostics facilities and increased the SARS-CoV-2 diagnostic capabilities in many regions. In addition, given that specific assays can be designed based on any DNA or RNA sequence, this technology is highly flexible and is relatively simple to apply to the detection of SARS-CoV-2 variants, and of course, for the detection of future infectious microbes. In a fashion similar to other nucleic acid detection methods, RT-qPCR-based detection heavily relies on the quantity, quality, and integrity of DNA or RNA in samples, therefore the development of simpler, cheaper, and more efficient sampling alternatives is a challenge for improving SARS-CoV-2 RT-qPCR assays.

3.2. Assays Based on Nucleic Acid Isothermal Amplification

Introduced in the early ‘90s, isothermal amplification offers an alternative to Polymerase Chain Reaction (PCR)-based detection methods [116,117]. In contrast to PCR, isothermal amplification can be performed without the need of thermocycling equipment. Instead, it can be carried at one reaction temperature and under relatively simple conditions (e.g., in a water bath) [118]. More than ten types of isothermal amplification methods have been developed [119]. Some are based on DNA replication, enzyme-free nucleic acid assembly, and the use of enzymes for digestion [118]. In this review, we focus on six different techniques, and, where specified, examples are given for its application in COVID-19 diagnostics.
Isothermal amplification detection tests have become popular because they can be performed with simpler laboratory equipment and produce results in shorter times than conventional RT-qPCR assays. Despite this, significant challenges should be addressed to fully exploit this testing technology as a massive and frequent SARS-CoV-2 (and other infectious agents) detection tool, including: (i) reducing the number of micropipetting steps needed and (ii) avoiding the need of cold chain conditions to transport or store reagents.

3.3. Nucleic Acid Sequence-Based Amplification (NASBA)

Self-sustained sequence replication, better known as Nucleic Acid Sequence-Based Amplification (NASBA) was originally designed for RNA amplification [120]. In this technique, RNA hybridizes to the forward primer for the synthesis of its complementary chain to obtain cDNA using RNAase H. The newly synthesized cDNA contains a promoter region for the formation of an intermediate double-stranded (ds) cDNA, which is formed by using a second primer [121]. Several copies of antisense RNA strands to the target RNA are produced when T7 DNA-dependent RNA polymerase carries out transcription from the dsDNA. Then, the antisense RNA and cDNA become templates for the continuous cycle carried out by reverse transcriptase. The reaction can happen at 41 °C and it usually takes 1.5 to 2 h to complete. Colorimetric assays, gel electrophoresis, and real-time fluorescence can be used for amplicon detection [118]. This method has been used in combination with next generation sequencing (NGS) for the detection of SARS-CoV-2 [121].

3.4. Loop-Mediated Isothermal Amplification (LAMP)

Loop-Mediated Isothermal Amplification or LAMP is a low-cost diagnostic technique which relies on the activity of a strand-displacement DNA polymerase and four to six primers to exponentially amplify the target sequence [122]. This method amplifies nucleic acid under isothermal conditions at about 65 °C [123]. The forward inner primer (FIP) and the backward inner primer (BIP) are crucial to the amplification. They both contain two functional sequences: the first one for priming at an early stage and the second for self-priming at a later stage. The functional sequences described for FIP and BIP correspond to the target sequence of the target DNA. The result of the amplification is stem-loop DNA that can be detected by real-time assays [118].
As a diagnostic method, LAMP has several advantages, such as avoiding the need of a thermal cycler, making LAMP simpler and cheaper [124]. Moreover, due to the use of several primers, this technique has a high specificity, allowing it to discriminate a single nucleotide difference. Besides, the amplification is highly efficient (DNA can be amplified 109–1010 times in 15–60 min) [125]. In addition, LAMP is fast due to the activity of the strand-displacement polymerase eliminating the need for a denaturation step, allowing the amplification and detection to occur in a single step. Additionally, in the case of low-resource settings, the amplification can be monitored by using a cheap turbid meter and, after, the addition of an intercalating agent [126]. LAMP has been used previously as a tool to diagnose diseases such as malaria [122] and tuberculosis [127].
In the context of COVID-19, RT-LAMP is the technology used by several assays of different companies (Table 1) worldwide. In addition, tests based on RT-LAMP have been approved by the FDA, CDC, and other regulatory agencies in many countries [128]. RT-LAMP has proven to be effective in detecting viral RNA, showing high specificity and sensitivity, thus being a potentially robust technique during the global COVID-19 pandemic [129].
In addition to its technical advantages for diagnostics, LAMP is a versatile method that has been coupled to other strategies to improve diagnostics. A clear example of this is STOP (SHERLOCK Testing in One Pot), a CRISPR -based method for the diagnosis of SARS-CoV-2 [130]. This method relies on the amplification performed by LAMP with further detection using CRISPR-Cas12 [130]. This is a very simple test that can be carried out at room temperature and, as the name implies, in a single pot.

3.5. Rolling Circle Amplification (RCA)

RCA is a simple and efficient amplification method that relies on the activity of highly processive DNA or RNA strand-displacing polymerases to generate a long single-stranded DNA (ssDNA) or RNA molecule [131]. RCA has attracted attention for diagnostics because of its simplicity and capability of being performed at a constant low temperature (room temperature to 37 °C) in a relatively short time (approximately 90 min). Importantly, due to its molecular principle, RCA avoids the generation of false-positive results, a common problem of other detection assays. This method is based the use of a circular probe (C-probe) formed by the interaction and subsequent ligation of the analyzed sequence with the 5′ and 3′ ends of a single-stranded DNA probe [132]. This C-probe hybridizes with the primer, which is elongated by the strand-displacement polymerase, resulting in the exponential amplification of the target DNA or RNA. It must be mentioned that DNA ligase catalyzes the ligation of the C-probe with the target only in the case of perfect complementarity of the 3′-and 5′-end sequences [133] avoiding false-positives [134].
RCA has been used before for detecting single nucleotide polymorphisms [135,136], M. tuberculosis genomic DNA [137], SARS-CoV [138], Ebola, and other pathogens [139]. Recently, a study described a modified RCA assay capable of detecting up to picomolar concentrations of synthetic viral DNA strands of SARS-CoV-2, Influenza A (H1N1pdm09), and Influenza B [140]. As in other nucleic acid amplification methods, RCA output can be detected by measuring fluorescence, by using colorimetric chemistry, or by coupling the results with other readout techniques [139].

3.6. Transcription-Mediated Amplification (TMA)

TMA is a technology based on isothermal amplification, that can be performed in a single tube [141] to detect target RNA sequences. This technology is more efficient than RT-PCR [141], as it can produce 10 billion amplicons in 15–30 min [142] and it does not require a thermal cycler [142,143], making TMA a more affordable alternative.
TMA employs two primers: one of them contains the promoter sequence for RNA-polymerase. The first step of the process consists of the hybridization of the primer–promoter to the target RNA. Then, a reverse transcriptase creates a copy of DNA based on the RNA target by extension. Later, an RNAse H degrades the RNA in the resulting RNA–DNA duplex. The second primer binds to the DNA copy, and the reverse transcriptase creates a new strand of DNA, resulting in a double-stranded DNA molecule. Then, an RNA polymerase initiates the transcription, since it recognizes the promoter sequence in DNA [142]. The detection is conducted by the HPA separation/detection procedure. First, the amplicon is hybridized with acridinium ester-labeled DNA probes. Then, a chemical reaction separates the hybridized probes from the unhybridized probes. The final step consists of the addition of the reagents into the reaction tubes to produce the chemiluminescent signal, which is measured by a luminometer [142].
TMA is an interesting alternative that is being used to diagnose COVID-19 [144] with a sensitivity of 98.15%, compared to RT-qPCR (96.25%) with a detection limit of 5.5 × 102 copies in 1 of 5 samples [145]. These results included 116 nasopharyngeal swabs [144,145]. The Aptima SARS-CoV-2 assay consists of target capture, TMA, and Dual Kinetic Assay (DKA). In the first step, the target RNA is isolated by capturing oligomers which have a complementary sequence to the target and a string of deoxyadenosine residues. Then, the oligomers:target complex is then captured out of the solution by decreasing the temperature of the reaction to room temperature. The deoxyadenosine region covalently attaches to magnetic microparticles which contain poly-deoxythymidine molecules. The microparticles are pulled to the side of the reaction tube by magnets. Then, the supernatant is aspirated, and the particles are washed. After the target is captured, the sample passes to TMA. This assay amplifies and detects two conserved regions of the ORF1ab gene in a single reaction [146]. The detection step is mediated by DKA detection technology, which is based on HPA. However, DKA uses two different acridinium ester molecules in two different nucleic acid probes [142].

3.7. Recombinase Polymerase Amplification (RPA)

Another polymerase-based amplification method is the recombinase polymerase amplification (RPA). This technology is an isothermal amplification method developed by Piepenburg et al., in 2006 by using proteins involved in recombination, synthesis, and DNA repair [147]. The mechanism in which this method works starts with a recombinase that binds to oligonucleotides in the presence of ATP and high molecular polyethylene glycol (also referred as crowding agent). This mixture forms a recombinase–primer complex which screens the template DNA until a homologous sequence is found. Then, a primer stop occurs so that the primer interacts with the template while single-stranded binding proteins stabilize the displaced DNA strand in order to prevent it from ejecting the complex from the target DNA. Finally, the recombinase breaks away and a polymerase binds to the 3′ end of the primer and starts to synthesize a new strand of DNA in the presence of dNTPs [148]. If this process is repeated in several cycles, linear (with one primer) and exponential (with a forward and a reverse primer) amplification can be achieved [147,149].
Certain aspects must be taken into account in the design/usage of RPA components. Primers used in RPA were thought to be 30–35 bases in length, however some studies have reported that PCR primers (≈19–21 mers) can be used, and that successful amplification can be achieved [143,150,151]. Additionally, 5-end guanines should be avoided while 3-end cytidines and guanines improve amplification performance, while GC content should be between 30% and 70%. Although RPA can work at a range of temperatures from 22–45 °C, optimal temperatures tend to be in the range of 37–42 °C [152,153], and multiple equipment can be employed to regulate the temperature such as incubators, heating blocks, or even body heat, facilitating its use as a point-of-care test in the field [154]. RPA can successfully amplify dsDNA, ssDNA, cDNA, and methylated DNA from a large variety of organisms such as viruses, bacteria, fungi, animals, and plants. Moreover, RPA has proven to be effective in different sample types such as body fluids and animal products [148].
Two main advantages of RPA over PCR have been pointed out by various studies: reaction temperature and reaction time. The latter is notorious if we compare the 15–40 min run time of RPA with the usual run times that PCR takes to show results [149]. Additionally, several detection methods have been developed to visualize results in RPA, such as solid phase RPA combined with electrochemical, ring resonators, and fluorescent and colorimetric detection [148,155,156].
Although RPA has been studied and used for the detection of various viral and bacterial pathogens, it appears to be confined to research-only activities for now [132]. Ebola detection by RPA studies has suggested that the diagnosis with recombinant polymerase amplification can be deployed in coming Ebola outbreaks due to the low amount of equipment required by this technique: a heating block and a centrifuge [157]. Some studies have used RPA as a step in CRISPR diagnostic systems, while RPA as the main detection technique has also been tested for COVID-19 [158,159].

4. Next Generation Sequencing (NGS) for Detection of SARS-CoV-2

First introduced as Massively Parallel Signature Sequencing (MPSS), next generation sequencing has been in the market since around 2004 [160]. The COVIDSeq Test is the first NGS test approved by the US Food and Drug Administration for COVID-19 diagnostics [161]. This test is designed for RNA detection of SARS-CoV-2 in nasopharyngeal, oropharyngeal, and mid-turbinate nasal swabs with a capacity of up to 3072 samples to be processed in 12 h [102].
Although NGS has proven to be a useful detection method, as in other diagnostic methods, as in the case of other diagnostic strategies, the test should not be used as a diagnostic strategy by itself, as results from the NGS test need to be confirmed with features such as clinical observation, patient history, and epidemiological information [102]. A critical disadvantage of NGS limiting its broad application for COVID diagnostics is the high cost of reagents and materials needed and the requirement of complex laboratory equipment [162,163].

5. CRISPR-Based Tests

Since the early days of the pandemic, the high transmission rate of SARS-CoV-2, as well as the high frequency of asymptomatic, but still infectious, cases, took the world by surprise. Even more than a year after the first cases, the lack of a high-throughput, high sensitivity, simple and low-cost diagnostic test still limits the massive detection of the positive individuals, inhibiting not only early medical intervention but also allowing for further disease transmission [164].
Rapid and reliable diagnostic challenges can be tackled by new-generation Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR)-based platforms that offer fast pathogen-tailored detection assays, as well as the possibility to simultaneously detect hundreds of pathogens per individual in a robust multiplex fashion [165] as a point-of-care molecular diagnostic test [166].
Clustered Regularly Interspaced Short Palindromic Repeats are a group of DNA sequences found in bacteria and archaea which serve as an immune system or an immune “hard drive” [167]. Although this immune hard drive was first discovered more than three decades ago [168], it was not until 2012 that their potential as genetic editing scalpels for prokaryotic, eukaryotic, and even mammalian cells was appreciated [169,170]. Therefore, it comes as no surprise that, after 2012, a group of academic laboratories working in collaboration with pharma counterparts leverage CRISPR-Cas systems to develop the DNA Endonuclease-Targeted CRISPR Trans Reporter (DETECTR) [171] and Specific High Sensitivity Enzymatic Reporter UnLOCKing (SHERLOCK) [172] developed the first pathogen diagnostic platforms based on this novel technology.
The DETECTR system uses “off-the-shelf” reagents, portable heat blocks, and a lateral flow readout, similar to pregnancy tests, which can achieve a point-of-care setup that can detect HPV16 and HPV18 in patient samples within 1 h [171]. The reported sensitivity of DETECTR assay is 95%, with a specificity close to 100% [171,173]. Due to the simple principle of targeting Cas proteins, CRISPR-Cas detection assays can be easily programmed to detect practically any pathogen by modifying the guide DNA or RNA that recognizes and directs the Cas enzyme to a specific sequence of the pathogen’s genetic material, including coronaviruses [171,174].
In a recent study, a research group from UCSF reported the design and validation of a SARS-CoV-2 DETECTR diagnostic assay [175]. As in other nucleic acid-based diagnostics, the assay’s initial steps are comprised of nasopharyngeal or oropharyngeal swab sampling and the ensuring of RNA extraction and a one-step reverse transcription and loop-amplification of viral genetic material (RT–LAMP) with primers that target the E and N genes of SARS-CoV-2 [141]. Then, the CRISPR step involves incubation, with Cas12-guide RNAs targeting sequences found in three SARS-like coronaviruses (SARS-CoV-2, bat-SL-CoVZC45, and SARS-CoV), and the detection of target sequences inducing a conformational change and cleavage by lbCas12a enzyme [175]. Activated lbCas12a starts cleaving nearby single-stranded DNA nonspecifically, including the fluorescent-quencher probe, releasing the fluorophore and increasing the fluorescence in the final readout [176]. The limit of detection of this method is reported to be around 10 copies per μL of reaction [175]. According to the authors, the assay showed no cross-reaction with other coronaviruses or influenza strains and showed a lower sensitivity than RT-qPCR; nine out of eleven RT-qPCR positive confirmed samples were positive using this assay [177,178]. Furthermore, the authors carried out an additional step, running 30 RT-qPCR positive and 30 negative samples, and demonstrated a positive predictive value of 95% and a negative predictive value of 100% for the SARS-CoV-2 DETECTR assay [175].
An alternative CRISPR-based detection method named All-In-One-Dual CRISPR-Cas12a (AIOD-CRISPR) was recently developed. The method uses a dual guide RNA approach to augment specificality and in an isothermic, one-pot amplification; the CRISPR readout reaction enables its use in a point-of-care setup, delivering results in 40, 20, or 1 min depending on the target and conditions [179]. As a proof of concept, HIV and SARS-CoV-2 RNA derived from plasmids were detected at a load of 1.2 and 4.6 copies/µL, respectively [180]. The method was further evaluated by detecting HIV RNA extracted from human plasma samples, exhibiting a performance comparable to RT-qPCR tests [180]. A study has used AIOD-CRISPR to test clinical swabs and the results showed to be consistent with the RT-qPCR methods approved by the CDC [179,181].
In 2017, a group of scientists from MIT and Broad Institute reported the development of Specific High Sensitivity Enzymatic Reporter UnLOCKing (SHERLOCK) [182]. This CRISPR-based technique for the detection of pathogens uses a similar approach as the DETECTR, while harnessing Cas13a and Cas13b (instead of Cas12a) enzymatic activity to produce a detectable fluorescence as a readout [182]. Cas13 orthologs recognize single-strand RNA sequences and exhibit a promiscuous RNase activity detected by including a fluorochrome-quencher probe. This method has been previously used to detect specific strains of Zika and Dengue virus to distinguish pathogenic bacteria, genotype human DNA, and to identify cell-free tumor DNA mutations [172].
Recently, Zhang and colleagues have presented a detailed SHERLOCK-based protocol for detecting SARS-CoV-2, starting with RNA purified from patient samples [158]. According to the presented data, the test can detect SARS-CoV-2 target sequences at a concentration of 10–100 copies per microliter of input within 60 min. In addition, once RNA has been extracted, the test can be performed using a dipstick without requiring elaborate instrumentation [158]. Currently, there is a SARS-CoV-2 SHERLOCK detection kit, as indicated in Table 1. Additionally, a high-throughput platform called CARMEN-Cas13 was developed by Pardis Sabeti’s lab. CARMEN-Cas13 simultaneously differentiates all 169 human-associated viruses, with at least 10 published genome sequences including SARS-CoV-2 [183]. Using a similar approach, the same lab developed a CRISPR-based surveillance tool for the detection of 67 viral species and subspecies, including SARS-CoV-2, phylogenetically related viruses, and viruses with similar clinical presentation [184]. This protocol has been tested with clinical samples [185].
Another CRISPR-based SARS-CoV-2 detection method worth mentioning is Cas13-based, Rugged, Equitable, Scalable Testing (CREST). CREST was designed to address three of the main hurdles—reagent accessibility, equipment availability, and cost—that limit the scalability of Cas13-based testing particularly in low-resource areas [186]. The rationale of CREST is using affordable reagents and equipment for RNA extraction, amplification, and detection. Final output is displayed in low-cost and easy-to-use fluorescent visualizers. According to the authors, the sensitivity of CREST is equivalent to RT-qPCR. This method has been successfully used as a rapid SARS-CoV-2 surveillance tool in asymptomatic college students [186].
Cas-based detection depends heavily on the quality of the sample and the integrity and quality of the extracted RNA. In addition, due to the involvement of various enzymatic steps, the robustness of the method still has room for improvement. To massively deploy sensitive, reliable, and easy-to-use Cas-based detection assays, it is important to optimize clinical performance and enhance the robustness of the assays. Given the simple principle of nucleic acid detection by Cas-based methods detection, it is expected that the next generations of SARS-CoV-2 tests will show enhanced clinical performance and will be optimized for the detection, not only of the original SARS-CoV-2 sequences, but of novel variants in the population as well.

