Next Article in Journal
Flavonone 3-hydroxylase Relieves Bacterial Leaf Blight Stress in Rice via Overaccumulation of Antioxidant Flavonoids and Induction of Defense Genes and Hormones
Next Article in Special Issue
Innate Immune Response to SARS-CoV-2 Infection: From Cells to Soluble Mediators
Previous Article in Journal
Evidence for Biological Age Acceleration and Telomere Shortening in COVID-19 Survivors
Previous Article in Special Issue
Could Histamine H1 Receptor Antagonists Be Used for Treating COVID-19?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nucleic Acid Testing of SARS-CoV-2

1
Microbiological Analysis Team, Biometrology Group, Korea Research Institute of Standards and Science (KRISS), Daejeon 34113, Korea
2
Department of Bio-Analytical Science, University of Science & Technology (UST), Daejeon 34113, Korea
3
Convergent Research Center for Emerging Virus Infection, Korea Research Institute of Chemical Technology, Daejeon 34114, Korea
*
Author to whom correspondence should be addressed.
Both authors are equally contributed.
Int. J. Mol. Sci. 2021, 22(11), 6150; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22116150
Submission received: 16 May 2021 / Revised: 25 May 2021 / Accepted: 4 June 2021 / Published: 7 June 2021
(This article belongs to the Special Issue Coronavirus Disease (COVID-19): Pathophysiology)

Abstract

:
The coronavirus disease 2019 (COVID-19) has caused a large global outbreak. It is accordingly important to develop accurate and rapid diagnostic methods. The polymerase chain reaction (PCR)-based method including reverse transcription-polymerase chain reaction (RT-PCR) is the most widely used assay for the detection of SARS-CoV-2 RNA. Along with the RT-PCR method, digital PCR has emerged as a powerful tool to quantify nucleic acid of the virus with high accuracy and sensitivity. Non-PCR based techniques such as reverse transcription loop-mediated isothermal amplification (RT-LAMP) and reverse transcription recombinase polymerase amplification (RT-RPA) are considered to be rapid and simple nucleic acid detection methods and were reviewed in this paper. Non-conventional molecular diagnostic methods including next-generation sequencing (NGS), CRISPR-based assays and nanotechnology are improving the accuracy and sensitivity of COVID-19 diagnosis. In this review, we also focus on standardization of SARS-CoV-2 nucleic acid testing and the activity of the National Metrology Institutes (NMIs) and highlight resources such as reference materials (RM) that provide the values of specified properties. Finally, we summarize the useful resources for convenient COVID-19 molecular diagnostics.

1. Background

Severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2) is a novel coronavirus that first appeared in Wuhan, Hubei Province, China in December 2019, connected to a seafood market [1,2]. Seven coronaviruses have been reported to infect humans; four of them, human CoV-NL63 [3], HCoV-OC43 [4,5,6], HCoV-229E [7,8], and HCoV-HKU [9,10], cause mild and seasonal respiratory tract disease, whereas SARS-CoV [11,12,13,14,15], MERS-CoV [16,17,18,19], and SARS-CoV-2 can cause severe symptoms. In particular, SARS- CoV-2 is suited to human-to-human transmission and spreads rapidly to other locations, causing lung injury, multiorgan failure, and death [20,21]. As of this date, the number of confirmed cases is still increasing, as is the number of deaths [22,23]. Therefore, an understanding of the SARS-CoV-2 host and pathogen biology is important to offer valuable insights into the diagnosis and treatment of the disease including the development of new therapies [24,25]. Here, we review the basic biology of SARS-CoV-2 including the origin, pathophysiology, and diagnosis methods.

1.1. Nomenclature of SARS-CoV-2

Currently, almost a million sequences of the SAR-CoV-2 genome are publicly available via the Global Initiative on Sharing All Influenza Data (GISAID) and GenBank [26,27]. Based on these genome sequences, the phylogenetic classification of SARS-CoV-2 was performed and the nomenclature of GISAID, Phylogenetic Assignment of Named Global Outbreak LINeages (PANGO lineage), and Nextstrain are widely used in scientific and clinical communities [26,28,29] The major lineages of each nomenclature are summarized in Table 1.
GISAID introduced the nomenclature system of SARS-CoV-2 based on marker mutations and named the clade with actual letters of marker mutations [30]. For example, the clade G has a characteristic mutation in the spike protein gene, D614G. In the nomenclature of GISAID, the initial strains of SARS-CoV-2 were grouped as S and L clades and the current strains of SARS-CoV-2 were classified as eight major clades (S, L, V, G, GH, GR, GV, and GRY) [30,31]. The L clade contains the reference strain WIV-04 and was the dominant lineage in early 2020. The L clade later diverged into clades V and G, and clade G diverged into clades GH, GR, GV, and GRY.
The PANGO nomenclature systems focused on the active virus lineage [29]. This nomenclature is dynamic and the lineages of the PANGO nomenclature are marked as three statuses: active, unobserved, or inactive. The lineages documented within a month are marked as active. The lineages documented within three months are marked as unobserved, and the lineages that were not documented for more than three months are regarded as inactive. The lineages of PANGO nomenclature are named with a letter and numerical values. The initial lineages are denoted as lineages A and B. Although clade B includes the first genome sequenced strain, the phylogenetic analysis suggested that the most recent common ancestor of SARS-CoV-2 was close to early lineage A [29]. The descendent lineages from initial lineages were assigned with numerical labels. The descendent lineages can be designated with the phylogenetic evidence that the descendent emerged from parental lineages and the descendants showed significant transmission to geographically distinct populations. The designated descendent lineages can also be parental of new emerging lineages and these new lineages have been labeled as parental lineages with additional numerical values. For example, a new emerging lineage from lineage B1 can be labeled as B1.1. The lineages can have a maximum of three sublevels and newly designated lineages emerging from a lineage with three sublevels will be labeled with new alphabetical letters. For example, the parental lineage of lineage C.1 is the lineage B.1.1.1.
The clades of Nextstrain nomenclature were initially named according to year–letter combinations [32]. Major clades were designated as the clade reached more than 20% of global frequency for more than two months. Based on this criteria, the initial clades were designated as 19A, 19B, 20A, 20B, and 20C. However, due to the global travel restriction, no more clades were designated according to the criteria. For this reason, Nextstrain updated their major clade designation criteria with regional frequency (>30%) and recognized variants of concern. Currently, 12 major clades (19A, 19B, 20A, 20B, 20C, 20D, 20E, 20F, 20G, 20H/501Y.V2, 20I/501Y.V1, 20J/501Y.V3) are designated in the nomenclature of Nextstrain.

1.2. Notable Variants of SARS-CoV-2

The first strain of SARS-CoV-2 was discovered in Wuhan, China and designated as Wuhan-Hu-1 or WIV-04 [1,33]. The comparison of whole genome sequences showed that the strain was closest to the SARS-like coronavirus RATG13 found in bats (Rhinolophus affinis) in China [2]. The overall genomic sequence similarity of RATG13 to SARS-CoV-2 was 96.1%. However, the spike protein gene of RATG13 lacked the furin cleavage site that is essential for the cell entry of SARS-CoV-2, indicating that RATG13 was not the immediate ancestor of SARS-CoV-2 [34,35]. After discovery of the first SARS-CoV-2 strains, SARS-CoV-2 like viruses were found in pangolins and bats [36,37,38]. The genome sequences of pangolin-derived CoVs also showed high similarity to those of SARS-CoV-2, but the furin cleavage site was missing in the spike gene sequences of pangolin-derived CoVs [35,37,38]. A bat-derived CoV, RmYN02, was identified and the genome of the virus showed high similarity to that of SARS-CoV-2 [36]. Although the sequence similarity of RmYN02 was slightly lower than those of RATG13 and pangolin-derived CoVs, the furin cleavage site was inserted, indicating that the addition of the cleavage site can occur naturally [35,36].
The D614G variants had a change in spike gene and replaced the initial strains of SARS-CoV-2 [39]. The studies on the variant D614G showed that the infectivity of the variant was increased without increased severity [39,40,41]. The engineered variants containing the D614G substitution showed more efficient infection in human cells and animal models without altering antibody neutralization and pathogenicity [41]. A population genetic analysis of COVID-19 also showed that the transmissibility of the variant was increased but there was no sign of increased mortality or clinical severity of the variants [40].
As new variants with increased pathogenicity, reduced neutralization, and/or increased transmissibility emerged, the U.S. Centers for Disease Control and Prevention (CDC) and Public Health England (PHE) classified some notable variants according to the attributes of the variants [42,43]. The CDC classified the variants according to the evidence and significance of the variants into Variant of Interest, Variant of Concern, and Variant of High Consequences. PHE classified variants as Variant Under Investigation (VUI) and Variants Of Concern (VOC). When the variants are considered to have concerning characteristics, they are designated as VUI. After a risk assessment of VUI is conducted, they can be re-designated as VOC. These notable variants are summarized in Table 2.
VOC-20DEC-01, also known as 20I/501Y.V1 or B.1.1.7, was first discovered in the United Kingdom in December 2020 [42] and is defined by 13 mutations [42]. Recent studies have estimated that the transmissibility of VOC-20DEC-01 is increased by 43–90% and a similar transmission increase was observed globally [44]. VOC-20DEC-01 was also detected in domestic cats and dogs, raising concern over human-to-animal transmission or vice versa [45]. Previously reported animal infections were asymptomatic to mild symptomatic, but VOC-20DEC-01 infection in animals showed relatively severe symptoms such as myocarditis [45]. In February 2021, the variants with a spike gene E484K mutation were reported and designated as VOC-202102/02 [42]. Another variant with N501Y mutation is VOC-20DEC-02 (20H/501Y.V2, or B.1.351), which was first discovered in South Africa. VOC-20DEC-02 is defined by 17 mutations including the E484K mutation, K417N mutation, and two deletions. The variant also showed increased transmissibility (approximately 50%) compared to previous variants [44]. The third variant with N501Y is VOC-21JAN-02 (P.1 or 20J/501Y.V3), discovered in Brazil [42,46]. The genome of VOC-21JAN-02 is defined with 17 non-synonymous mutations, four synonymous mutations, three deletions, and four insertions [42]. VOC-202101/02 almost fully replaced its parental variant within two months, indicating increased transmissibility of VOC-21JAN-02 [47,48]. Molecular clock analysis showed that the variants emerged in mid-November 2020 at which time hospitalizations rapidly increased [49].
The characteristic mutations (N501Y, E484K, and K417N) of the variants with N501Y are mutations in binding sites to viral receptor ACE2 and were already a concern prior to the discovery of these variants [50,51,52,53]. The studies on these variants showed that they had impacts on neutralization by immunity [54,55,56,57,58,59,60]. However, recent research showed that the residual immunity still provided protection, although variants reduced the efficacy of the vaccine [61].
There were also emerging variants without N501Y, E484K, and/or K417N. The characteristic mutations of A.23.1 are F157L, V367F, Q613H, and P681R [62]. A.23.1 with E484K was designated as VUI-21FEB-01 in the United Kingdom. These strains were first identified in Uganda and are spreading. One of the characteristic mutations, Q613H, is regarded as functionally equivalent to the D614G mutation of ‘G’ clade strains. B.1.427 and B.1.429 were first discovered and designated as Variants of Concern in the United States [43]. The characteristic mutations of both lineages are L452R and D614G; these variants showed increased transmissibility and reduced neutralization by convalescent and post-vaccination sera [63].
B.1.617 was the emerging lineage in India and also designated as VUI-21APR-01 in the United Kingdom [42]. B.1.167 has two characteristic mutations of different lineages: L452R and E484Q [64]. The variants were neutralized with convalescent sera of COVID-19 patients and vaccine of BBV152, although the efficacy was low [64].

2. PCR-Based SARS-CoV-2 Detection

2.1. Reverse Transcription Quantitative PCR (RT-qPCR) Method

Detection of the SARS-CoV-2 viral genome, consisting of single-stranded RNA, is effectively done by reverse transcription quantitative polymerase chain reaction (RT-qPCR), which is the gold standard technique widely used in molecular diagnostics [65,66]. There are several practical considerations when performing diagnostic assays using RT-qPCR.
(1) Sample quality: RT-qPCR tests are presently being used for the identification of SARS-CoV-2 in clinical specimens such as upper respiratory tract specimens (saliva, oropharyngeal swab-OPS, nasopharyngeal swab-NPS, nasal swabs), lower respiratory specimens (sputum, bronchoalveolar lavage-BAL, endotracheal aspirate-ET, fibrobronchoscope brush biopsy-FBB), blood (serum, plasma), urine, feces, rectal/anal swabs, stool, and corneal secretion [67,68]. To check the sample quality of clinical specimens from different origins, an RNA isolation procedure is required to obtain purified high-quality RNA from the samples, which then needs to be analyzed using chip-based capillary electrophoresis (such as the Agilent Bioanalyzer system), electrophoretic separation on a high-resolution agarose gel, and spectrophotometry [69].
(2) Reference curve: Data processing can critically affect the analysis of RT-qPCR results [70]. PCR data processing is based on standard curves or on PCR efficiency assessments [70]. Standard curves are used to assess RT-qPCR efficiency without theoretical and practical problems [70]. The estimation of RT-qPCR efficiency using standard curves usually involves the serial dilution of a concentrated stock solution, after which standard samples are analyzed through RT-qPCR by measuring the quantification cycle (Cq) using standard procedures [70]. The most widely used Cq value is the threshold cycle (Ct), the cycle at which the expression of a target gene first exceeds a calculated fluorescence threshold level [71]. For example, to detect low amounts of SARS-CoV-2 RNA, a series of diluted RNA templates are used to determine the Ct value, which can provide a standard curve for evaluating the reaction efficiency [72]. However, the Ct value itself cannot be directly explained as viral load without a standard curve using reference materials [73]. When interpreting the results of SARS-CoV-2 RT-qPCR, the validity of the standard curve should be proved using reference materials with accurate viral copy numbers to interpret Ct values as viral loads [73].
(3) Viral load: The success of virus isolation depends on the viral load [74]. Viral loads in sputum samples and throat swabs are high when obtained within seven days after initial symptoms are observed, ranging from 104 to 107 copies per mL. This pattern is broken as low quantity of virus are obtained from samples taken after day 8 [75]. In general, sputum samples show higher viral loads than throat swab samples, whereas low viral RNA is detected in urine or stool samples [75]. The two main factors that influence the quantitative measurement of viral roads are Cq values that are repeatable with acceptable uncertainty and a reliable means of converting from the Ct value to viral load [76,77,78]. For molecular diagnostic assays, a limit of detection (LoD) and a limit of quantification (LoQ) are also considered the lowest concentrations of target RNA that can be detected by RT-qPCR [79].
(4) Sampling methods: RT-qPCR tests for SARS-CoV-2 have shown a high variation of false-negative rates (FNR) and false-positive rates (FPR) [80,81]. Numerous methods have been developed with the goal of improving the sensitivity, safety, and rapidity of COVID-19 tests by RT-qPCR. For example, one group tested the efficiency and sensitivity of SARS-CoV-2 detection of clinical specimens collected directly in nucleic acid stabilization and lysis buffer (NSLB), a mixture of lysis buffer and RNA preservative, instead of a viral transport medium (VTM), thus inactivating the virus immediately after sampling [82].
(5) Sample source: To improve the expandability of SARS-CoV-2 testing, several sampling approaches have been developed including nasal, pooled nasal, and throat (oropharyngeal) swabs as well as saliva. Different clinical sampling methods affect the diagnostic performance of SARS-CoV-2 infection tests by RT-qPCR including sensitivity and specificity, and thus should be carefully considered [83,84,85,86]. The combined swab is largely recommended as a more appropriate specimen for diagnosis by RT-qPCR [87,88,89].
(6) Sensitivity: The conserved regions, ORF 1ab (RNA-dependent RNA polymerase, RdRp), envelope (E), and nucleocapsid (N) genes of SARS-CoV-2, are usually selected as the standard target genes for primer and probe design [90,91]. However, initial reports of SARS-CoV-2 and other coronavirus sequences gave rise to an incorrect degenerate base that did not align with the SARS-CoV-2 RNA sequence found, and there were reports regarding the decreased sensitivity of using RdRp as a target gene for RT-qPCR assays [90,92]. As the pandemic continues, many laboratories around the world rely on routine diagnostic primers and probes. Thus, proper assays can increase the sensitivity of SARS-CoV-2 detection and help prevent the further spread of the virus [92,93,94,95].
(7) Pooling technologies: The pooling of multiple swab samples before RNA isolation and RT-qPCR analysis has been proposed as a promising solution to reduce costs and time as well as elevate the throughput of SARS-CoV-2 tests for large-scale testing as in the case of schools [96,97,98,99]. For example, batch testing of over 100,000 hospital-collected nasopharyngeal swab samples from patients alleviated three quarters of testing reactions with a minor reduction in sensitivity, indicating the effectiveness of the pooling approach in the field [100,101]. Current studies suggest that the pooling of individual samples before testing should be considered to increase the reliability of SARS-CoV-2 testing throughput.
Once all practical considerations have been evaluated, there are two ways that RT-qPCR can be performed. The two-step RT-qPCR method is required to convert RNA into complementary DNA (cDNA) [102]. On the other hand, the one-step RT-qPCR method combines reverse transcription and PCR in a single tube and uses reverse transcriptase as well as a DNA polymerase [103]. The schematic procedure of RT-qPCR is shown in Figure 1A.
A critical need for rapid and accurate diagnostic methods has emerged in the clinic and public health organizations. Several PCR-based assays have been developed and are currently being used in clinical, research, and public health laboratories [104,105,106]. However, it is not clear which PCR condition they should adopt or whether the data are comparable. In response to the growing need and the lack of publicly available information, several research groups have optimized real-time PCR-based primer sets, protocols, and PCR conditions [107,108].
Independent evaluations of the designed primer–probe sets used in SARS-CoV-2 RT–qPCR detection assays are necessary to compare and select appropriate assays [90,109]. Additionally, several studies have utilized serum and stool specimens for the RT-qPCR-based detection method [110,111,112,113].

