Next Article in Journal
The Impact of Load Style Variation on Gait Recognition Based on sEMG Images Using a Convolutional Neural Network
Previous Article in Journal
Multi-Cloud Resource Management Techniques for Cyber-Physical Systems
Previous Article in Special Issue
Biosensors as Nano-Analytical Tools for COVID-19 Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Human Body Performance with COVID-19 Affectation According to Virus Specification Based on Biosensor Techniques

by
Mohammed Jawad Ahmed Alathari
1,
Yousif Al Mashhadany
2,
Mohd Hadri Hafiz Mokhtar
1,
Norhafizah Burham
1,3,
Mohd Saiful Dzulkefly Bin Zan
1,
Ahmad Ashrif A Bakar
1 and
Norhana Arsad
1,*
1
Department of Electrical, Electronic and Systems Engineering, Faculty of Engineering and Built Environment, Universiti Kebangsaan Malaysia (UKM), Bangi 43600, Malaysia
2
Department of Electrical Engineering, College of Engineering, University of Anbar, Anbar 00964, Iraq
3
School of Electrical Engineering, College of Engineering, Universiti Teknologi MARA, Shah Alam 40450, Malaysia
*
Author to whom correspondence should be addressed.
Sensors 2021, 21(24), 8362; https://doi.org/10.3390/s21248362
Submission received: 11 October 2021 / Revised: 8 December 2021 / Accepted: 9 December 2021 / Published: 15 December 2021
(This article belongs to the Special Issue Detection and Diagnosis of the New Coronavirus)

Abstract

:
Life was once normal before the first announcement of COVID-19’s first case in Wuhan, China, and what was slowly spreading became an overnight worldwide pandemic. Ever since the virus spread at the end of 2019, it has been morphing and rapidly adapting to human nature changes which cause difficult conundrums in the efforts of fighting it. Thus, researchers were steered to investigate the virus in order to contain the outbreak considering its novelty and there being no known cure. In contribution to that, this paper extensively reviewed, compared, and analyzed two main points; SARS-CoV-2 virus transmission in humans and detection methods of COVID-19 in the human body. SARS-CoV-2 human exchange transmission methods reviewed four modes of transmission which are Respiratory Transmission, Fecal–Oral Transmission, Ocular transmission, and Vertical Transmission. The latter point particularly sheds light on the latest discoveries and advancements in the aim of COVID-19 diagnosis and detection of SARS-CoV-2 virus associated with this disease in the human body. The methods in this review paper were classified into two categories which are RNA-based detection including RT-PCR, LAMP, CRISPR, and NGS and secondly, biosensors detection including, electrochemical biosensors, electronic biosensors, piezoelectric biosensors, and optical biosensors.

1. Introduction and Overview of Coronaviruses:

The Coronaviridae family to which the Coronavirus belongs is also a part of the Nidovirales order and is the subfamily of Orthocoronavirinae. The subfamily of Orthocoronavirinae consists of four types which are delta (δ) beta (β), alpha (α), and gamma (γ) coronavirus as shown in Figure 1, these viruses share qualities where they are all enveloped, contain single-stranded RNA, have positive-sense and are not segmented viruses that cause minor or critical illnesses in some breathing creatures which includes human beings. The name of Coronavirus is driven from its club-shaped spikes, protruding from the virion surface resulting in a solar corona shape [1,2,3].
Preliminary studies on Coronaviruses were reported in the 1930s. In contrast, the human type of the virus named (HCoVs) was first discovered in the 1960s. Until recently, seven types of the virus have been discovered including the HCoVs-NL63 and HCoVs-229E 229E (α-Corona viruses) and HCoVs-OC43, HCoVs-HKU1 (β-Corona viruses), severe acute respiratory syndrome-CoV (SARS-CoV), and Middle East respiratory syndrome-CoV (MERS-CoV), and SARS-CoV-2 which was discovered in 2019 [4,5,6].

1.1. Coronaviruses That Infect the Human Body Respiratory System

Coronaviruses were mostly known to infect animals of mammal species and birds; it is only recently that they have infected humans. The latest research reported seven types of the Coronavirus that can infect humans, four of which are (229E, NL63 HKU1, and OC43). These viruses are reported to cause a respiratory system infection, in which symptoms include coughing, sore throat and coryza. Other reported symptoms that are less likely to occur are pneumonia, bronchiolitis, and bronchitis, whereas three of them MERS-CoV, SARS-CoV and SARS-CoV-2 result in illnesses with fluctuating severity to the human respiratory system. The severity ranges from the common cold to incurable pneumonia [7,8,9,10].
In 2002, severe acute respiratory syndrome (SARS) was reported first in China and soon after affected other countries worldwide. The virus caused infection in human respiratory system organs that were incurable and therefore fatal in most cases [11,12,13]. The virus’s fatality range is 50% higher in seniors than younger adults, with young adults having a ~3–6% fatality percentage [14,15]. A coronavirus known as SARS-CoV was found as the etiological agent of SARS [16].
A decade later a new variation of the virus known as (MERS-CoV) Middle East respiratory syndrome started spreading in the Middle East in 2012; the first fatality reported was at the hospital in Jeddah, Kingdom of Saudi Arabia [17,18,19]. This case was the starting point for a series of infections in surrounding countries and later worldwide [20]. The virus causes lower respiratory system infection and has a 35% fatality percentage. Since the virus outbreak, 2279 cases were reported, 806 of which were fatal. After reports of an outbreak in 27 countries, the WHO listed MERS-CoV as a priority disease requiring urgent research. Notably, this virus’s nature is continuously evolving. Hence, it can efficiently fight human antiviral responses. In retrospect, there are only supportive treatments for the virus [21,22].
In December 2019, seven years after the appearance of MERS-CoV, a new case of virus infection was reported in Wuhan, China. The virus was named by the World Health Organization (WHO) the 2019 novel Coronavirus (2019-nCoV). The name was later changed to SARS-CoV-2 (severe acute respiratory syndrome coronavirus 2), based on its similarities to SARS-CoV. The WHO named the disease caused by the virus infection as COVID-19 [23]. Based on recent data, over 216 M people have reported positive for COVID-19 worldwide by August 2021 [24]. This highly dangerous pandemic has affected all aspects of human day-to-day activities and impeded the most routine activities to a stage that was not possible to predict [25]. In retrospect, international transportations were limited and quarantine and social distancing rules were enforced in most nations [26].
However, the enforced strategies are not a permanent solution as they will have unfavorable impacts on the economy, education, food system including mental wellbeing [27]. In response to the enforced strategies, many have lost their income sources, and others are threatened to lose their job at any minute. It is notable that enforcing strict rules is not controllable and governments are yet to handle this long term [28]. The transmission speed of the virus is soaring and it is a challenge to manage the issues it prevails [29]. This issue is even more crucial when associated with healthcare providers [30]. Having rapid diagnostic technologies is vital to navigating this fatal outbreak [31,32]. Table 1 compares coronaviruses that infect the human body respiratory system; the table compares coronavirus names, year of finding, emergence, type, host, cellular receptor, incubation period, respiratory system infection, symptoms, and mortality rate. The most recently discovered virus in India was the Delta Variant which is a mutation of the SARS-CoV-2 virus. The Delta Variant virus is hosted in humans. The cellular receptor of the virus is ACE2 and the incubation period is 5–6 days. The symptoms accompanied are headaches, sore throat, runny nose, cough, loss of taste and loss of smell.