6. Serology-Based Tests

Since the early days of epidemiology, serological tests have been a very useful tool to assess previous infections and track the community spreading of diseases [187], including SARS-CoV-2 [188]. In this context, serological tests can be useful for (1) estimating epidemiological variables such as the degree of transmission in the community and its burden [189], (2) performing mass testing to reduce the exposure of susceptible individuals to the virus [190], and (3) identifying individuals with a strong immune response to the virus and whose isolated antibodies can, potentially, be used for a convalescent plasma trial [4,191].
Several antibody detection platforms have been developed over the last 30 years, including enzyme-linked immunosorbent (ELISA) assays and lateral-flow assays [192]. Serological tests are designed to detect antibodies produced by the host in response to specific pathogens, in this case SARS-CoV-2, and, ideally, to minimize cross-reactivity to antibodies generated in response to other viruses or bacterial pathogens [193]. However, given the nature of the antigen–antibody interactions and the structural similarities between antigens from related viruses, cross-reactivity may exist.
Although the most common serologic tests available measure antibodies that bind to viral S and N proteins, not all of these antibodies are considered neutralizing/protective, therefore results from these kind of tests should not be used for identifying individuals who are protected against SARS-CoV-2 infection. In general, serological tests have lower sensitivity and specificity than average RT-qPCR-based tests. Negative results should be interpreted with caution, as negative serology does not definitively rule out acute infection [194].
Serologic assays for SARS-CoV-2 are currently being cleared by the U.S. Food and Drug Administration (FDA) for emergency use and are available and include indirect or lateral flow ELISA assays [195].
During the current pandemic, serological tests have been important for detecting the previous infection of SARS-CoV-2 and became popular because some are cheaper and simple to use. Despite this, they should not be used to substitute molecular testing for detecting acute infections. Some of the advantages of this kind of test are its low-price and simplicity to use. To improve SARS-CoV-2 serological testing, the main challenges are improving the sensitivity and specificity of the assays to enhance its clinical performance. Table 2 shows some serological tests developed to detect SARS-CoV-2.

Antigen Testing

Antigen testing is a type of immunoassay designed to detect the presence of one or more specific proteins of the pathogen, allowing the crucial identification of infected individuals (symptomatic or asymptomatic), so that they can receive the appropriate care, follow public health measures, such as self-isolation, and so that contact tracing can be implemented without delay [226].
In the case of COVID-19, The FDA has issued an EUA for the first antigen assay, the Sofia 2 SARS Antigen FIA test, which employs immunofluorescence-based lateral flow technology in a sandwich design used to detect viral nucleocapsid protein in nasopharyngeal and nasal swab samples with test results within 15 min [227].
Antigen tests may not be as accurate as molecular tests, but they are more accessible in terms of availability and ease of use, and can be used to amplify tests without laboratories, including frequent retesting if necessary. Modeling studies have shown that an availability of results within 15–20 min and the frequency of testing are more important than sensitivity [226].
SARS-CoV-2 antigen tests developed during 2020 have comparable clinical performance to RT-qPCR [228], being useful for the detection of symptomatic and asymptomatic individuals [227]. Moreover, since antigen testing is significantly cheaper (5 to 10 times cheaper than the average RT-qPCR test) [229], simpler, and faster to use (approximately 15 min) than other diagnostic tests, this technology has become a powerful tool for reducing the spread of COVID-19, particularly in community settings such as universities, schools, and offices [230]. Moreover, a frequent method of antigen testing has been implemented as a safe way to reopen economies while controlling the spread [226]. In a recent study, Chaimayo et al., demonstrated that SARS-CoV-2 antigen detection (Standard™ Q COVID-19 Ag kit) shows a sensitivity (98.33%; 95% CI, 91.06–99.96%) and specificity (98.73%; 95% CI, 97.06–99.59%) comparable with the real-time RT-PCR assay [228]. Another advantage of antigen testing is that this type of test is the most likely to detect a positive person when at the peak of their infection, where people are most able to transmit the virus to others.
Regarding its use, WHO [8] has suggested the direct detection of SARS-CoV-2 viral proteins by nasal swabs and other respiratory secretions by lateral flow immunoassay, giving results in less than 30 min, as they offer the possibility of rapid, inexpensive, and early detection of the most infectious cases of COVID-19 in suitable settings. At the time of revision of this article, 13 tests have been granted Emergency Use Authorization from the FDA [231] (Table 3).

7. Nanotechnology for COVID-19 Diagnostics

The synthesis of nanoparticles (NPs), the development of nanosensors and microchips, and the use of devices have become essential tools for the diagnosis of various diseases, including COVID-19 [244]. In recent months, an increasing number of studies have shown that nucleic acid sequences can be detected and play an essential role in diagnosis. However, the detection of single nucleotide variations, such as polymorphisms or small nucleotide changes in viral variants, has proven to be challenging [245].
A particularly relevant technology useful for pathogen and antibody detection purposes are gold NP-based immunoassays. In these assays, antibodies are coupled to gold NPs that, in the presence of a viral antigen or antibody, form a tertiary complex, leading to the immobilization and aggregation of the complex [246]. Other examples of the use of nanotechnology are: Biosensors based on graphene field effect transistors (FETs) coupled to a specific antibody against SARS-CoV-2 spike protein [247]; Dual-function plasmonic biosensors combining plasmonic photothermal effect (PPT) and localized surface plasmon resonance to detect Au nanomaterials coupled to complementary DNA sequences to detect SARS-CoV-2 hybridized cDNA [248]; Dendrimer-like DNA nanostructures (DL-DNA) for DNA detection (Li Y) system-on-a-chip (SoC) technology, whose detection mechanism is based on RNA aptamer conjugated to QDs [249] and materials such as silica that have been used to identify bioimaging and biodetection, and pathogen detection [250] (Figure 3).
In addition, gold nanoparticles coated with sulfated ligands, silver nanoparticles, hybrid silver, and copper nanoparticles can bind to the HIV envelope glycoprotein gp120 and inhibit HIV-1 infection in vitro in cell models [251]. Single wall carbon nanotubes (SWCNT) were also proposed as antiviral carriers, testing the increase in antiviral activity of hybrid materials compared to that of antiviral drugs [252]. Even antiviral and antibacterial textiles were based on nanomaterials; however, the commercialization of new functional nanomaterials is still not clear yet, and seems to be limited by nanotoxicology concerns or various other practical aspects (off-target and/or any other unpredictable effects, costs, performance production, durability, environmental impact, etc.). The emergence of new diseases has generated increased development and innovative tools in various areas of biotechnology and nanomedicine. The development of Au nanoparticles has mainly been useful for detecting different coronaviruses and respiratory infections, as presented in Table 4.

8. High-Throughput and Automated Screening for COVID-19

Given that test, track, and trace are crucial to tackle the COVID-19 pandemic, these processes are the core recommendation by CDC, European Centre for Disease Prevention and Control (ECDC), WHO, and other health agencies [262]. The rapid spread of COVID-19 has led to a huge demand for diagnostic screening [263], and to meet this goal while increasing the testing capacity of cities and countries, automated workflows have become key to enable high-throughput screening at an increased speed while improving diagnostic precision and reducing labor costs [264].
Several alliances between academic laboratories and companies such as Analytik Jena, Beckman Coulter, Hamilton, Tecan, Roche, Opentrons, among others, have focused their time and efforts in an impressive way to provide automated testing solutions for COVID-19 diagnosis [265]. These automated solutions were designed to cater to the various alternatives of molecular diagnosis protocols, such as the SHERLOCK [130] and LAMP [266] assays, although the RT-qPCR assay is still considered the “gold standard” for automated for SARS-CoV-2 testing [267].

8.1. Automation Platforms for SARS-CoV-2 Testing

An impressive example of automation support for COVID-19 testing and screening was the Huo-Yan Air Laboratory or “Fire Eye Lab’’ built in Wuhan, where the COVID-19 outbreak started in China. The Huo-Yan Lab uses automated nucleic acid extraction as a part of the RT-qPCR testing workflow, reaching 14.000 testing capacity per day on February 9 in a 2000 m2 site. This testing capacity was further increased to 20.000 per day by March 1 [268]. To achieve this, MGISP-960 platforms produced by the Chinese biotech company, MGI, were used to automate RNA extraction using the MGIEasy Magnetic Beads Virus DNA/RNA Extraction Kit to achieve high-throughput and standardized clinical testing, as shown in Figure 4 [81,269].
Manual RNA extractions using the QIAamp Viral RNA Mini Kit were compared to the platform automation, showing that the manual RNA extraction of 24 samples took on average 1 h and 50 min, whereas the automated extraction could finish processing 192 samples in 1 h and 8 min with a single run. The Ct values between the manual automated extraction resulted in no significant statistical difference, which was necessary to validate the automated platform [269].
In parallel to Huo-Yan Laboratory and MGI automated alternatives, several other biotech companies focusing on automation platforms rapidly adapted their technologies for COVID-19 screening. For instance, Hamilton offered an automated RNA extraction solution with its MagEx STARlet platform, enabling it to process 96 samples simultaneously [270]. Moreover, this platform is compatible with five commercially available magnetic bead-based RNA extraction kits, including MagMAX™ Viral/Pathogen NA Extraction Kit (Thermo Fisher, Waltham, MA, USA), Quick-DNA/RNA™ Viral MagBead Kit (Zymo Research, Irvine, CA, USA), and the Maxwell® HT Viral TNA Kit (Promega, Madison, WI, USA). In addition to RNA extraction, PCR Prep STARlet for automated qPCR mix prep solution is also offered by the company as a compatible automation platform with FDA-approved kits such as TaqPath COVID-19 Combo Kit (Thermo Fisher), 2019-nCoV Kit, (Integrated DNA Technologies, Coralville, IA, USA), and the 2019-nCoV CDC Probe and Primer Kit (BioSearch Technologies, Hoddesdon, UK) [270].
The popular Fluent and Freedom EVO platforms of Tecan were adapted for use in automated RNA extraction and PCR preparation processes [271]. A Freedom EVO platform allowing for a fully automated qPCR reaction set-up works in compatibility with the real-time thermocyclers of Thermo Fisher, Bio-Rad, and Roche [271].
These technologies are currently being improved at a very rapid scale through innovative partnerships such as the collaboration of Zymo Research, focusing on molecular biology tools with Tecan and Opentrons to optimize the magnetic bead-based automated RNA extraction process [272,273,274,275].

8.2. Biofoundries Role against COVID-19

Biofoundries, which are non-profit organizations with state-of-art automation platforms and high-throughput equipment focusing on accelerating and prototyping biological designs for bioengineering applications [276]. Such facilities were uniquely positioned to contribute to the development of novel automated platforms for SARS-CoV-2.
The London Biofoundry configured its existing automated liquid handling infrastructure to COVID-19 screening platforms using three different workflows, encompassing RT-qPCR, CRISPR-Cas, and LAMP detection methods [276]. These three methods were automated and validated using various liquid handling platforms [276]. Figure 5 illustrates the automated workflows for the three different SARS-CoV-2 assays [276].
To test the platform performance, the researchers also tested two different RNA extraction kits from Analytik Jena and Promega and three different qPCR master mixes from ThermoFisher (TaqPath and Fast Virus) and NEB (Luna). Ct values of 103 copies/mL of virus-like particles cwere observed around 30 cycles, while Ct values of 104 copies/mL of VLP were approximately 27 cycles in both RNA extraction kits. For the LAMP workflow, however, it was observed that at least 42.5 copies of VLP were necessary for appropriate detection. The RT-qPCR workflow was validated with 173 patient samples obtained from the North West London Pathology using two different RNA extraction kits, obtaining an adequate R2 correlation of 0.8310 and R2 = 0.9357 for the innuPREP Virus DNA/RNA Kit (Analytik Jena) and Maxwell HT Viral RNA extraction kit (Promega), respectively. This automated platform employing the RT-qPCR workflow became operational in two London hospitals from May 2020 with a 2000 testing capacity per day [276].
There are also commercially available alternatives from several suppliers such as Hologic and Roche. For example, the Cobas® 8800 (Roche, Basel, Switzerland) platform can process 1056 samples in eight hours or 384 samples in 4 h [59]. On the other hand, Panther Fusion from Hologic can run 336 samples simultaneously when 28 cartridges are used at the same time [278]. However, these platforms are also relatively expensive, and they mainly use proprietary reagents and cartridges which are tailored to their own platforms.

8.3. Free and Open Source Scientific and Medical Hardware (FOSH)

Even though the combination of the FeliX and Echo 525 were proven to be effective automation platforms for SARS-CoV-2, their accessibility by many laboratories is limited due to the higher costs of the equipment. To counteract this, the free open-source hardware and software community has played a prominent role for COVID-19 diagnostics, contributing to ease the burden on the global health systems. This has been achieved by the development and sharing of tools freely available for the community for further study, customization, and commercialization at a lower cost [279]. The best example of this for automation is Opentrons’ platform OT-2, which is an affordable and open-source lab automation system allowing to scale-up and automate hundreds of assays and workflows.
The OT-2 has gained popularity in the scientific community day by day, thanks to its flexibility and affordability [280]. While the Felix and Echo 525 cost are in the order of hundreds of thousands of pounds, the customizable OT-2 costs between 5 and 15 thousand, depending on the selected configuration.
Researchers at the Biomedical Diagnostic Center (CBD) of the Hospital Clinic of Barcelona have developed an automated COVID-19 screening circuit named ROBOCOV, using four OT-2 robots from Opentrons, a KingFisher Flex extraction instrument from ThermoFisher, and an ABI 7500 qPCR device [280]. In this circuit, the initial setup, sample preparation, plate filling for RNA extraction, and qPCR mix preparation steps were carried out by four OT-2 robots while RNA extraction was performed by the KingFisher Flex and real-time qPCR process was done by the ABI 7500 qPCR device, as shown in Figure 6.
This study run preparation was done by using open-source Python codes. Initial sample setup, sample preparation, and plate filling are performed by OT-2 Station A, B1, and B2, respectively. RNA extraction is processed by KingFisher Flex [280]. Then, qPCR mic preparation is done by OT-2 Station C, and, finally, the qPCR process is run by ABI 7500 [280]. The analysis results are exported as a user-friendly R file. Due to this parallel and automated molecular testing circuit consisting of six platforms, the researchers tram obtained a high screening capacity completing the testing process of 96 samples in 4 h while starting each new run every 70 min. This set-up was further assessed and validated by the European Molecular Genetics Quality Network, which provides external quality assessment, and the findings revealed a comparable performance and consistent Ct levels compared with other validated automated platforms from Roche and Hamilton. A similar ROBOCOV is currently being used for high-throughput COVID-19 screening as an FDA-approved method at the Biomedical Diagnostic Center of the Hospital Clinic of Barcelona, with a 2400 test capacity per day [281]. The CLIAHUB facility, which intends to aid on COVID-19 testing in the San Francisco Bay Area, have also demonstrated that research institutions can contribute to COVID-19 testing and that the workflow provided by their report can process a vast number of tests by using eight automated liquid handles into the RNA extraction step [282].
The studies mentioned in this section highlight the key role that the automation of molecular diagnosis has had in the scaling and speeding up of COVID-19 testing and screening. It also showcases the need to develop novel and affordable automated diagnostic workflows to be deployed in lower income settings where testing is currently done predominantly manually. The automated platforms developed during the COVID-19 outbreak are very relevant, as they can easily be adapted for diagnostics for many other types of biological threats and infectious diseases, allowing the global community to be better prepared for future pandemics.

9. Conclusions

The COVID-19 pandemic has been and continues to be a threat to economies and health systems worldwide. Currently, no standard treatment for COVID-19 is globally available. Despite the enormous efforts that have allowed the unprecedented development of multiple highly effective vaccines against COVID-19 in record time, the broad access and distribution of vaccines in many countries is still a significant challenge. Therefore, the implementation of efficient and versatile COVID-19 testing policies remains a key strategy to tackle the effects of COVID-19, to its epidemiological features, and to prevent further damage from this virus and from additional SARS-CoV-2 variants.
Given the structure of the SARS-CoV-2 virus, and due to the previous development of technologies for the detection of nucleic acids and proteins, the major diagnostic strategies for COVID-19 are based on the detection of these molecules. The main objective of this review is to present a clear and useful diagnostic landscape from the most common strategies and technologies to novel diagnostic alternatives, describing their principles, as well as highlighting their advantages and disadvantages. It is also intended to review some of the most relevant automation platforms useful for boosting diagnostics capabilities, as well as diagnostic alternatives that are based on the application of nanotechnology principles.

Author Contributions

C.B.-O. and L.P.G.-B. were responsibles for the conceptualization of this work. C.B.-O., E.M.-U. and C.G.O.-A. were in charge of drafting the document in all of the stages. E.M.-U. and C.G.O.-A. compiled the information and organized the Table 1, Table 2, Table 3. M.T.-A. completed the Section 6 (Serology-based tests) and Section 7 (Nanotechnology for COVID-19 Diagnostics). N.C.K. and L.P.G.-B. completed the Section 5 (CRISPR-Based Tests) and the antigen testing section. K.M. and L.R.-S. completed the Section 8 (Automation Platforms for SARS-CoV-2 Testing). C.B.-O. and N.J.-B. completed the Section 3 and Section 4. A.L.-C. and E.R.-M. completed the Section 1 (Introduction) and Section 2 (Diagnosis). A.L.-C. and C.B.-O. completed all the figures for this work. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by The Royal Academy of Engineering, Pandemic Preparedness Phase 1 grant EXPP2021\1\207, The University of Edinburgh SFC-GCRF Covid-19 Fund, COV_27 and the PAPIIT project, UNAM number: IN216720.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare having no competing or financial interests or personal relationships related to this manuscript.