2.2. Reverse Transcription Digital PCR (RT-dPCR) Method

In recent years, we have seen the advance of digital PCR (dPCR) as a complementary approach for measuring nucleic acids, a technique that is highly accurate and reproducible when targeting the viral genes of SARS-CoV-2 [114,115,116,117]. The advantages of digital PCR compared to quantitative PCR include quantification without the need for calibration curves, higher accuracy, and sensitivity that may arise from sub-optimal amplification efficacy because dPCR can detect low amounts of nucleic acid [118,119]. The schematic procedure of dPCR is shown in Figure 1B.
Reverse transcriptase quantitative PCR (RT-qPCR) and digital PCR (dPCR) have been widely used for quantitative analyses of clinical samples. Recently, many groups have developed a reverse transcription droplet digital PCR (RT-ddPCR) assay for sensitive detection of the SARS-CoV-2 virus [120,121,122,123]. Optimization of the primer. and probe assays is necessary to remove false negatives or positives for clinical diagnosis of viral infection [72,124]. Multiple molecular diagnostic kits have been developed and validated for use nationwide [125]. However, the analytical sensitivity and the relative sensitivity of different kits to detect low copy number of SARS-CoV-2 viral RNA are variable [126,127].

3. Isothermal Nucleic Acid Amplification Methods

Although the RT-qPCR method is considered the ‘gold standard’ for SARS-CoV-2 detection [128], its limitations have stimulated the development of simple, rapid yet sensitive nucleic acid detection methods [129]. As a result, isothermal nucleic acid amplification has emerged as an alternative detection method for SARS-CoV-2 viral RNA from clinical samples [130]. In general, isothermal amplification techniques increase the analytical signal by increasing the target nucleic acid concentration through enzymatic activities at a fixed temperature, and simultaneously detecting the signal with colorimetric or fluorescence indicators [131]. Changes in color, fluorescence level, or turbidity indicate the presence of SARS-CoV-2 RNA or DNA [131]. Therefore, unlike RT-qPCR, isothermal amplification methods do not require thermal cycling instruments or specialized technicians for disease diagnosis [132]. Current isothermal nucleic acid amplification methods used for SARS-CoV-2 detection include, but are not limited to, loop-mediated isothermal amplification (LAMP), recombinase polymerase amplification (RPA), nucleic acid sequence-based amplification (NASBA), strand-displacement amplification (SDA), and rolling circle amplification (RCA) [133,134,135,136,137] (Figure 1C). Herein, we describe the general procedures and components of isothermal amplification methods commonly used for diagnosis of SARS-CoV-2.

3.1. Loop-Mediated Isothermal Amplification (LAMP)

The loop-mediated isothermal amplification method, coupled with reverse transcription (RT-LAMP), is the most widely used isothermal amplification technique for SARS-CoV-2 nucleic acid detection. First described by Notomi et al. [138], this method uses strand displacement activity of DNA polymerase and a set of inner and outer primers (four or six specific primer sequences) to amplify the target nucleic acids. LAMP is carried out at a single temperature between 60 and 65 °C, and generates up to 109 copies of DNA in less than an hour [133,137,139]. The LAMP procedure is initiated by hybridization of the forward inner primer (FIP) toward the target DNA template, which synthesizes the complementary strand. Then, the outer primer hybridizes to the target DNA, which initiates DNA synthesis by strand displacement. Subsequently, a FIP-hybridized complementary strand is released and forms a loop structure at one end of the sequence. The corresponding sequence becomes the template for the backward inner primer (BIP), which initiates another DNA synthesis by strand displacement, and then produces a ‘dumb-bell’ like DNA structure. Self-primed DNA synthesis of the corresponding sequence then converts the ‘dumb-bell’ like structure into a ‘stem-loop’ like DNA structure. Corresponding stem-loop DNA then becomes the template for LAMP cycling, and the target DNA sequence exponentially amplifies until the reaction is completed [138]. Amplified products are detected by changes of color as the accumulation of DNA changes, pH levels, or by changes in turbidity as magnesium pyrophosphate level increases [140,141,142]. Amplified products are also detected by Calcein fluorescent dye or fluorescent intercalating dye [129,139]. The schematic procedure of RT-LAMP is shown as Figure 2.
Researchers have made efforts to optimize RT-LAMP for the development of rapid and sensitive detection of SARS-CoV-2. Several studies have evaluated the experimental parameters for RT-LAMP such as incubation temperature, incubation time, LoD, target genes, and primer sequences [143,144,145]. Aside from optimizing the experimental parameters, researchers have developed modified RT-LAMP procedures including methods without prior RNA extraction steps and high-throughput colorimetric assay methods using a 96-well plate format [146,147]. Modified RT-LAMP procedures also include methods coupled with Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR) technology, a nanoparticle-based biosensor, and artificial intelligence [148,149,150,151].

3.2. Recombinase Polymerase Amplification (RPA)

Recombinase polymerase amplification is another isothermal amplification method that is widely used for SARS-CoV-2 detection. First described by Piepenburg et al. [152], RPA uses a complex of recombinase and two target specific primers (forward and reverse primers) to amplify the target nucleic acids [152]. Once the target nucleic acids are identified, recombinase-primer complex unwinds the target DNA and allows forward and reverse primers to hybridize [153]. The displaced DNA strand is amplified in the presence of DNA polymerase as primers elongate, and the template DNA is exponentially amplified until the reaction is completed [153]. RPA reaction is carried out at a single temperature between 37 and 42 °C, and the reaction is completed when ATPs are depleted, typically in less than an hour [154]. Amplified products are detected by gel electrophoresis, antigenic tags on primers and tag-specific antibodies, or fluorescent signals produced by a conjugated fluorophore and quencher on primers [152,153,154]. The schematic procedure of reverse transcription RPA (RT-RPA) is shown as Figure 2.
For SARS-CoV-2 detection, researchers have optimized RPA procedures by testing various experimental parameters that can now detect less than five viral copies from patient samples within 45 min from the sample collection [134]. RPA methods have also been optimized for SARS-CoV-2 detection by coupling RPA-based amplification with various CRISPR-based detection methods [155,156,157].

3.3. Other Isothermal Nuleic Acid Amplification Methods

Other than LAMP and RPA, isothermal amplification methods such as NASBA, SDA, and RCA have been used for the detection of SARS-CoV-2 [136,137,158]. Although we will not describe each technique in detail here, Table 3 presents the general features and components of each isothermal amplification method.

4. Non-Conventional Methods

4.1. Genome Sequencing

Unprecedentedly massive genome sequencing has been undertaken with SARS-CoV-2 strains. The total number of sequenced genomes is approximately a million at present [30]. As the number of genomes increases rapidly, the World Health Organization (WHO) has provided guidelines for the genome sequencing of SARS-CoV-2 [159]. According to this guideline, the genome sequencing of SARS-CoV-2 can be used for understanding the emergence of SARS-CoV-2, understanding the biology of SARS-CoV-2, improving diagnostics and therapeutics, investigating virus transmission and spread, and inferring epidemiological parameters [159]. Based on the accumulative genome sequences of SARS-CoV-2, the emergence of the variants of concern [43,44,48,49,57,62], the origin of SARS-CoV-2 [2,31], and the mutation frequency of RT-qPCR primer/probe sites [30] can be known to humanity. Currently, various whole genome sequencing methods of the virus are being developed [160,161,162,163,164,165,166]. The sequencing methods of viruses can be categorized into the metagenomics approach and target enrichment based methods. In the metagenomics approach, the viral genome can be extracted from clinical samples and the extracted nucleic acid are sequenced. This can also be done with cultured viruses. These approaches have a clear advantage over target enrichment based methods. The metagenomics approach can be used even if there is no information of the pathogen or there are novel pathogens that were not previously known. However, a high proportion of host cell genetic materials can be found, which should be removed or reduced for sequencing. The removal or depletion methods of the host genetic materials vary by the type of sample or the virus [167,168,169,170,171,172,173,174]. Due to the nature of the metagenomics approaches, the clinical samples should ideally have a high titer of the pathogens. The metagenomics approach also can be done with cultured pathogens. However, the isolation and culturing of the pathogen are very time-consuming and labor-intensive work. In some cases, the isolation and culturing of some pathogens are not possible or are very difficult [175,176]. Alternatively, target enrichment methods can be used for the genome sequencing. The genetic materials of specific pathogens can be enriched through hybrid capture probes [177,178]. The sequences of the probes are complementary to the genome sequences of specific pathogens and these target enrichments effectively remove not-target sequences and increase the proportion of the target sequences. One of the advantages of hybrid capture approaches is the tolerance of sequence mismatch, allowing the capture of divergent variants. However, the hybrid capture approaches are relatively more expensive and complicated than other approaches. Another group of target specific enrichment approaches is the amplicon based approaches. The amplicon-based approaches are mainly dependent on PCR reactions. The PCR reaction can selectively enrich the genome of the target pathogens in the presence of non-target nucleic acids just like host genetic materials. Due to the nature of PCR, the amplicon based approaches are relatively more inexpensive, sensitive, and specific than other approaches. The WHO guidelines suggest that the complete genome sequencing can be done from the sample with Ct values of up to 30 and the partial genome sequencing can be done from the sample with Ct values of 30–35, although the Ct value can vary with various factors [159]. Currently, the most widely used primer panels for SARS-CoV-2 are ARTIC network amplicon sets [179]. At least three commercially available SARS-CoV-2 primer panels (CleanPlex SARS-CoV-2 Panel; Paragon genomics, QIAseq SARS-CoV-2 Primer Panel; Qiagen, and NEBNext ARTIC SARS-CoV-2 Library Prep Kit; NEB) are based on ARTIC network amplicon sets. However, there are also limitations. The design of primers requires prior knowledge of full sequence information of the genome. In addition, the primer intolerance to the sequence mismatch hinders the genome sequencing of the variants. The amplicon appears that the amplicon based approaches can be applied only to previously well-known pathogens. Alternatively, the genomic materials can be amplified sequence-independently [163,180,181]. Single primer isothermal amplification (SPIA) can amplify the genomic materials in a sequence-independent manner. As SPIA can amplify the genomic materials, prior knowledge of the pathogens is not required for a target enrichment approach. However, removing non-target genomic materials is mandatory for sequencing based SPIA as SPIA can also amplify non-target genomic material. Due to these characteristics, a high proportion of target genomic materials in the samples is crucial for successful SPIA based sequencing. SPIA based sequencing with low viral input showed very low coverage compared to other methods [180].
For the genome sequencing of pathogens, various sequencing technology can be used. While the conventional Sanger sequencing can still be used for viral genome sequencing [182], most SARS-CoV-2 genome sequencing is done with NGS sequencing technology. Currently, the most widely used NGS sequencing technology is the sequencing platforms of Illumina. Although the sequence length of individual reads is relatively short (paired-end 150 bp), the throughput and the accuracy of the individual reads are outstanding. Ion Torrent is another short reads sequencing platform technology, where the length of individual reads is 400 bp or 600 bp. The running time of the Ion Torrent sequencer is shorter than that of the Illumina sequencer. Long read alternatives are also available. The lengths of individual reads from PacBio and Oxford Nanopore Technology sequencers are tens of kilo base pairs or more. The individual reads from these long read sequences can cover most of the viral genome. However, the throughput of the long read sequencers is relatively lower than that of short read sequencers such as Illumina sequencing platforms. Furthermore, the accuracy of the individual reads is relatively lower than that of Illumina sequencing platforms. The sequencing platforms of Oxford Nanopore Technology maximize the benefits and drawbacks of the long read sequencer. The maximum length of the individual read is recorded up to a megabase scale [183]. However, due to the relative low accuracy of the individual reads, the WHO guidelines do not recommend these for SARS-CoV-2 genome sequencing unless the sequencing is replicated [159,184]. The coverage and depth of viral genome sequencing can be varied by the number of samples in single runs. Generally, most multiplex sequencing library kits for NGS support up to 384 samples per single runs. However, the production scale sequencers of Illumina (Hiseq, NextSeq, and NovaSeq) generate massive reads for small genome of viruses even with multiplex libraries. Though short individual reads require relatively more depth for high coverage, massive generation of sequence reads and low sequencing error rates of individual reads can compensate for short individual read length. Due to the massive sequence reads generation of the Illumina sequencer, the metagenomics approach of viral genome sequencing is practically available to only Illumina sequencer or similar platforms. Even if most of the sequence reads are non-target sequence reads (host genetic materials, contamination, etc.), a high quality genome assembly of the virus can be produced from the small remaining fraction of the target sequences. The long read sequencers such as sequencers by Pacbio and Oxford Nanopore Technology can generate very long individual reads that can cover most of the viral genome. However, due to the relatively low yield of total sequence reads and low accuracy of individual reads, the long read sequencers are not adequate for the metagenomics approach of viral genome sequencing. Instead, the long sequencers are more suitable for the amplicon-based approach. The target specific amplification can overcome the drawbacks of long read sequencers such as the low yield of total sequence reads and improving the accuracy of individual reads. Moreover, long sequencers can use long amplicons unlike short read sequencers. The schematic procedure of genome sequencing is shown in Figure 3.