1.2. SARS-CoV and SARS-CoV-2 Genomic Structure and Proteins

The genomic structure and proteins of SARS-CoV and SARS-CoV-2 viruses have similarities more than differences. Both viruses reported being spherical single-stranded RNA viruses that are known to have protein spikes that are protruding from the virion surface. This characteristic in the virus shape gave it its name, Coronavirus, driven from the Latin word corona due to its resemblance to the crown shape [38,39].
SARS-CoV-2 genome consists of 29,903 nucleotides that are significantly similar to SARS such as SARS-CoV that infects bats by 81% nucleotides resemblance. Both viruses have a unique replication approach considering their large RNA genome with none segmented positive-sense RNA genome featuring a shape of 5′cap and 3′ poly (A) tail. The unique structure of the virus allows the replicase polyproteins to understand the genome for translating 2/3 of the genome encodes nonstructural proteins (NSPs). In contrast, the remaining 1/3 encodes for structural and auxiliary proteins. There are four main structural proteins known as (S) spike, (M) membrane, (E) envelope and nucleocapsid (N) proteins. Furthermore, PP1a and PP1b are viral replicas created by the virus are later customized into 16 mature NSPs [40,41]. The pre-mentioned proteins of coronavirus and genomic structure of SARS-CoV and SARS-Cov-2 are illustrated in Figure 2.
The S protein in the virus that weighs about 180 kDa molecular units and is 20 nm long is possibly a target for inhibition of viral entrance and the growth of antibody-based therapeutics to prevent the disease. The protein comprises two subunits which are S1 and S2. The S1 receptor is for binding, which contains the receptor-binding domain (RBD), which identifies and binds to its receptor angiotensin-converting enzyme 2 (ACE2), which is found in the epithelial cells of the lungs. The S2 subunit for membrane fusion, by undertaking specific conformational modifications, induces membrane fusion, which enables viral entry into the host cell. This protein is the chief focus of neutralizing antibodies, which in response prevents infection and further dissemination by stopping binding to ACE2 [42,43].
The E protein is a small membrane protein that manages virion construction and has added effects on the affected host cells. During the host cell infection, the E protein controls the stress response in the cell which, as a result, modifies the virulence of the virus leading to caspase-mediated cell death. The M protein is the amplest in the viral envelope, this protein is essential to the virion particles creation. N protein helps in creating the virial nucleocapsid by forming complexes with the genome RNA within the viral membrane. Both M and N proteins play an important role in RNA replication [44].

2. Background of COVID-19 Virus Detection

With the emergence of COVID-19, different detection techniques were developed in the pursuit of containing the rapidly spreading virus by identifying prospective patients using image processing, artificial intelligence, biosensors and others.
Image processing was utilized for the detection of COVID-19; an Image processing technique based on interferometry that uses the phenomenon of interference of waves was developed to detect COVID-19 [45]. Research reported that the computerized tomography (CT) technique used to detect COVID-19 had high accuracy percentage [46]. Another technique based on image processing, Fast COVID-19 Detector (FCOD), was introduced with a promising detection time limit using X-ray images [47]. Infection maps generated by CXR images were used to detect COVID-19 as well as joint localization and severity grading [48]. Additionally, CXR images were employed for the COVID-19 detection via a system called (ACoS) Automatic COVID-19 screening. It uses radiomic texture descriptors by mining CXR images. The system can successfully identify healthy or infected and non-COVID-19 patients [49].
Artificial intelligence has also been widely used in the medical community and, without doubt, was used to detect COVID-19 patients; AI-based techniques were successfully used to diagnose and classify COVID-19 by studying and analyzing X-ray and CT scan images [50]. AI has been helping to prevent the spread of COVID-19 by tracking, diagnosis and social control [51]. CNN deep learning techniques are used to identify potential patients by classifying X-ray images [52]. A decision-making system based on CNN deep learning was developed to assist the radiologist in identifying suspected patients [53].
Deep learning techniques were used to automatically diagnose affected patients which helped radiologists with suspected patients’ diagnoses [54]. A COVNet detection system based on a neural network deep learning model was developed to take out ritualistic characteristics from volumetric chest CT scans for the detection of COVID-19 [55]. A deep learning medical system named (COVIDetction-Net) was developed to automatically identify and detect COVID-19 by using chest radiography images (CRIs) [56]. Popular architectures of deep learning were used to develop a Coronavirus diagnostic system. The architectures are VGG16, DenseNet121, Xception, NASNet, and EfficientNet [57]. A deep convolutional neural network-based architecture was proposed for the COVID-19 detection utilizing chest radiographs [58]. The virus was detected using cough data through artificial intelligence (AI) [59].
An automatic COVID-19 detection approach based on deep learning that applies multilayer-Spatial Convolutional Neural Network CNN was successfully proposed, which can identify affected patients by scanning chest X-ray images and CT scans [60]. Another novel CNN model named CoroDet was developed to detect COVID-19 automatically via CT-scan images and chest X-rays [61]. The deep learning CNN method was also applied to validate and classify chest X-ray images of SARS-CoV-2 suspected patients [62]. Another CNN using profound learning algorithms was proposed to efficiently classify the SARS-CoV-2 virus through CT scans and chest images [63].
Detection techniques merged AI with other approaches for COVID-19 detection; a study presented a joint approach including electrochemical biosensing of SARS-CoV-2 empowered by AI to produce bioinformatics that is needed for an early patient’s diagnosis [64]. A hybrid deep neural network (HDNNs) was developed using computed tomography (CT) and X-ray imaging, to foresee the risk of the onset of disease in patients infected with SARS-CoV-2 [65]. Similarly, a hybrid model joining both deep learning with machine learning was introduced in an effort to arrange possible affected patients’ chest images and classify them as SARS-CoV-2 positive or negative [66]. Additionally, a generative adversarial network (GAN) mixed with in-depth coronavirus detection learning using X-ray chest images was proposed [67]. Another deep learning model that uses a multileveled pipeline was proposed to scan X-ray images to identify the SARS-CoV-2 virus along with other potential chest diseases [68].
Similarly, biosensors were used to detect COVID-19 patients; the MERS-CoV detection approach was developed using an LSPR biosensor with surface-enhanced Raman scattering (SERS) multiplex [69]. Another modern approach was developed called Dual functioning SPR, this approach works by merging the photo-thermal effect in a biosensor LSPR for COVID-19 virus detection [70]. Additionally, a biosensor based on a field-effect transistor (FET) method was developed for the detection and diagnosis of the COVID-19 virus [71]. An overview of the optical biosensors used to identify the COVID-19 virus was published [72]. Researchers successfully developed a COVID-19 detection device based on an optical biosensor that uses SPR with a gold nanoparticle coating [73]. Another Optical biosensor LSPR was demonstrated for the possible detection of coronavirus disease [74]. A rapid and accurate COVID-19 detection using a point-of-care nanophotonic biosensor was developed, it can be used with both human patient samples and animal reservoirs [75]. Promising strategies were proposed for the development of CP-based electrochemical biosensors for COVID-19 detection [76]. An approach of noble metal nanomaterials and associated biosensors was proposed for detecting viruses causing human respiratory system diseases, including COVID-19, using a combination of electrochemical and optical detection techniques [77]. An electrochemical immunoassay was developed for fast and smart COVID-19 detection using samples of saliva [78].
Other strategies for COVID-19 detection were also recorded; a device was developed to detect COVID-19 in a single reaction using a multiples RT-qPCR assay [79]. Another PCR-based optimization using digital droplets was used to identify the COVID-19 virus. This approach exhibited a significantly lower limit of detection compared to RT-PCR [80]. A portable COVID-19 detector tool that is bifunctional electrical with transistors based on graphene field-effect was developed, the tool detects the virus via either acid hybridization or antigen–antibody-protein interaction [81]. An instantaneous and hypersensitive approach to detect anti- SARS-CoV-2 IgG in human serum was developed using lateral flow immunoassay (LFIA) that uses lanthanide-doped polystyrene nanoparticles (LNPs) [82]. On the other hand, a low-frequency Raman (LFR) spectroscopy was used to detect the COVID-19 virus [83].

3. Taxonomy of Literature Reviews on COVID-19 Viral Virus

This review studied and analyzed a collection of literature focused on the current viral virus known as COVID-19, which have been divided into two main categories: Transmission COVID-19 Human Exchange and detection techniques of COVID-19 virus. The first category has four subcategories which are Respiratory Transmission, Fecal–Oral Transmission, Ocular Transmission and Vertical Transmission. The second category has two subcategories for detection techniques which are: a. Based on an RNA method that includes next generation sequencing, RT-PCR, LAMP and CRISPR; b. Based on a biosensor that includes electrochemical biosensor, electronic biosensor, piezoelectric biosensor, and optical biosensor as shown below in Figure 3.
The literature was collected using several search engines which were PubMed, WOS, Scopus, Science direct, IEEE Xplore, as well as Google scholar.