References

  1. Zhu, N.; Zhang, D.; Wang, W.; Li, X.; Yang, B.; Song, J.; Zhao, X.; Huang, B.; Shi, W.; Lu, R.; et al. A Novel Coronavirus from Patients with Pneumonia in China, 2019. N. Engl. J. Med. 2020, 382, 727–733. [Google Scholar] [CrossRef] [PubMed]
  2. Cui, J.; Li, F.; Shi, Z.-L. Origin and Evolution of Pathogenic Coronaviruses. Nat. Rev. Microbiol. 2019, 17, 181–192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Weiss, S.R.; Navas-Martin, S. Coronavirus Pathogenesis and the Emerging Pathogen Severe Acute Respiratory Syndrome Coronavirus. Microbiol. Mol. Biol. Rev. 2005, 69, 635–664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Amanat, F.; Krammer, F. SARS-CoV-2 Vaccines: Status Report. Immunity 2020, 52, 583–589. [Google Scholar] [CrossRef]
  5. World Health Organization. WHO Consensus Document on Theepidemiology of Severe Acuterespiratory Syndrome (SARS); Department of Communicable Disease Surveillance and Response: Geneva, Switzerland, 2003. [Google Scholar]
  6. CDC Fact Sheet: Basic Information about SARS. 2004. Available online: https://www.cdc.gov/sars/about/fs-SARS.pdf (accessed on 1 October 2020).
  7. De Wit, E.; van Doremalen, N.; Falzarano, D.; Munster, V.J. SARS and MERS: Recent Insights into Emerging Coronaviruses. Nat. Rev. Microbiol. 2016, 14, 523–534. [Google Scholar] [CrossRef]
  8. WHO. Coronavirus Disease (COVID-19) Dashboard. Available online: https://covid19.who.int (accessed on 7 November 2020).
  9. Baud, D.; Qi, X.; Nielsen-Saines, K.; Musso, D.; Pomar, L.; Favre, G. Real Estimates of Mortality Following COVID-19 Infection. Lancet Infect. Dis. 2020, 20, 773. [Google Scholar] [CrossRef] [Green Version]
  10. Wilson, N.; Kvalsvig, A.; Barnard, L.T.; Baker, M.G. Case-Fatality Risk Estimates for COVID-19 Calculated by Using a Lag Time for Fatality. Emerging Infect. Dis. 2020, 26, 1339–1441. [Google Scholar] [CrossRef]
  11. Rajgor, D.D.; Lee, M.H.; Archuleta, S.; Bagdasarian, N.; Quek, S.C. The Many Estimates of the COVID-19 Case Fatality Rate. Lancet Infect. Dis. 2020, 20, 776–777. [Google Scholar] [CrossRef] [Green Version]
  12. Kim, D.D.; Goel, A. Estimating Case Fatality Rates of COVID-19. Lancet Infect. Dis. 2020, 20, 773–774. [Google Scholar] [CrossRef]
  13. Spychalski, P.; Błażyńska-Spychalska, A.; Kobiela, J. Estimating Case Fatality Rates of COVID-19. Lancet Infect. Dis. 2020, 20, 774–775. [Google Scholar] [CrossRef]
  14. Lipsitch, M. Estimating Case Fatality Rates of COVID-19. Lancet Infect. Dis. 2020, 20, 775. [Google Scholar] [CrossRef]
  15. Un Describing COVID-19 Pandemic as Wake-Up Call, Dress Rehearsal for Future Challenges, Secretary-General Opens Annual General Assembly Debate with Vision for Solidarity. Available online: https://www.un.org/press/en/2020/ga12268.doc.htm (accessed on 1 November 2020).
  16. Chan, J.F.-W.; Kok, K.-H.; Zhu, Z.; Chu, H.; To, K.K.-W.; Yuan, S.; Yuen, K.-Y. Genomic Characterization of the 2019 Novel Human-Pathogenic Coronavirus Isolated from a Patient with Atypical Pneumonia after Visiting Wuhan. Emerg. Microbes Infect. 2020, 9, 221–236. [Google Scholar] [CrossRef] [Green Version]
  17. V’kovski, P.; Kratzel, A.; Steiner, S.; Stalder, H.; Thiel, V. Coronavirus Biology and Replication: Implications for SARS-CoV-2. Nat. Rev. Microbiol. 2020, 19, 155–170. [Google Scholar] [CrossRef]
  18. Wang, C.; Liu, Z.; Chen, Z.; Huang, X.; Xu, M.; He, T.; Zhang, Z. The Establishment of Reference Sequence for SARS-CoV-2 and Variation Analysis. J. Med. Virol. 2020, 92, 667–674. [Google Scholar] [CrossRef]
  19. Andersen, K.G.; Rambaut, A.; Lipkin, W.I.; Holmes, E.C.; Garry, R.F. The Proximal Origin of SARS-CoV-2. Nat. Med. 2020, 26, 450–452. [Google Scholar] [CrossRef] [Green Version]
  20. Yoshimoto, F.K. The Proteins of Severe Acute Respiratory Syndrome Coronavirus-2 (SARS CoV-2 or n-COV19), the Cause of COVID-19. Protein J. 2020, 39, 198–216. [Google Scholar] [CrossRef] [PubMed]
  21. Nakagawa, S.; Miyazawa, T. Genome Evolution of SARS-CoV-2 and Its Virological Characteristics. Inflamm. Regen. 2020, 40, 17. [Google Scholar] [CrossRef] [PubMed]
  22. Calligari, P.; Bobone, S.; Ricci, G.; Bocedi, A. Molecular Investigation of SARS-CoV-2 Proteins and Their Interactions with Antiviral Drugs. Viruses 2020, 12, 445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Prajapat, M.; Sarma, P.; Shekhar, N.; Avti, P.; Sinha, S.; Kaur, H.; Kumar, S.; Bhattacharyya, A.; Kumar, H.; Bansal, S.; et al. Drug Targets for Corona Virus: A Systematic Review. Indian J. Pharmacol. 2020, 52, 56–65. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, X.-H.; Zhang, X.; Lu, Z.-H.; Zhu, Y.-S.; Wang, T. Potential Molecular Targets of Nonstructural Proteins for the Development of Antiviral Drugs against SARS-CoV-2 Infection. Biomed. Pharmacother. 2020, 133, 111035. [Google Scholar] [CrossRef] [PubMed]
  25. Wondmkun, Y.T.; Mohammed, O.A. A Review on Novel Drug Targets and Future Directions for COVID-19 Treatment. Biologics 2020, 14, 77–82. [Google Scholar] [CrossRef] [PubMed]
  26. Liu, D.X.; Fung, T.S.; Chong, K.K.-L.; Shukla, A.; Hilgenfeld, R. Accessory Proteins of SARS-CoV and Other Coronaviruses. Antiviral Res. 2014, 109, 97–109. [Google Scholar] [CrossRef] [PubMed]
  27. Ren, Y.; Shu, T.; Wu, D.; Mu, J.; Wang, C.; Huang, M.; Han, Y.; Zhang, X.-Y.; Zhou, W.; Qiu, Y.; et al. The ORF3a Protein of SARS-CoV-2 Induces Apoptosis in Cells. Cell. Mol. Immunol. 2020, 17, 881–883. [Google Scholar] [CrossRef] [PubMed]
  28. Hu, B.; Guo, H.; Zhou, P.; Shi, Z.-L. Characteristics of SARS-CoV-2 and COVID-19. Nat. Rev. Microbiol. 2020, 19, 141–154. [Google Scholar] [CrossRef] [PubMed]
  29. Ortiz-Prado, E.; Simbaña-Rivera, K.; Gomez-Barreno, L.; Rubio-Neira, M.; Guaman, L.P.; Kyriakidis, N.; Muslin, C.; Gomez-Jaramillo, A.M.; Barba, C.; Cevallos, D.; et al. Clinical, Molecular and Epidemiological Characterization of the SARS-CoV2 Virus and the Coronavirus Disease 2019 (COVID-19): A Comprehensive Literature Review. Diagn. Microbiol. Infect. Dis. 2020. [Google Scholar] [CrossRef] [PubMed]
  30. Rodrigues, R.; Costa de Oliveira, S. The Impact of Angiotensin-Converting Enzyme 2 (ACE2) Expression Levels in Patients with Comorbidities on COVID-19 Severity: A Comprehensive Review. Microorganisms 2021, 9, 1692. [Google Scholar] [CrossRef] [PubMed]
  31. Ni, W.; Yang, X.; Yang, D.; Bao, J.; Li, R.; Xiao, Y.; Hou, C.; Wang, H.; Liu, J.; Yang, D.; et al. Role of Angiotensin-Converting Enzyme 2 (ACE2) in COVID-19. Crit. Care 2020, 24, 422. [Google Scholar] [CrossRef] [PubMed]
  32. Bian, J.; Li, Z. Angiotensin-Converting Enzyme 2 (ACE2): SARS-CoV-2 Receptor and RAS Modulator. Acta Pharm. Sin. B 2020, 11, 1–12. [Google Scholar] [CrossRef] [PubMed]
  33. Seidu, A.-A.; Hagan, J.E.; Ameyaw, E.K.; Ahinkorah, B.O.; Schack, T. The Role of Testing in the Fight against COVID-19: Current Happenings in Africa and the Way Forward. Int. J. Infect. Dis. 2020, 98, 237–240. [Google Scholar] [CrossRef] [PubMed]
  34. Vandenberg, O.; Martiny, D.; Rochas, O.; van Belkum, A.; Kozlakidis, Z. Considerations for Diagnostic COVID-19 Tests. Nat. Rev. Microbiol. 2020, 19, 171–183. [Google Scholar] [CrossRef] [PubMed]
  35. Sethuraman, N.; Jeremiah, S.S.; Ryo, A. Interpreting Diagnostic Tests for SARS-CoV-2. JAMA 2020, 323, 2249–2251. [Google Scholar] [CrossRef]
  36. Nalla, A.K.; Casto, A.M.; Huang, M.-L.W.; Perchetti, G.A.; Sampoleo, R.; Shrestha, L.; Wei, Y.; Zhu, H.; Jerome, K.R.; Greninger, A.L. Comparative Performance of SARS-CoV-2 Detection Assays Using Seven Different Primer-Probe Sets and One Assay Kit. J. Clin. Microbiol. 2020, 58, e00557-20. [Google Scholar] [CrossRef] [Green Version]
  37. Xiang, F.; Wang, X.; He, X.; Peng, Z.; Yang, B.; Zhang, J.; Zhou, Q.; Ye, H.; Ma, Y.; Li, H.; et al. Antibody Detection and Dynamic Characteristics in Patients with COVID-19. Clin. Infect. Dis. 2020, 71, 1930–1934. [Google Scholar] [CrossRef]
  38. CDC Using Antibody Tests for COVID-19. Available online: https://www.cdc.gov/coronavirus/2019-ncov/lab/resources/antibody-tests.html (accessed on 6 April 2021).
  39. FDA Antibody (Serology) Testing for COVID-19: Information for Patients and Consumers. Available online: https://www.fda.gov/medical-devices/coronavirus-covid-19-and-medical-devices/antibody-serology-testing-covid-19-information-patients-and-consumers (accessed on 6 April 2021).
  40. 1drop, 1drop 1copy™ COVID-19 QPCR Multi Kit. Available online: https://www.fda.gov/media/137935/download (accessed on 20 August 2020).
  41. TRUPCR, TRUPCR® SARS-CoV-2 KIT. Available online: https://www.fda.gov/media/139296/download (accessed on 23 August 2020).
  42. 3D Med ANDiS SARS-CoV-2 and Influenza A/BRT-QPCR Detection Kit. Available online: http://www.3dmedcare.com/UploadImage/covid/04%20DA%20for%20ANDiS%20SARS-CoV-2%20and%20Influenza%20A&B%20RT-qPCR%20Detection%20Kit.pdf (accessed on 24 August 2020).
  43. Abbott. ABBOTT REALTIME SARS-COV-2 ASSAY. Available online: https://www.molecular.abbott/int/en/products/infectious-disease/RealTime-SARS-CoV-2-Assay#order (accessed on 23 August 2020).
  44. CareStart™ COVID-19 Antigen. Available online: https://carestartantigen.com (accessed on 1 April 2021).
  45. Acupath Laboratories Acupath COVID-19 RT-PCR Assay EUA Summary. Available online: https://www.fda.gov/media/139672/download (accessed on 24 August 2020).
  46. ABL ULTRAGENE COMBO2SCREEN SARS-COV-2 ASSAY. Available online: https://www.ablsa.com/laboratory-applications/ultragene-combo2screen (accessed on 24 August 2020).
  47. Altona Diagnostics RealStar®SARS-CoV-2 RT-PCR Kit 1.0. Available online: https://altona-diagnostics.com/files/public/Content%20Homepage/-%2002%20RealStar/MAN%20-%20CE%20-%20EN/RealStar%20SARS-CoV-2%20RT-PCR%20Kit%201.0_WEB_CE_EN-S03.pdf (accessed on 24 August 2020).
  48. Altru Diagnostics Thermo Fisher TaqMan 2019-NCoV Assay Kit v1 (Singleplex). Available online:https://www.fda.gov/media/137546/download (accessed on 23 August 2020).
  49. Anatolia geneworks Bosphore Novel Coronavirus (2019-NCoV) Detection Kit. Available online:http://www.anatoliageneworks.com/en/kitler.asp?id=360&baslik=Bosphore%20Novel%20Coronavirus%20(2019-nCoV)%20Detection%20Kit&bas=Bosphore%20Novel%20Coronavirus%20(2019-nCoV)%20Detection%20Kit (accessed on 24 August 2020).
  50. Applied BioCode BioCode® SARS-CoV-2 Assay. Available online: https://www.fda.gov/media/139049/download (accessed on 24 August 2020).
  51. LineaCOVID-19 LineaTM COVID-19Assay Kit. Available online: https://www.fda.gov/media/138059/download (accessed on 24 August 2020).
  52. Aspirus Laboratory ASPIRUS SARS-COV-2 RRT-PCR ASSAY. Available online: https://www.fda.gov/media/138526/download (accessed on 24 August 2020).
  53. Assurance Scientific Laboratories ASSURANCE SARS-COV-2 PANEL. Available online: https://www.fda.gov/media/138154/download (accessed on 24 August 2020).
  54. Accelarate Technologies A*STAR FORTITUDE KIT 2.0COVID-19 Real-Time RT-PCR Test. Available online: https://www.biovendor.com/file/13373/AW00032-02%20IFU%20ASTAR%20(MiRXES)Fortitude%202.0Watermarked.pdf (accessed on 23 August 2020).
  55. Avera Institute for Human Genetics AVERA INSTITUTE for HUMAN GENETICS SARS-CoV-2 RT-PCR. Available online: https://www.fda.gov/media/138332/download (accessed on 24 August 2020).
  56. Avellino Avellino SARS-CoV-2/COVID-19 Test. Available online: https://avellinocoronatest.com/product/ (accessed on 24 August 2020).
  57. Bag Diagnostics with Secured Results against COVID-19. Available online: https://www.bag-diagnostics.com/files/downloads/Produkt_Flyer/ViroQ_SARS-CoV-2_Broschuere.pdf (accessed on 23 August 2020).
  58. ThermoFisher Scientific Applied Biosystems™ TaqPath™ COVID-19 CE-IVD RT-PCR Kit. Available online: https://aslm.org/wp-content/uploads/2020/05/ThermoFisher_TaqPathCOVID19_CE-IVD_KIT.pdf (accessed on 24 August 2020).
  59. Roche Cobas® 8800 System. Available online: https://diagnostics.roche.com/gb/en/products/instruments/cobas-8800.html (accessed on 8 April 2021).
  60. Eurofins GSD NovaPrime® SARS-CoV-2 (COVID-19) RT-PCR (96 Reactions). Available online: https://www.eurofins-technologies.com/gsd-novaprimer-sars-cov-2-covid-19-real-time-pcr.html (accessed on 24 August 2020).
  61. QIAGEN QIAstat-Dx® Respiratory SARS-CoV2 Panel Instructions ForUse (Handbook). Available online: https://www.fda.gov/media/136571/download (accessed on 24 August 2020).
  62. R-Biopharm RIDA® GENE SARS-CoV-2. Available online: https://clinical.r-biopharm.com/products/ridagene-sars-cov-2 (accessed on 24 August 2020).
  63. SD Biosensor STANDARD M NCoV Real-Time Detection Kit. Available online: http://sdbiosensor.com/xe/product/7653 (accessed on 24 August 2020).
  64. TBG Biotechnology Corp ExProbe SARS-CoV-2 Testing Kit. Available online: https://www.fda.gov/media/138819/download (accessed on 24 August 2020).
  65. Zymo Research Quick SARs-Cov-2 RRT-PCR Kit. Available online: https://www.fda.gov/media/137780/download (accessed on 23 August 2020).
  66. Fulgent Therapeutics COVID-19 by RT-PCR TEST. Available online: https://www.fda.gov/media/138150/download (accessed on 24 August 2020).
  67. CENTOGENE SARS-CoV-2 Detection Based on E and RdRp Genes. Available online: https://www.fda.gov/media/139725/download (accessed on 24 August 2020).
  68. GeneXpert Xpert® Xpress SARS-CoV-2. Available online: https://www.fda.gov/media/136314/download (accessed on 24 August 2020).
  69. ChromaCode HDPCRTM SARS-CoV-2 Assay. Available online: https://www.fda.gov/media/138786/download (accessed on 24 August 2020).
  70. SpectronRx HymonTM SARS-CoV-2 Test KitInstructionsforUse (Handbook). Available online: https://www.fda.gov/media/138345/download (accessed on 23 August 2020).
  71. Diagnostic Solutions COVID-19 | SARS-CoV-2. Available online: https://www.fda.gov/media/139516/download (accessed on 23 August 2020).
  72. DiaSorin Molecular Enabling Swift Action Against TheCOVID-19 Pandemic. Available online: https://molecular.diasorin.com/international/wp-content/uploads/2020/03/OUSC19BR0720-C19-Direct-Brochure-A4-APPROVED-OUS-Only.pdf (accessed on 24 August 2020).
  73. Eli Lilly Clinical Diagnostics Laboratory LILLY SARS-CoV-2 ASSAY. Available online: https://www.fda.gov/media/140543/download (accessed on 24 August 2020).
  74. Enzo Life Sciences AMPIPROBE® SARS-CoV-2 Test System. Available online: https://www.fda.gov/media/139828/download (accessed on 24 August 2020).
  75. EuroRealTime EURORealTime SARS-CoV-2. Available online: https://www.fda.gov/media/138761/download (accessed on 24 August 2020).
  76. Fosun Pharma Fosun COVID-19 RT-PCR Detection Kit. Available online: https://www.fda.gov/media/137120/download (accessed on 24 August 2020).