4.2. CRISPR Based COVID-19 Detection

The clustered regularly interspaced short palindromic repeats (CRISPR)-Cas technique has been repurposed for diagnostics and is one of the widely used nucleic acid detection methods [185,186]. Many types of Cas proteins have been developed to create highly accurate and sensitive diagnostic methods [187]. Cas9 has been widely used for genome editing while DNA-targeting Cas12 (also known as Cpf1 or C2c1) effectors and RNA-targeting Cas 13a are more suitable for disease diagnosis [188].
Compared to conventional diagnostic methods such as RT-qPCR, CRISPR-based approaches can quickly provide rapid, visual, highly sensitive, and specific detection due to the collateral cleavage of a reporter dye in the presence of a target [189].
Numerous CRISPR-Cas detection systems have been developed. For example, techniques of a CRISPR–Cas12-based assay have been developed for the detection of SARS-CoV-2 from patient sample RNA, called SARS-CoV-2 DNA Endonuclease-Targeted CRISPR Trans Reporter (DETECTR) [190]. This assay includes simultaneous reverse transcription and isothermal amplification using loop-mediated amplification (RT–LAMP) [191]. An RNA targeting Cas13a dependent platform [156], the SHERLOCK (Specific High Sensitivity Enzymatic Reporter UnLOCKing) technique, offers a simplified test and has a limit of detection of 100 copies of the viral genome [192,193].
Recently, a lyophilized CRISPR-Cas12 assay for SARS-CoV-2 detection (Lyo-CRISPR SARS-CoV-2 kit) has been developed based on reverse transcription (RT), isothermal amplification, and CRISPR-Cas12 reaction [194]. The schematic procedure of CRISPR detection systems is shown in Figure 1D.

4.3. Nanotechnology Based Methods

Nanotechnology has already proven its value through its diagnostic, vaccine, and therapeutic applications that have expanded into clinical applications [195]. Scientists have shown that nucleic acid detection using nanomaterials for viral infectious diseases now have various advantages in the diagnostic field [196]. Moreover, nanomaterials are powerful tools for the diagnosis, prevention, and treatment of COVID-19 [197].
Magnetic nanoparticles have been used in RT-qPCR diagnosis for the extraction of viral RNA from SARS-CoV-2. This method merges the lysis and binding steps, and the poly (amino ester) with carboxyl groups (PC)-coated magnetic nanoparticles (pcMNPs)-RNA complexes can be directly introduced into RT-qPCR reactions [198]. In addition, a test has been developed to diagnose SARS-CoV-2 that can rapidly detect the virus. The test is performed using gold nanoparticles to detect specific proteins such as nucleocapsid phosphoprotein. In the presence of gold nanoparticles, the test is positive upon the color of the liquid reagent changing from purple to blue [199]. Nanoparticles that can be applied in the RT-qPCR diagnosis are shown in Figure 1E.

5. Resources and Standardization for SARS-CoV-2 Nucleic Acid Testing

5.1. International Activity for Standardization

The Consultative Committee on the Quantity of Material (CCQM) is responsible for developing and documenting the equivalence of national standards such as certified reference materials (CRMs)/reference materials (RMs) and reference methods for biological and chemical measurements [200]. It advises the International Committee for Weights and Measures (Comité international des poids et mesures, CIPM) on matters related to biological and chemical measurements including advice on the International Bureau of Weights and Measures’ (Bureau international des poids et mesures, BIPM) scientific program activities [200]. One of the responsibilities of the CCQM is to contribute to the development of a worldwide admitted system of national measurement standards, reference methods, and facilities for biological and chemical measurements [201]. Several National Metrology Institutes (NMIs) and expert laboratories from many countries have performed highly sensitive and accurate measurements of the amount of the SARS-CoV-2 viral RNA tested using reverse transcription-digital PCR (RT-dPCR) [202]. The capability to accurately measure the amount of COVID-19 causing viral nucleic acid with equivalence globally will remarkably improve diagnostic testing confidence and support countries in effectually tackling the pandemic situation [202]. Moreover, the biological, chemical, and physical measurement ability of NMIs are enabling industry and laboratories to effectively and quickly face the COVID-19 challenge [203].
NMIs support quality assurance by developing and providing CRM/RM for COVID-19 and ensures equivalent pathology testing, while minimizing false negative and positive test outcomes.
Non-NMIs including both public and private sectors have developed standards and reference materials for SARS-CoV-2 RNA. The National Institute for Biological Standards and Control (NIBSC) distributed the 1st WHO International Standards for SARS-CoV-2 RNA in 2020, which is used for the standardization of nucleic acid amplification technique (NAT)-based diagnostic assays. The U.S. Food and Drug Administration developed the SARS-CoV-2 Reference Panel in 2020 to precisely compare the performance of NAT-based assays for SARS-CoV-2 detection. Non-profit organizations such as American Type Culture Collection (ATCC) and private industries such as LGC SeraCare have also developed and distributed reference materials for NAT-based SARS-CoV-2 diagnosis.

5.2. Reference Materials

A RM can be defined as a “material, sufficiently homogeneous and stable with respect to one or more specified properties, which has been established to be fit for its intended use in a measurement process” [204]. According to the general requirements for the competence of reference material producers (ISO 17034:2016), a CRM is defined as a “reference material characterized by a metrologically valid procedure for one or more specified properties, accompanied by a reference material certificate that provides the value of the specified property, its associated uncertainty, and a statement of metrological traceability” [204]. Generally, major producers of RMs are NMIs [205]. However, RMs of SARS-CoV-2 are produced by not only NMIs, but also by commercial institutes and culture collections. The first RM of SARS-CoV-2 was produced by the National Institute of Metrology of China (NIMC). The RM was synthetic RNA based on the genome sequence of Wuhan-Hu-1, the first identified SARS-CoV-2 strain. The RM of NIMC was quantified with the ddPCR method and contained N, E, and partial genes of RdRp genes that can cover the WHO announced in-house assay for SARS-CoV-2 [91]. Other NMIs such as the Joint Research Center in Europe (JRC), Korea Research Institute of Standards and Science (KRISS), National Measurement Institute of Australia (NMIA), National Institute of Standards and Technology (NIST), and The National Metrology Institute of Turkey (UME) have also produced RMs for SARS-CoV-2 nucleic acid testing. NIBSC also produced a RM for SARS-CoV-2 nucleic acid testing and it has been used as WHO reference materials. Type culture collection also produced RMs. ATCC and BEI resources produced RMs with inactivated SARS-CoV-2 strain (USA-WA1/2020). Not only public institutes but also commercial institutes have produced RMs. Seracare, which is a subsidiary of LGC, produced RMs in the early phase of the pandemic. The RMs were SARS-CoV-2 RNA covered with viral proteins, which can be used for the WHO announced in-house assays. Twist Bioscience produced RMs with various SARS-CoV-2 strains including emerging variants such B.1.1.7, B.1.351, and P.1.
The majority of these RMs were quantitated with dPCR. The comparison of qPCR and dPCR methods showed that dPCR can quantitate specific genes regardless of primer sequences [72], indicating that the dPCR assay is a suitable method for the measurement of RMs. The reference materials of SARS-CoV-2 nucleic acid testing are listed in Table 4 and Table S1.

5.3. Other Resources

The genomic sequences of SARS-CoV-2 are mainly deposited in GISAID databases and are also available at the NCBI GenBank [27,30]. The metadata of strains such as collection date, patient age, gender, sequencing methods, etc. are also available in GISAID. There are also other specialized databases. Virus Pathogen Resources (ViPR) provides detailed genome sequence information and analysis tools [206]. Nextstrain provides real-time tracking evolution and spreading of SARS-CoV-2 [28]. Nextstrain visualizes tracking of SARS-CoV-2 lineages by integrating geographic information and sequence information. CoV-GLUE is a mutation dedicated database for SARS-CoV-2 [207]. CoV-GLUE summarizes the mutations of amino acid replacements, insertions, and deletion.
Although more than a million SARS-CoV-2 genome sequences are publicly available, live SARS-CoV-2 sources are relatively limited. As SARS-CoV-2 is regarded as risk group 3 and handling of live SARS-CoV-2 requires biosafety level 3 laboratories, only a few culture collections distribute SARS-CoV-2 strains. BEI resources have been established by the National Institute of Allergy and Infectious Diseases (NIAID) and provide live SARS-CoV-2 strains and derivatives of SARS-CoV-2. The National Culture Collection for Pathogens of Korea (NCCP) also provides live viruses and derivatives from SARS-CoV-2 isolated in Korea. European Virus-Archive Global (EVAg) is a network between 25 laboratories including 16 EU member state institutions and nine non-EU institutions. It also provides isolated strains and derivatives from the EU and related countries. Training courses such as SARS-CoV-2 diagnostic training are also available at EVAg.

6. Conclusions

Currently, representative diagnostic methods of SARS-CoV-2 are RT-qPCR assays. Most countries use the RT-qPCR assay as a primary method for diagnostics. Though alternative methods are available, their sensitivity, specificity, or costs are not comparable to RT-qPCR assays. However, RT-qPCR assays require relatively expensive instruments and highly trained personnel. These requirements restrict the expansion of diagnostics capacity in some countries. To overcome these drawbacks, future diagnostics methods should be inexpensive and simple, which can be used for point-of-care testing. For example, if inexpensive whole viral genome sequencing methods are developed in the future, they will be capable of replacing RT-qPCR assays as a standard disease diagnostic test. The whole genome sequencing of the virus can provide more information including variant information of the viruses. The reference materials for the diagnostics are also important to assess newly developed diagnostic methods.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ijms22116150/s1, Table S1: Extended table of Reference materials and resources for SARS-CoV-2 nucleic acid testing.

Author Contributions

Conceptualization, S.K.; Investigation, H.M.Y., I.-H.K. and S.K.; Writing—original draft preparation, H.M.Y., I.-H.K. and S.K.; Writing—review and editing H.M.Y., I.-H.K. and S.K.; Funding acquisition, H.M.Y. and S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Council of Science and Technology grant by the Ministry of Science and ICT (Grant No. CRC-16-01-KRICT). This work was also supported by the “Establishment of measurement standards for microbiology”, grant number GP2021-0003-08, funded by the Korea Research Institute of Standards and Science.

Acknowledgments

The professional illustrations were done by Scapiens.

Conflicts of Interest

Seil Kim and Hee-Min Yoo are producers of KRISS SARS-CoV-2 reference materials.

Abbreviations

COVID-19Coronavirus disease 2019
PCRpolymerase chain reaction
RT-PCRReverse transcription PCR
RT-qPCRReverse transcription quantitative PCR
RT-dPCRReverse transcription digital PCR
RT-ddPCRReverse transcription droplet digital PCR
RT-LAMPReverse transcription loop-mediated isothermal amplification
RT-RPAReverse transcription recombinase polymerase amplification
NGSNext-generation sequencing
NMINational Metrology Institutes
RMReference materials
SARS-CoV-2Severe acute respiratory syndrome coronavirus-2
GISAIDGlobal Initiative on Sharing All Influenza Data
PANGO lineagePhylogenetic Assignment of Named Global Outbreak LINeages
CDCUS Centers for Disease Control and Prevention
PHEPublic Health England
VUIVariant Under Investigation
VOCVariants Of Concern
BALBronchoalveolar lavage
ETEndotracheal aspirate
FBBFibrobronchoscope brush biopsy
CqQuantification cycle
CtThreshold cycle
LoDLimit of detection
LoQLimit of quantification
FNRFalse-negative rates
FPRFalse-positive rates
NSLBNucleic acid stabilization and lysis buffer
VTMViral transport medium
RdRpRNA-dependent RNA polymerase
EEnvelope
NNucleocapsid
cDNAComplementary DNA
LAMPLoop-mediated isothermal amplification
RPARecombinase polymerase amplification
NASBANucleic acid sequence-based amplification
SDAStrand-displacement amplification
RCARolling circle amplification
FIPForward inner primer
BIPBackward inner primer
CRISPRClustered Regularly Interspaced Short Palindromic Repeats
WHOWorld Health Organization
SPIASingle primer isothermal amplification
DETECTRDNA Endonuclease-Targeted CRISPR Trans Reporter
SHERLOCKSpecific High Sensitivity Enzymatic Reporter UnLOCKing
RTReverse transcription
CCQMConsultative Committee on the Quantity of Material
CRMsCertified reference materials
CIPMComité international des poids et mesures
BIPMBureau international des poids et mesures
NMINational Metrology Institutes
NIBSCNational Institute for Biological Standards and Control
NATNucleic acid amplification technique
ATCCAmerican Type Culture Collection
NIMCNational Institute of Metrology of China
JRCJoint Research Centre in Europe
KRISSKorea Research Institute of Standards and Science
NMIANational Measurement Institute of Australia
NISTNational Institute of Standards and Technology
UMEThe National Metrology Institute of Turkey
ViPRVirus Pathogen Resources
NIAIDNational Institute of Allergy and Infectious Diseases
NCCPNational Culture Collection for Pathogens of Korea
EVAgEuropean Virus-Archive Global