4. Transmission COVID-19 Human Exchange

With the emergence of the COVID-19 virus, much research has proved that it is transmittable between humans. In order to contain the pandemic, the need to understand the virus’s various modes of transmission is crucial. This section elaborates four human-to-human transmission modes, which are respiratory transmission and fecal–oral transmission. Both are the most popular modes of transmission. Ocular transmission and vertical transmission are also possible modes of transmission; the first one is less common, while the last one is not yet confirmed.

4.1. Respiratory Transmission

Collectively the recently discovered coronaviruses which are SARS-CoV, MERS-CoV, and SARS-CoV-2 share the effect of infection in the respiratory tract [84]. Primary SARS-CoV-2 identification was carried out by studying the biofluids of the lower respiratory tract and bronchoalveolar lavage (BAL). The bio fluid was taken from the onset patients present in the Hunan wet market in Wuhan, China on 21 December 2019. All patients had the same symptoms, unknown etiology pneumonia which ranged from the common cold, high temperature, dry coughing to dyspnea. In rare cases, mostly in elderly infected patients, the prementioned symptoms evolved to severe acute respiratory syndrome (SARS), which had similarities to the famous acute respiratory distress syndrome (ARDS) [85]. COVID-19 can transmit from human to human predominantly through respiratory secretions. It can be transmitted either directly through inhalation of the infected human droplets produced from sneezing or coughing; infection transmission can occur through contact by oral, nasal, or ocular mucous membranes, or via contact with surfaces that are infected with human bodily serum. A less possible transmission is via aerosols. The exhaled air produced from the human respiratory system normally contains a high number of droplets that are extremely small. The droplets produced from infected humans will contaminate another human if inhaled or digested [86,87,88]. Previous studies have shown that besides sneezing and coughing, normal breathing and speech can also generate aerosol ranging from 0.75 to 1.1 μm, which is smaller than the aerosol generated by sneezing or coughing (~5 μm) and hence it can transmit the virus up to greater ranges [89]. Another notable transmission method is through airborne dust as it may transport virus-laden through air inhalation causing infection in deeper bronchial and alveolar regions [90]. In the early stages of the current pandemic, the possibility of surface infection with infected droplets being a cause of transmission was a subject of concern to many researchers. Nevertheless, it was proven that it is unlikely to be a method of transmission even though the virus can survive on surfaces for days. Attempts to revive the virus off the contaminated surfaces were in vain. In summary, the most likely transmission of the virus occurs from proportionally large infected droplets produced from infected humans by coughing, sneezing and breathing in close proximity to another person as reported by the infection control guideline [91]. The human exchange of the SARS-CoV-2 virus through respiratory transmission is summarized in Figure 4.

4.2. Fecal–Oral Transmission

Although the spread of COVID-19 through droplets, surface contact and aerosolized transmission has been well-publicized, the fecal–oral route is yet another identified method of transmission. According to a meta-analysis by the New England Journal of Medicine, it was found that fecal viral shedding continues throughout the disease, even after nasopharyngeal tests appear negative. Moreover, gastrointestinal symptoms seem to be common for COVID-19 patients, with a prevalence of approximately 18% [92,93]. Research reported that patients who showed symptoms in the digestive system experience significantly longer hospital stays [94]. SARS-CoV-2 and MERS-CoV share 82% genome sequence are reported to trigger respiratory and gastrointestinal (GI) symptoms [95,96]. With that in mind, patients’ stools that are infected with the virus and may be viable under circumstances that can enable fecal–oral transmission; it is possible that COVID-19 could also be transmitted via this route [96]. Fecal–oral transmission occurs through direct or indirect contact with pathogens from contaminated fecal excreta [97]. Holshue et al., (2020) reported case zero of fecal–oral transmission by detecting the patients’ stool RNA of COVID-19 in the US. It was reported that ACE2 is profusely expressed in esophageal, gastric, duodenal and rectal epithelial cells and absorptive enterocytes of ileum and colon allowing the gastrointestinal tract to be prone to COVID-19 infection and therefore suggestive of possible fecal transmission [98,99,100]. Proof of gastrointestinal infection by SARS-CoV-2 was presented by isolating the viral RNA from the epithelial cells and intracellular staining of the nucleocapsid protein [100]. Research in COVID-19 transmission on a family subject in China found two young adults in the same family were diagnosed with diarrhea, proposing that fecal–oral transmission may be a major pathway for virus transmission [36]. COVID-19 virus fecal–oral transmission is illustrated in Figure 5, infected patients pass feces with COVID-19 virus, then the virus is transmitted via unwashed hands, insects and food to other humans.

4.3. Ocular Transmission

Recent studies indicate that tears are a potential source for this infection and that the conjunctiva may maintain viral replication for a long period [101]. Another study reported the detection of the SARS-CoV-2 virus in tears of infected patients by RT-PCR [102]. Studies have shown that ACE2 and TMPRSS2 are present in the affected human body and are concentrated in the nasal secretory cells [103]. Several studies have reported high concentrations of ACE2 and TMPRSS2 in the eyes. Immunohistochemical analysis revealed its expression in the conjunctiva, limbus, and cornea [104,105]. With the onset of the SARS-CoV-2 outbreak, conjunctivitis was a rare symptom of the plague. Several patients reported conjunctivitis and other related ocular diseases before the onset of common symptoms of the virus, such as high temperature and dry cough [106].
The human eye could be the entry portal for the virus to enter the body, the eye’s cornea, and conjunctiva form the ocular mucosa surface, which is prone to infection in close proximity with infected people or contaminated hands. Considering the human body anatomy where the eyes mucus is linked to the respiratory system through the nasolacrimal duct, the eyes could be the entry point of COVID-19 [107]. A Chinese expert became infected with COVID-19 and reported the first case of conjunctivitis suggesting a route of conjunctival infection and tear transmission [108].

4.4. Vertical Transmission

SARS-CoV-2, resulting in COVID-19 disease, emerged as a third coronavirus epidemic at the end of 2019. Despite the fact that scientific evidence suggests that women are less likely than men to become infected with COVID-19 [109]. Pregnant females were proven to have higher chances of infection with COVID-19 compared to non-pregnant women, which is credited to immunological and anatomical changes. Thus, increasing the risk of developing severe illness [110]. In the case of pregnant females that are infected with the virus, the infection increases morbidity and poses a potential threat to the fetus development due to the spread of COVID-19 receptors known as ACE2 and TMPRSS-2 in fetal organs along with the cells of the maternal–fetal interface [111,112]. In addition, a COVID-19 positive neonate was found in China, the test was confirmed using RT-PCR to test a sample of pharyngeal swabs thirty-six hours after birth; nevertheless, vertical transmission, in this case, was not confirmed [113]. Other risks faced by infected pregnant females are premature birth, intrauterine growth restriction and spontaneous abortion. Yet, there is no strong evidence of COVID-19 virus transmission. Therefore it appears that the threats posed on the fetus are caused by the direct virus effect on the mother. While the current confirmation of vertical transmission during pregnancy is limited, the possible risk of vertical transmission should not be overlooked yet [114,115,116].

5. Comparison for Transmission COVID-19 Human Exchange

Transmissions through the human exchange are Respiratory, Fecal–Oral, Ocular and Vertical. Each transmission is compared in Table 2 by probability, possibility, confirmed cases, virus entry organ, transmissibility approach and gender.
Both Respiratory and Fecal–Oral Transmission probability are more likely to happen, they also have a higher possibility of transmission. The entry organs are mouth and nose in Respiratory Transmission via direct and indirect transmissibility approach, while in Fecal–Oral Transmission it is the mouth via indirect transmissibility approach. Both transmissions as well as Ocular transmission can happen for both genders men and women and both have confirmed cases as well as Ocular transmission.
Ocular and Vertical transmission is less likely to happen while Vertical Transmission has the rarest possibility and is yet to be confirmed. The virus entry organ in the ocular transmission is the Eyes via direct and indirect contact whereas in Vertical transmission it is the uterus via direct contact and it can only affect pregnant females.