  77. Genematrix NeoPlex COVID-19 Detection Kit (PID: 3218179). Available online: https://www.buykorea.org/bk/byr/product/GOODS_DETAIL-3218179.do (accessed on 24 August 2020).
  78. GB GB SARS-CoV-2 Real-Time RT-PCR. Available online: https://www.gbc.com.tw/products/real-time-pcr/gb-sars-cov-2-real-time-rt-pcr (accessed on 24 August 2020).
  79. Genetron Genetron SARS-CoV-2 RNA Test. Available online: https://www.fda.gov/media/138685/download (accessed on 24 August 2020).
  80. Helix Laboratory Helix COVID-19 Test. Available online: https://www.fda.gov/media/140420/download (accessed on 24 August 2020).
  81. Hologic SARS-CoV-2 Assay (Panther Fusion™ System). Available online: https://www.hologic.com/sites/default/files/2020-05/AW-21388-001_002_01.pdf (accessed on 24 August 2020).
  82. InBios Smart Detect™ SARS-CoV-2 RRT-PCR Kit. Available online: https://inbios.com/smart-detecttm-sars-cov-2-rrt-pcr-kit/ (accessed on 23 August 2020).
  83. IDT SARS-CoV-2 Probes and Other COVID-19 Research Reagents. Available online: https://www.idtdna.com/pages/landing/coronavirus-research-reagents#media (accessed on 24 August 2020).
  84. Integrity Laboratories SARS-CoV-2 RT-PCR Assay. Available online: https://www.fda.gov/media/136942/download#:~:text=The%20Integrity%20Laboratories%20SARS%2DCoV,detect%20cases%20of%20COVID%2D19. (accessed on 24 August 2020).
  85. JN Medsys ProTect™ COVID-19 RT-QPCR Kit. Available online: https://jnmedsys.com/covid19 (accessed on 24 August 2020).
  86. BioPerfectus technologies COVID-19 Coronavirus Real Time PCR Kit. Available online: https://www.fda.gov/media/139279/download (accessed on 24 August 2020).
  87. KorvaLabs Inc. Clinical Laboratory CURATIVE SARS-COV-2 ASSAY. Available online: https://www.fda.gov/media/137089/download (accessed on 24 August 2020).
  88. IDYLLA IDYLLA™ SARS-COV-2 TEST. 2020. Available online: https://www.biocartis.com/en-US/meet-idylla/idylla-infectious-disease-assays/idylla-sars-cov-2-test-eua-pending (accessed on 2 September 2020).
  89. BioFire BioFire® Respiratory Panel 2.1-EZ (RP2.1-EZ). Available online: https://www.fda.gov/media/142696/download (accessed on 21 December 2020).
  90. Medicon VitaPCRTM Platform. Available online: https://www.mediconire.com/vitapcrtm-platform (accessed on 21 December 2020).
  91. RCA Laboratory Services GENETWORx Covid-19 Nasal Swab Test. Available online: https://www.fda.gov/media/144553/download (accessed on 21 December 2020).
  92. Euroimmun EURORealTime SARS-CoV-2/InfluenzaA/B. Available online: https://www.coronavirus-diagnostics.com/documents/Indications/Infections/Coronavirus/MP_2606_D_UK_B.pdf (accessed on 21 December 2020).
  93. Agena Bioscience Multiplex RT-PCR/MALDI-TOF Test Intended for the Qualitativedetection of Nucleic Acid from SARS-CoV-2. Available online: https://agenabio.com/wp-content/uploads/2020/07/SC2_Panel_IFU-CUS-001_R03.pdf (accessed on 24 August 2020).
  94. Bio-Rad Qualitative Assay for Use on the QX200TM AndQXDx™ Droplet Digital™ PCR Systems. Available online: https://www.fda.gov/media/137579/download (accessed on 24 August 2020).
  95. Atila BioSystems IAMP® COVID-19 Detection Kit. Available online: https://www.fda.gov/media/136870/download (accessed on 24 August 2020).
  96. ID NOW ID NOW COVID-19. Available online: https://www.fda.gov/media/136525/download (accessed on 24 August 2020).
  97. Color SARS-CoV-2 LAMP Diagnostic Assay. Available online: https://www.color.com/wp-content/uploads/2020/05/LAMP-Diagnostic-Assay.pdf (accessed on 24 August 2020).
  98. BioSpace Rendu Gets NMPA’s Emergency Approval for Coronavirus Nucleic Acid Detection System. Available online: https://www.biospace.com/article/releases/rendu-gets-nmpa-s-emergency-approval-for-coronavirus-nucleic-acid-detection-system (accessed on 24 August 2020).
  99. Seasun Biomaterials AQ-TOP™ COVID-19 Rapid Detection Kit. Available online: https://www.fda.gov/media/138307/download (accessed on 24 August 2020).
  100. Lucira Lucira™ COVID-19All-In-One Test Kit. Available online: https://www.fda.gov/media/143808/download (accessed on 21 December 2020).
  101. Poplar Healthcare SARS-COV-2 TMA POOLING ASSAY. Available online: https://www.fda.gov/media/140792/download (accessed on 24 August 2020).
  102. Illumina Illumina COVIDSeq Test. Available online: https://www.illumina.com/products/by-type/ivd-products/covidseq.html (accessed on 24 August 2020).
  103. Fulgent COVID-19 Testing Solutions. Available online: https://www.fulgentgenetics.com/covid19/molecular (accessed on 23 August 2020).
  104. Sherlock Biosciences Sherlock Biosciences Receives FDA Emergency Use Authorization for CRISPR SARS-CoV-2 Rapid Diagnostic. Available online: https://sherlock.bio/sherlock-biosciences-receives-fda-emergency-use-authorization-for-crispr-sars-cov-2-rapid-diagnostic/ (accessed on 20 August 2020).
  105. UCSF SARS-COV-2 RNA DETECTR ASSAY. Available online: https://www.fda.gov/media/139937/download (accessed on 24 August 2020).
  106. Tahamtan, A.; Ardebili, A. Real-Time RT-PCR in COVID-19 Detection: Issues Affecting the Results. Expert Rev. Mol. Diagn. 2020, 20, 453–454. [Google Scholar] [CrossRef] [Green Version]
  107. Brennan-Krohn, T. Making Sense of Respiratory Viral Panel Results. Available online: https://asm.org/Articles/2020/March/Making-Sense-of-Respiratory-Viral-Panel-Results (accessed on 6 July 2020).
  108. Lin, C.-Y.; Hwang, D.; Chiu, N.-C.; Weng, L.-C.; Liu, H.-F.; Mu, J.-J.; Liu, C.-P.; Chi, H. Increased Detection of Viruses in Children with Respiratory Tract Infection Using PCR. Int. J. Environ. Res. Public Health 2020, 17, 564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Mahony, J.B. Detection of Respiratory Viruses by Molecular Methods. Clin. Microbiol. Rev. 2008, 21, 716–747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. C D C Interim Guidelines for Collecting, Handling, and Testing Clinical Specimens from Persons for Coronavirus Disease 2019 (COVID-19). Available online: https://www.cdc.gov/coronavirus/2019-ncov/lab/guidelines-clinical-specimens.html#specimen (accessed on 22 December 2020).
  111. Azghandi, M.; Kerachian, M.A. Detection of Novel Coronavirus (SARS-CoV-2) RNA in Peripheral Blood Specimens. J. Transl. Med. 2020, 18, 412. [Google Scholar] [CrossRef] [PubMed]
  112. Kim, J.-M.; Kim, H.M.; Lee, E.J.; Jo, H.J.; Yoon, Y.; Lee, N.-J.; Son, J.; Lee, Y.-J.; Kim, M.S.; Lee, Y.-P.; et al. Detection and Isolation of SARS-CoV-2 in Serum, Urine, and Stool Specimens of COVID-19 Patients from the Republic of Korea. Osong Public Health Res. Perspect. 2020, 11, 112–117. [Google Scholar] [CrossRef]
  113. Wozniak, A.; Cerda, A.; Ibarra-Henríquez, C.; Sebastian, V.; Armijo, G.; Lamig, L.; Miranda, C.; Lagos, M.; Solari, S.; Guzmán, A.M.; et al. A Simple RNA Preparation Method for SARS-CoV-2 Detection by RT-QPCR. Sci. Rep. 2020, 10, 16608. [Google Scholar] [CrossRef]
  114. Mo, Y.; Wan, R.; Zhang, Q. Application of Reverse Transcription-PCR and Real-Time PCR in Nanotoxicity Research. Methods Mol. Biol. 2012, 926, 99–112. [Google Scholar] [CrossRef] [Green Version]
  115. Barreto, H.G.; de Pádua Milagres, F.A.; de Araújo, G.C.; Daúde, M.M.; Benedito, V.A. Diagnosing the Novel SARS-CoV-2 by Quantitative RT-PCR: Variations and Opportunities. J. Mol. Med. 2020, 98, 1727–1736. [Google Scholar] [CrossRef] [PubMed]
  116. Joon, D.; Nimesh, M.; Saluja, D. Loop-Mediated Isothermal Amplification as Alternative to PCR for the Diagnosis of Extra-Pulmonary Tuberculosis. Int. J. Tuberc. Lung Dis. 2015, 19, 986–991. [Google Scholar] [CrossRef]
  117. Salamin, O.; Kuuranne, T.; Saugy, M.; Leuenberger, N. Loop-Mediated Isothermal Amplification (LAMP) as an Alternative to PCR: A Rapid on-Site Detection of Gene Doping. Drug Test. Anal. 2017, 9, 1731–1737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Zhao, Y.; Chen, F.; Li, Q.; Wang, L.; Fan, C. Isothermal Amplification of Nucleic Acids. Chem. Rev. 2015, 115, 12491–12545. [Google Scholar] [CrossRef] [PubMed]
  119. Zanoli, L.M.; Spoto, G. Isothermal Amplification Methods for the Detection of Nucleic Acids in Microfluidic Devices. Biosensors 2013, 3, 18–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Deiman, B.; van Aarle, P.; Sillekens, P. Characteristics and Applications of Nucleic Acid Sequence-Based Amplification (NASBA). Mol. Biotechnol. 2002, 20, 163–179. [Google Scholar] [CrossRef]
  121. Fakruddin, M. Nucleic acid sequence-based amplification (NASBA)-prospects and applications. Int. J. Life Sci. Pharma Res. 2020, 2, 106–121. [Google Scholar]
  122. Selvarajah, D.; Naing, C.; Htet, N.H.; Mak, J.W. Loop-Mediated Isothermal Amplification (LAMP) Test for Diagnosis of Uncomplicated Malaria in Endemic Areas: A Meta-Analysis of Diagnostic Test Accuracy. Malar. J. 2020, 19, 211. [Google Scholar] [CrossRef] [PubMed]
  123. Sahoo, P.R.; Sethy, K.; Mohapatra, S.; Panda, D. Loop Mediated Isothermal Amplification: An Innovative Gene Amplification Technique for Animal Diseases. Vet. World 2016, 9, 465–469. [Google Scholar] [CrossRef] [Green Version]
  124. Notomi, T.; Okayama, H.; Masubuchi, H.; Yonekawa, T.; Watanabe, K.; Amino, N.; Hase, T. Loop-Mediated Isothermal Amplification of DNA. Nucleic Acids Res. 2000, 28, E63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Fakruddin, M.; Mannan, K.S.B.; Chowdhury, A.; Mazumdar, R.M.; Hossain, M.N.; Islam, S.; Chowdhury, M.A. Nucleic Acid Amplification: Alternative Methods of Polymerase Chain Reaction. J. Pharm. Bioallied Sci. 2013, 5, 245–252. [Google Scholar] [CrossRef]
  126. Dixit, K.K.; Verma, S.; Singh, O.P.; Singh, D.; Singh, A.P.; Gupta, R.; Negi, N.S.; Das, P.; Sundar, S.; Singh, R.; et al. Validation of SYBR Green I Based Closed Tube Loop Mediated Isothermal Amplification (LAMP) Assay and Simplified Direct-Blood-Lysis (DBL)-LAMP Assay for Diagnosis of Visceral Leishmaniasis (VL). PLoS Negl. Trop. Dis. 2018, 12, e0006922. [Google Scholar] [CrossRef] [Green Version]
  127. Toonkomdang, S.; Phinyo, P.; Phetsuksiri, B.; Patumanond, J.; Rudeeaneksin, J.; Klayut, W. Pragmatic Accuracy of an In-House Loop-Mediated Isothermal Amplification (LAMP) for Diagnosis of Pulmonary Tuberculosis in a Thai Community Hospital. PLoS ONE 2020, 15, e0236496. [Google Scholar] [CrossRef]
  128. Thompson, D.; Lei, Y. Mini Review: Recent Progress in RT-LAMP Enabled COVID-19 Detection. Sensors and Actuators Reports 2020, 2, 100017. [Google Scholar] [CrossRef]
  129. Kashir, J.; Yaqinuddin, A. Loop Mediated Isothermal Amplification (LAMP) Assays as a Rapid Diagnostic for COVID-19. Med. Hypotheses 2020, 141, 109786. [Google Scholar] [CrossRef]
  130. Joung, J.; Ladha, A.; Saito, M.; Segel, M.; Bruneau, R.; Huang, M.-L.W.; Kim, N.-G.; Yu, X.; Li, J.; Walker, B.D.; et al. Point-of-Care Testing for COVID-19 Using SHERLOCK Diagnostics. medRxiv 2020, 1–21. [Google Scholar] [CrossRef]
  131. Hao, M.; Qiao, J.; Qi, H. Current and Emerging Methods for the Synthesis of Single-Stranded DNA. Genes 2020, 11, 116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Goo, N.-I.; Kim, D.-E. Rolling Circle Amplification as Isothermal Gene Amplification in Molecular Diagnostics. BioChip J. 2016, 10, 262–271. [Google Scholar] [CrossRef]
  133. Zhang, D.; Wu, J.; Ye, F.; Feng, T.; Lee, I.; Yin, B. Amplification of Circularizable Probes for the Detection of Target Nucleic Acids and Proteins. Clin. Chim. Acta 2006, 363, 61–70. [Google Scholar] [CrossRef]
  134. Pokhrel, P.; Hu, C.; Mao, H. Detecting the Coronavirus (COVID-19). ACS Sens. 2020, 5, 2283–2296. [Google Scholar] [CrossRef]
  135. Qi, X.; Bakht, S.; Devos, K.M.; Gale, M.D.; Osbourn, A. L-RCA (Ligation-Rolling Circle Amplification): A General Method for Genotyping of Single Nucleotide Polymorphisms (SNPs). Nucleic Acids Res. 2001, 29, E116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Kashkin, K.N.; Strizhkov, B.N.; Gryadunov, D.A.; Surzhikov, S.A.; Grechishnikova, I.V.; Kreindlin, E.Y.; Chupeeva, V.V.; Evseev, K.B.; Turygin, A.Y.; Mirzabekov, A.D. Detection of Single-Nucleotide Polymorphisms in the P53 Gene by LDR/RCA in Hydrogel Microarrays. Mol. Biol. 2005, 39, 26–34. [Google Scholar] [CrossRef]
  137. Chen, X.; Wang, B.; Yang, W.; Kong, F.; Li, C.; Sun, Z.; Jelfs, P.; Gilbert, G.L. Rolling Circle Amplification for Direct Detection of RpoB Gene Mutations in Mycobacterium Tuberculosis Isolates from Clinical Specimens. J. Clin. Microbiol. 2014, 52, 1540–1548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Wang, B.; Potter, S.J.; Lin, Y.; Cunningham, A.L.; Dwyer, D.E.; Su, Y.; Ma, X.; Hou, Y.; Saksena, N.K. Rapid and Sensitive Detection of Severe Acute Respiratory Syndrome Coronavirus by Rolling Circle Amplification. J. Clin. Microbiol. 2005, 43, 2339–2344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Ciftci, S.; Neumann, F.; Abdurahman, S.; Appelberg, K.S.; Mirazimi, A.; Nilsson, M.; Madaboosi, N. Digital Rolling Circle Amplification-Based Detection of Ebola and Other Tropical Viruses. J. Mol. Diagn. 2020, 22, 272–283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Huang, W.; Hsu, H.; Su, J.; Clapper, J.; Hsu, J. Room Temperature Isothermal Colorimetric Padlock Probe Rolling Circle Amplification for Viral RNA Detection. BioRxiv 2020. [Google Scholar] [CrossRef]
  141. Carter, L.J.; Garner, L.V.; Smoot, J.W.; Li, Y.; Zhou, Q.; Saveson, C.J.; Sasso, J.M.; Gregg, A.C.; Soares, D.J.; Beskid, T.R.; et al. Assay Techniques and Test Development for COVID-19 Diagnosis. ACS Cent. Sci. 2020, 6, 591–605. [Google Scholar] [CrossRef] [PubMed]
  142. Hill, C.S. Molecular Diagnostic Testing for Infectious Diseases Using TMA Technology. Expert Rev. Mol. Diagn. 2001, 1, 445–455. [Google Scholar] [CrossRef] [PubMed]
  143. Shen, C.-H. Amplification of nucleic acids. In Diagnostic Molecular Biology; Elsevier: Amsterdam, The Netherlands, 2019; pp. 215–247. ISBN 9780128028230. [Google Scholar]
  144. Gorzalski, A.J.; Tian, H.; Laverdure, C.; Morzunov, S.; Verma, S.C.; VanHooser, S.; Pandori, M.W. High-Throughput Transcription-Mediated Amplification on the Hologic Panther Is a Highly Sensitive Method of Detection for SARS-CoV-2. J. Clin. Virol. 2020, 129, 104501. [Google Scholar] [CrossRef] [PubMed]
  145. Sheikhzadeh, E.; Eissa, S.; Ismail, A.; Zourob, M. Diagnostic Techniques for COVID-19 and New Developments. Talanta 2020, 220, 121392. [Google Scholar] [CrossRef] [PubMed]
  146. Hologic Aptima® SARS-CoV-2 Assay (Panther® System). Available online: https://www.fda.gov/media/138096/download (accessed on 1 September 2020).
  147. Piepenburg, O.; Williams, C.H.; Stemple, D.L.; Armes, N.A. DNA Detection Using Recombination Proteins. PLoS Biol. 2006, 4, e204. [Google Scholar] [CrossRef] [PubMed]