References

  1. Wu, F.; Zhao, S.; Yu, B.; Chen, Y.M.; Wang, W.; Song, Z.G.; Hu, Y.; Tao, Z.W.; Tian, J.H.; Pei, Y.Y.; et al. A New Coronavirus Associated with Human Respiratory Disease in China. Nature 2020, 579, 265–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Zhou, P.; Yang, X.L.; Wang, X.G.; Hu, B.; Zhang, L.; Zhang, W.; Si, H.R.; Zhu, Y.; Li, B.; Huang, C.L.; et al. A Pneumonia Outbreak Associated with a New Coronavirus of Probable Bat Origin. Nature 2020, 579, 270–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Van Der Hoek, L.; Pyrc, K.; Jebbink, M.F.; Vermeulen-Oost, W.; Berkhout, R.J.M.; Wolthers, K.C.; Wertheim-Van Dillen, P.M.E.; Kaandorp, J.; Spaargaren, J.; Berkhout, B. Identification of a New Human Coronavirus. Nat. Med. 2004, 10, 368–373. [Google Scholar] [CrossRef]
  4. Vabret, A.; Mourez, T.; Gouarin, S.; Petitjean, J.; Freymuth, F. An Outbreak of Coronavirus OC43 Respiratory Infection in Normandy, France. Clin. Infect. Dis. 2003, 36, 985–989. [Google Scholar] [CrossRef] [PubMed]
  5. Vijgen, L.; Keyaerts, E.; Moës, E.; Thoelen, I.; Wollants, E.; Lemey, P.; Vandamme, A.-M.; Van Ranst, M. Complete Genomic Sequence of Human Coronavirus OC43: Molecular Clock Analysis Suggests a Relatively Recent Zoonotic Coronavirus Transmission Event. J. Virol. 2005, 79, 1595–1604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Owczarek, K.; Szczepanski, A.; Milewska, A.; Baster, Z.; Rajfur, Z.; Sarna, M.; Pyrc, K. Early Events during Human Coronavirus OC43 Entry to the Cell. Sci. Rep. 2018, 8, 7124. [Google Scholar] [CrossRef] [PubMed]
  7. Pene, F.; Merlat, A.; Vabret, A.; Rozenberg, F.; Buzyn, A.; Dreyfus, F.; Cariou, A.; Freymuth, F.; Lebon, P. Coronavirus 229E-Related Pneumonia in Immunocompromised Patients. Clin. Infect. Dis. 2003, 37, 929–932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Hendley, J.O.; Fishburne, H.B.; Gwaltney, J.M. Coronavirus Infections in Working Adults. Eight-Year Study with 229 E and OC 43. Am. Rev. Respir. Dis. 1972, 105, 805–811. [Google Scholar] [PubMed]
  9. Woo, P.C.Y.; Lau, S.K.P.; Chu, C.; Chan, K.; Tsoi, H.; Huang, Y.; Wong, B.H.L.; Poon, R.W.S.; Cai, J.J.; Luk, W.; et al. Characterization and Complete Genome Sequence of a Novel Coronavirus, Coronavirus HKU1, from Patients with Pneumonia. J. Virol. 2005, 79, 884–895. [Google Scholar] [CrossRef] [Green Version]
  10. Lau, S.K.P.; Woo, P.C.Y.; Yip, C.C.Y.; Tse, H.; Tsoi, H.W.; Cheng, V.C.C.; Lee, P.; Tang, B.S.F.; Cheung, C.H.Y.; Lee, R.A.; et al. Coronavirus HKU1 and Other Coronavirus Infections in Hong Kong. J. Clin. Microbiol. 2006, 44, 2063–2071. [Google Scholar] [CrossRef] [Green Version]
  11. Nitsche, A.; Schweiger, B.; Ellerbrok, H.; Niedrig, M.; Pauli, G. SARS Coronavirus Detection. Emerg. Infect. Dis. 2004, 10, 1300–1303. [Google Scholar] [CrossRef]
  12. Drosten, C.; Günther, S.; Preiser, W.; van der Werf, S.; Brodt, H.-R.; Becker, S.; Rabenau, H.; Panning, M.; Kolesnikova, L.; Fouchier, R.A.M.; et al. Identification of a Novel Coronavirus in Patients with Severe Acute Respiratory Syndrome. N. Engl. J. Med. 2003, 348, 1967–1976. [Google Scholar] [CrossRef]
  13. Poutanen, S.M.; Low, D.E.; Henry, B.; Finkelstein, S.; Rose, D.; Green, K.; Tellier, R.; Draker, R.; Adachi, D.; Ayers, M.; et al. Identification of Severe Acute Respiratory Syndrome in Canada. N. Engl. J. Med. 2003, 348, 1995–2005. [Google Scholar] [CrossRef]
  14. Ksiazek, T.G.; Erdman, D.; Goldsmith, C.S.; Zaki, S.R.; Peret, T.; Emery, S.; Tong, S.; Urbani, C.; Comer, J.A.; Lim, W.; et al. A Novel Coronavirus Associated with Severe Acute Respiratory Syndrome. N. Engl. J. Med. 2003, 348, 1953–1966. [Google Scholar] [CrossRef]
  15. Peiris, J.S.M.; Lai, S.T.; Poon, L.L.M.; Guan, Y.; Yam, L.Y.C.; Lim, W.; Nicholls, J.; Yee, W.K.S.; Yan, W.W.; Cheung, M.T.; et al. Coronavirus as a Possible Cause of Severe Acute Respiratory Syndrome. Lancet 2003, 361, 1319–1325. [Google Scholar] [CrossRef] [Green Version]
  16. Killerby, M.E.; Biggs, H.M.; Midgley, C.M.; Gerber, S.I.; Watson, J.T. Middle East Respiratory Syndrome Coronavirus Transmission. Emerg. Infect. Dis. 2020, 26, 191–198. [Google Scholar] [CrossRef] [Green Version]
  17. Zaki, A.M.; van Boheemen, S.; Bestebroer, T.M.; Osterhaus, A.D.M.E.; Fouchier, R.A.M. Isolation of a Novel Coronavirus from a Man with Pneumonia in Saudi Arabia. N. Engl. J. Med. 2012, 367, 1814–1820. [Google Scholar] [CrossRef]
  18. Al Johani, S.; Hajeer, A.H. MERS-CoV Diagnosis: An Update. J. Infect. Public Health 2016, 9, 216–219. [Google Scholar] [CrossRef] [Green Version]
  19. Seddiq, N.; Al-Qahtani, M.; Al-Tawfiq, J.A.; Bukamal, N. First Confirmed Case of Middle East Respiratory Syndrome Coronavirus Infection in the Kingdom of Bahrain: In a Saudi Gentleman after Cardiac Bypass Surgery. Case Rep. Infect. Dis. 2017, 2017, 1–4. [Google Scholar] [CrossRef] [Green Version]
  20. Zheng, F.; Liao, C.; Fan, Q.H.; Chen, H.B.; Zhao, X.G.; Xie, Z.G.; Li, X.L.; Chen, C.X.; Lu, X.X.; Liu, Z.S.; et al. Clinical Characteristics of Children with Coronavirus Disease 2019 in Hubei, China. Curr. Med. Sci. 2020, 40, 275–280. [Google Scholar] [CrossRef] [Green Version]
  21. Tay, M.Z.; Poh, C.M.; Rénia, L.; MacAry, P.A.; Ng, L.F.P. The Trinity of COVID-19: Immunity, Inflammation and Intervention. Nat. Rev. Immunol. 2020, 20, 363–374. [Google Scholar] [CrossRef] [PubMed]
  22. COVID-19 Map - Johns Hopkins Coronavirus Resource Center. Available online: https://0-coronavirus-jhu-edu.brum.beds.ac.uk/map.html (accessed on 31 May 2021).
  23. Toyoshima, Y.; Nemoto, K.; Matsumoto, S.; Nakamura, Y.; Kiyotani, K. SARS-CoV-2 Genomic Variations Associated with Mortality Rate of COVID-19. J. Hum. Genet. 2020, 65, 1075–1082. [Google Scholar] [CrossRef] [PubMed]
  24. 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]
  25. Dong, Y.; Dai, T.; Wei, Y.; Zhang, L.; Zheng, M.; Zhou, F. A Systematic Review of SARS-CoV-2 Vaccine Candidates. Signal Transduct. Target. Ther. 2020, 5, 1–14. [Google Scholar] [CrossRef]
  26. Elbe, S.; Buckland-Merrett, G. Data, Disease and Diplomacy: GISAID’s Innovative Contribution to Global Health. Glob. Challenges 2017, 1, 33–46. [Google Scholar] [CrossRef] [Green Version]
  27. Agarwala, R.; Barrett, T.; Beck, J.; Benson, D.A.; Bollin, C.; Bolton, E.; Bourexis, D.; Brister, J.R.; Bryant, S.H.; Canese, K.; et al. Database Resources of the National Center for Biotechnology Information. Nucleic Acids Res. 2018, 46, D8–D13. [Google Scholar]
  28. Hadfield, J.; Megill, C.; Bell, S.M.; Huddleston, J.; Potter, B.; Callender, C.; Sagulenko, P.; Bedford, T.; Neher, R.A. NextStrain: Real-Time Tracking of Pathogen Evolution. Bioinformatics 2018, 34, 4121–4123. [Google Scholar] [CrossRef]
  29. Rambaut, A.; Holmes, E.C.; O’Toole, Á.; Hill, V.; McCrone, J.T.; Ruis, C.; du Plessis, L.; Pybus, O.G. A Dynamic Nomenclature Proposal for SARS-CoV-2 Lineages to Assist Genomic Epidemiology. Nat. Microbiol. 2020, 5, 1403–1407. [Google Scholar] [CrossRef]
  30. GISAID Clade and Lineage Nomenclature Aids in Genomic Epidemiology Studies of Active HCoV-19 Viruses. Available online: https://www.gisaid.org/references/statements-clarifications/clade-and-lineage-nomenclature-aids-in-genomic-epidemiology-of-active-hcov-19-viruses/ (accessed on 31 May 2021).
  31. Tang, X.; Wu, C.; Li, X.; Song, Y.; Yao, X.; Wu, X.; Duan, Y.; Zhang, H.; Wang, Y.; Qian, Z.; et al. On the Origin and Continuing Evolution of SARS-CoV-2. Natl. Sci. Rev. 2020, 7, 1012–1023. [Google Scholar] [CrossRef] [Green Version]
  32. Hodcroft, E.B.; Hadfield, J.; Neher, R.A.; Bedford, T. Year-Letter Genetic Clade Naming for SARS-CoV-2 on Nextstain.Org. Available online: https://nextstrain.org/blog/2020-06-02-SARSCoV2-clade-naming (accessed on 31 May 2021).
  33. 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]
  34. Xia, S.; Lan, Q.; Su, S.; Wang, X.; Xu, W.; Liu, Z.; Zhu, Y.; Wang, Q.; Lu, L.; Jiang, S. The Role of Furin Cleavage Site in SARS-CoV-2 Spike Protein-Mediated Membrane Fusion in the Presence or Absence of Trypsin. Signal Transduct. Target. Ther. 2020, 5, 92. [Google Scholar] [CrossRef]
  35. Hakim, M.S. SARS-CoV-2, Covid-19, and the Debunking of Conspiracy Theories. Rev. Med. Virol. 2021, e2222. [Google Scholar] [CrossRef]
  36. Zhou, H.; Chen, X.; Hu, T.; Li, J.; Song, H.; Liu, Y.; Wang, P.; Liu, D.; Yang, J.; Holmes, E.C.; et al. A Novel Bat Coronavirus Closely Related to SARS-CoV-2 Contains Natural Insertions at the S1/S2 Cleavage Site of the Spike Protein. Curr. Biol. 2020, 30, 2196–2203.e3. [Google Scholar] [CrossRef]
  37. Lam, T.T.Y.; Jia, N.; Zhang, Y.W.; Shum, M.H.H.; Jiang, J.F.; Zhu, H.C.; Tong, Y.G.; Shi, Y.X.; Ni, X.B.; Liao, Y.S.; et al. Identifying SARS-CoV-2-Related Coronaviruses in Malayan Pangolins. Nature 2020, 583, 282–285. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, T.; Wu, Q.; Zhang, Z. Probable Pangolin Origin of SARS-CoV-2 Associated with the COVID-19 Outbreak. Curr. Biol. 2020, 30, 1346–1351.e2. [Google Scholar] [CrossRef]
  39. Korber, B.; Fischer, W.M.; Gnanakaran, S.; Yoon, H.; Theiler, J.; Abfalterer, W.; Hengartner, N.; Giorgi, E.E.; Bhattacharya, T.; Foley, B.; et al. Tracking Changes in SARS-CoV-2 Spike: Evidence That D614G Increases Infectivity of the COVID-19 Virus. Cell 2020, 182, 812–827.e19. [Google Scholar] [CrossRef]
  40. Volz, E.; Hill, V.; McCrone, J.T.; Price, A.; Jorgensen, D.; O’Toole, Á.; Southgate, J.; Johnson, R.; Jackson, B.; Nascimento, F.F.; et al. Evaluating the Effects of SARS-CoV-2 Spike Mutation D614G on Transmissibility and Pathogenicity. Cell 2021, 184, 64–75.e11. [Google Scholar] [CrossRef]
  41. Hou, Y.J.; Chiba, S.; Halfmann, P.; Ehre, C.; Kuroda, M.; Dinnon, K.H.; Leist, S.R.; Schäfer, A.; Nakajima, N.; Takahashi, K.; et al. SARS-CoV-2 D614G Variant Exhibits Efficient Replication Ex Vivo and Transmission in Vivo. Science 2021, 370, 1464–1468. [Google Scholar]
  42. PHE SARS-CoV-2 Variants of Concern and Variants under Investigation in England. Available online: https://www.gov.uk/government/publications/covid-19-variants-genomically-confirmed-case-numbers/variants-distribution-of-cases-data (accessed on 31 May 2021).
  43. CDC SARS-CoV-2 Variant Classifications and Definitions. Available online: https://www.cdc.gov/coronavirus/2019-ncov/variants/variant-info.html (accessed on 31 May 2021).
  44. Davies, N.G.; Abbott, S.; Barnard, R.C.; Jarvis, C.I.; Kucharski, A.J.; Munday, J.D.; Pearson, C.A.B.; Russell, T.W.; Tully, D.C.; Washburne, A.D.; et al. Estimated Transmissibility and Impact of SARS-CoV-2 Lineage B.1.1.7 in England. Science 2021, 372, eabg3055. [Google Scholar] [CrossRef]
  45. Ferasin, L.; Fritz, M.; Ferasin, H.; Legros, V.; Leroy, E.M. Myocarditis in Naturally Infected Pets with the British Variant of COVID-19. bioRxiv 2021. [Google Scholar] [CrossRef]
  46. Fujino, T.; Nomoto, H.; Kutsuna, S.; Ujiie, M.; Suzuki, T.; Sato, R.; Fujimoto, T.; Kuroda, M.; Wakita, T.; Ohmagari, N. Novel SARS-CoV-2 Variant Identified in Travelers from Brazil to Japan. Emerg. Infect. Dis. 2021, 27, 1243–1245. [Google Scholar] [CrossRef] [PubMed]
  47. Sabino, E.C.; Buss, L.F.; Carvalho, M.P.S.; Prete, C.A.; Crispim, M.A.E.; Fraiji, N.A.; Pereira, R.H.M.; Parag, K.V.; da Silva Peixoto, P.; Kraemer, M.U.G.; et al. Resurgence of COVID-19 in Manaus, Brazil, despite High Seroprevalence. Lancet 2021, 397, 452–455. [Google Scholar] [CrossRef]