6. Detection Techniques of COVID-19 Virus

The recent pandemic of COVID-19 has drawn attention to the high importance of rapid and accurate diagnostic assays. In parallel to the outbreak, researchers from different areas worldwide have worked together for such assays, concurrently to other many assays that are approved along with the ones that are yet clinically validated. Various techniques of Coronavirus detection have been studied to detect potential patients, in this paper we elaborate on two main methods which are Ribonucleic Acid (RNA)-based techniques and Biosensor techniques.

6.1. Based on Ribonucleic Acid (RNA) Method

Molecular diagnostic assays are the leading group of tests used to diagnose COVID-19, mainly using RNA to detect the SARS-CoV-2 virus in patients. The latest detection techniques of COVID-19 RNA, including those still in the development stage, are CRISPR and next-generation sequencing (NGS). Another detection method which is the current golden standard technique is reverse transcription-polymerase chain reaction (RT-PCR) as well as the second most used method which is known as loop-mediated isothermal amplification (LAMP). This section thoroughly explains the previously mentioned techniques and their applications in the efforts of COVID-19 detection.

6.1.1. Next-Generation Sequencing (NGS)

Technologies along with bioinformatics played a significant role in changing the conventional research on viral pathogens and attracting attention to the field of virus diagnosis [117]. Recent research and developments in next-generation sequencing (NGS), alternatively named high-throughput sequencing (HTS) offered to the science field countless applications, and the current outbreak is urging these applications to be wildly used. An advantage of applying NGS for the detection of infectious diseases in the clinical diagnosis is that it is neither culturing nor clinical hypothesis dependent. NGS-based testing discloses all kinds of existing microorganisms in the sample such as fungi, bacteria, parasites and of course viruses, unlike conventional testing methods that require clinicians to do the extra work of addressing patients’ symptoms with possible explanations and requests for testing those particular pathogens [118].
NGS is based on massively parallel sequencing, meaning that billions of short DNA fragments are sequenced simultaneously producing short sequence “reads” rendering dramatically reduced time and cost of sequencing [119]. This technology is reported to be one of the best applications in the current outbreak. Since the emergence of the virus, false-negative results were reported for patients admitted with acute respiratory distress syndrome. It was also reported that this technology was the only one that discovered the etiological pathogen by applying metagenomic RNA sequencing and analyzing the phylogenetic of the complete generated genome allowing us to conclude that the founded new strain of RNA belonged to the Coronaviridae family and was later specified as COVID-19 after nucleotide similarity and genome matching with the existing pathogen’s genome [117]. As a result, NGS technology shows promising results in the efforts of SARS-CoV-2 virus detection.

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

RT-PCR is the most reliable and gold standard method for the identification of COVID-19 infection [120]. because of its advantages as a precise, and simple quantitative assay [121]. Moreover, real-time RT-PCR showed higher sensitivity compared to RT-PCR assay, which helped greatly in early infection diagnosis [122]. Based on laboratory testing, using this method requires collecting samples for RNA extraction followed by reverse transcription from suspected patients’ upper respiratory tract fluids such as nasal aspirate, nasopharyngeal swab, or pharyngeal swab or the lower respiratory tract (sputum, tracheal aspirate) [123]. After completing the reverse transcription, the cDNA regions are amplified to sufficient levels for pathogen presence detection. This process depends on DNA primer–probe sets complementary to specific regions of the SARS-CoV-2 cDNA, as well as scientists worldwide, who are competing to create these sets since the first SARS-CoV-2 genome was publicly shared. Tests were constructed by finding a number of SARS-CoV-2 particular regions, such as (N/S/E) genes, ORF1ab, and RNA-dependent RNA polymerase (RdRp) [118].
There are two types of RT-PCR testing; the first type uses a single tube containing the required primers to execute the whole process of RT-PCR reaction, the other type is known as two-step, whereby it requires more than one tube to complete the test by doing separate reverse transcription and amplification reactions while offering higher flexibility and accuracy than the other test type. It demands fewer starting materials as well as allowing multiple targets quantification using cDNA stocks. Nevertheless, the first type is more appealing for its quick setup and limited sample handling [124].
Accuracy and specificity are the reason why RT-PCR is widely used despite its expensive price, time-consuming test duration as well as the need for experienced staff [125].

6.1.3. Loop-Mediated Isothermal Amplification (LAMP)

The second in popularity for COVID-19 detection after PCR testing is LAMP testing, PCR testing requires two primers (one in the front and one in the back section), which presents a challenge for LAMP testing that occurs in designing primers and their conflict prevention. Whereas, the LAMP technique needs four or six primers that divide the target sequence into three regions in the forward section and the other three in the back section making a total of six regions. LAMP process is favored because it is executed in an isothermal condition which is an affordable and rapid method. Furthermore, numerous researches reported the application of the LAMP method for coronaviruses detection [126,127,128].
LAMP works on molecular amplification techniques, by rapidly amplifying the genomic material using high efficiency. This technique primarily uses targeted DNA that is synthesized at a stable temperature of 60–65 °C achieved by using explicitly modeled primer sand enzyme (DNA polymerase) that uses strand displacement activity rather than heat denaturation which is used in PCR techniques and within 60 min or less the targeted sequence is amplified to more than 109 copies resulting in the shape of cauliflower including a stem and a loop form of DNA with many inverted repetitions [129].
Researchers combined the techniques of reverse transcription and LAMP (RT-LAMP) and produced a detection method that is characterized by being one-step with the high-throughput method of detection for finding the RNA of SARS-CoV-2 with 100 copies/reaction of an RNA virus. By applying this method the results will be given in 30 min which makes it suitable for POC test and screening application for its simplicity and rapidness compared to RT-qPCR and its needlessness for complicated equipment [130].

6.1.4. Clusters of Regularly Interspaced Short Palindromic Repeats (CRISPR)

Clusters of regularly interspaced short palindromic repeats (CRISPR) possess a frequent series of nucleotides and tiny-scaled spacer sequences. CRISPR-associated proteins known as CAS acts as the nuclease enzymes. Both are used as a bacterial defense system to protect from unknown invaders and are extensively used in RNA modification, therapy using gene alteration, and viral genome detection. Over the last few years, CRISPR has been widely utilized in the field of Vitro diagnostics due to its allele accuracy which is significant in its successful implementation and delivering high accuracy detection and treatment [131].
Several types of research were conducted using CRISPER-based detection systems for SARS-CoV-2 virus detection. For example, a study reported the successful development of a CRISPR-Cas13a-based mobile phone assay that is non-amplified to detect SARS-CoV-2 by testing extracts of RNA taken from nasal swaps. The assay attained ~100 copies/µL sensitivity in less than an hour [132]. Additionally, researches showed detection of SARS-CoV-2 using CRISPR approaches in optimal conditions provides rapid and efficient results. Hence, CRISPR-based approaches are a promising, robust and precise method for SARS-CoV-2 diagnosis [133].

6.1.5. RNA Corona Virus Detection Methods Analysis Based on RT-PCR, LAMP, and CRISPR

This paper studied the detection of Coronavirus using RNA via RT-PCR, LAMP, NGS, and CRISPR Table 3 compares the four techniques by analyzing eleven reported studies. LAMP technique reported fastest detection time of 20–25 min using test sample of Synthetic RNA solution followed by rRT-PCR which takes 30 min using test sample of Plasmids containing the complete N gene. RT-PCR, LAMP, RT-LAMP, and rRT-PCR collectively use Polymerase, (N) Gene, (E) Gene, and ORF1b as target genes, whereas, whereas CRISPR uses (N) and (E) Gene. On the other hand, NGS uses (S) Gene. NGS Detection method has not had much information due to its recent discovery and not many tests were conducted using it.