  148. Lobato, I.M.; O’Sullivan, C.K. Recombinase Polymerase Amplification: Basics, Applications and Recent Advances. Trends Analyt. Chem. 2018, 98, 19–35. [Google Scholar] [CrossRef] [PubMed]
  149. Lee, J.; Heo, S.; Bang, D. Applying a Linear Amplification Strategy to Recombinase Polymerase Amplification for Uniform DNA Library Amplification. ACS Omega 2019, 4, 19953–19958. [Google Scholar] [CrossRef] [Green Version]
  150. Mayboroda, O.; Gonzalez Benito, A.; Sabaté del Rio, J.; Svobodova, M.; Julich, S.; Tomaso, H.; O’Sullivan, C.K.; Katakis, I. Isothermal Solid-Phase Amplification System for Detection of Yersinia Pestis. Anal. Bioanal. Chem. 2016, 408, 671–676. [Google Scholar] [CrossRef]
  151. Wu, L.; Ye, L.; Wang, Z.; Cui, Y.; Wang, J. Utilization of Recombinase Polymerase Amplification Combined with a Lateral Flow Strip for Detection of Perkinsus Beihaiensis in the Oyster Crassostrea Hongkongensis. Parasites Vectors 2019, 12, 360. [Google Scholar] [CrossRef] [Green Version]
  152. Li, J.; Ma, B.; Fang, J.; Zhi, A.; Chen, E.; Xu, Y.; Yu, X.; Sun, C.; Zhang, M. Recombinase Polymerase Amplification (RPA) Combined with Lateral Flow Immunoassay for Rapid Detection of Salmonella in Food. Foods 2019, 9, 27. [Google Scholar] [CrossRef] [Green Version]
  153. Nguyen, C.T.; Nguyen, U.D.; Le, T.T.; Bui, H.T.; Nguyen, A.N.T.; Thi Nguyen, A.N.; Trieu, N.T.; Trieu, L.P.; Bui, S.T.; Nguyen, C.; et al. Establishment of Recombinase Polymerase Amplification Assay for Rapid and Sensitive Detection of Orientia Tsutsugamushi in Southeast Asia. Acta Trop. 2020, 210, 105541. [Google Scholar] [CrossRef]
  154. Wang, R.; Zhang, F.; Wang, L.; Qian, W.; Qian, C.; Wu, J.; Ying, Y. Instant, Visual, and Instrument-Free Method for On-Site Screening of GTS 40-3-2 Soybean Based on Body-Heat Triggered Recombinase Polymerase Amplification. Anal. Chem. 2017, 89, 4413–4418. [Google Scholar] [CrossRef] [PubMed]
  155. Rostron, P.; Pennance, T.; Bakar, F.; Rollinson, D.; Knopp, S.; Allan, F.; Kabole, F.; Ali, S.M.; Ame, S.M.; Webster, B.L. Development of a Recombinase Polymerase Amplification (RPA) Fluorescence Assay for the Detection of Schistosoma Haematobium. Parasites Vectors 2019, 12, 514. [Google Scholar] [CrossRef]
  156. Yang, Y.; Qin, X.; Wang, G.; Zhang, Y.; Shang, Y.; Zhang, Z. Development of a Fluorescent Probe-Based Recombinase Polymerase Amplification Assay for Rapid Detection of Orf Virus. Virol. J. 2015, 12, 206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. James, A.S.; Todd, S.; Pollak, N.M.; Marsh, G.A.; Macdonald, J. Ebolavirus Diagnosis Made Simple, Comparable and Faster than Molecular Detection Methods: Preparing for the Future. Virol. J. 2018, 15, 75. [Google Scholar] [CrossRef]
  158. Zhang, F.; Abudayyeh, O.; Goothenberg, J. A Protocol for Detection of COVID-19 Using CRISPR Diagnostics. Available online: https://www.broadinstitute.org/files/publications/special/COVID-19%20detection%20(updated).pdf (accessed on 10 September 2020).
  159. Xia, S.; Chen, X. Single-Copy Sensitive, Field-Deployable, and Simultaneous Dual-Gene Detection of SARS-CoV-2 RNA via Modified RT-RPA. Cell Discov. 2020, 6, 37. [Google Scholar] [CrossRef] [PubMed]
  160. Barba, M.; Czosnek, H.; Hadidi, A. Historical Perspective, Development and Applications of next-Generation Sequencing in Plant Virology. Viruses 2014, 6, 106–136. [Google Scholar] [CrossRef] [PubMed]
  161. First NGS-Based COVID-19 Diagnostic. Nat. Biotechnol. 2020, 38, 777. [CrossRef] [PubMed]
  162. Canadian Agency for Drugs and Technologies in Health. CADTH Summary of findings—Cost-effectiveness of next generation sequencing. In CADTH Rapid Response Service; Canadian Agency for Drugs and Technologies in Health: Ottawa, ON, Canada, 2014. [Google Scholar]
  163. Illumina Illumina COVIDSeq Test Instructions for Use. Available online: https://www.fda.gov/media/138776/download (accessed on 22 December 2020).
  164. Deckert, A.; Bärnighausen, T.; Kyei, N.N. Simulation of Pooled-Sample Analysis Strategies for COVID-19 Mass Testing. Bull. World Health Organ. 2020, 98, 590–598. [Google Scholar] [CrossRef] [PubMed]
  165. Wang, X.; Shang, X.; Huang, X. Next-Generation Pathogen Diagnosis with CRISPR/Cas-Based Detection Methods. Emerg. Microbes Infect. 2020, 9, 1682–1691. [Google Scholar] [CrossRef] [PubMed]
  166. Furmanski, L.; Kellar, J.; Mathewson, S.; Verma, M. Crispr Catalyzes Point-of-Care Testing; BCG: Boston, MA, USA, 2020. [Google Scholar]
  167. Mendoza, B.J.; Trinh, C.T. In Silico Processing of the Complete CRISPR-Cas Spacer Space for Identification of PAM Sequences. BioRxiv 2018. [Google Scholar] [CrossRef]
  168. Mojica, F.J.; Juez, G.; Rodríguez-Valera, F. Transcription at Different Salinities of Haloferax Mediterranei Sequences Adjacent to Partially Modified PstI Sites. Mol. Microbiol. 1993, 9, 613–621. [Google Scholar] [CrossRef]
  169. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A Programmable Dual-RNA-Guided DNA Endonuclease in Adaptive Bacterial Immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef] [PubMed]
  170. Cong, L.; Ran, F.A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P.D.; Wu, X.; Jiang, W.; Marraffini, L.A.; et al. Multiplex Genome Engineering Using CRISPR/Cas Systems. Science 2013, 339, 819–823. [Google Scholar] [CrossRef] [Green Version]
  171. Chen, J.S.; Ma, E.; Harrington, L.B.; Da Costa, M.; Tian, X.; Palefsky, J.M.; Doudna, J.A. CRISPR-Cas12a Target Binding Unleashes Indiscriminate Single-Stranded DNase Activity. Science 2018, 360, 436–439. [Google Scholar] [CrossRef] [Green Version]
  172. Gootenberg, J.S.; Abudayyeh, O.O.; Lee, J.W.; Essletzbichler, P.; Dy, A.J.; Joung, J.; Verdine, V.; Donghia, N.; Daringer, N.M.; Freije, C.A.; et al. Nucleic Acid Detection with CRISPR-Cas13a/C2c2. Science 2017, 356, 438–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Brandsma, E.; Verhagen, H.J.; van de Laar, T.J.W.; Claas, E.C.J.; Cornelissen, M.; van den Akker, E. Rapid, Sensitive and Specific SARS Coronavirus-2 Detection: A Multi-Center Comparison between Standard QRT-PCR and CRISPR Based DETECTR. J. Infect. Dis. 2020, 223, 206–213. [Google Scholar] [CrossRef]
  174. Kumar, P.; Malik, Y.S.; Ganesh, B.; Rahangdale, S.; Saurabh, S.; Natesan, S.; Srivastava, A.; Sharun, K.; Yatoo, M.I.; Tiwari, R.; et al. CRISPR-Cas System: An Approach With Potentials for COVID-19 Diagnosis and Therapeutics. Front. Cell. Infect. Microbiol. 2020, 10, 576875. [Google Scholar] [CrossRef] [PubMed]
  175. Broughton, J.P.; Deng, X.; Yu, G.; Fasching, C.L.; Servellita, V.; Singh, J.; Miao, X.; Streithorst, J.A.; Granados, A.; Sotomayor-Gonzalez, A.; et al. CRISPR-Cas12-Based Detection of SARS-CoV-2. Nat. Biotechnol. 2020, 38, 870–874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Nguyen, L.T.; Smith, B.M.; Jain, P.K. Enhancement of Trans-Cleavage Activity of Cas12a with Engineered CrRNA Enables Amplified Nucleic Acid Detection. Nat. Commun. 2020, 11, 4906. [Google Scholar] [CrossRef]
  177. Benzigar, M.R.; Bhattacharjee, R.; Baharfar, M.; Liu, G. Current Methods for Diagnosis of Human Coronaviruses: Pros and Cons. Anal. Bioanal. Chem. 2020, 413, 2311–2330. [Google Scholar] [CrossRef]
  178. Arnaout, R.; Lee, R.A.; Lee, G.R.; Callahan, C.; Yen, C.F.; Smith, K.P.; Arora, R.; Kirby, J.E. SARS-CoV2 Testing: The Limit of Detection Matters. BioRxiv 2020. [Google Scholar] [CrossRef]
  179. Ding, X.; Yin, K.; Li, Z.; Lalla, R.V.; Ballesteros, E.; Sfeir, M.M.; Liu, C. All-in-One Dual CRISPR-Cas12a (AIOD-CRISPR) Assay: A Case for Rapid, Ultrasensitive and Visual Detection of Novel Coronavirus SARS-CoV-2 and HIV Virus at the Point of Care. BioRxiv 2020. [Google Scholar] [CrossRef]
  180. Javalkote, V.S.; Kancharla, N.; Bhadra, B.; Shukla, M.; Soni, B.; Goodin, M.; Bandyopadhyay, A.; Dasgupta, S. CRISPR-Based Assays for Rapid Detection of SARS-CoV-2. Methods 2020. [Google Scholar] [CrossRef]
  181. Ding, X.; Yin, K.; Li, Z.; Lalla, R.V.; Ballesteros, E.; Sfeir, M.M.; Liu, C. Ultrasensitive and Visual Detection of SARS-CoV-2 Using All-in-One Dual CRISPR-Cas12a Assay. Nat. Commun. 2020, 11, 4711. [Google Scholar] [CrossRef] [PubMed]
  182. Kellner, M.J.; Koob, J.G.; Gootenberg, J.S.; Abudayyeh, O.O.; Zhang, F. Author Correction: SHERLOCK: Nucleic Acid Detection with CRISPR Nucleases. Nat. Protoc. 2020, 15, 1311. [Google Scholar] [CrossRef] [Green Version]
  183. Ackerman, C.M.; Myhrvold, C.; Thakku, S.G.; Freije, C.A.; Metsky, H.C.; Yang, D.K.; Ye, S.H.; Boehm, C.K.; Kosoko-Thoroddsen, T.-S.F.; Kehe, J.; et al. Massively Multiplexed Nucleic Acid Detection with Cas13. Nature 2020, 582, 277–282. [Google Scholar] [CrossRef]
  184. Metsky, H.C.; Freije, C.A.; Kosoko-Thoroddsen, T.-S.F.; Sabeti, P.C.; Myhrvold, C. CRISPR-Based COVID-19 Surveillance Using a Genomically-Comprehensive Machine Learning Approach. BioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  185. MIT News CRISPR-Based Diagnostic Chips Perform Thousands of Tests Simultaneously to Detect Viruses. Available online: https://news.mit.edu/2020/crispr-diagnostic-chips-test-viruses-0429 (accessed on 4 September 2021).
  186. Rauch, J.N.; Valois, E.; Solley, S.C.; Braig, F.; Lach, R.S.; Baxter, N.J.; Kosik, K.S.; Arias, C.; Acosta-Alvear, D.; Wilson, M.Z. A Scalable, Easy-to-Deploy, Protocol for Cas13-Based Detection of SARS-CoV-2 Genetic Material. J. Clin. Microbiol. 2020, 59, 2–20. [Google Scholar] [CrossRef]
  187. Fernández-Barat, L.; López-Aladid, R.; Torres, A. The Value of Serology Testing to Manage SARS-CoV-2 Infections. Eur. Respir. J. 2020, 56, 2002411. [Google Scholar] [CrossRef]
  188. Bao, L.; Deng, W.; Gao, H.; Xiao, C.; Liu, J.; Xue, J.; Lv, Q.; Liu, J.; Yu, P.; Xu, Y.; et al. Reinfection Could Not Occur in SARS-CoV-2 Infected Rhesus Macaques. BioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  189. Hoffman, T.; Nissen, K.; Krambrich, J.; Rönnberg, B.; Akaberi, D.; Esmaeilzadeh, M.; Salaneck, E.; Lindahl, J.; Lundkvist, Å. Evaluation of a COVID-19 IgM and IgG Rapid Test; an Efficient Tool for Assessment of Past Exposure to SARS-CoV-2. Infect. Ecol. Epidemiol. 2020, 10, 1754538. [Google Scholar] [CrossRef] [Green Version]
  190. Taipale, J.; Romer, P.; Linnarsson, S. Population-Scale Testing Can Suppress the Spread of COVID-19. medRxiv 2020. [Google Scholar] [CrossRef]
  191. Chen, L.; Xiong, J.; Bao, L.; Shi, Y. Convalescent Plasma as a Potential Therapy for COVID-19. Lancet Infect. Dis. 2020, 20, 398–400. [Google Scholar] [CrossRef]
  192. Imai, K.; Tabata, S.; Ikeda, M.; Noguchi, S.; Kitagawa, Y.; Matuoka, M.; Miyoshi, K.; Tarumoto, N.; Sakai, J.; Ito, T.; et al. Clinical Evaluation of an Immunochromatographic IgM/IgG Antibody Assay and Chest Computed Tomography for the Diagnosis of COVID-19. J. Clin. Virol. 2020, 128, 104393. [Google Scholar] [CrossRef] [PubMed]
  193. Beretta, A.; Cranage, M.; Zipeto, D. Is Cross-Reactive Immunity Triggering COVID-19 Immunopathogenesis? Front. Immunol. 2020, 11, 567710. [Google Scholar] [CrossRef] [PubMed]
  194. Ward, S.; Lindsley, A.; Courter, J.; Assa’ad, A. Clinical Testing for COVID-19. J. Allergy Clin. Immunol. 2020, 146, 23–34. [Google Scholar] [CrossRef] [PubMed]
  195. FDA Independent Evaluations of COVID-19 Serological Tests. Available online: https://open.fda.gov/apis/device/covid19serology (accessed on 6 April 2021).
  196. Babson Diagnostics BABSON DIAGNOSTICS AC19G1. Available online: https://www.fda.gov/media/139446/download (accessed on 8 September 2020).
  197. Beckman Coulter SARS-CoV-2 IgG. Available online: https://www.beckmancoulter.com/products/immunoassay/access-sars-cov-2-igg-antibody-test#/documentos (accessed on 8 September 2020).
  198. Diazyme DIAZYME DZ-LITE SARS-CoV-2 IgG CLIA KIT. Available online: https://www.fda.gov/media/139865/download (accessed on 8 September 2020).
  199. InBios SCoV-2 DetectTM IgM ELISA. Available online: https://www.fda.gov/media/139730/download (accessed on 8 September 2020).
  200. Emory Medical Laboratories SARS-COV-2 RBD IGG FOR ANTIBODY DETECTION. Available online: https://www.fda.gov/media/139053/download (accessed on 8 September 2020).
  201. Luminex XMAP® SARS-CoV-2 Multi-AntigenIgG Assay Package Insert. Available online: https://www.fda.gov/media/140256/download (accessed on 8 September 2020).
  202. Siemens Healthineers SARS-CoV-2 IgG (COV2G). Available online: https://www.fda.gov/media/140699/download (accessed on 8 September 2020).
  203. Siemens SARS-CoV-2 IgG (COV2G). Available online: https://www.fda.gov/media/140704/download (accessed on 8 September 2020).
  204. EUROIMMUN Anti-SARS-CoV-2 ELISA (IgG). Available online: https://www.fda.gov/media/137609/download (accessed on 8 September 2020).
  205. Abbott Laboratories SARS-CoV-2 IgG. Available online: https://www.fda.gov/media/137383/download (accessed on 8 September 2020).
  206. Dia-Sorin LIAISON® SARS-CoV-2 S1/S2 IgG. Available online: https://www.fda.gov/media/137359/download (accessed on 8 September 2020).
  207. Ortho-Clinical Diagnostics VITROS Immunodiagnostic Products Anti-SARS-CoV-2 IgGReagent Pack. Available online: https://www.fda.gov/media/137363/download (accessed on 8 September 2020).
  208. Mount Sinai Laboratory COVID-19 ELISA IGG ANTIBODY TEST. Available online: https://www.fda.gov/media/137029/download (accessed on 7 September 2020).
  209. BIOMÉRIEUX VIDAS® SARS-COV-2 IgG. Available online: https://www.fda.gov/media/140937/download (accessed on 8 September 2020).
  210. Healgen COVID-19 IgG/IgM Rapid Test Cassette. Available online: https://www.fda.gov/media/138438/download (accessed on 8 September 2020).
  211. Megna Health RAPID COVID-19 IgM/IgG COMBO TEST KIT. Available online: https://www.fda.gov/media/140297/download (accessed on 9 September 2020).
  212. Assure Assure COVID-19 IgG/IgM Rapid Test Device. Available online: https://www.fda.gov/media/139792/download (accessed on 8 September 2020).
  213. Sienna SiennaTM-Clarity COVIBLOCK™. Available online: https://www.fda.gov/media/140082/download (accessed on 8 September 2020).
  214. AccessBio COVID-19 IgM/IgG COVID-19 IgM/IgG. Available online: https://www.apacor.com/wp-content/uploads/2020/05/IFU-RCIM71-E_Rev.A_20200417.pdf (accessed on 8 September 2020).
  215. Xiamen Biotime Biotechnology Co., Ltd. BIOTIME SARS-CoV-2 IgG/IgM Rapid Qualitative Test. Available online: https://www.fda.gov/media/140443/download (accessed on 8 September 2020).
  216. Ab Clinical Labs Vibrant COVID-19 Ab Assay. Available online: https://www.vibrant-america.com/wp-content/uploads/2020/04/VA-COV-001-CovidAb-Analytical-and-Clinical-Studies-Report-FDA-Rev-2.pdf (accessed on 8 September 2020).
  217. Autobio Diagnostics Anti-SARS-CoV-2 Rapid Test. Available online: https://www.fda.gov/media/137367/download (accessed on 8 September 2020).