  48. Naveca, F.; Souza, V.; Corado, A.; Nascimento, F.; Silva, G.; Costa, Á.; Duarte, D.; Pessoa, K.; Mejía, M.; Brandão, M.; et al. COVID-19 Epidemic in the Brazilian State of Amazonas Was Driven by Long-Term Persistence of Endemic SARS-CoV-2 Lineages and the Recent Emergence of the New Variant of Concern P.1. Res. Sq. 2021. [Google Scholar] [CrossRef]
  49. Faria, N.R.; Mellan, T.A.; Whittaker, C.; Claro, I.M.; Candido, D.d.S.; Mishra, S.; Crispim, M.A.E.; Sales, F.C.S.; Hawryluk, I.; McCrone, J.T.; et al. Genomics and Epidemiology of the P.1 SARS-CoV-2 Lineage in Manaus, Brazil. Science 2021, 372, 815–821. [Google Scholar] [CrossRef]
  50. Lan, J.; Ge, J.; Yu, J.; Shan, S.; Zhou, H.; Fan, S.; Zhang, Q.; Shi, X.; Wang, Q.; Zhang, L.; et al. Structure of the SARS-CoV-2 Spike Receptor-Binding Domain Bound to the ACE2 Receptor. Nature 2020, 581, 215–220. [Google Scholar] [CrossRef] [Green Version]
  51. Shang, J.; Ye, G.; Shi, K.; Wan, Y.; Luo, C.; Aihara, H.; Geng, Q.; Auerbach, A.; Li, F. Structural Basis of Receptor Recognition by SARS-CoV-2. Nature 2020, 581, 221–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Starr, T.N.; Greaney, A.J.; Hilton, S.K.; Ellis, D.; Crawford, K.H.D.; Dingens, A.S.; Navarro, M.J.; Bowen, J.E.; Tortorici, M.A.; Walls, A.C.; et al. Deep Mutational Scanning of SARS-CoV-2 Receptor Binding Domain Reveals Constraints on Folding and ACE2 Binding. Cell 2020, 182, 1295–1310.e20. [Google Scholar] [CrossRef]
  53. Tian, F.; Tong, B.; Sun, L.; Shi, S.; Zheng, B.; Wang, Z.; Dong, X.; Zheng, P. Mutation N501Y in RBD of Spike Protein Strengthens the Inter-Action between COVID-19 and Its Receptor ACE2. bioRxiv 2021. [Google Scholar] [CrossRef]
  54. FDA Fact Sheet for Health Care Providers Emergency Use Authorization (EUA) of Bamlanivimab and Etesevimab. Available online: https://www.fda.gov/media/145802/download (accessed on 31 May 2021).
  55. Wang, P.; Nair, M.S.; Liu, L.; Iketani, S.; Luo, Y.; Guo, Y.; Wang, M.; Yu, J.; Zhang, B.; Kwong, P.D.; et al. Antibody Resistance of SARS-CoV-2 Variants B.1.351 and B.1.1.7. Nature 2021, 593, 130–135. [Google Scholar] [CrossRef]
  56. FDA Fact Sheet for Health Care Providers Emergency Use Authorization (EUA) of Regen-Covtm (Casirivimab with Imdevimab). Available online: https://www.fda.gov/media/145611/download (accessed on 31 May 2021).
  57. Wang, P.; Casner, R.G.; Nair, M.S.; Wang, M.; Yu, J.; Cerutti, G.; Liu, L.; Kwong, P.D.; Huang, Y.; Shapiro, L.; et al. Increased Resistance of SARS-CoV-2 Variant P.1 to Antibody Neutralization. Cell Host Microbe 2021, 29, 747–751.e4. [Google Scholar] [CrossRef]
  58. Edara, V.V.; Floyd, K.; Lai, L.; Gardner, M.; Hudson, W.; Piantadosi, A.; Waggoner, J.J.; Babiker, A.; Ahmed, R.; Xie, X.; et al. Infection and MRNA-1273 Vaccine Antibodies Neutralize SARS-CoV-2 UK Variant. medRxiv 2021. [Google Scholar] [CrossRef]
  59. Wu, K.; Werner, A.P.; Moliva, J.I.; Koch, M.; Choi, A.; Stewart-Jones, G.B.E.; Bennett, H.; Boyoglu-Barnum, S.; Shi, W.; Graham, B.S.; et al. mRNA-1273 Vaccine Induces Neutralizing Antibodies against Spike Mutants from Global SARS-CoV-2 Variants. bioRxiv 2021. [Google Scholar] [CrossRef]
  60. Collier, D.A.; De Marco, A.; Ferreira, I.A.T.M.; Meng, B.; Datir, R.; Walls, A.C.; Kemp S, S.A.; Bassi, J.; Pinto, D.; Fregni, C.S.; et al. SARS-CoV-2 B.1.1.7 Sensitivity to MRNA Vaccine-Elicited, Convalescent and Monoclonal Antibodies. medRxiv 2021. [Google Scholar] [CrossRef]
  61. Altmann, D.M.; Boyton, R.J.; Beale, R. Immunity to SARS-CoV-2 Variants of Concern. Science 2021, 371, 1103–1104. [Google Scholar] [CrossRef]
  62. Lule Bugembe, D.; VTPhan, M.; Ssewanyana, I.; Semanda, P.; Dhaala, B.; Nabadda, S.; Niamh, Á.O.; Rambaut, A.; Kaleebu, P.; Cotten, M. A SARS-CoV-2 Lineage A Variant (A.23.1) with Altered Spike Has Emerged and Is 1 Dominating the Current Uganda Epidemic 2. medRxiv 2021. [Google Scholar] [CrossRef]
  63. Deng, X.; Garcia-knight, M.A.; Khalid, M.M.; Servellita, V.; Morris, M.K.; Sotomayor-gonzález, A.; Glasner, D.R.; Reyes, R.; Gliwa, A.S.; Reddy, N.P.; et al. Transmission, Infectivity, and Neutralization of a Spike L452R SARS-CoV-2 Variant. Cell 2021, in press. [Google Scholar] [CrossRef]
  64. Yadav, P.D.; Sapkal, G.N.; Abraham, P.; Ella, R.; Deshpande, G.; Patil, D.Y.; Nyayanit, D.A.; Gupta, N.; Sahay, R.R.; Shete, A.M.; et al. Neutralization of Variant under Investigation B.1.617 with Sera of BBV152 Vaccinees. Clin. Infect. Dis. 2021, in press. [Google Scholar] [CrossRef]
  65. Rahbari, R.; Moradi, N.; Abdi, M. RRT-PCR for SARS-CoV-2: Analytical Considerations. Clin. Chim. Acta 2021, 516, 1–7. [Google Scholar] [CrossRef]
  66. Eftekhari, A.; Alipour, M.; Chodari, L.; Dizaj, S.M.; Ardalan, M.R.; Samiei, M.; Sharifi, S.; Vahed, S.Z.; Huseynova, I.; Khalilov, R.; et al. A Comprehensive Review of Detection Methods for SARS-CoV-2. Microorganisms 2021, 9, 232. [Google Scholar] [CrossRef]
  67. Sharma, K.; Aggarwala, P.; Gandhi, D.; Mathias, A.; Singh, P.; Sharma, S.; Negi, S.S.; Bhargava, A.; Das, P.; Gaikwad, U.; et al. Comparative Analysis of Various Clinical Specimens in Detection of SARS-CoV-2 Using rRT-PCR in New and Follow up Cases of COVID-19 Infection: Quest for the Best Choice. PLoS ONE 2021, 16, e0249408. [Google Scholar] [CrossRef]
  68. Tong, Y.; Bao, A.; Chen, H.; Huang, J.; Lv, Z.; Feng, L.; Cheng, Y.; Wang, Y.; Bai, L.; Rao, W.; et al. Necessity for Detection of SARS-CoV-2 RNA in Multiple Types of Specimens for the Discharge of the Patients with COVID-19. J. Transl. Med. 2020, 18, 1–8. [Google Scholar] [CrossRef]
  69. Udvardi, M.K.; Czechowski, T.; Scheible, W.R. Eleven Golden Rules of Quantitative RT-PCR. Plant Cell 2008, 20, 1736–1737. [Google Scholar] [CrossRef] [Green Version]
  70. Larionov, A.; Krause, A.; Miller, W.R. A Standard Curve Based Method for Relative Real Time PCR Data Processing. BMC Bioinform. 2005, 6, 1–16. [Google Scholar] [CrossRef] [Green Version]
  71. Sherina, V.; McMurray, H.R.; Powers, W.; Land, H.; Love, T.M.T.; McCall, M.N. Statistical Approaches to Decreasing the Discrepancy of Non-detects in qPCR Data. bioRxiv 2017. [Google Scholar] [CrossRef]
  72. Park, C.; Lee, J.; Hassan, Z.U.; Ku, K.B.; Kim, S.J.; Kim, H.G.; Park, E.C.; Park, G.S.; Park, D.; Baek, S.H.; et al. Comparison of Digital PCR and Quantitative PCR with Various SARS-CoV-2 Primer-Probe Sets. J. Microbiol. Biotechnol. 2021, 31, 358–367. [Google Scholar] [CrossRef]
  73. Han, M.S.; Byun, J.H.; Cho, Y.; Rim, J.H. RT-PCR for SARS-CoV-2: Quantitative versus Qualitative. Lancet Infect. Dis. 2021, 21, 165. [Google Scholar] [CrossRef]
  74. Wölfel, R.; Corman, V.M.; Guggemos, W.; Seilmaier, M.; Zange, S.; Müller, M.A.; Niemeyer, D.; Jones, T.C.; Vollmar, P.; Rothe, C.; et al. Virological Assessment of Hospitalized Patients with COVID-2019. Nature 2020, 581, 465–469. [Google Scholar] [CrossRef] [Green Version]
  75. Pan, Y.; Zhang, D.; Yang, P.; Poon, L.L.M.; Wang, Q. Viral Load of SARS-CoV-2 in Clinical Samples. Lancet Infect. Dis. 2020, 20, 411–412. [Google Scholar] [CrossRef]
  76. Arnaout, R.; Lee, R.A.; Lee, G.R.; Callahan, C.; Cheng, A.; Yen, C.F.; Smith, K.P.; Arora, R.; Kirby, J.E. The Limit of Detection Matters: The Case for Benchmarking Severe Acute Respiratory Syndrome Coronavirus 2 Testing. Clin. Infect. Dis. 2021, in press. [Google Scholar] [CrossRef] [PubMed]
  77. Platts, A.E.; Johnson, G.D.; Linnemann, A.K.; Krawetz, S.A. Real-Time PCR Quantification Using a Variable Reaction Efficiency Model. Anal. Biochem. 2008, 380, 315–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Lievens, A.; Van Aelst, S.; Van Den Bulcke, M.; Goetghebeur, E. Enhanced Analysis of Real-Time PCR Data by Using a Variable Efficiency Model: FPK-PCR. Nucleic Acids Res. 2012, 40, e10. [Google Scholar] [CrossRef]
  79. Forootan, A.; Sjöback, R.; Björkman, J.; Sjögreen, B.; Linz, L.; Kubista, M. Methods to Determine Limit of Detection and Limit of Quantification in Quantitative Real-Time PCR (QPCR). Biomol. Detect. Quantif. 2017, 12, 1–6. [Google Scholar] [CrossRef]
  80. Wikramaratna, P.S.; Paton, R.S.; Ghafari, M.; Lourenço, J. Estimating the False-Negative Test Probability of SARSCoV- 2 by RT-PCR. Eurosurveillance 2020, 25, 2000568. [Google Scholar] [CrossRef]
  81. Xiao, A.T.; Tong, Y.X.; Zhang, S. False Negative of RT-PCR and Prolonged Nucleic Acid Conversion in COVID-19: Rather than Recurrence. J. Med. Virol. 2020, 92, 1755–1756. [Google Scholar] [CrossRef] [Green Version]
  82. Erster, O.; Shkedi, O.; Benedek, G.; Zilber, E.; Varkovitzky, I.; Shirazi, R.; Shorka, D.O.; Cohen, Y.; Bar, T.; Yechieli, R.; et al. Improved Sensitivity, Safety, and Rapidity of COVID-19 Tests by Replacing Viral Storage Solution with Lysis Buffer. PLoS ONE 2021, 16, e0249149. [Google Scholar] [CrossRef]
  83. Tsang, N.N.Y.; So, H.C.; Ng, K.Y.; Cowling, B.J.; Leung, G.M.; Ip, D.K.M. Diagnostic Performance of Different Sampling Approaches for SARS-CoV-2 RT-PCR Testing: A Systematic Review and Meta-Analysis. Lancet Infect. Dis. 2021, in press. [Google Scholar] [CrossRef]
  84. Karligkiotis, A.; Arosio, A.; Castelnuovo, P. How to Obtain a Nasopharyngeal Swab Specimen. N. Engl. J. Med. 2020, 383, e14. [Google Scholar]
  85. Pinninti, S.; Trieu, C.; Pati, S.K.; Latting, M.; Cooper, J.; Seleme, M.C.; Boppana, S.; Arora, N.; Britt, W.J.; Boppana, S.B. Comparing Nasopharyngeal and Midturbinate Nasal Swab Testing for the Identification of Severe Acute Respiratory Syndrome Coronavirus 2. Clin. Infect. Dis. 2021, 72, 1253–1255. [Google Scholar] [CrossRef]
  86. Mohammadi, A.; Esmaeilzadeh, E.; Li, Y.; Bosch, R.J.; Li, J.Z. SARS-CoV-2 Detection in Different Respiratory Sites: A Systematic Review and Meta-Analysis. EBioMedicine 2020, 59, 102903. [Google Scholar] [CrossRef]
  87. Fani, M.; Teimoori, A.; Ghafari, S. Comparison of the COVID-2019 (SARS-CoV-2) Pathogenesis with SARS-CoV and MERS-CoV Infections. Future Virol. 2020, 15, 317–323. [Google Scholar] [CrossRef]
  88. 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 Heal. Res. Perspect. 2020, 11, 112–117. [Google Scholar] [CrossRef]
  89. Wang, W.; Xu, Y.; Gao, R.; Al, E. Detection of SARS-CoV-2 in Different Types of Clinical Specimens. JAMA 2020, 323, 1843–1844. [Google Scholar] [CrossRef] [Green Version]
  90. Vogels, C.B.F.; Brito, A.F.; Wyllie, A.L.; Fauver, J.R.; Ott, I.M.; Kalinich, C.C.; Petrone, M.E.; Casanovas-Massana, A.; Catherine Muenker, M.; Moore, A.J.; et al. Analytical Sensitivity and Efficiency Comparisons of SARS-CoV-2 RT–QPCR Primer–Probe Sets. Nat. Microbiol. 2020, 5, 1299–1305. [Google Scholar] [CrossRef]
  91. Jung, Y.; Park, G.S.; Moon, J.H.; Ku, K.; Beak, S.H.; Lee, C.S.; Kim, S.; Park, E.C.; Park, D.; Lee, J.H.; et al. Comparative Analysis of Primer-Probe Sets for RT-qPCR of COVID-19 Causative Virus (SARS-CoV-2). ACS Infect. Dis. 2020, 6, 2513–2523. [Google Scholar] [CrossRef]
  92. Pillonel, T.; Scherz, V.; Jaton, K.; Greub, G.; Bertelli, C. Letter to the Editor: SARS-CoV-2 Detection by Real-Time RT-PCR. Eurosurveillance 2020, 25, 1–2. [Google Scholar] [CrossRef]
  93. Klungthong, C.; Chinnawirotpisan, P.; Hussem, K.; Phonpakobsin, T.; Manasatienkij, W.; Ajariyakhajorn, C.; Rungrojcharoenkit, K.; Gibbons, R.V.; Jarman, R.G. The Impact of Primer and Probe-Template Mismatches on the Sensitivity of Pandemic Influenza A/H1N1/2009 Virus Detection by Real-Time RT-PCR. J. Clin. Virol. 2010, 48, 91–95. [Google Scholar] [CrossRef]
  94. Nayar, G.; Seabolt, E.E.; Kunitomi, M.; Agarwal, A.; Beck, K.L.; Mukherjee, V.; Kaufman, J.H. Analysis and Forecasting of Global Real Time RT-PCR Primers and Probes for SARS-CoV-2. Sci. Rep. 2021, 11, 8988. [Google Scholar] [CrossRef]