6.2. Based on Biosensor Techniques

Nowadays, researchers are focusing on developing detection tools using biosensors techniques due to their sensitivity, mobility and miniaturization. This section explains four widely used biosensors techniques which are electrochemical biosensors, electronic biosensors, piezoelectric biosensors and optical biosensors.

6.2.1. Electrochemical Biosensors

Electrochemical biosensors are biological concentrations of information that are converted into an analytically relevant signal by using a current or voltage [143]. These biosensing devices can read biochemical information to detect biological materials such as protein and nucleic acid. It also holds qualities such as simple instrumentation, high sensitivity, economic, and the capacity for miniaturization [144,145]. Applications using this type of biosensing device are utilized widely in many areas, whereby it represents a standardized platform for constructing biosensors that include semiconductors and screen-printed electrodes [146]. In brief, these biosensors observe the dielectric changes in properties depicted in dimension, shape, and charge distribution. At the same time, the antibody–antigen complex is formed on the electrode surface, which is categorized into four main groups: potentiometric, amperometric, cyclic voltammetry, and impedimetric transducers [147].
In a recent study focused on SARS-CoV-2 detection, researchers proposed the detection of COVID-19 N-gene using an antifouling electrochemical biosensor. The biosensor is assembled based on electropolymerized polyaniline (PANI) nanowires combined with lately designed peptides. The biosensor detects the N-gene by using biotin-labeled probes that are immobilized onto peptide-coated PANI nanowires, creating an electrochemical interface that is antifouling and susceptible. The recorded detection limit of this biosensor was shallow at (3.5 fm) [148].

6.2.2. Electronic Biosensors

A biosensing device based on field-effect transistors (FETs) consists of a three-electrode structure containing the source, drain, and gate and was developed for the detection of small molecules and the diagnosis of viral diseases. The FET is an electric biosensor that detects changes in surface potential after the target molecule binds to the biorecognition element immobilized on the highly conductive chip surfac [149]. For instance, graphene, zinc oxide, gallium nitride, disulfide, and molybdenum are used in FET-based biosensors where heterogeneous analyte concentrations can be detected rapidly via probes fixed on conducting channels [150]. This type of biosensor is recently used extensively in creating assays for SARS-CoV-2 detection using spike membrane protein] [151]. An example of this method is a graphene FET-based biosensor that can detect COVID-19 related viruses through its spike proteins in only 120 s, also in another way by employing spike protein-specific antibodies, with a 0.2 pM detection limit for the assay [152].
An electronic biosensor device based on FET was developed for the detection of COVID-19 related viruses in clinical samples. The sensor was prepared by covering graphene sheets used in the FET with a unique antibody to the spike protein of SARS-CoV-2. The three determinants of the device’s performance are using a cultured virus, antigen protein, and nasopharyngeal swab specimens from infected patients. The resulting device is a super-sensitive immunological diagnostic method for COVID-19 with minimum sample preparations [71].

6.2.3. Piezoelectric Biosensors

Piezoelectric biosensors are a collection of analytical devices operated based on recording affinity interaction. A piezoelectric platform or piezoelectric crystal is known as a sensor part that works on the basis of oscillations change that results from the mass tied on the piezoelectric crystal surface [153].
Piezoelectric biosensors technology is commonly used to detect hormones, cells, bacteria, viruses, and to study a wide range of interactions on a biomolecular level. This tech offers immediate and unlabelled transduction with high susceptibility, simplicity, and velocity [154,155]. In diagnosing SARS, the piezoelectric biosensor was utilized to measure a type of coronavirus using a sputum sample. To get the antibodies to stick to the piezoelectric crystals, the experiment bound the SARS-CoV horse polyclonal antibody from protein A to the surface. Changes in mass from the crystal, due to viral binding, recorded a shift in frequency [156]. COVID-19 samples are being directly detected by the piezoelectric microcantilever biosensor, which was built without requiring further processing. The biosensor functions as a transducer, and it is coated with an antibody that is relevant to the substance being detected. Because of the mass change caused by the SARS-CoV-2 antigens’ spike proteins, the microcantilever’s surface would experience surface stress and exhibit a quantifiable tip deflection and floating voltage [157]. Compared to other biosensors, piezoelectric biosensors seem to display an enhanced level of performance [158,159]. There is still more study to be carried out before the technology can be applied; while the piezoelectric sensors can detect viral frequency changes, they can also detect them using output voltage directly. On top of that, piezoelectric energy-harvesting devices are anticipated to be used in the IoT, where it is possible to detect viruses by monitoring mechanical vibration [160].

6.2.4. Optical Biosensors

High sensitivity and selectivity are benefits of optical biosensors. They can provide precise detection based on a variety of signals, including absorption, refraction, reflection, dispersion, infrared, polarisation, chemiluminescence, fluorescence, phosphorescence, and so on [161]. In the industry and publications, there are various types of optical biosensors, including fiber-optic biosensors, such as the optrode biosensor and the evanescent wave biosensor, as well as time-resolved fluorescence, the resonant mirror biosensor, interferometric biosensors, and surface plasmon resonance biosensors. They are capable of detecting various biomolecules on both medical and biological specimens, with an impressive window of use [162,163].
Four types of optical biosensors (fluorescence, surface plasmon resonance, localized surface plasmon coupled fluorescence (LSPCF), and fiber optic) will be thoroughly described in this section.
Starting with the up to date known largest group of sensors Fluorescence-based optical biosensors, its popularity is credited to the accessibility of countless fluorescent probes, high quality and fitting optical instruments [164]. This biosensor is characterized by having a variety of intensity, lifetime, energy, transfer and quantum yield that offers opportunities for further exploration [165].
A detection stripe assay based on fluorescent immunochromatographic was built to detect N proteins in a duration of 10 min using samples of nasopharyngeal aspirate and urine with a 68–98% sensitivity rate [166]. Another rapid and quantitative approach of anti-COVID-19 IgG antibody based on the fluorescence biosensor optofluidic POC testing was built. It is an easily handled portable system that is suited for instantaneous results for detecting anti-COVID-19 IgG antibodies in samples. Given perfect conditions, the testing duration can be brought down to 25 min with a detection limit of 12.5 ng/mL that surely meets the diagnostic requirements [167].
Another auspicious platform for pathogens, as well as COVID-19 N gene detection biosensors, is SPR-based biosensors, owing to their characteristic of label-free real-time sensing [151]. An additional label-free SPR-based pioneering aptasensore was developed for COVID-19 N gene detection through thiol-modified niobium carbide MXene quantum dots (Nb2C SH QDs). The resulting tool had a low limit of detection (LOD) of 4.9 pg·mL−1. Thus, it exhibited exceptional selectivity in the existence of different respiratory viruses in human bio serum [168].
In addition, the localized surface plasm coupling fluorescence (LSPCF) fiberoptic biosensor was developed to combine a sandwich immunoassay with the SARS coronavirus LSP detection technique. The SARS-CoV N protein can detect very low concentrations (~1 pg/mL) in serum from the LSPCF fiber-optic biosensor [169].
Biosensors that use fiber optics are another type of optical biosensor. A device made from fibers is used in the field of optical science to measure biological species (cells, proteins, DNA, and so on) [161]. To detect the COVID-19 N protein at the point of care, researchers created an easy-to-use plasmonic fiber optic absorbance biosensor (PFAB) that does not need washing. To perform the P-FAB test, the fiber-optic sensor probe is U-bent, and has high EWA (Evanescent Wave Absorbance) sensibility. The COVID-19 N-protein is measured as the light propagates by the U-bent sensor probe linked with the green LED and the photodetector. The P-FAB approach resulted in the lodging of ~2.5 ng/mL or less within 10 min of reading time by using a GOF fused silica/glass optical fiber (GOF) with citrate-capped AuNP labels (size ~40 nm) [170].