  218. ChemBio Serology Test Evaluation Report for “DPP COVID-19IgM/IgG System. Available online: https://www.accessdata.fda.gov/cdrh_docs/presentations/maf/maf3265-a001.pdf (accessed on 8 September 2020).
  219. Celex Cellex QSARS-CoV-2 IgG/IgM Rapid Test. Available online: https://www.fda.gov/media/136625/download (accessed on 8 September 2020).
  220. SIEMENS SARSCoV2 Total Antibody Assay (CV2T). Available online: https://www.fda.gov/media/138757/download (accessed on 8 September 2020).
  221. Beijing Wantai Biological Pharmacy Enterprise Co., Ltd. WANTAI SARS-CoV-2 Ab ELISA. Available online: https://www.fda.gov/media/140929/download (accessed on 7 September 2020).
  222. WANTAI Wantai SARS-CoV-2 Diagnostics. Available online: https://www.fda.gov/media/140030/download (accessed on 8 September 2020).
  223. Wadsworth Center NEW YORK SARS-COV MICROSPHERE IMMUNOASSAY. Available online: https://www.fda.gov/media/137541/download (accessed on 8 September 2020).
  224. Bio-Rad Platelia SARS-CoV-2 Total Ab. Available online: https://www.fda.gov/media/137493/download (accessed on 8 September 2020).
  225. Roche Diagnosis Elecsys Anti-SARS-CoV-2. Available online: https://www.fda.gov/media/137605/download (accessed on 8 September 2020).
  226. Peeling, R.W.; Olliaro, P.L.; Boeras, D.I.; Fongwen, N. Scaling up COVID-19 Rapid Antigen Tests: Promises and Challenges. Lancet Infect. Dis. 2021, 21, E290–E295. [Google Scholar] [CrossRef]
  227. Chau, C.H.; Strope, J.D.; Figg, W.D. COVID-19 Clinical Diagnostics and Testing Technology. Pharmacotherapy 2020, 40, 857–868. [Google Scholar] [CrossRef] [PubMed]
  228. Chaimayo, C.; Kaewnaphan, B.; Tanlieng, N.; Athipanyasilp, N.; Sirijatuphat, R.; Chayakulkeeree, M.; Angkasekwinai, N.; Sutthent, R.; Puangpunngam, N.; Tharmviboonsri, T.; et al. Rapid SARS-CoV-2 Antigen Detection Assay in Comparison with Real-Time RT-PCR Assay for Laboratory Diagnosis of COVID-19 in Thailand. Virol. J. 2020, 17, 177. [Google Scholar] [CrossRef] [PubMed]
  229. Hirotsu, Y.; Maejima, M.; Shibusawa, M.; Nagakubo, Y.; Hosaka, K.; Amemiya, K.; Sueki, H.; Hayakawa, M.; Mochizuki, H.; Tsutsui, T.; et al. Comparison of Automated SARS-CoV-2 Antigen Test for COVID-19 Infection with Quantitative RT-PCR Using 313 Nasopharyngeal Swabs, Including from Seven Serially Followed Patients. Int. J. Infect. Dis. 2020, 99, 397–402. [Google Scholar] [CrossRef] [PubMed]
  230. Porte, L.; Legarraga, P.; Vollrath, V.; Aguilera, X.; Munita, J.M.; Araos, R.; Pizarro, G.; Vial, P.; Dittrich, S.; Weitzel, T. Evaluation of Novel Antigen-Based Rapid Detection Test for the Diagnosis of SARS-CoV-2 in Respiratory Samples. SSRN J. 2020. [Google Scholar] [CrossRef]
  231. FDA Emergency Use Authorizations for Medical Devices. Available online: https://www.fda.gov/medical-devices/emergency-situations-medical-devices/emergency-use-authorizations-medical-devices#covid19ivd (accessed on 13 March 2021).
  232. Ortho Clinical Diagnostics Introducing Ortho’s VITROS® SARS-CoV-2 Antigen Test. Available online: https://www.orthoclinicaldiagnostics.com/global/covid19 (accessed on 31 March 2021).
  233. LumiraDx LumiraDx SARS-CoV-2 Ag Test. Available online: https://www.lumiradx.com/uk-en/what-we-do/diagnostics/test-technology/antigen-test (accessed on 6 April 2021).
  234. BD Veritor™ System for Rapid Detection of SARS-CoV-2. Available online: https://www.fda.gov/media/139755/download (accessed on 8 September 2020).
  235. Luminostics. Available online: https://luminostics.com (accessed on 1 April 2021).
  236. Ellume COVID-19 Home Test. Available online: https://www.ellumehealth.com/products/consumer-products/covid-home-test (accessed on 31 March 2021).
  237. Quidel Sofia SARS Antigen FIA. Available online: https://www.fda.gov/media/137885/download (accessed on 8 September 2020).
  238. Quidel QuickVue SARS Antigen Test. Available online: https://www.quidel.com/immunoassays/quickvue-sars-antigen-test (accessed on 31 March 2021).
  239. Abbott BinaxNOW™ COVID-19 Ag Card Will Help You Feel More Confident about Your COVID-19 Status. Available online: https://www.abbott.com/BinaxNOW-Test-NAVICA-App.html (accessed on 30 March 2021).
  240. Abbott Diagnostics Scarborough, Inc. BinaxNOWTM COVID-19 Ag CARD. Available online: https://www.fda.gov/media/141570/download (accessed on 8 September 2020).
  241. Quidel Sofia 2 Flu + SARS Antigen Fluorescent Immunoassay (FIA). Available online: https://www.quidel.com/immunoassays/sofia-2-flu-sars-antigen-fia (accessed on 6 April 2021).
  242. SAMPINUTE. Available online: https://www.celltrion.com/en-us/kit/sampinute (accessed on 1 April 2021).
  243. Quanterix SARS-CoV-2 N Protein Antigen. Available online: https://www.quanterix.com/simoa-assay-kits/sars-cov-2-n-protein-antigen (accessed on 31 March 2021).
  244. Nikaeen, G.; Abbaszadeh, S.; Yousefinejad, S. Application of Nanomaterials in Treatment, Anti-Infection and Detection of Coronaviruses. Nanomedicine 2020, 15, 1501–1512. [Google Scholar] [CrossRef] [PubMed]
  245. Yang, S.; Rothman, R.E. PCR-Based Diagnostics for Infectious Diseases: Uses, Limitations, and Future Applications in Acute-Care Settings. Lancet Infect. Dis. 2004, 4, 337–348. [Google Scholar] [CrossRef]
  246. Yadavalli, T.; Shukla, D. Role of Metal and Metal Oxide Nanoparticles as Diagnostic and Therapeutic Tools for Highly Prevalent Viral Infections. Nanomedicine 2017, 13, 219–230. [Google Scholar] [CrossRef] [Green Version]
  247. Seo, G.; Lee, G.; Kim, M.J.; Baek, S.-H.; Choi, M.; Ku, K.B.; Lee, C.-S.; Jun, S.; Park, D.; Kim, H.G.; et al. Rapid Detection of COVID-19 Causative Virus (SARS-CoV-2) in Human Nasopharyngeal Swab Specimens Using Field-Effect Transistor-Based Biosensor. ACS Nano 2020, 14, 5135–5142. [Google Scholar] [CrossRef] [Green Version]
  248. Qiu, G.; Gai, Z.; Tao, Y.; Schmitt, J.; Kullak-Ublick, G.A.; Wang, J. Dual-Functional Plasmonic Photothermal Biosensors for Highly Accurate Severe Acute Respiratory Syndrome Coronavirus 2 Detection. ACS Nano 2020, 14, 5268–5277. [Google Scholar] [CrossRef] [Green Version]
  249. Roh, C.; Jo, S.K. Quantitative and Sensitive Detection of SARS Coronavirus Nucleocapsid Protein Using Quantum Dots-Conjugated RNA Aptamer on Chip. J. Chem. Technol. Biotechnol. 2011, 86, 1475–1479. [Google Scholar] [CrossRef] [PubMed]
  250. Tallury, P.; Malhotra, A.; Byrne, L.M.; Santra, S. Nanobioimaging and Sensing of Infectious Diseases. Adv. Drug Deliv. Rev. 2010, 62, 424–437. [Google Scholar] [CrossRef]
  251. Di Gianvincenzo, P.; Marradi, M.; Martínez-Avila, O.M.; Bedoya, L.M.; Alcamí, J.; Penadés, S. Gold Nanoparticles Capped with Sulfate-Ended Ligands as Anti-HIV Agents. Bioorg. Med. Chem. Lett. 2010, 20, 2718–2721. [Google Scholar] [CrossRef]
  252. Zhu, B.; Liu, G.-L.; Ling, F.; Wang, G.-X. Carbon Nanotube-Based Nanocarrier Loaded with Ribavirin against Grass Carp Reovirus. Antiviral Res. 2015, 118, 29–38. [Google Scholar] [CrossRef] [PubMed]
  253. Sekimukai, H.; Iwata-Yoshikawa, N.; Fukushi, S.; Tani, H.; Kataoka, M.; Suzuki, T.; Hasegawa, H.; Niikura, K.; Arai, K.; Nagata, N. Gold Nanoparticle-Adjuvanted S Protein Induces a Strong Antigen-Specific IgG Response against Severe Acute Respiratory Syndrome-Related Coronavirus Infection, but Fails to Induce Protective Antibodies and Limit Eosinophilic Infiltration in Lungs. Microbiol. Immunol. 2020, 64, 33–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Wang, K.; Zhu, J.; Dong, H.; Pei, Z.; Zhou, T.; Hu, G. Rapid Detection of Variant and Classical Porcine Epidemic Diarrhea Virus by Nano-Nest PCR. Pak. Vet. J. 2017, 37, 225–229. [Google Scholar]
  255. Teengam, P.; Siangproh, W.; Tuantranont, A.; Vilaivan, T.; Chailapakul, O.; Henry, C.S. Multiplex Paper-Based Colorimetric DNA Sensor Using Pyrrolidinyl Peptide Nucleic Acid-Induced AgNPs Aggregation for Detecting MERS-CoV, MTB, and HPV Oligonucleotides. Anal. Chem. 2017, 89, 5428–5435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Layqah, L.A.; Eissa, S. An Electrochemical Immunosensor for the Corona Virus Associated with the Middle East Respiratory Syndrome Using an Array of Gold Nanoparticle-Modified Carbon Electrodes. Mikrochim. Acta 2019, 186, 224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Liu, I.-L.; Lin, Y.-C.; Lin, Y.-C.; Jian, C.-Z.; Cheng, I.-C.; Chen, H.-W. A Novel Immunochromatographic Strip for Antigen Detection of Avian Infectious Bronchitis Virus. Int. J. Mol. Sci. 2019, 20, 2216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Qiao, J.; Li, Y.; Wei, C.; Yang, H.; Yu, J.; Wei, H. Rapid Detection of Viral Antibodies Based on Multifunctional Staphylococcus Aureus Nanobioprobes. Enzyme Microb. Technol. 2016, 95, 94–99. [Google Scholar] [CrossRef] [PubMed]
  259. Weng, X.; Neethirajan, S. Immunosensor Based on Antibody-Functionalized MoS2 for Rapid Detection of Avian Coronavirus on Cotton Thread. IEEE Sens. J. 2018, 18, 4358–4363. [Google Scholar] [CrossRef] [PubMed]
  260. Ahmed, S.R.; Kang, S.W.; Oh, S.; Lee, J.; Neethirajan, S. Chiral Zirconium Quantum Dots: A New Class of Nanocrystals for Optical Detection of Coronavirus. Heliyon 2018, 4, e00766. [Google Scholar] [CrossRef] [Green Version]
  261. Ahmed, S.R.; Nagy, É.; Neethirajan, S. Self-Assembled Star-Shaped Chiroplasmonic Gold Nanoparticles for an Ultrasensitive Chiro-Immunosensor for Viruses. RSC Adv. 2017, 7, 40849–40857. [Google Scholar] [CrossRef] [Green Version]
  262. Pham, J.; Meyer, S.; Nguyen, C.; Williams, A.; Hunsicker, M.; McHardy, I.; Gendlina, I.; Goldstein, D.Y.; Fox, A.S.; Hudson, A.; et al. Performance Characteristics of a High-Throughput Automated Transcription-Mediated Amplification Test for SARS-CoV-2 Detection. J. Clin. Microbiol. 2020, 58, 10. [Google Scholar] [CrossRef] [PubMed]
  263. Mayer, F.J.; Ratzinger, F.; Schmidt, R.L.J.; Greiner, G.; Landt, O.; Am Ende, A.; Corman, V.M.; Perkmann-Nagele, N.; Watkins-Riedel, T.; Petermann, D.; et al. Development of a Fully Automated High Throughput PCR for the Detection of SARS-CoV-2: The Need for Speed. Virulence 2020, 11, 964–967. [Google Scholar] [CrossRef]
  264. LGC Biosearch Technologies. Biosearch Tech Fully Automated PCR System for COVID-19. Available online: https://www.biosearchtech.com/covid-19/fully-automated-pcr-system-for-covid-19-detection (accessed on 11 March 2021).
  265. Pan, M.; Mensah, F.; Kirk, D. 21st Century Technology Offers Hope in Covid-19 Pandemic. Available online: https://www.computeraidedbiology.com/cab-companies-on-covid19 (accessed on 13 June 2020).
  266. Zhang, Y.; Odiwuor, N.; Xiong, J.; Sun, L.; Nyaruaba, R.O.; Wei, H.; Tanner, N.A. Rapid Molecular Detection of SARS-CoV-2 (COVID-19) Virus RNA Using Colorimetric LAMP. medRxiv 2020. [Google Scholar] [CrossRef]
  267. Arumugam, A.; Wong, S.S. The Potential Use of Unprocessed Sample for RT-QPCR Detection of COVID-19 without an RNA Extraction Step. BioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  268. Xie, Q.; Wang, J.; You, J.; Zhu, S.; Zhou, R.; Tian, Z.; Wu, H.; Lin, Y.; Chen, W.; Xiao, L.; et al. Effect of Large-Scale Testing Platform in Prevention and Control of the COVID-19 Pandemic: An Empirical Study with a Novel Numerical Model. medRxiv 2020. [Google Scholar] [CrossRef]
  269. Liu, C.; Huang, B.; Zhou, R.; Hao, W.; Qiao, M.; Chen, C.; Zhang, W.; Dong, J.; Zhu, S.; Huang, J.; et al. The Application of a Testing Platform in High-Throughput Nucleic Acid Detection of SARS-CoV-2. OSF Prepr. 2020. [Google Scholar] [CrossRef]
  270. Hamilton Automated Liquid Handling Equipment: Four Platforms, Unlimited Solutions. Available online: https://www.hamiltoncompany.com/automated-liquid-handling (accessed on 21 August 2020).
  271. TECAN Lab Automation Solutions for Coronavirus Testing and Research. Available online: https://www.tecan.com/covid19 (accessed on 19 August 2020).
  272. PR-Newswire. Zymo Research Announces Partnership with Tecan for COVID-19 Viral RNA Extraction, CISION. 2020. Available online: https://www.prnewswire.com/news-releases/zymo-research-announces-partnership-with-tecan-for-covid-19-viral-rna-extraction-301033021.html (accessed on 1 September 2020).
  273. SynbioBeta Opentrons Partners with Zymo Research To Offer an Affordable, Automated COVID-19 Testing Platform. Available online: https://synbiobeta.com/opentrons-partners-with-zymo-research-to-offer-an-affordable-automated-covid-19-testing-platform (accessed on 21 August 2020).
  274. Zymo Research Announces Partnership with Tecan for COVID-19 Viral RNA. Available online: https://www.zymoresearch.com/blogs/press-releases/zymo-research-announces-partnership-with-tecan-for-covid-19-viral-rna-extraction (accessed on 8 April 2021).
  275. Zymo Research Opentrons Partners with Zymo Research to Offer an Affordable Automate. Available online: https://www.zymoresearch.com/blogs/press-releases/opentrons-partners-with-zymo-research-to-offer-an-affordable-automated-covid-19-testing-platform (accessed on 7 April 2021).
  276. Crone, M.A.; Priestman, M.; Ciechonska, M.; Jensen, K.; Sharp, D.J.; Randell, P.; Storch, M.; Freemont, P. A New Role for Biofoundries in Rapid Prototyping, Development, and Validation of Automated Clinical Diagnostic Tests for SARS-CoV-2. medRxiv 2020. [Google Scholar] [CrossRef]
  277. London Bio Foundry Covid-19 Resources. Available online: https://www.londonbiofoundry.org/covid19 (accessed on 3 September 2020).
  278. Hologic Panther Fusion Assays. Available online: https://www.hologic.com/hologic-products/diagnostic-solutions/panther-fusionr-assays (accessed on 8 April 2021).
  279. Maia Chagas, A.; Molloy, J.C.; Prieto-Godino, L.L.; Baden, T. Leveraging Open Hardware to Alleviate the Burden of COVID-19 on Global Health Systems. PLoS Biol. 2020, 18, e3000730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Villanueva-Cañas, J.L.; Gonzalez-Roca, E.; Gastaminza Unanue, A.; Titos, E.; Martínez Yoldi, M.J.; Vergara Gómez, A.; Puig Butillé, J.A. ROBOCOV: An Affordable Open-Source Robotic Platform for COVID-19 Testing by RT-QPCR. BioRxiv 2020. [Google Scholar] [CrossRef]
  281. Opentrons Hospital Clinic of Barcelona Uses Opentrons To Scale COVID-19 Testing. Available online: https://blog.opentrons.com/barcelona-hospital-clinic-uses-opentrons-robot-for-scalable-covid-19-testing (accessed on 21 August 2020).
  282. Crawford, E.D.; Acosta, I.; Ahyong, V.; Anderson, E.C.; Arevalo, S.; Asarnow, D.; Axelrod, S.; Ayscue, P.; Azimi, C.S.; Azumaya, C.M.; et al. Rapid Deployment of SARS-CoV-2 Testing: The CLIAHUB. PLoS Pathog. 2020, 16, e1008966. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structure of SARS-CoV-2 and diagnostic targets: SARS-CoV-2 structure has two components: viral genome (RNA) and viral proteins. Tests based on the detection of nucleic acids (RT-qPCR, RT-LAMP, CRISPR-Cas-based, NGS) target different regions of the viral RNA. Antigen tests aim at the detection of viral proteins such as spike protein (S). Serological tests detect host antibodies produced in response to SARS-COV-2 infection.