  95. van Kasteren, P.B.; van der Veer, B.; van den Brink, S.; Wijsman, L.; de Jonge, J.; van den Brandt, A.; Molenkamp, R.; Reusken, C.B.E.M.; Meijer, A. Comparison of Seven Commercial RT-PCR Diagnostic Kits for COVID-19. J. Clin. Virol. 2020, 128, 104412. [Google Scholar] [CrossRef]
  96. Simas, A.M.; Crott, J.W.; Sedore, C.; Rohrbach, A.; Monaco, A.P.; Gabriel, S.B.; Lennon, N.; Blumenstiel, B.; Genco, C.A. Pooling for SARS-CoV2 Surveillance: Validation and Strategy for Implementation in K-12 Schools. medRxiv 2020. [Google Scholar] [CrossRef]
  97. Sun, Q.; Li, J.; Ren, H.; Pastor, L.; Loginova, Y.; Madej, R.; Taylor, K.; Wong, J.K.; Zhang, Z.; Zhang, A.; et al. Saliva as a Testing Specimen with or without Pooling for SARS-CoV-2 Detection by Multiplex RT-PCR Test. PLoS ONE 2021, 16, e0243183. [Google Scholar] [CrossRef]
  98. de Salazar, A.; Aguilera, A.; Trastoy, R.; Fuentes, A.; Alados, J.C.; Causse, M.; Galán, J.C.; Moreno, A.; Trigo, M.; Pérez-Ruiz, M.; et al. Sample Pooling for SARS-CoV-2 RT-PCR Screening. Clin. Microbiol. Infect. 2020, 26, 1687.e1–1687.e5. [Google Scholar] [CrossRef] [PubMed]
  99. Barak, N.; Ben-Ami, R.; Sido, T.; Perri, A.; Shtoyer, A.; Rivkin, M.; Licht, T.; Peretz, A.; Magenheim, J.; Fogel, I.; et al. Lessons from Applied Large-Scale Pooling of 133,816 SARS-CoV-2 RT-PCR Tests. Sci. Transl. Med. 2021, 13, eabf2823. [Google Scholar] [CrossRef] [PubMed]
  100. Cleary, B.; Hay, J.A.; Blumenstiel, B.; Harden, M.; Cipicchio, M.; Bezney, J.; Simonton, B.; Hong, D.; Senghore, M.; Sesay, A.K.; et al. Using Viral Load and Epidemic Dynamics to Optimize Pooled Testing in Resource-Constrained Settings. Sci. Transl. Med. 2021, 13, eabf1568. [Google Scholar] [CrossRef] [PubMed]
  101. Kim, S.Y.; Lee, J.; Sung, H.; Lee, H.; Han, M.G.; Yoo, C.K.; Lee, S.W.; Hong, K.H. Pooling Upper Respiratory Specimens for Rapid Mass Screening of COVID-19 by Real-Time RT-PCR. Emerg. Infect. Dis. 2020, 26, 2469–2472. [Google Scholar] [CrossRef]
  102. Wacker, M.J.; Godard, M.P. Analysis of One-Step and Two- Step Real-Time RT-PCR Using Superscript III. J. Biomol. Tech. 2005, 16, 266–271. [Google Scholar]
  103. Dharavath, B.; Yadav, N.; Desai, S.; Sunder, R.; Mishra, R.; Ketkar, M.; Bhanshe, P.; Gupta, A.; Redhu, A.K.; Patkar, N.; et al. A One-Step, One-Tube Real-Time RT-PCR Based Assay with an Automated Analysis for Detection of SARS-CoV-2. Heliyon 2020, 6, e04405. [Google Scholar] [CrossRef]
  104. Yu, C.Y.; Chan, K.G.; Yean, C.Y.; Ang, G.Y. Nucleic Acid-Based Diagnostic Tests for the Detection SARS-CoV-2: An Update. Diagnostics 2021, 11, 53. [Google Scholar] [CrossRef]
  105. Miranda, J.P.; Osorio, J.; Videla, M.; Angel, G.; Camponovo, R.; Henríquez-Henríquez, M. Analytical and Clinical Validation for RT-qPCR Detection of SARS-CoV-2 Without RNA Extraction. Front. Med. 2020, 7, 567572. [Google Scholar] [CrossRef]
  106. Das Mukhopadhyay, C.; Sharma, P.; Sinha, K.; Rajarshi, K. Recent Trends in Analytical and Digital Techniques for the Detection of the SARS-Cov-2. Biophys. Chem. 2021, 270, 106538. [Google Scholar] [CrossRef]
  107. Park, M.; Won, J.; Choi, B.Y.; Lee, C.J. Optimization of Primer Sets and Detection Protocols for SARS-CoV-2 of Coronavirus Disease 2019 (COVID-19) Using PCR and Real-Time PCR. Exp. Mol. Med. 2020, 52, 963–977. [Google Scholar] [CrossRef]
  108. Chen, C.J.; Hsieh, L.L.; Lin, S.K.; Wang, C.F.; Huang, Y.H.; Lin, S.Y.; Lu, P.L. Optimization of the CDC Protocol of Molecular Diagnosis of COVID-19 for Timely Diagnosis. Diagnostics 2020, 10, 333. [Google Scholar] [CrossRef]
  109. Bustin, S.A.; Nolan, T. RT-qPCR Testing of SARS-COV-2: A Primer. Int. J. Mol. Sci. 2020, 21, 3004. [Google Scholar] [CrossRef]
  110. Wyllie, A.L.; Fournier, J.; Casanovas-Massana, A.; Campbell, M.; Tokuyama, M.; Vijayakumar, P.; Warren, J.L.; Geng, B.; Muenker, M.C.; Moore, A.J.; et al. Saliva or Nasopharyngeal Swab Specimens for Detection of SARS-CoV-2. N. Engl. J. Med. 2020, 383, 1283–1286. [Google Scholar] [CrossRef]
  111. Takeuchi, Y.; Furuchi, M.; Kamimoto, A.; Honda, K.; Matsumura, H.; Kobayashi, R. Saliva-Based PCR Tests for SARS-CoV-2 Detection. J. Oral Sci. 2020, 62, 350–351. [Google Scholar] [CrossRef]
  112. Gallo, G.; La Torre, M.; Pietroletti, R.; Bianco, F.; Altomare, D.F.; Pucciarelli, S.; Gagliardi, G.; Perinotti, R. Italian Society of Colorectal Surgery Recommendations for Good Clinical Practice in Colorectal Surgery during the Novel Coronavirus Pandemic. Tech. Coloproctol. 2020, 24, 501–505. [Google Scholar] [CrossRef] [Green Version]
  113. Mizumoto, K.; Kagaya, K.; Zarebski, A.; Chowell, G. Estimating the Asymptomatic Proportion of Coronavirus Disease 2019 (COVID-19) Cases on Board the Diamond Princess Cruise Ship, Yokohama, Japan, 2020. Eurosurveillance 2020, 25, 2000180. [Google Scholar] [CrossRef] [Green Version]
  114. Zhu, H.; Zhang, H.; Xu, Y.; Laššáková, S.; Korabečná, M.; Neužil, P. PCR Past, Present and Future. Biotechniques 2020, 69, 317–325. [Google Scholar] [CrossRef]
  115. Kojabad, A.A.; Farzanehpour, M.; Galeh, H.E.G.; Dorostkar, R.; Jafarpour, A.; Bolandian, M.; Nodooshan, M.M. Droplet Digital PCR of Viral DNA/RNA, Current Progress, Challenges, and Future Perspectives. J. Med. Virol. 2021, 93, 4182–4197. [Google Scholar] [CrossRef]
  116. Sedlak, R.H.; Jerome, K.R. Viral Diagnostics in the Era of Digital Polymerase Chain Reaction. Diagn. Microbiol. Infect. Dis. 2013, 75, 1–4. [Google Scholar] [CrossRef] [Green Version]
  117. Salipante, S.J.; Jerome, K.R. Digital PCR—An Emerging Technology with Broad Applications in Microbiology. Clin. Chem. 2020, 66, 117–123. [Google Scholar] [CrossRef]
  118. Quan, P.L.; Sauzade, M.; Brouzes, E. DPCR: A Technology Review. Sensors 2018, 18, 1271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Kuypers, J.; Jerome, K.R. Applications of Digital PCR for Clinical Microbiology. J. Clin. Microbiol. 2017, 55, 1621–1628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Alteri, C.; Cento, V.; Antonello, M.; Colagrossi, L.; Merli, M.; Ughi, N.; Renica, S.; Matarazzo, E.; Ruscio, F.D.; Tartaglione, L.; et al. Detection and Quantification of SARS-CoV-2 by Droplet Digital PCR in Real-Time PCR Negative Nasopharyngeal Swabs from Suspected COVID-19 Patients. PLoS ONE 2020, 15. [Google Scholar] [CrossRef] [PubMed]
  121. Falzone, L.; Musso, N.; Gattuso, G.; Bongiorno, D.; Palermo, C.I.; Scalia, G.; Libra, M.; Stefani, S. Sensitivity Assessment of Droplet Digital PCR for SARS-CoV-2 Detection. Int. J. Mol. Med. 2020, 46, 957–964. [Google Scholar] [CrossRef]
  122. Suo, T.; Liu, X.; Feng, J.; Guo, M.; Hu, W.; Guo, D.; Ullah, H.; Yang, Y.; Zhang, Q.; Wang, X.; et al. DdPCR: A More Accurate Tool for SARS-CoV-2 Detection in Low Viral Load Specimens. Emerg. Microbes Infect. 2020, 9, 1259–1268. [Google Scholar] [CrossRef]
  123. Dang, Y.; Liu, N.; Tan, C.; Feng, Y.; Yuan, X.; Fan, D.; Peng, Y.; Jin, R.; Guo, Y.; Lou, J. Comparison of Qualitative and Quantitative Analyses of COVID-19 Clinical Samples. Clin. Chim. Acta 2020, 510, 613–616. [Google Scholar] [CrossRef]
  124. Liu, X.; Feng, J.; Zhang, Q.; Guo, D.; Zhang, L.; Suo, T.; Hu, W.; Guo, M.; Wang, X.; Huang, Z.; et al. Analytical Comparisons of SARS-COV-2 Detection by qRT-PCR and ddPCR with Multiple Primer/Probe Sets. Emerg. Microbes Infect. 2020, 9, 1175–1179. [Google Scholar] [CrossRef]
  125. Kevadiya, B.D.; Machhi, J.; Herskovitz, J.; Oleynikov, M.D.; Blomberg, W.R.; Bajwa, N.; Soni, D.; Das, S.; Hasan, M.; Patel, M.; et al. Diagnostics for SARS-CoV-2 Infections. Nat. Mater. 2021, 20, 593–605. [Google Scholar] [CrossRef]
  126. Gatti, M.; Calandri, M.; Biondo, A.; Geninatti, C.; Piatti, C.; Ruggirello, I.; Santonocito, A.; Varello, S.; Bergamasco, L.; Bironzo, P.; et al. Emergency Room Comprehensive Assessment of Demographic, Radiological, Laboratory and Clinical Data of Patients with COVID-19: Determination of Its Prognostic Value for in-Hospital Mortality. Intern. Emerg. Med. 2021, in press. [Google Scholar] [CrossRef]
  127. Mostafa, H.H.; Hardick, J.; Morehead, E.; Miller, J.A.; Gaydos, C.A.; Manabe, Y.C. Comparison of the Analytical Sensitivity of Seven Commonly Used Commercial SARS-CoV-2 Automated Molecular Assays. J. Clin. Virol. 2020, 130, 104578. [Google Scholar] [CrossRef]
  128. D’Cruz, R.J.; Currier, A.W.; Sampson, V.B. Laboratory Testing Methods for Novel Severe Acute Respiratory Syndrome-Coronavirus-2 (SARS-CoV-2). Front. Cell Dev. Biol. 2020, 8, 1–11. [Google Scholar] [CrossRef]
  129. Ganguli, A.; Mostafa, A.; Berger, J.; Aydin, M.Y.; Sun, F.; Stewart de Ramirez, S.A.; Valera, E.; Cunningham, B.T.; King, W.P.; Bashir, R. Rapid Isothermal Amplification and Portable Detection System for SARS-CoV-2. Proc. Natl. Acad. Sci. USA 2020, 117, 22727–22735. [Google Scholar] [CrossRef]
  130. 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]
  131. Bodulev, O.L.; Sakharov, I.Y. Isothermal Nucleic Acid Amplification Techniques and Their Use in Bioanalysis. Biochemistry 2020, 85, 147–166. [Google Scholar] [CrossRef]
  132. Udugama, B.; Kadhiresan, P.; Kozlowski, H.N.; Malekjahani, A.; Osborne, M.; Li, V.Y.C.; Chen, H.; Mubareka, S.; Gubbay, J.B.; Chan, W.C.W. Diagnosing COVID-19: The Disease and Tools for Detection. ACS Nano 2020, 14, 3822–3835. [Google Scholar] [CrossRef] [Green Version]
  133. Mautner, L.; Baillie, C.K.; Herold, H.M.; Volkwein, W.; Guertler, P.; Eberle, U.; Ackermann, N.; Sing, A.; Pavlovic, M.; Goerlich, O.; et al. Rapid Point-of-Care Detection of SARS-CoV-2 Using Reverse Transcription Loop-Mediated Isothermal Amplification (RT-LAMP). Virol. J. 2020, 17, 1–14. [Google Scholar] [CrossRef]
  134. Qian, J.; Boswell, S.A.; Chidley, C.; Lu, Z.X.; Pettit, M.E.; Gaudio, B.L.; Fajnzylber, J.M.; Ingram, R.T.; Ward, R.H.; Li, J.Z.; et al. An Enhanced Isothermal Amplification Assay for Viral Detection. Nat. Commun. 2020, 11, 1–10. [Google Scholar] [CrossRef]
  135. Chaibun, T.; Puenpa, J.; Ngamdee, T.; Boonapatcharoen, N.; Athamanolap, P.; O’Mullane, A.P.; Vongpunsawad, S.; Poovorawan, Y.; Lee, S.Y.; Lertanantawong, B. Rapid Electrochemical Detection of Coronavirus SARS-CoV-2. Nat. Commun. 2021, 12, 1–10. [Google Scholar] [CrossRef]
  136. Zhang, C.; Zheng, T.; Fan, H.; Zhang, T.; Han, D. Aligner-Mediated Cleavage-Based Isothermal Amplification for SARS-CoV-2 RNA Detection. ACS Appl. Bio Mater. 2021, 4, 3805–3810. [Google Scholar] [CrossRef]
  137. Xing, W.; Liu, Y.; Wang, H.; Li, S.; Lin, Y.; Chen, L.; Zhao, Y.; Chao, S.; Huang, X.; Ge, S.; et al. A High-Throughput, Multi-Index Isothermal Amplification Platform for Rapid Detection of 19 Types of Common Respiratory Viruses Including SARS-CoV-2. Engineering 2020, 6, 1130–1140. [Google Scholar] [CrossRef]
  138. 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]
  139. Tomita, N.; Mori, Y.; Kanda, H.; Notomi, T. Loop-Mediated Isothermal Amplification (LAMP) of Gene Sequences and Simple Visual Detection of Products. Nat. Protoc. 2008, 3, 877–882. [Google Scholar] [CrossRef] [PubMed]
  140. Mori, Y.; Kitao, M.; Tomita, N.; Notomi, T. Real-Time Turbidimetry of LAMP Reaction for Quantifying Template DNA. J. Biochem. Biophys. Methods 2004, 59, 145–157. [Google Scholar] [CrossRef] [PubMed]
  141. Yu, L.; Wu, S.; Hao, X.; Dong, X.; Mao, L.; Pelechano, V.; Chen, W.-H.; Yin, X. Rapid Detection of COVID-19 Coronavirus Using a Reverse Transcriptional Loop-Mediated Isothermal Amplification (RT-LAMP) Diagnostic Platform. Clin. Chem. 2020, 66, 975–977. [Google Scholar] [CrossRef] [PubMed]