7. Coronavirus Detection Methods Analysis Based on Biosensor Usage

COVID-19 detection methods based on biosensor usage were thoroughly analyzed by comparing 11 research reported studies from the year 2004 up till now, including the detection of SARS-CoV, MERS-CoV and COVID-19. Table 4 analyzes and compares the development of four biosensor-based techniques with the aim of virus detection. Different materials were used in the biosensors such as PANI (Electrochemical), Gold (Electrochemical), Gold nanoparticles (Optical (P-FAB)), Graphene (Electrical (FET)), Crystal with quartz wafer (Piezoelectric), Nb2C-SH QD (Optical (SPR)) and Polymethyl methacrylate (Optical (LSPCF)). Detection of COVID-19 using electrochemical biosensor reported the lowest temperature of 4 °C and fastest detection time of 10–30 s using test samples of spiked saliva.

8. COVID-19 Detection Techniques Advantages and Limitations

The detection techniques that were extensively explained previously are like any other techniques have their advantage and limitations. RNA-based detection techniques (RT-PCR, LAMP, NGS, and CRISPR) advantages include specificity and sensitivity. In comparison, limitations include the need for extracting RNA from clinical samples and the inability to detect already recovered patients. On the other hand, biosensors advantages include affordability for both electrochemical and electronic sensors, whereas for piezoelectric and optical advantages include rapidness, specificity and sensitivity. The common limitation for electrochemical and electronic is that both are time-consuming. Below, Table 5 illustrates both aspects of each detection technique.

9. Conclusions

The COVID-19 pandemic is ongoing globally and is yet to be contained; simultaneously, research in enhancing methods of preventing the spread is persistent. This review article compiled and studied research concerning the recently discovered COVID-19 virus that is rapidly morphing and highly transmittable between humans. This article addressed the pre-mentioned issue by comparing and analyzing SARS-CoV-2 transmission methods between humans and detection methods utilized to detect COVID-19. Human-to-human exchange transmission through the respiratory system proved to be the most efficient method of transmission, followed by fecal–oral transmission. Rare cases were recorded for the ocular transmission method, whereas vertical transmission has not yet been confirmed. In synchronization with the virus’s rapid spread, advancements in the detection techniques with effective diagnoses were urgently needed by clinicians. This paper gives a comprehensive review of COVID-19 virus detection techniques based on RNA and biosensors. RNA-based methods included the golden standard testing set by WHO known as RT-PCR and RT-LAMP. The latter is deemed a possible proper substitute due to it being more economical and rapid compared to RT-PCR. Last but not least, the paper reviewed the latest advancements in detection techniques using biosensors for SARS-CoV-2 detection, by using target genes in clinical samples which are the Spike (S) protein, Nucleocapsid (N) protein as well as Antigen sputum, and IgM antibody. Biosensors are proven to be rapid, sensitive, precise, mobile, economical, and have the ability for miniaturization. Therefore, COVID-19 detection using biosensors is innovative and promising.

Author Contributions

M.J.A.A., wrote the initial manuscript draft, Y.A.M., conceptualization and methodology, M.H.H.M.; M.S.D.B.Z.; N.B. and A.A.A.B. review and editing, N.A., supervision and validation. All authors have read and agreed to the published version of the manuscript.

Funding

Universiti Kebangsaan Malaysia (UKM) funded this study through the grant code GP-2019-K013436, GUP-2019-010, dan Dana Pecutan Penerbitan UKM.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

This manuscript has not been previously released and is not now under consideration by any journal for publication.

Acknowledgments

This paper’s authors highly appreciate support from the Department of Electrical, Electronic and Systems Engineering/Faculty of Engineering and Built Environment/Universiti Kebangsaan Malaysia (UKM).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AbbreviationDescription
AuNPsGold nanoparticles
COVID-19Coronavirus disease
DNADeoxyribonucleic acid
EWAEvanescent Wave Absorbance
FETField-effect transistor
GOFGlass optical fiber
LAMPLoop-mediated isothermal amplification
LSPCFLocalized surface plasmon coupled fluorescence
MERS-CoVMiddle East Respiratory Syndrome
NGSNext-generation sequencing
PFABPlasmonic fiber-optic absorbance biosensor
RNARibonucleic Acid
CRISPRClusters of regularly interspaced short palindromic repeats
RT-PCRReverse Transcription-Polymerase Chain Reaction
SARS-CoVCoronavirus Severe Acute Respiratory Syndrome
SARS-CoV-2Coronavirus Severe Acute Respiratory Syndrome 2
SPRSurface plasmon resonance