Figure 1. Structure of SARS-CoV-2 and diagnostic targets: SARS-CoV-2 structure has two components: viral genome (RNA) and viral proteins. Tests based on the detection of nucleic acids (RT-qPCR, RT-LAMP, CRISPR-Cas-based, NGS) target different regions of the viral RNA. Antigen tests aim at the detection of viral proteins such as spike protein (S). Serological tests detect host antibodies produced in response to SARS-COV-2 infection.
Diagnostics 11 01981 g001
Figure 2. Timeline of SARS-CoV-2 detection using nucleic acid-based, antigen and serological tests: Detection of viral nucleic acids is useful from 5 to 15 days after the onset of symptoms. Because many nucleic acid detection tests include nucleic acid amplification, these tests can give positive results for prolonged periods. Due to the nature of antibody production and the variability of immune responses between individuals, serological tests can be used for seroprevalence studies and surveillance from the second week after the onset of symptoms. Antigen tests are useful for the detection of infected individuals from 5 to 15 days after the onset of symptoms.
Figure 2. Timeline of SARS-CoV-2 detection using nucleic acid-based, antigen and serological tests: Detection of viral nucleic acids is useful from 5 to 15 days after the onset of symptoms. Because many nucleic acid detection tests include nucleic acid amplification, these tests can give positive results for prolonged periods. Due to the nature of antibody production and the variability of immune responses between individuals, serological tests can be used for seroprevalence studies and surveillance from the second week after the onset of symptoms. Antigen tests are useful for the detection of infected individuals from 5 to 15 days after the onset of symptoms.
Diagnostics 11 01981 g002
Figure 3. Nanomaterial tools for fighting future pandemics: (1) diagnostics, (2) vaccine delivery, (3) cell tramps, and (4) microfluidics are the future to fight against the pandemic. Created with BioRender.com (accessed date 15 January 2021).
Figure 3. Nanomaterial tools for fighting future pandemics: (1) diagnostics, (2) vaccine delivery, (3) cell tramps, and (4) microfluidics are the future to fight against the pandemic. Created with BioRender.com (accessed date 15 January 2021).
Diagnostics 11 01981 g003
Figure 4. General RT-qPCR test workflow and comparison of manual and automated nucleic acid extraction in Huo-Yan Lab. The use of MGISP-960 platform to automate nucleic acid extraction increased the testing capacity, simplified the steps for extraction, and reduced the time per test. Adapted from [146,269].
Figure 4. General RT-qPCR test workflow and comparison of manual and automated nucleic acid extraction in Huo-Yan Lab. The use of MGISP-960 platform to automate nucleic acid extraction increased the testing capacity, simplified the steps for extraction, and reduced the time per test. Adapted from [146,269].
Diagnostics 11 01981 g004
Figure 5. Schematic illustrations of three different diagnostic workflows from patient samples: A thermal cycler device is necessary in the case of RT-qPCR tests, while plate reader or other alternatives are available for measuring the readout in the case of CRISPR-Cas and LAMP tests. Adapted from [277].
Figure 5. Schematic illustrations of three different diagnostic workflows from patient samples: A thermal cycler device is necessary in the case of RT-qPCR tests, while plate reader or other alternatives are available for measuring the readout in the case of CRISPR-Cas and LAMP tests. Adapted from [277].
Diagnostics 11 01981 g005
Figure 6. ROBOCOV circuit: Run preparation is done by using open-source Python codes. Initial sample setup, sample preparation, and plate filling are performed by OT-2 Station A, B1, and B2, respectively. RNA extraction is processed by KingFisher Flex. Then, qPCR mic preparation is done by OT-2 Station C, and, finally, the qPCR process is run by ABI 7500. The analysis results are exported as a user-friendly R file. Adapted from [280].
Figure 6. ROBOCOV circuit: Run preparation is done by using open-source Python codes. Initial sample setup, sample preparation, and plate filling are performed by OT-2 Station A, B1, and B2, respectively. RNA extraction is processed by KingFisher Flex. Then, qPCR mic preparation is done by OT-2 Station C, and, finally, the qPCR process is run by ABI 7500. The analysis results are exported as a user-friendly R file. Adapted from [280].
Diagnostics 11 01981 g006
Table 1. Diagnostic tests based on detection of nucleic acids.
Table 1. Diagnostic tests based on detection of nucleic acids.
Test NameManufacture CountryTechnologyLoDSampleApprovalRef
1copy COVID-19 qPCR Multi KitKorea
1 drop
RT-qPCR0.2 copies/µLNasopharyngeal and nasal swab and washEUA 1/5/2020[40]
TRUPCR SARS-CoV-2 KitIndia
3B Blackbio Biotech India Kilpest India subsidiary
RT-qPCR10 copies/µLNasopharyngeal and oropharyngeal swabs, anterior nasal swab, and mid-turbinate nasal swabs, nasopharyngeal aspirates/washes or nasal aspirates, and bronchoalveolar lavage EUA 18/6/2020[41]
SARS-CoV-2 and Influenza A and B RT-qPCR Detection KitChina
3D Medicines
RT-qPCR5 copies/reaction-CE mark 3/2020[42]
Abbott RealTime SARS-CoV-2 EUA testUSA
Abbott
RT-qPCR0.1 copies/µLNasopharyngeal swabs, Oropharyngeal swabsEUA 18/3/2020[43]
CareStart COVID-19 MDx RT-PCRUSA
Access Bio
RT-qPCR5.4 NDU/µLNasopharyngeal, oropharyngeal and nasal swabs, and nasopharyngeal wash/aspirate or nasal aspirate and bronchoalveolar lavageEUA 7/7/2020[44]
Acupath COVID-19 Real-Time (RT-PCR) AssayUSA
Acupath Laboratories
RT-qPCR25 copies/µLNasopharyngeal swab, nasopharyngeal aspirate, and bronchoalveolar lavageEUA 29/6/2020[45]
UltraGene Combo2Screen SARS-CoV-2 assayLuxemburg
Advanced Biological Laboratories
RT-qPCR10E-6 TCID50/mLNasopharyngeal swabEUA submission pending, CE mark 5/2020[46]
RealStar SARS-CoV-2 RT-PCR Kits (1.0 and U.S versions)Germany
Altona Diagnostics
RT-qPCR1.0E-01 PFU/mLHuman respiratory swabs EUA 22/4/2020, CE mark 4/2020[47]
Altru Dx SARS-CoV-2 RT-PCR assayUSA
Altru Diagnostics
RT-qPCR0.625 copies/µLNasal, midturbinate, nasopharyngeal, and oropharyngeal swabEUA 30/4/2020[48]
Bosphore Novel Coronavirus (2019-nCoV) Detection KitTurkey
Anatolia Geneworks
RT-qPCR-Human respiratory sampleCE mark 2/2020[49]
BioCode SARS-CoV-2 AssayUSA
Applied BioCode
RT-qPCR1.7E-2 TCID50/mLNasopharyngeal swabs (NPS), oropharyngeal swabs (OPS), nasal swabs or bronchoalveolar lavage EUA 6/15/2020[50]
Linea COVID-19 RT-PCR testUSA
Applied DNA Sciences
RT-qPCR1.25 copies/µLNasopharyngeal and oropharyngeal swabs, mid-turbinate nasal swabs, nasopharyngeal washes/aspirates or nasal aspirates, and bronchoalveolar lavageEUA 5/13/2020[51]
Aspirus SARS-CoV rRT-PCR AssayUSA
Aspirus Reference Laboratory
RT-qPCR0.5 copies/µLNasal, mid-turbinate, nasopharyngeal, oropharyngeal swab specimens and bronchoalveolar lavage specimensEUA 6/1/2020[52]
Assurance SARS-CoV-2 PanelUSA
Assurance Scientific Laboratories
RT-qPCR9 copies/µLNasal, nasopharyngeal, or oropharyngeal swabsEUA 5/15/2020[53]
A*STAR Fortitude 2.0Singapore A*STAR, Tan Tock Seng Hospital of SingaporeRT-qPCR25 copies/reaction Nasal pharyngeal swabSingapore Health Sciences Authority provisional authorization[54]
Avera Institute for Human Genetics SARS-CoV-2 AssayUSA
Avera Institute for Human Genetics
RT-qPCR1.6 copies/µLNasopharyngeal, nasal, and oropharyngeal swab specimensEUA 5/22/2020[55]
Avellino SARS-CoV-2/COVID-19 (AvellinoCoV2)USA
Avellino Lab
RT-qPCR18 NDU/µLNasopharyngeal swab and oropharyngeal swabEUA 3/25/2020[56]
Viro-Q SARS-CoV-2 kitGermany
BAG Diagnostics
RT-qPCR--CE mark 4/2020[57]
ThermoFisher—Applied Biosystems TaqPath COVID-19 Combo KitUSA
Rutgers University Clinical Genomics Laboratory
RT-qPCR0.250 GCE/µLNasopharyngeal swabs, nasopharyngeal aspirate (nasal aspirate), and nasopharyngeal aspirate (nasal aspirate), and bronchoalveolar lavage (BAL)EUA 4/10/2020[58]
Cobas SARS-CoV-2 Test USA
Roche
RT-qPCR0.003 TCID50/mLNasopharyngeal and oropharyngeal swab samplesEUA 3/12/2020, CE mark 2020[59]
GSD NovaPrime SARS-CoV-2 (COVID-19) Real-Time PCR testHungary
Gold Standard Diagnostics/Eurofins Technologies
RT-qPCR3.75 copies/reactionNasal wash/swab, nasopharyngeal wash/swab, oropharyngeal swab and bronchoalveolar lavageCE mark 5/2020[60]
QiaStat-Dx Respiratory SARS-CoV-2 PanelUSA
Qiagen
RT-qPCR180 NDU/µLNasopharyngeal swabunder CDC’s EUA[61]
Rida Gene SARS-CoV-2Germany
R-Biopharma
RT-qPCR50 copies/reactionHuman throat and nasopharyngeal swabsCE mark 5/2020[62]
Standard M nCoV Real-Time Detection KitKorea
SD Biosensor
RT-qPCR0.25 copies/μLNasoropharyngeal, nasal, and mid-turbinate nasal swab, and sputumEUA 4/23/2020[63]
ExProbe SARS-CoV-2 Testing KitTaiwan
TBG Biotechnology
RT-qPCR-Nasopharyngeal and oropharyngeal swabs, anterior nasal and mid-turbinate nasal swabsEUA 6/10/2020[64]
Quick SARS-CoV-2rRT-PCR KitUSA
Zymo Research
RT-qPCR2.5 E3 GEC/μLNasal, nasopharyngeal, mid-turbinate or oropharyngeal swabs), and lower respiratory specimens (such as sputum, tracheal aspirates, and bronchoalveolar lavage)EUA 5/7/2020[65]
Fulgent COVID-19 by RT-PCR TestUSA
Fulgent Genetics/Fulgent Therapeutics
RT-qPCR20 copies/µLNasal, nasopharyngeal, and oropharyngeal swabsEUA 5/15/2020[66]
SARS-CoV-2 RT-PCR testGermany
Centogene
RT-qPCR5 copies/µlDry oropharyngeal swabsEUA 7/1/2020[67]
Xpert Xpress SARS-CoV-2 testUSA
Cepheid
RT-qPCR0.0001 PFU/µLNasopharyngeal, oropharyngeal, nasal, or mid-turbinate swab and/or nasal wash/aspirateEUA 3/20/2020[68]
HDPCR SARS-CoV-2 real-time PCR assayUSA
ChromaCode
RT-qPCR0.250 copies/µLNasopharyngeal swabs oropharyngeal swabs, anterior nasal swabs, midturbinate nasal swabs, nasal aspirate, nasal wash, and bronchoalveolar lavage (BAL) specimens EUA 6/9/2020[69]
Hymon SARS-CoV-2 Test KitChina
Dba SpectronRx
RT-qPCR1.2 copies/µlNasal, mid-turbinate, nasopharyngeal, and oropharyngeal swab specimens)EUA 5/22/2020[70]
DSL COVID-19 AssayUSA
Diagnostic Solutions Laboratory
RT-qPCR18 NDU/µLNasopharyngeal swabEUA 6/25/2020[71]
Simplexa COVID-19 DirectUSA
DiaSorin Molecular
RT-qPCR6 NDU/µLNasal swab, nasopharyngeal swab, nasal wash/aspirate, and BALEUA 3/19/2020, CE mark 4/2020[72]
Lilly SARS-CoV-2 AssayUSA
Eli Lilly
RT-qPCR1 copy/µLNasopharyngeal swabs, oropharyngeal (throat) swabs, anterior nasal swabs, mid-turbinate nasal swabs, nasal aspirates, nasal washes and bronchoalveolar lavageEUA 7/27/2020[73]
Ampiprobe SARS-CoV-2 Test SystemUSA
Enzo Biochem/Enzo Life Sciences
RT-qPCR0.280 copies/µLNasopharyngeal swabsEUA 7/7/2020[74]
Euroimmun/PerkinElmerUSA
EuroRealTime SARS-CoV-2
RT-qPCR0.150 copies/µLNasal, mid-turbinate, nasopharyngeal, oropharyngeal swabs and bronchioalveolar lavageEUA 6/8/2020, CE mark 3/2020[75]
Fosun COVID-19 RT-PCR Detection KitChina
Fosun Pharma USA
RT-qPCR0.3 copies/µLAnterior nasal swabs, mid-turbinate nasal swabs, nasopharyngeal swabs, oropharyngeal swabs, sputum, lower respiratory tract aspirates, bronchoalveolar lavage, and nasopharyngeal wash/aspirate or nasal aspirateEUA 4/17/2020[76]
NeoPlex COVID-19 Detection KitKorea
GeneMatrix
RT-qPCR5.4 NDU/µLNasopharyngeal swabsEUA 5/14/2020[77]
GB SARS-CoV-2 Real-Time RT-PCRTaiwan
General Biologicals
RT-qPCR0.1 copies/µLNasopharyngeal swabsCE mark 4/2020[78]
Genetron SARS-CoV-2 RNA TestChina
Genetron
RT-qPCR10 copies/µLOropharyngeal, nasopharyngeal, anterior nasal and mid-turbinate nasal swabEUA 6/5/2020[79]
Helix COVID-19 testUSA
Helix
RT-qPCR1 GCE/µLNasopharyngeal and oropharyngeal swabsEUA 7/23/2020[80]
Panther Fusion SARS-CoV-2 assayUSA
Hologic
RT-qPCR1 × 10−5 TCID50/µLNasopharyngeal (NP), nasal, oropharyngeal (OP) swab specimens and lower respiratory tract (LRT) specimensEUA 3/16/2020[81]
Smart Detect SARS-CoV-2 rRT-PCR KitUSA
InBios International
RT-qPCR12 GCE/reactionNasopharyngeal swab, anterior nasal swab and mid-turbinate nasal swabEUA 4/7/2020[82]
IDT 2019-novel coronavirus kitUSA
Integrated DNA Technologies/Danaher
RT-qPCR-Oropharyngeal, nasopharyngeal, anterior nasal and mid-turbinate nasal swabunder CDC’s EUA[83]
SARS-CoV-2 AssayUSA
Integrity Laboratories
RT-qPCR2.5 copies/µLNasal, nasopharyngeal and oropharyngeal swab EUA 4/13/2020[84]
ProTect Covid-19 kitSingapour
JN Medsys
RT-qPCR<2%Upper respiratory nasopharyngeal swabsUnder CDC’s EUA, CE mark 4/2020, Singapore HAS provisional authorization; Philippines FDA[85]
COVID-19 Coronavirus Real Time PCR KitChina
Jiangsu Bioperfectus Technologies
RT-qPCR0.350 copies/µLNasopharyngeal swabs, oropharyngeal (throat) swabs, anterior nasal swabs, mid-turbinate nasal swabs, nasal aspirates, nasal washes, bronchoalveolar lavage (BAL) fluid and sputumEUA 18/6/2020[86]
Curative-Korva SARS-Cov-2 AssayUSA
KorvaLabs
RT-qPCR0.200 copies/µLOropharyngeal (throat) swab, nasopharyngeal swab, nasal swab, and oral fluid specimensEUA 16/4/2020[87]
Idylla SARS-CoV-2 TestBelgium
Biocartis
RT-qPCR0.5 copies/µLNasopharyngeal swabCE mark 11/2020[88]
BioFire Respiratory Panel 2.1-EZ (RP2.1-EZ)USA
BioMérieux/BioFire Diagnostics
RT-qPCR6 NDU/µLNasopharyngeal swabEUA 2/10/2020 [89]
VitaPCR Influenza A and B/SARS-CoV-2 assay Singapore
Credo Diagnostics Biomedical
RT-qPCR2.73 copies/μlNasopharyngeal swabCE mark 10/2020[90]
Genetworx Covid-19 Nasal Swab TestUSA
Genetworx
RT-qPCR0.274 copies/µLNasal swabEUA 15/12/2020[91]
EuroRealTime SARS-CoV-2/Influenza A/BGermany
Euroimmun/PerkinElmer
RT-qPCR1.8 NDU/µLThroat swabCE mark 12/2020[92]
MassArray SARS-CoV-2 PanelUSA
Agena Bioscience
RT-PCR/MALDI-TOF0.3 copies/µLNasopharyngeal swab, oropharyngeal swab, and BALEUA 10/26/2020 CE mark 9/2020[93]
Bio-Rad SARS-CoV-2 ddPCR TestUSA
Bio-Rad Laboratories
ddPCR0.625 copies/µLNasopharyngeal, anterior nasal and mid-turbinate swab specimens as well as nasopharyngeal wash/aspirate and nasal aspirate specimensEUA 1/5/2020[94]
iAMP COVID-19 Detection KitUSA
Atila BioSystems
Isothermal amplification (OMEGA amplification)10 copies/µLNasal, nasopharyngeal (NP), and oropharyngeal (OP) swabsEUA 10/4/2020[95]
ID Now COVID-19USA
Abbott
Isothermal amplification (proprietary enzymes)0.125 GCE/µLDirect nasal, nasopharyngeal or throat swabsEUA 27/3/2020[96]
Color SARS-CoV-2 LAMP Diagnostic AssayUSA
Color
RT-LAMP0.75 copies/µLNasopharyngeal (NP) swabs, oropharyngeal (OP) swabs, anterior nares (AN) swabs, mid-turbinate nasal (MTN) swabs, NP wash/aspirate or nasal aspirates, and bronchoalveolar lavage specimensEUA 5/18/2020, amended 24/7/2020[97]