  142. Dao Thi, V.L.; Herbst, K.; Boerner, K.; Meurer, M.; Kremer, L.P.M.; Kirrmaier, D.; Freistaedter, A.; Papagiannidis, D.; Galmozzi, C.; Stanifer, M.L.; et al. A Colorimetric RT-LAMP Assay and LAMP-Sequencing for Detecting SARS-CoV-2 RNA in Clinical Samples. Sci. Transl. Med. 2020, 12, eabc7075. [Google Scholar] [CrossRef] [PubMed]
  143. Lu, R.; Wu, X.; Wan, Z.; Li, Y.; Jin, X.; Zhang, C. A Novel Reverse Transcription Loop-Mediated Isothermal Amplification Method for Rapid Detection of Sars-Cov-2. Int. J. Mol. Sci. 2020, 21, 2826. [Google Scholar] [CrossRef] [Green Version]
  144. Park, G.S.; Ku, K.; Baek, S.H.; Kim, S.J.; Kim, S.I.; Kim, B.T.; Maeng, J.S. Development of Reverse Transcription Loop-Mediated Isothermal Amplification Assays Targeting Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2). J. Mol. Diagnostics 2020, 22, 729–735. [Google Scholar] [CrossRef]
  145. Baek, Y.H.; Um, J.; Antigua, K.J.C.; Park, J.H.; Kim, Y.; Oh, S.; Kim, Y.I.; Choi, W.S.; Kim, S.G.; Jeong, J.H.; et al. Development of a Reverse Transcription-Loop-Mediated Isothermal Amplification as a Rapid Early-Detection Method for Novel SARS-CoV-2. Emerg. Microbes Infect. 2020, 9, 998–1007. [Google Scholar] [CrossRef] [Green Version]
  146. Wei, S.; Kohl, E.; Djandji, A.; Morgan, S.; Whittier, S.; Mansukhani, M.; Hod, E.; D’Alton, M.; Suh, Y.; Williams, Z. Direct Diagnostic Testing of SARS-CoV-2 without the Need for Prior RNA Extraction. Sci. Rep. 2021, 11, 1–6. [Google Scholar] [CrossRef]
  147. Lalli, M.A.; Langmade, J.S.; Chen, X.; Fronick, C.C.; Sawyer, C.S.; Burcea, L.C.; Wilkinson, M.N.; Fulton, R.S.; Heinz, M.; Buchser, W.J.; et al. Rapid and Extraction-Free Detection of SARS-CoV-2 from Saliva by Colorimetric Reverse-Transcription Loop-Mediated Isothermal Amplification. Clin. Chem. 2021, 67, 415–424. [Google Scholar] [CrossRef]
  148. Pang, B.; Xu, J.; Liu, Y.; Peng, H.; Feng, W.; Cao, Y.; Wu, J.; Xiao, H.; Pabbaraju, K.; Tipples, G.; et al. Isothermal Amplification and Ambient Visualization in a Single Tube for the Detection of SARS-CoV-2 Using Loop-Mediated Amplification and CRISPR Technology. Anal. Chem. 2020, 92, 16204–16212. [Google Scholar] [CrossRef]
  149. Ali, Z.; Aman, R.; Mahas, A.; Rao, G.S.; Tehseen, M.; Marsic, T.; Salunke, R.; Subudhi, A.K.; Hala, S.M.; Hamdan, S.M.; et al. ISCAN: An RT-LAMP-Coupled CRISPR-Cas12 Module for Rapid, Sensitive Detection of SARS-CoV-2. Virus Res. 2020, 288, 198129. [Google Scholar] [CrossRef]
  150. Zhu, X.; Wang, X.; Han, L.; Chen, T.; Wang, L.; Li, H.; Li, S.; He, L.; Fu, X.; Chen, S.; et al. Multiplex Reverse Transcription Loop-Mediated Isothermal Amplification Combined with Nanoparticle-Based Lateral Flow Biosensor for the Diagnosis of COVID-19. Biosens. Bioelectron. 2020, 166, 112437. [Google Scholar] [CrossRef]
  151. Rohaim, M.A.; Clayton, E.; Sahin, I.; Vilela, J.; Khalifa, M.E.; Al-natour, M.Q.; Bayoumi, M.; Poirier, A.C.; Branavan, M.; Tharmakulasingam, M.; et al. Artificial Intelligence-Assisted Loop Mediated Isothermal Amplification (AI-LAMP) for Rapid Detection of SARS-CoV-2. Viruses 2020, 12, 972. [Google Scholar] [CrossRef]
  152. Piepenburg, O.; Williams, C.H.; Stemple, D.L.; Armes, N.A. DNA Detection Using Recombination Proteins. PLoS Biol. 2006, 4, 1115–1121. [Google Scholar] [CrossRef]
  153. Kersting, S.; Rausch, V.; Bier, F.F.; von Nickisch-Rosenegk, M. Rapid Detection of Plasmodium Falciparum with Isothermal Recombinase Polymerase Amplification and Lateral Flow Analysis. Malar. J. 2014, 13, 99. [Google Scholar] [CrossRef] [Green Version]
  154. Lobato, I.M.; O’Sullivan, C.K. Recombinase Polymerase Amplification: Basics, Applications and Recent Advances. TrAC Trends Anal. Chem. 2018, 98, 19–35. [Google Scholar] [CrossRef]
  155. Zhang, W.S.; Pan, J.; Li, F.; Zhu, M.; Xu, M.; Zhu, H.; Yu, Y.; Su, G. Reverse Transcription Recombinase Polymerase Amplification Coupled with CRISPR-Cas12a for Facile and Highly Sensitive Colorimetric SARS-CoV-2 Detection. Anal. Chem. 2021, 93, 4126–4133. [Google Scholar] [CrossRef]
  156. Patchsung, M.; Jantarug, K.; Pattama, A.; Aphicho, K.; Suraritdechachai, S.; Meesawat, P.; Sappakhaw, K.; Leelahakorn, N.; Ruenkam, T.; Wongsatit, T.; et al. Clinical Validation of a Cas13-Based Assay for the Detection of SARS-CoV-2 RNA. Nat. Biomed. Eng. 2020, 4, 1140–1149. [Google Scholar] [CrossRef]
  157. Arizti-Sanz, J.; Freije, C.A.; Stanton, A.C.; Petros, B.A.; Boehm, C.K.; Siddiqui, S.; Shaw, B.M.; Adams, G.; Kosoko-Thoroddsen, T.S.F.; Kemball, M.E.; et al. Streamlined Inactivation, Amplification, and Cas13-Based Detection of SARS-CoV-2. Nat. Commun. 2020, 11, 5921. [Google Scholar] [CrossRef]
  158. Kim, H.S.; Abbas, N.; Shin, S. A Rapid Diagnosis of SARS-CoV-2 Using DNA Hydrogel Formation on Microfluidic Pores. Biosens. Bioelectron. 2021, 177, 113005. [Google Scholar] [CrossRef]
  159. WHO. Genomic Sequencing of SARS-CoV-2; WHO: Geneva, Switzerland, 2021; ISBN 9789240018440. [Google Scholar]
  160. Quick, J.; Loman, N.J.; Duraffour, S.; Simpson, J.T.; Severi, E.; Cowley, L.; Bore, J.A.; Koundouno, R.; Dudas, G.; Mikhail, A.; et al. Real-Time, Portable Genome Sequencing for Ebola Surveillance. Nature 2016, 530, 228–232. [Google Scholar] [CrossRef] [Green Version]
  161. Hoenen, T.; Groseth, A.; Rosenke, K.; Fischer, R.J.; Hoenen, A.; Judson, S.D.; Martellaro, C.; Falzarano, D.; Marzi, A.; Squires, R.B.; et al. Nanopore Sequencing as a Rapidly Deployable Ebola Outbreak Tool. Emerg. Infect. Dis. 2016, 22, 331–334. [Google Scholar] [CrossRef] [Green Version]
  162. Houldcroft, C.J.; Beale, M.A.; Breuer, J. Clinical and Biological Insights from Viral Genome Sequencing. Nat. Rev. Microbiol. 2017, 15, 183–192. [Google Scholar] [CrossRef]
  163. Myrmel, M.; Oma, V.; Khatri, M.; Hansen, H.H.; Stokstad, M.; Berg, M.; Blomström, A.-L. Single Primer Isothermal Amplification (SPIA) Combined with next Generation Sequencing Provides Complete Bovine Coronavirus Genome Coverage and Higher Sequence Depth Compared to Sequence-Independent Single Primer Amplification (SISPA). PLoS ONE 2017, 12, e0187780. [Google Scholar] [CrossRef] [Green Version]
  164. Reyes, G.R.; Kim, J.P. Sequence-Independent, Single-Primer Amplification (SISPA) of Complex DNA Populations. Mol. Cell. Probes 1991, 5, 473–481. [Google Scholar] [CrossRef]
  165. Maurier, F.; Beury, D.; Fléchon, L.; Varré, J.S.; Touzet, H.; Goffard, A.; Hot, D.; Caboche, S. A Complete Protocol for Whole-Genome Sequencing of Virus from Clinical Samples: Application to Coronavirus OC43. Virology 2019, 531, 141–148. [Google Scholar] [CrossRef] [PubMed]
  166. Kim, K.W.; Choi, S.; Shin, S.K.; Lee, I.; Ku, K.B.; Kim, S.J.; Kim, S.; Yi, H. A Half-Day Genome Sequencing Protocol for Middle East Respiratory Syndrome Coronavirus. Front. Microbiol. 2021, 12, 1–8. [Google Scholar] [CrossRef] [PubMed]
  167. Hall, R.J.; Wang, J.; Todd, A.K.; Bissielo, A.B.; Yen, S.; Strydom, H.; Moore, N.E.; Ren, X.; Huang, Q.S.; Carter, P.E.; et al. Evaluation of Rapid and Simple Techniques for the Enrichment of Viruses Prior to Metagenomic Virus Discovery. J. Virol. Methods 2014, 195, 194–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Deng, L.; Silins, R.; Castro-Mejía, J.L.; Kot, W.; Jessen, L.; Thorsen, J.; Shah, S.; Stokholm, J.; Bisgaard, H.; Moineau, S.; et al. A Protocol for Extraction of Infective Viromes Suitable for Metagenomics Sequencing from Low Volume Fecal Samples. Viruses 2019, 11, 667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Sathiamoorthy, S.; Malott, R.J.; Gisonni-Lex, L.; Ng, S.H.S. Selection and Evaluation of an Efficient Method for the Recovery of Viral Nucleic Acid Extraction from Complex Biologicals. Npj Vaccines 2018, 3, 1–6. [Google Scholar]
  170. Tan, C.C.S.; Maurer-Stroh, S.; Wan, Y.; Sessions, O.M.; de Sessions, P.F. A Novel Method for the Capture-Based Purification of Whole Viral Native RNA Genomes. AMB Express 2019, 9, 45. [Google Scholar] [CrossRef] [Green Version]
  171. Manso, C.F.; Bibby, D.F.; Mbisa, J.L. Efficient and Unbiased Metagenomic Recovery of RNA Virus Genomes from Human Plasma Samples. Sci. Rep. 2017, 7, 1–13. [Google Scholar] [CrossRef] [Green Version]
  172. Zhang, D.; Lou, X.; Yan, H.; Pan, J.; Mao, H.; Tang, H.; Shu, Y.; Zhao, Y.; Liu, L.; Li, J.; et al. Metagenomic Analysis of Viral Nucleic Acid Extraction Methods in Respiratory Clinical Samples 11 Medical and Health Sciences 1108 Medical Microbiology. BMC Genomics 2018, 19, 1–8. [Google Scholar]
  173. Temmam, S.; Monteil-Bouchard, S.; Robert, C.; Pascalis, H.; Michelle, C.; Jardot, P.; Charrel, R.; Raoult, D.; Desnues, C. Host-Associated Metagenomics: A Guide to Generating Infectious RNA Viromes. PLoS ONE 2015, 10, e0139810. [Google Scholar] [CrossRef] [Green Version]
  174. Ponce-Rojas, J.C.; Costello, M.S.; Proctor, D.A.; Kosik, K.S.; Wilson, M.Z.; Arias, C.; Acosta-Alvear, D. A Fast and Accessible Method for the Isolation of RNA, DNA, and Protein to Facilitate the Detection of SARS-CoV-2. J. Clin. Microbiol. 2020, 59, e02403-20. [Google Scholar]
  175. Jones, M.K.; Grau, K.R.; Costantini, V.; Kolawole, A.O.; de Graaf, M.; Freiden, P.; Graves, C.L.; Koopmans, M.; Wallet, S.M.; Tibbetts, S.A.; et al. Human Norovirus Culture in B Cells. Nat. Protoc. 2015, 10, 1939–1947. [Google Scholar] [CrossRef] [Green Version]
  176. Pyrc, K.; Sims, A.C.; Dijkman, R.; Jebbink, M.; Long, C.; Deming, D.; Donaldson, E.; Vabret, A.; Baric, R.; van der Hoek, L.; et al. Culturing the Unculturable: Human Coronavirus HKU1 Infects, Replicates, and Produces Progeny Virions in Human Ciliated Airway Epithelial Cell Cultures. J. Virol. 2010, 84, 11255–11263. [Google Scholar] [CrossRef] [Green Version]
  177. Charre, C.; Ginevra, C.; Sabatier, M.; Regue, H.; Destras, G.; Brun, S.; Burfin, G.; Scholtes, C.; Morfin, F.; Valette, M.; et al. Evaluation of NGS-Based Approaches for SARS-CoV- 2 Whole Genome Characterisation. Virus Evol. 2020, 6, 1–8. [Google Scholar] [CrossRef]
  178. Gaudin, M.; Desnues, C. Hybrid Capture-Based next Generation Sequencing and Its Application to Human Infectious Diseases. Front. Microbiol. 2018, 9, 1–9. [Google Scholar] [CrossRef] [Green Version]
  179. Tyson, J.R.; James, P.; Stoddart, D.; Sparks, N.; Wickenhagen, A.; Hall, G.; Choi, J.H.; Lapointe, H.; Kamelian, K.; Smith, A.D.; et al. Improvements to the ARTIC Multiplex PCR Method for SARS-CoV-2 Genome Sequencing Using Nanopore. bioRxiv 2020. [Google Scholar] [CrossRef]
  180. Liu, T.; Chen, Z.; Chen, W.; Chen, X.; Hosseini, M.; Yang, Z.; Li, J.; Ho, D.; Turay, D.; Gheorghe, C.; et al. A Benchmarking Study of SARS-CoV-2 Whole-Genome Sequencing Protocols Using COVID-19 Patient Samples. bioRxiv 2020. [Google Scholar] [CrossRef]
  181. Maurano, M.T.; Ramaswami, S.; Zappile, P.; Dimartino, D.; Boytard, L.; Ribeiro-dos-Santos, A.M.; Vulpescu, N.A.; Westby, G.; Shen, G.; Feng, X.; et al. Sequencing Identifies Multiple Early Introductions of SARS-CoV-2 to the New York City Region. Genome Res. 2020, 30, 1781–1788. [Google Scholar] [CrossRef] [PubMed]
  182. Shaibu, J.O.; Onwuamah, C.K.; James, A.B.; Okwuraiwe, A.P.; Amoo, O.S.; Salu, O.B.; Ige, F.A.; Liboro, G.; Odewale, E.; Okoli, L.C.; et al. Full Length Genomic Sanger Sequencing and Phylogenetic Analysis of Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) in Nigeria. PLoS ONE 2021, 16, 1–14. [Google Scholar] [CrossRef]