References

  1. Harapan, H.; Itoh, N.; Yufika, A.; Winardi, W.; Keam, S.; Te, H.; Megawati, D.; Hayati, Z.; Wagner, A.L.; Mudatsir, M. Coronavirus disease 2019 (COVID-19): A literature review. J. Infect. Public Health 2020, 13, 667–673. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, Y.; Qianyun, L.; Deyin, G. Emerging coronaviruses: Genome structure, replication, and pathogenesis. J. Med. Virol. 2020, 92, 418–423. [Google Scholar] [CrossRef] [PubMed]
  3. Shereen, M.A.; Khan, S.; Kazmi, A.; Bashir, N.; Siddique, R. COVID-19 infection: Origin, transmission, and characteristics of human coronaviruses. J. Adv. Res. 2020, 24, 91. [Google Scholar] [PubMed]
  4. Estola, T. Coronaviruses, a new group of animal RNA viruses. Avian Dis. 1970, 14, 330–336. [Google Scholar] [CrossRef]
  5. Wu, D.; Tiantian, W.; Qun, L.; Zhicong, Y. The SARS-CoV-2 outbreak: What we know. Int. J. Infect. Dis. 2020, 94, 44–48. [Google Scholar]
  6. Bchetnia, M.; Girard, C.; Duchaine, C.; Laprise, C. The outbreak of the novel severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2): A review of the current global status. J. Infect. Public Health 2020, 13, 1601–1610. [Google Scholar] [CrossRef]
  7. Weiss, S.R.; Navas-Martin, S. Coronavirus pathogenesis and the emerging pathogen severe acute respiratory syndrome coronavirus. Microbiol. Mol. Biol. Rev. 2005, 69, 635–664. [Google Scholar] [CrossRef] [Green Version]
  8. Wevers, B.A.; van der Hoek, L. Recently discovered human coronaviruses. Clin. Lab. Med. 2009, 29, 715–724. [Google Scholar] [CrossRef]
  9. Lau, S.K.; Woo, P.C.; Li, K.S.; Huang, Y.; Tsoi, H.W.; Wong, B.H.; Wong, S.Y.W.; Leung, S.-Y.; Chan, K.-H.; Yuen, K.-Y. Severe acute respiratory syndrome coronavirus-like virus in Chinese horseshoe bats. Proc. Natl. Acad. Sci. USA 2005, 102, 14040–14045. [Google Scholar] [CrossRef] [Green Version]
  10. Antiochia, R. Nanobiosensors as new diagnostic tools for SARS, MERS and COVID-19: From past to perspectives. Microchim. Acta 2020, 187, 1–13. [Google Scholar] [CrossRef]
  11. Tsang, T.; Lai-Yin, T.; Pak-Yin, L.; Lee, M. Update: Outbreak of severe acute respiratory syndrome-worldwide, 2003. MMWR Morb. Mortal. Wkly. Rep. 2003, 52, 241. [Google Scholar]
  12. 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]
  13. Lee, N.; Hui, D.; Wu, A.; Chan, P.; Cameron, P.; Joynt, G.M.; Ahuja, A.; Yung, M.Y.; Leung, C.; To, K.; et al. A major outbreak of severe acute respiratory syndrome in Hong Kong. N. Engl. J. Med. 2003, 348, 1986–1994. [Google Scholar] [CrossRef]
  14. 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]
  15. Holmes, K.V. SARS coronavirus: A new challenge for prevention and therapy. J. Clin. Investig. 2003, 111, 1605–1609. [Google Scholar] [CrossRef]
  16. Drosten, C.; Günther, S.; Preiser, W.; Van Der Werf, S.; Brodt, H.R.; Becker, S.; Holger, R.; 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]
  17. Mohd, H.A.; Memish, Z.A.; Alfaraj, S.H.; McClish, D.; Altuwaijri, T.; Alanazi, M.S.; Aloqiel, S.A.; Alenzi, A.M.; Bafaqeeh, F.; Mohammed, A.M.; et al. Predictors of MERS-CoV infection: A large case control study of patients presenting with ILI at a MERS-CoV referral hospital in Saudi Arabia. Travel Med. Infect. Dis. 2016, 14, 464–470. [Google Scholar] [CrossRef]
  18. Eifan, S.A.; Nour, I.; Hanif, A.; Zamzam, A.M.; Al Johani, S.M. A pandemic risk assessment of Middle East respiratory syndrome coronavirus (MERS-CoV) in Saudi Arabia. Saudi J. Biol. Sci. 2017, 24, 1631–1638. [Google Scholar] [CrossRef]
  19. Azhar, E.I.; Hui, D.S.; Memish, Z.A.; Drosten, C.; Zumla, A. The middle east respiratory syndrome (MERS). Infect. Dis. Clin. 2019, 33, 891–905. [Google Scholar] [CrossRef]
  20. Alqahtani, A.S.; Rashid, H.; Basyouni, M.H.; Alhawassi, T.M.; BinDhim, N.F. Public response to MERS-CoV in the Middle East: iPhone survey in six countries. J. Infect. Public Health 2017, 10, 534–540. [Google Scholar] [CrossRef]
  21. Yang, Y.; Zhang, L.; Geng, H.; Deng, Y.; Huang, B.; Guo, Y.; Zhao, Z.; Tan, W. The structural and accessory proteins M, ORF 4a, ORF 4b, and ORF 5 of Middle East respiratory syndrome coronavirus (MERS-CoV) are potent interferon antagonists. Protein Cell 2013, 4, 951–961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Siu, K.L.; Yeung, M.L.; Kok, K.H.; Yuen, K.S.; Kew, C.; Lui, P.Y.; Chan, C.P.; Tse, H.; Woo, P.C.; Yuen, K.Y. Middle east respiratory syndrome coronavirus 4a protein is a double-stranded RNA-binding protein that suppresses PACT-induced activation of RIG-I and MDA5 in the innate antiviral response. J. Virol. 2014, 88, 4866–4876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Chang, L.; Yan, Y.; Wang, L. Coronavirus disease 2019: Coronaviruses and blood safety. Transfus. Med. Rev. 2020, 34, 75–80. [Google Scholar] [CrossRef] [PubMed]
  24. Worldometer. 2021. Available online: https://www.worldometers.info/coronavirus/ (accessed on 29 August 2021).
  25. Haleem, A.; Javaid, M.; Vaishya, R. Effects of COVID-19 pandemic in daily life. Curr. Med. Res. Pract. 2020, 10, 78. [Google Scholar] [CrossRef]
  26. Cheval, S.; Adamescu, C.M.; Georgiadis, T.; Herrnegger, M.; Piticar, A.; Legates, D.R. Observed and Potential Impacts of the COVID-19 Pandemic on the Environment. Int. J. Environ. Res. Public Health 2020, 17, 4140. [Google Scholar] [CrossRef]
  27. Li, H.O.-Y.; Huynh, D. Long-term social distancing during COVID-19: A social isolation crisis among seniors? CMAJ 2020, 192, E588. [Google Scholar] [CrossRef]
  28. Nicola, M.; Alsafi, Z.; Sohrabi, C.; Kerwan, A.; Al-Jabir, A.; Iosifidis, C.; Agha, M.; Agha, R. The socio-economic implications of the coronavirus pandemic (COVID-19): A review. Int. J. Surg. 2020, 78, 185–193. [Google Scholar] [CrossRef]
  29. Perez, G.I.P.; Abadi, A.T.B. Ongoing challenges faced in the global control of COVID-19 pandemic. Arch. Med. Res. 2020, 51, 574–576. [Google Scholar] [CrossRef]
  30. Ehrlich, H.; McKenney, M.; Elkbuli, A. Protecting our healthcare workers during the COVID-19 pandemic. Am. J. Emerg. Med. 2020, 38, 1527–1528. [Google Scholar] [CrossRef]
  31. Zhai, P.; Ding, Y.; Wu, X.; Long, J.; Zhong, Y.; Li, Y. The epidemiology, diagnosis and treatment of COVID-19. Int. J. Antimicrob. Agents. 2020, 55, 105955. [Google Scholar] [CrossRef]
  32. Taha, B.A.; Al Mashhadany, Y.; Bachok, N.N.; Bakar, A.A.; Hafiz Mokhtar, M.H.; Dzulkefly Bin Zan, M.S.; Arsad, N. Detection of COVID-19 Virus on Surfaces Using Photonics: Challenges and Perspectives. Diagnostics. 2021, 11, 1119. [Google Scholar] [CrossRef]
  33. Mohapatra, R.K.; Pintilie, L.; Kandi, V.; Sarangi, A.K.; Das, D.; Sahu, R.; Perekhoda, L. The recent challenges of highly contagious COVID-19, causing respiratory infections: Symptoms, diagnosis, transmission, possible vaccines, animal models, and immunotherapy. Chem. Biol. Drug Design. 2020, 96, 1187–1208. [Google Scholar] [CrossRef]
  34. Abdul-Rasool, S.; Fielding, B.C. Understanding human coronavirus HCoV-NL63. Open Virol. J. 2010, 4, 76. [Google Scholar] [CrossRef]
  35. Liu, D.X.; Liang, J.Q.; Fung, T.S. Human Coronavirus-229E,-OC43,-NL63, and-HKU1 (Coronaviridae). Encycl. Virol. 2021, 428. [Google Scholar]
  36. Chan, J.F.W.; Yuan, S.; Kok, K.H.; To, K.K.W.; Chu, H.; Yang, J.; Xing, F.; Liu, J.; Yip, C.C.; Poon, R.W.; et al. A familial cluster of pneumonia associated with the 2019 novel coronavirus indicating person-to-person transmission: A study of a family cluster. Lancet 2020, 395, 514–523. [Google Scholar] [CrossRef] [Green Version]