2019-nCoV detection kitChina
Rendu Biotechnology
RT-LAMP-Nasal, nasopharyngeal, and oropharyngeal swabChina NMPA 3/2020[98]
AQ-TOP COVID-19 Rapid Detection KitKorea
Seasun Biomaterials
RT-LAMP7 copies/µLOropharyngeal and nasopharyngeal swab specimens, anterior nasal and mid-turbinate nasal swabs, nasopharyngeal wash/aspirate or nasal aspirate specimens, bronchoalveolar lavage (BAL) and sputum from individualsEUA 21/5/2020[99]
Lucira HealthUSA
Lucira COVID-19 All-In-One Test Kit
RT-LAMP0.9 copies/µLNasal swabEUA 17/11/2020[100]
Poplar SARS-CoV-2 TMA Pooling assayUSA
Poplar Healthcare
TMA-Nasal, nasopharyngeal, and oropharyngeal swab EUA 3/8/2020[101]
COVIDSeq TestUSA
Illumina
Next generation gene sequencing5.4 NDU/µLNasopharyngeal (NP), oropharyngeal (OP), and mid-turbinate (MT) nasal swabsEUA 9/6/2020[102]
Fulgent COVID-19 by NGSUSA
Fulgent Genetics/MedScan Laboratory
Next generation gene sequencing3.6 NDU/µLNasal, nasopharyngeal, and oropharyngeal swabsEUA submission pending[103]
Sherlock CRISPR SARS-CoV-2 kitUSA
Sherlock Biosciences
RT-LAMP and CRISPR-Cas 136.75 copies/µL Nasopharyngeal and oropharyngeal swab samplesEUA 6/5/2020[104]
SARS-CoV-2 RNA DETECTR AssayUSA
UCSF Clinical Labs at China Basin
CRISPR-Cas1220 copies/µLNasopharyngeal swabs, oropharyngeal (throat) swabs, mid-turbinate nasal swabs, anterior nasal swabs, nasopharyngeal wash/aspirate or nasal aspirateEUA 9/7/2020[105]
LoD: Limit of detection. TCID50: Median Tissue Culture Infectious Dose. GCE: Genome copy equivalents. NDU: RNA NAAT detectable units. PFU: plaque-forming unit.
Table 2. Serological tests that have some certification to date, FDA. Include: % sensitivity, %, specificity, cross reactivity, interferences, mark, country.
Table 2. Serological tests that have some certification to date, FDA. Include: % sensitivity, %, specificity, cross reactivity, interferences, mark, country.
DiagnosticCountry
Manufacturer
TechSpecificitySensitivityCross ReactivityInterferencesDate EUA IssuedRef
Babson Diagnostics aC19G1USA
Babson Diagnostics, Inc.
IgG CLIA100%100%Anti-HIV 1 + 2, Anti-HCV, CMV IgG, Anti-HBs, Anti-HAV EIAN/A23/6/2020[196]
Access SARS-CoV-2 IgGUSA
Beckman Coulter, Inc.
IgG CLIA99.6%96.8%Human chorionic gonadotropin (hCG), HIV antibody, Influenza antibody, Influenza A antibody, Influenza B antibody, Measles antibody, Mycoplasma pneumoniae IgG, Parvovirus B19 antibody, Respiratory pathogen antibodies, Respiratory syncytial virus (RSV) antibodyHemoglobin, Bilirubin (conjugated), Bilirubin (unconjugated), Triglycerides (Intralipid)26/6/2020[197]
Diazyme DZ-Lite SARS-CoV-2 IgG CLIA KitUSA, Diazyme Laboratories, Inc.IgG CLIA97.4%100%Influenza A H1N1 IgM/IgG, Influenza A H7N9 IgM/IgG, Rhinovirus Type A IgM/IgG, Rotavirus IgM/IgG, Human coronavirus HKU1 IgM/IgG, Human coronavirus NL63 IgM/IgG, ANATriglycerides, Hemoglobin, Rheumatoid Factor, Anti-Mitochondrial, HAMA, Total IgG, Total IgM, Interferon α, Ribavirin8/7/2020[198]
SCoV-2 Detect IgG ELISAUSA, InBios International, Inc.IgG ELISA98.9%97.8%VIH 1 y 2, hepatitis C and BHemoglobin Bilirubin Triglycerides Cholesterol10/6/2020[199]
SARS-CoV-2 RBD IgG testUSA, Emory Medical LaboratoriesIgG ELISA96.4%100%Anti-Influenza B, Anti-HCV, Anti-HBV, Anti-Haemophilus, Anti-Rhinovirus influenzae, ANA, Anti-HIVN/A15/6/2020[200]
xMAP SARS-CoV-2 Multi-Antigen IgG AssayUSA, Luminex CorporationIgG FMIA99.3%96.3%N/A(dipotassium EDTA)16/7/2020[201]
ADVIA Centaur SARS-CoV-2 IgG (COV2G)Germany, Siemens Healthcare Diagnostics Inc.IgG Semi-quantitative99.9%100%Chlamydia trachomatis IgM, Cytomegalovirus (CMV) IgM, Epstein Barr virus (EBV) IgG, Epstein Barr virus (EBV) IgM, Hepatitis A virus (HAV) IgM, Hepatitis B core (Anti-HBc) IgM, Hepatitis B core (Anti-HBc) total antibody, Hepatitis C virus (HCV) antibody, Herpes simplex virus (HSV) IgMHemoglobin, Bilirubin (conjugated), Bilirubin (unconjugated), Triglycerides (Intralipid), Biotin31/7/2020[202]
Atellica IM SARS-CoV-2 IgG (COV2G)Germany, Siemens Healthcare Diagnostics Inc.IgG Semi-quantitative99.9%100%Human chorionic gonadotropin (hCG), Human immunodeficiency virus (HIV) antibody, Influenza antibody, Influenza A antibody, Influenza B antibody, Measles antibodyHemoglobin, Bilirubin (conjugated), Bilirubin (unconjugated), Triglycerides (Intralipid), Biotin, Cholesterol, Protein, total7/31/2020[203]
Anti-SARS-CoV-2 ELISA (IgG)Germany, EUROIMMUN US Inc.Serology IgG99.6%86.7%Influenza, Acute EBV infection, Rheumatoid factor, other Human coronavirusHemoglobin, triglycerides, and bilirubin 4/5/2020[204]
SARS-CoV-2 IgG assayUSA, Abbott Laboratories Inc.Serology IgG99%100%CMV, Hepaitis A,B, RSV, Rubella, Herpes virusAdenovirus, pregnant female, lupus26/4/2020[205]
LIAISON SARS-CoV-2 S1/S2 IgGItaly, DiaSorin Inc.Serology IgG99.3%97.6%Anti-Human CoV OC43; Anti-Human CoV HKU1, 4 Anti-Human CoV unknown strain.Triglycerides, Hemoglobin, Unconjugated bilirubin24/4/2020[206]
VITROS Immunodiagnostic Products Anti-SARS-CoV-2 IgG Reagent PackUSA, Ortho-Clinical Diagnostics, Inc.Serology IgG100%83.3%Adenovirus Antibody, Influenza A IgG, Influenza A IgM, Influenza B IgG, Influenza B IgM, Coxsackie Virus Antibody, Echovirus Antibody, Polio Virus, Anti-respiratory syncytial virus (RSV), HCV Antibody, Anti Nuclear AntibodyBilirubin (conjugated), Bilirubin (unconjugated), Biotin, Hemoglobin, Intralipid24/4/2020[207]
COVID-19 ELISA IgG Antibody TestUSA, Mount Sinai LaboratorySerology IgG100%92%Varicella, Infleunza, Hepatitis, HIV, CMVAscorbic Acid, Hemoglobin, Bilirubin, Albumin, Triglyceride15/4/2020[208]
SCoV-2 Detect IgM ELISAUSA, InBios International, Inc.IgM ELISA100%100%Anti-Influenza A/B, Anti-Hepatitis B, Anti-Hepatitis C, Anti-Nuclear Antibody, Rheumatoid Factor, Human Anti-Mouse Antibody, Anti-HIV, Anti-Respiratory Syncytial Virus, Normal Human SeraHemoglobin, Bilirubin (conjugated and unconjugated), Triglycerides, Buffer (SDB), Cholesterol30/6/2020[199]
VIDAS SARS-CoV-2 IgMFrance, bioMérieux SAIgM ELFA99.4%100%SARS-CoV(-1) Infection (2005), SARS-CoV(-1) Infection (2020), HCoV-NL63 Infection, HCoV-229E Infection, HCoV-OC43 Infection HCoV-HKU1 Infection MERS-CoV Infection, Acute EBV infections with heterophile antibodies, ANAHemoglobin, Lipids, Albumin, Bilirubin (conjugated), Bilirubin (unconjugated)6/8/2020[209]
RightSign COVID-19 IgG/IgM Rapid Test CassetteChina, Hangzhou Biotest Biotech Co., Ltd.IgM and IgG Lateral Flow100%93.3%Anti-FLU A, Anti-FLU B, Anti-Respiratory Syncytial, Virus Anti-Adenovirus, Anti-HBsAg, Anti-Syphilis, Anti-H. Pylori, Anti-HIV, Anti-HCV, Anti-SARS-COVAcetaminophen, Caffeine, Albumin, Acetylsalicylic Acid, Gentisic Acid, Ethanol, Ascorbic Acid4/6/2020[210]
LYHER Novel Coronavirus (2019-nCoV) IgM/IgG Antibody Combo Test Kit (Colloidal Gold)China, Hangzhou Laihe Biotech Co., Ltd.IgM and IgG Lateral Flow98.8%96.7%H1N1-1~H1N1-3, H7N9-1~H7N9-2, ANA-1, Staphyl. -1~Staphyl.-2, EBV-1~EBV-5, RSV-1~RSV-2, Chlamydia-1~Chlamydia-3N/A19/6/2020[211]
Assure COVID-19 IgG/IgM Rapid Test DeviceChina, Assure Tech. (Hangzhou Co., Ltd.)IgM and IgG Lateral Flow100%98.8%Human coronavirus HKU1 IgM/IgG, Human coronavirus NL63 IgM/IgGIcteric (Bilirubin), Lipemicicines Acetylsalicylic acid Ascorbic acid (Vitamin C), Amoxicillin6/7/2020[212]
Sienna-Clarity COVIBLOCK COVID-19 IgG/IgM Rapid Test CassetteFinland, Salofa OyIgM and IgG Lateral Flow99.2%94.9%HCV, HBV, ANA MetapneumovirusAscorbic Acid, Hemoglobin, Bilirubin, Albumin, Triglyceride13/7/2020[213]
Rapid COVID-19 IgM/IgG Combo Test KitUSA, Megna Health, Inc.IgM and IgG Lateral Flow97.5%100%Influenza A Virus, Influenza B Virus, Adenovirus, Rotavirus and Mycoplasma Pneumoniae.Rheumatoid Factor, Bilirubin, Triglyceride, Hemoglobin17/7/2020[211]
CareStart COVID-19 IgM/IgGUSA, Access Bio, Inc.IgM and IgG Lateral Flow100%98.8%Anti-Influenza A, Anti-HCV, Anti-229E (alpha coronavirus), Anti-OC43 (beta coronavirus), Anti-HKU1 (beta coronavirus), Antinuclear antibodies (ANA), Anti-respiratory syncytial virus, Anti-RhinovirusAcetaminophen, HAMA, Acetylsalicylic acid, Hemoglobin, Albendazole, Ibuprofen, Chloroquine diphosphate, Rifampicin24/7/2020[214]
BIOTIME SARS-CoV-2 IgG/IgM Rapid Qualitative TestChina, Xiamen Biotime Biotechnology Co., Ltd.IgM and IgG Lateral Flow98.8%100%Anti-OC43 (beta coronavirus), Anti-HKU1 (beta coronavirus), Antinuclear antibodies (ANA)N/A24/7/2020[215]
Vibrant COVID-19 Ab AssayUSA, Vibrant America Clinical LabsIgM and IgG CLIA98.3%97.1%Anti-Influenza A, Anti- Influenza B, Anti-HCV, Anti-HBV, ANA, Anti-respiratory syncytial virus, Anti-Haemophilus influenzaeBilirubin, Triglycerides Hemoglobin Rheumatoid Factor (RF) Cholesterol, HAMA, Ribavirin, Levofloxacin, Azithromycin, Ceftriaxone sodium4/6/2020[216]
Anti-SARS-CoV-2 Rapid TestChina, Autobio Diagnostics Co. Ltd.Serology IgM and IgG97%93%Other human coronavirus, InfluenzaHAMA positive sample, Rheumatoid factor, Antinuclear antibody (ANA), Anti-mitochondrial antibody (AMA), Bilirubin24/4/2020[217]
DPP COVID-19 IgM/IgG SystemUsa, Chembio Diagnostic System, Inc.Serology IgM and IgG100%95%Human coronavirus, Zika, Influenza A, B, Mononucleosis, Chikungunya, Yellow fever virus, Dengue virusBiotin Hemoglobin14/4/2020[218]
qSARS-CoV-2 IgG/IgM Rapid TestUSA, Cellex Inc.Serology IgM and IgG96.0%93.8%Human coronavirus, Adenovirus, Influenza A, B, Rhinovirus, Chlamydia pneumoniae, Streptococcus pneumoniae, Mycobacterium tuberculosis, Mycoplasma pneumoniaeHemoglobin, Bilirubin (conjugated), Triglycerides, Cholesterol1/4/2020[219]
COVID-19 IgG/IgM Rapid Test Cassette (Whole Blood/Serum/Plasma)USA, Healgen Scientific LLCSerology IgM and IgG96.7%97%Influenza A virus IgG, Influenza B virus IgG, Respiratory syncytial virus IgG, Adenovirus IgG, Rhinovirus IgG, Human metapneumovirus IgG, Mycoplasma pneumoniae IgG, Chlamydia pneumoniae IgG, HCV IgG, Haemophilus influenza IgG, HBV core antibody IgGAscorbic Acid, Hemoglobin, Bilirubin, Albumin, Triglyceride29/5/2020[210]
Dimension EXL SARS-CoV-2 Total antibody assay (CV2T)USA, Siemens Healthcare Diagnostics Inc.Total Antibody CLIA99.9%100%Anti-influenza A, Anti-influenza B, AntiHBV Antinuclear, antibody (ANA), Hepatitis B core antigen (anti-HBc) IgM, Hepatitis B surface antigen (HBs Ag), Hepatitis C virus (HCV) antibody, HIV antibody, influenza antibodyHemoglobin, Bilirubin, conjugated, Bilirubin, Lipemia (Intralipid)8/6/2020[220]
WANTAI SARS-CoV-2 Ab ELISAChina, Beijing Wantai Biological Pharmacy Enterprise Co., Ltd.Total Antibody ELISA100%94.5%HBsAg and antibodies to HIV 1/2, HCV, TPN/A5/8/2020[221]
WANTAI SARS-CoV-2 Ab Rapid TestChina, Beijing Wantai Biological Pharmacy Enterprise Co., Ltd.Total Antibody Lateral Flow98.8%100%alpha COV 229E, alpha COV NL63, beta COV OC43, beta COV HKU1 Flu A, Flu B, HCV, HBV, ANAN/A10/7/2020[222]
New York SARS-CoV Microsphere Immunoassay for Antibody DetectionUSA, Wadsworth Center, New York State Department of HealthSerology Total Antibody99.1%99%HCV, HIV, Measles, Mumps, Rubella, Varicella Zoster virus, WNV, Herpes Simplex virus, ZIKV, Enterovirus, Rheumatoid factorBiotin Hemoglobin30/4/2020[223]
Platelia SARS-CoV-2 Total Ab assayUSA, Bio-Rad Laboratories, Inc.Serology Total Antibody98%99%Influenza, Mycoplasma pneumoniae, Rheumatoid factor, Other human coronavirus 29/4/2020[224]
Elecsys Anti-SARS-CoV-2Germany, Roche DiagnosticsSerology Antibody99.8%100%Other human coronavirus, InfluenzaBiotin Hemoglobin2/5/2020[225]
Table 3. FDA-approved antigen-based diagnostic tests.
Table 3. FDA-approved antigen-based diagnostic tests.
DiagnosticCountry,
Manufacturer
TechnologySpecificitySensitivityRef
Lateral Flow, Visual ReadUSA, Ortho clinical Diagnostic, Inc.Chemiluminescence immunoassay98.50%97.10%[232]
LumiraDx SARS·CoV-2 Ag TestSpain, LumiraDx UK Ltd.Chromatographic Digital immunoassay96.50%98%[233]
BD Veritor System for Rapid DetectionEngland, Becton DickinsonChromatographic immunoassay100%84%[234]
Clip COVID Rapid Antigen TestUSA, Luminostics, Inc.Lateral Flow, Immunoluminescent Antigen100%100%[235]
CareStart COVID-19 Antigen testUSA, Access Bio, Inc.Lateral Flow, Antigen98%99%[44]
Ellume COVID-19 Home TestAustralia, Ellume Limited TestLateral Flow, Fluorescence100%100%[236]
Sofia 2 SARS Antigen FIAGermany, Quidel CorporationLateral Flow, Fluorescence Antigen100%80%[237]
QuickVue SARS Antlgen TestUSA, Quidel CorporatlonLateral Flow, Visual Read98%98%[238]
BinaxNOW COVID-19 Ag Card HomeUSA, Abbott DiagnosticsLateral Flow, Visual Read98%98%[239]
BinaxNOW COVID-19 Ag cardUSA, Abbott Diagnostics Scarborough, Inc.Lateral Flow, Visual Read98.50%97.10%[240]
Sofía 2 Flu+ SARS Antlgen FIAGermany, Quidel CorporatlonLateral Flow, Fluorescence, Instrument Read, Multl-Analyte100%80%[241]
Sampinute COVID-19 Antlgen MIAUSA, Celltrion, Inc.Magnetic force assisted electrochemical100%100%[242]
Simoa SARS-CoV-2 N Protein Antigen testUSA, QuanterixParamagnetic Microbead-based98.50%97.10%[243]
Table 4. Nanoparticle-based tests to detect several coronaviruses.
Table 4. Nanoparticle-based tests to detect several coronaviruses.
VirusNPsReadout/OutputLoDT (min)Ref
SARS-CoVAuNPsImmunosensing10 fmol30[253]
PEDVAU NPsNano nest PCR2.21 × 10−7 ngμl−145[254]
Mers CoVAgNPsColorimetric1.53 nM20[255]
Mers CoVAgNPsElectrochemiluminescence1.0 pg/mL−120[256]
Mers CoVS.aureus nanobioparticlesAgglutination test0.9305555620[257]
IBVMoS2 nanosheetsImmunosensing4.6 × 102 EID50 per mL20[258]
IBVQd-MP NPs and Zr NPsPhotoluminesce79.15 EID/50 mL20[259]
IBVCAu NPsChiroimmunosensing47.91 EID/50 mL20[260]
IBVColloidal Au NPsICS10 4.4 EID/50 mL45[257]
HCoVAgNPsElectrochemiluminescence0.4 pgml−120[257,261]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guaman-Bautista, L.P.; Moreta-Urbano, E.; Oña-Arias, C.G.; Torres-Arias, M.; Kyriakidis, N.C.; Malcı, K.; Jonguitud-Borrego, N.; Rios-Solis, L.; Ramos-Martinez, E.; López-Cortés, A.; et al. Tracking SARS-CoV-2: Novel Trends and Diagnostic Strategies. Diagnostics 2021, 11, 1981. https://doi.org/10.3390/diagnostics11111981

AMA Style

Guaman-Bautista LP, Moreta-Urbano E, Oña-Arias CG, Torres-Arias M, Kyriakidis NC, Malcı K, Jonguitud-Borrego N, Rios-Solis L, Ramos-Martinez E, López-Cortés A, et al. Tracking SARS-CoV-2: Novel Trends and Diagnostic Strategies. Diagnostics. 2021; 11(11):1981. https://doi.org/10.3390/diagnostics11111981

Chicago/Turabian Style

Guaman-Bautista, Linda P., Erick Moreta-Urbano, Claudia G. Oña-Arias, Marbel Torres-Arias, Nikolaos C. Kyriakidis, Koray Malcı, Nestor Jonguitud-Borrego, Leonardo Rios-Solis, Espiridion Ramos-Martinez, Andrés López-Cortés, and et al. 2021. "Tracking SARS-CoV-2: Novel Trends and Diagnostic Strategies" Diagnostics 11, no. 11: 1981. https://doi.org/10.3390/diagnostics11111981

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