  183. Payne, A.; Holmes, N.; Rakyan, V.; Loose, M. Bulkvis: A Graphical Viewer for Oxford Nanopore Bulk FAST5 Files. Bioinformatics 2019, 35, 2193–2198. [Google Scholar] [CrossRef]
  184. Grubaugh, N.D.; Gangavarapu, K.; Quick, J.; Matteson, N.L.; De Jesus, J.G.; Main, B.J.; Tan, A.L.; Paul, L.M.; Brackney, D.E.; Grewal, S.; et al. An Amplicon-Based Sequencing Framework for Accurately Measuring Intrahost Virus Diversity Using PrimalSeq and IVar. Genome Biol. 2019, 20, 8. [Google Scholar] [CrossRef] [Green Version]
  185. Rezaei, M.; Razavi Bazaz, S.; Zhand, S.; Sayyadi, N.; Jin, D.; Stewart, M.P.; Ebrahimi Warkiani, M. Point of Care Diagnostics in the Age of COVID-19. Diagnostics 2020, 11, 9. [Google Scholar] [CrossRef]
  186. Wang, M.; Zhang, R.; Li, J. CRISPR/Cas Systems Redefine Nucleic Acid Detection: Principles and Methods. Biosens. Bioelectron. 2020, 165, 112430. [Google Scholar] [CrossRef]
  187. Bin Moon, S.; Yon Kim, D.; Ko, J.-H.; Kim, Y.-S. Recent Advances in the CRISPR Genome Editing Tool Set. Exp. Mol. Med. 2019, 51, 1–11. [Google Scholar] [CrossRef] [Green Version]
  188. Jiang, F.; Doudna, J.A. CRISPR-Cas9 Structures and Mechanisms. Annu. Rev. Biophys. 2017, 46, 505–529. [Google Scholar] [CrossRef] [Green Version]
  189. Zhang, W.; Liu, K.; Zhang, P.; Cheng, W.; Li, L.L.; Zhang, F.; Yu, Z.; Li, L.L.; Zhang, X. CRISPR-Based Approaches for Efficient and Accurate Detection of SARS-CoV-2. Lab. Med. 2021, 52, 116–121. [Google Scholar] [CrossRef]
  190. 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] [Green Version]
  191. Broughton, J.P.; Deng, X.; Yu, G.; Fasching, C.L.; Singh, J.; Streithorst, J.; Granados, A.; Sotomayor-Gonzalez, A.; Zorn, K.; Gopez, A.; et al. Rapid Detection of 2019 Novel Coronavirus SARS-CoV-2 Using a CRISPR-Based DETECTR Lateral Flow Assay. medRxiv 2020. [Google Scholar] [CrossRef]
  192. Kellner, M.J.; Koob, J.G.; Gootenberg, J.S.; Abudayyeh, O.O.; Zhang, F. SHERLOCK: Nucleic Acid Detection with CRISPR Nucleases. Nat. Protoc. 2019, 14, 2986–3012. [Google Scholar] [CrossRef]
  193. 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. [Google Scholar] [CrossRef]
  194. Kudr, J.; Michalek, P.; Ilieva, L.; Adam, V.; Zitka, O. COVID-19: A Challenge for Electrochemical Biosensors. Trends Anal. Chem. 2021, 136, 116192. [Google Scholar] [CrossRef]
  195. Varahachalam, S.P.; Lahooti, B.; Chamaneh, M.; Bagchi, S.; Chhibber, T.; Morris, K.; Bolanos, J.F.; Kim, N.Y.; Kaushik, A. Nanomedicine for the SARS-CoV-2: State-of-Theart and Future Prospects. Int. J. Nanomed. 2021, 16, 539–560. [Google Scholar] [CrossRef] [PubMed]
  196. Cheng, X.; Chen, G.; Rodriguez, W.R. Micro- and Nanotechnology for Viral Detection. Anal. Bioanal. Chem. 2009, 393, 487–501. [Google Scholar] [CrossRef]
  197. Tang, Z.; Kong, N.; Zhang, X.; Liu, Y.; Hu, P.; Mou, S.; Liljeström, P.; Shi, J.; Tan, W.; Kim, J.S.; et al. A Materials-Science Perspective on Tackling COVID-19. Nat. Rev. Mater. 2020, 5, 847–860. [Google Scholar] [CrossRef]
  198. Zhao, Z.; Cui, H.; Song, W.; Ru, X.; Zhou, W.; Yu, X. A Simple Magnetic Nanoparticles-Based Viral RNA Extraction Method for Efficient Detection of SARS-CoV-2. bioRxiv 2020. [Google Scholar] [CrossRef]
  199. Moitra, P.; Alafeef, M.; Alafeef, M.; Alafeef, M.; Dighe, K.; Frieman, M.B.; Pan, D.; Pan, D.; Pan, D. Selective Naked-Eye Detection of SARS-CoV-2 Mediated by N Gene Targeted Antisense Oligonucleotide Capped Plasmonic Nanoparticles. ACS Nano 2020, 14, 7617–7627. [Google Scholar] [CrossRef] [PubMed]
  200. Consultative Committee for Amount of Substance: Metrology in Chemistry and Biology (CCQM). Available online: https://www.bipm.org/en/committees/cc/ccqm (accessed on 31 May 2021).
  201. The Asia Pacific Metrology Programme (APMP). Available online: http://www.apmpweb.org/index.html (accessed on 31 May 2021).
  202. National Measurement Institutes Demonstrate High Accuracy Reference Measurement System for SARS-CoV-2 Testing. Available online: https://www.bipm.org/en/-/2020-nmi-covid (accessed on 31 May 2021).
  203. National Measurement Institute Helping Industry Respond to COVID-19. Available online: https://www.industry.gov.au/news/national-measurement-institute-helping-industry-respond-to-covid-19 (accessed on 31 May 2021).
  204. ISO. ISO 17034:2016 General Requirements for the Competence of Reference Material Producers; ISO: Geneva, Switzerland, 2016. [Google Scholar]
  205. Wise, S.A. What Is Novel about Certified Reference Materials? Anal. Bioanal. Chem. 2018, 410, 2045–2049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Pickett, B.E.; Sadat, E.L.; Zhang, Y.; Noronha, J.M.; Squires, R.B.; Hunt, V.; Liu, M.; Kumar, S.; Zaremba, S.; Gu, Z.; et al. ViPR: An Open Bioinformatics Database and Analysis Resource for Virology Research. Nucleic Acids Res. 2012, 40, 593–598. [Google Scholar] [CrossRef]
  207. Singer, J.; Gifford, R.; Cotten, M.; Robertson, D. CoV-GLUE: A Web Application for Tracking SARS-CoV-2 Genomic Variation. Preprints 2020. [Google Scholar] [CrossRef]
Figure 1. Overview of nucleic acid testing for SARS-CoV-2. The schematic procedure of RT-qPCR (A), and dPCR (B). Current isothermal amplification methods (C), CRISPR detection systems (D), and nanoparticles (E) are also shown.
Figure 1. Overview of nucleic acid testing for SARS-CoV-2. The schematic procedure of RT-qPCR (A), and dPCR (B). Current isothermal amplification methods (C), CRISPR detection systems (D), and nanoparticles (E) are also shown.
Ijms 22 06150 g001
Figure 2. Schematic procedure of reverse transcription loop-mediated isothermal amplification (RT-LAMP) and reverse transcription recombinase polymerase amplification (RT-RPA). FIP = Forward Inner Primer, FOP = Forward Outer Primer, BIP = Backward Inner Primer, BOP = Backward Outer Primer.
Figure 2. Schematic procedure of reverse transcription loop-mediated isothermal amplification (RT-LAMP) and reverse transcription recombinase polymerase amplification (RT-RPA). FIP = Forward Inner Primer, FOP = Forward Outer Primer, BIP = Backward Inner Primer, BOP = Backward Outer Primer.
Ijms 22 06150 g002
Figure 3. Schematic procedure of genome sequencing.
Figure 3. Schematic procedure of genome sequencing.
Ijms 22 06150 g003
Table 1. Nomenclature of SARS-CoV-2.
Table 1. Nomenclature of SARS-CoV-2.
GISAID CladesPANGO LineageNextstrain CladesNotable Variants
SA19BA.23.1
LB19AWuhan-Hu-1
V
GB.120AB.1.525, B.1.627
GHB.120CB.1.427, B.1.429, B.1.526
B.1.220G
B.1.596
B.1.35120H/501.Y.V2B.1.351
GRB.1.1.120B
P.3 P.3
C20D
D.220F
P.120J/501.Y.V3P.1
GVB.1.17720E (EU1)B.1.177
GRYB.1.1.720I/501.Y.V1B.1.1.7
Table 2. Notable variants of SARS-CoV-2.
Table 2. Notable variants of SARS-CoV-2.
PANGO LineageCDC DesignationPHE DesignationFirst DetectedSpike Protein Substitutions
B.1.1.7VOCVOC-20DEC-01, VOC-21FEB-02 *United Kingdom69del, 70del, 144del, (E484K), (S494P), N501Y, A570D, D614G, P681H, T716I, S982A, D1118H (K1191N)
B.1.351VOCVOC-20DEC-02South Africa D80A, D215G, 241del, 242del, 243del, K417N, E484K, N501Y, D614G, A701V
P.2VOIVUI-21JAN-01BrazilE484K, (F565L), D614G, V1176F
P.1VOCVOC-21JAN-02BrazilL18F, T20N, P26S, D138Y, R190S, K417T, E484K, N501Y, D614G, H655Y, T1027I
A.23.1-VUI-21FEB-01 *UgandaF157L, V367F, (E484K), Q613H, P681R
B.1.525VOIVUI-21FEB-03United KingdomA67V, 69del, 70del, 144del, E484K, D614G, Q677H, F888L
B.1.1.318-VUI-21FEB-04United KingdomD614G, D796H, E484K, P681H, T95I, 144del
P.3-VUI-21MAR-02PhilippinesE484K, N501Y, P681H
B.1.617VOIVUI-21APR-01India L452R, E484Q, D614G
B.1.617.2VOIVOC-21APR-02IndiaT19R, (G142D), 156del, 157del, R158G, L452R, T478K, D614G, P681R, D950N
B.1.617.3VOIVUI-21APR-03IndiaT19R, G142D, L452R, E484Q, D614G, P681R, D950N
AV.1-VUI-21MAY-01United KingdomD80G, T95I, G142D, 144del, N439K, E484K, D614G, P681H, I1130V, D1139H
B.1.617.1VOI-India(T95I), G142D, E154K, L452R, E484Q, D614G, P681R, Q1071H
B.1.526VOI-United States(L5F), T95I, D253G, (S477N), (E484K), D614G, (A701V)
B.1.526.1VOI-United StatesD80G, 144del, F157S, L452R, D614G, (T791I), (T859N), D950H
B.1.427VOC-United StatesL452R, D614G
B.1.429VOC-United StatesS13I, W152C, L452R, D614G
* with E484K; ( ) detected in some sequences but not all; VOC; Variant of Concern, VOI; Variant of Interest, VUI; Variant under Investigation.
Table 3. General features of the isothermal amplification techniques for SARS-CoV-2 detection.
Table 3. General features of the isothermal amplification techniques for SARS-CoV-2 detection.
MethodComponentsTemperatureTimeDetection MethodAdvantages *Disadvantages *
Loop-mediated isothermal amplification (LAMP)DNA polymerase, forward inner primer, backward inner primer, forward outer primer, backward outer primer60–65 °C>1 hColorimetric, turbimetric, fluorescence probe, intercalating dyeHigh specificity. Less sensitive to inhibitors in biological samplesFalse positive in negative control
Recombinase polymerase amplification (RPA)Recombinase, single stranded binding protein, DNA polymerase, forward primer, reverse primer37–42 °C>1 hFluorescence, antigenic-tag (antibody)Performed in the presence of PCR inhibitors. Fast and sensitiveInhibited by detergents (SDS and CTAB). Non-specific/high background signal
Nucleic acid sequence-based amplification (NASBA)RNase H, reverse transcriptase, T7 DNA-dependent RNA polymerase, forward primer with T7 promoter sequence, reverse primer41 °C>2 hFluorescenceMore sensitive and less time-consumingNon-specific reactions/false positives
Strand-displacement amplification (SDA)DNA polymerase, restriction endonuclease, primers, dCTP, dTTP, dGTP, dATPα37–49 °C>2 hFluorescenceHigh specificity. Detection of large RNA moleculesNon-specific reaction/high background signal
Rolling circle amplification (RCA)DNA ligase, DNA polymerase, primer, padlock probe30–37 °C>1.5 hFluorescenceHigh specificityFalse negatives and false positives
* Advantages and disadvantages in comparison with RT-qPCR methods.
Table 4. Reference materials and resources for SARS-CoV-2 nucleic acid testing.
Table 4. Reference materials and resources for SARS-CoV-2 nucleic acid testing.
InstituteTypeNumbers
ATCCheat-inactivated3
Synthetic RNA5
Bio-RadSynthetic RNA1
JRCSynthetic RNA1
KRISSSynthetic RNA1
Virus Like Particle1
NIBSC Heat-inactivated1
NMIAInactivated1
NIMCSynthetic RNA1
NISTSynthetic RNA1
Randox QnosticsHeat-inactivated3
SeracareVirus Like Particle4
Thermo ScientificInactivated1
Genomic RNA1
Synthetic RNA2
Twist BioscienceSynthetic RNA20
UMESynthetic RNA2
ZeptoMetrixChemical-inactivated1
ATCC; American Type Culture Collection, JRC; Joint Research Center, KRISS; Korea Research Institute of Standards and Science, NIBSC; National Institute for Biological Standards and Control, NMIA; National Measurement Institute Australia, NIMC; National Institute of Metrology of China, NIST; National Institute of Standards and Technology, UME; The National Metrology Institute of Turkey.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yoo, H.M.; Kim, I.-H.; Kim, S. Nucleic Acid Testing of SARS-CoV-2. Int. J. Mol. Sci. 2021, 22, 6150. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22116150

AMA Style

Yoo HM, Kim I-H, Kim S. Nucleic Acid Testing of SARS-CoV-2. International Journal of Molecular Sciences. 2021; 22(11):6150. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22116150

Chicago/Turabian Style

Yoo, Hee Min, Il-Hwan Kim, and Seil Kim. 2021. "Nucleic Acid Testing of SARS-CoV-2" International Journal of Molecular Sciences 22, no. 11: 6150. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22116150

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