  37. Alexandar, S.; Ravisankar, M.; Kumar, R.S.; Jakkan, K. A comprehensive review on Covid-19 Delta variant. Int. J. Pharmacol. Clin. Res. (IJPCR) 2021, 5, 83–85. [Google Scholar]
  38. Bárcena, M.; Oostergetel, G.T.; Bartelink, W.; Faas, F.G.; Verkleij, A.; Rottier, P.J. Koster, A.J.; Bosch, B.J. Cryo-electron tomography of mouse hepatitis virus: Insights into the structure of the coronavirion. Proc. Natl. Acad. Sci. USA 2009, 106, 582–587. [Google Scholar] [CrossRef] [Green Version]
  39. Neuman, B.W.; Adair, B.D.; Yoshioka, C.; Quispe, J.D.; Orca, G.; Kuhn, P.; Milligan, R.A.; Yeager, M.; Buchmeier, M.J. Supramolecular architecture of severe acute respiratory syndrome coronavirus revealed by electron cryomicroscopy. J. Virol. 2006, 80, 7918–7928. [Google Scholar] [CrossRef] [Green Version]
  40. Rossi, J.J.; Rossi, D. Oligonucleotides and the COVID-19 pandemic: A perspective. Nucleic Acid Ther. 2020, 30, 129–132. [Google Scholar] [CrossRef] [Green Version]
  41. Zhong, Y.; Tan, Y.W.; Liu, D.X. Recent progress in studies of arterivirus-and coronavirus-host interactions. Viruses 2012, 4, 980–1010. [Google Scholar]
  42. Sun, C.; Chen, L.; Yang, J.; Luo, C.; Zhang, Y.; Li, J.; Yang, J.; Zhang, J.; Xie, L. SARS-CoV-2 and SARS-CoV spike-RBD structure and receptor binding comparison and potential implications on neutralizing antibody and vaccine development. Biorxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  43. Jiang, S.; Hillyer, C.; Du, L. Neutralizing antibodies against SARS-CoV-2 and other human coronaviruses. Trends Immunol. 2020, 41, 355–359. [Google Scholar] [CrossRef]
  44. Zhang, F.; Walters, M.D. Pathogen Genomics and Host Cellular Susceptibility Factors of COVID-19. Glob. Clin. Transl. Res. 2020, 2, 107–126. [Google Scholar] [CrossRef]
  45. Hamed, A.M. Image processing of corona virus using interferometry. Opt. Photonics J. 2016, 6, 75. [Google Scholar] [CrossRef] [Green Version]
  46. Xu, B.; Xing, Y.; Peng, J.; Zheng, Z.; Tang, W.; Sun, Y.; Xu, C.; Peng, F. Chest CT for detecting COVID-19: A systematic review and meta-analysis of diagnostic accuracy. Eur. Radiol. 2020, 30, 5720–5727. [Google Scholar] [CrossRef]
  47. Panahi, A.H.; Rafiei, A.; Rezaee, A. FCOD: Fast COVID-19 Detector based on deep learning techniques. Inform. Med. Unlocked 2021, 22, 100506. [Google Scholar] [CrossRef]
  48. Degerli, A.; Ahishali, M.; Yamac, M.; Kiranyaz, S.; Chowdhury, M.E.H.; Hameed, K.; Hamid, T.; Mazhar, R.; Gabbouj, M. COVID-19 infection map generation and detection from chest X-ray images. Health Inf. Sci. Syst. 2021, 9, 1–6. [Google Scholar] [CrossRef]
  49. Chandra, T.B.; Verma, K.; Singh, B.K.; Jain, D.; Netam, S.S. Coronavirus disease (COVID-19) detection in chest X-ray images using majority voting based classifier ensemble. Expert Syst. Appl. 2021, 165, 113909. [Google Scholar] [CrossRef]
  50. Shi, F.; Wang, J.; Shi, J.; Wu, Z.; Wang, Q.; Tang, Z.; He, K.; Shi, Y.; Shen, D. Review of artificial intelligence techniques in imaging data acquisition, segmentation, and diagnosis for COVID-19. IEEE Rev. Biomed. Eng. 2020, 14, 4–15. [Google Scholar] [CrossRef] [Green Version]
  51. Naudé, W. Artificial Intelligence against COVID-19: An Early Review; Institute of Labor Economics (IZA): Bonn, Germany, 2020. [Google Scholar]
  52. Sethy, P.K.; Behera, S.K. Detection of coronavirus disease (COVID-19) based on deep features. Int. J. Math. Eng. Manag. Sci. 2020. [Google Scholar] [CrossRef]
  53. Majeed, T.; Rashid, R.; Ali, D.; Asaad, A. Problems of deploying cnn transfer learning to detect covid-19 from chest x-rays. MedRxiv 2020. [Google Scholar] [CrossRef]
  54. Hemdan, E.E.D.; Shouman, M.A.; Karar, M.E. Covidx-net: A framework of deep learning classifiers to diagnose covid-19 in x-ray images. arXiv 2020, arXiv:2003.11055. [Google Scholar]
  55. Li, L.; Qin, L.; Xu, Z.; Yin, Y.; Wang, X.; Kong, B.; Bai, J.; Lu, Y.; Fang, Z.; Song, Q.; et al. Using artificial intelligence to detect COVID-19 and community-acquired pneumonia based on pulmonary CT: Evaluation of the diagnostic accuracy. Radiology 2020, 296, E65–E71. [Google Scholar] [CrossRef] [PubMed]
  56. Elkorany, A.S.; Elsharkawy, Z.F. COVIDetection-Net: A tailored COVID-19 detection from chest radiography images using deep learning. Optik 2021, 231, 166405. [Google Scholar] [CrossRef] [PubMed]
  57. Nigam, B.; Nigam, A.; Jain, R.; Dodia, S.; Arora, N.; Annappa, B. COVID-19: Automatic detection from X-ray images by utilizing deep learning methods. Expert Syst. Appl. 2021, 176, 114883. [Google Scholar] [CrossRef] [PubMed]
  58. Agrawal, T.; Choudhary, P. FocusCovid: Automated COVID-19 detection using deep learning with chest X-ray images. Evol. Syst. 2021, 9, 1–5. [Google Scholar] [CrossRef]
  59. Gökcen, A.; Karadağ, B.; Riva, C.; Boyacı, A. Artificial intelligence–based COVID-19 detection using cough records. Electrica 2021, 21, 203–208. [Google Scholar] [CrossRef]
  60. Khattak, M.I.; Al-Hasan, M.A.; Jan, A.; Saleem, N.; Verdú, E.; Khurshid, N. Automated Detection of COVID-19 using Chest X-Ray Images and CT Scans through Multilayer-Spatial Convolutional Neural Networks. Int. J. Interact. Multimed. Artif. Intell. 2021, 6, 15–24. [Google Scholar] [CrossRef]
  61. Hussain, E.; Hasan, M.; Rahman, M.A.; Lee, I.; Tamanna, T.; Parvez, M.Z. CoroDet: A deep learning based classification for COVID-19 detection using chest X.-ray images. Chaos Solitons Fractals 2021, 142, 110495. [Google Scholar] [CrossRef]
  62. Hashmi, H.A.; Asif, H.M. Early Detection Of COVID-19. medRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  63. Mei, X.; Lee, H.C.; Diao, K.Y.; Huang, M.; Lin, B.; Liu, C.; Xie, Z.; Ma, Y.; Robson, P.M.; Chung, M.; et al. Artificial intelligence–enabled rapid diagnosis of patients with COVID-19. Nat. Med. 2020, 26, 1224–1228. [Google Scholar] [CrossRef]
  64. Kaushik, A.K.; Dhau, J.S.; Gohel, H.; Mishra, Y.K.; Kateb, B.; Kim, N.Y.; Goswami, D.Y. Electrochemical SARS-CoV-2 sensing at point-of-care and artificial intelligence for intelligent COVID-19 management. ACS Appl. Bio Mater. 2020, 3, 7306–7325. [Google Scholar] [CrossRef]
  65. Irfan, M.; Iftikhar, M.A.; Yasin, S.; Draz, U.; Ali, T.; Hussain, S.; Bukhari, S.; Alwadie, A.S.; Rahman, S.; Glowacz, A.; et al. Role of Hybrid Deep Neural Networks (HDNNs), Computed Tomography, and Chest X-rays for the Detection of COVID-19. Int. J. Environ. Res. Public Health 2021, 18, 3056. [Google Scholar] [CrossRef]
  66. Alqudah, A.M.; Qazan, S.; Alquran, H.; Qasmieh, I.A.; Alqudah, A. Covid-2019 detection using x-ray images and artificial intelligence hybrid systems. X-Ray Imaging 2020, 2, 1. [Google Scholar]
  67. Loey, M.; Smarandache, F.; Khalifa, M.N.E. Within the lack of chest COVID-19 X-ray dataset: A novel detection model based on GAN and deep transfer learning. Symmetry 2020, 12, 651. [Google Scholar] [CrossRef] [Green Version]
  68. Albahli, S.; Yar, G.N. Fast and Accurate Detection of COVID-19 Along With 14 Other Chest Pathologies Using a Multi-Level Classification: Algorithm Development and Validation Study. J. Med. Internet Res. 2021, 23, e23693. [Google Scholar] [CrossRef]
  69. Kim, H.; Hwang, J.; Kim, J.H.; Lee, S.; Kang, M. Sensitive detection of multiple fluoresence probes based on surface-enhanced raman scattering (sers) for mers-cov. In Proceedings of the 2019 IEEE 14th International Conference on Nano/Micro Engineered and Molecular Systems (NEMS), Bangkok, Thailand, 11–14 April 2019; pp. 498–501. [Google Scholar]
  70. Qiu, G.; Gai, Z.; Tao, Y.; Schmitt, J.; Kullak-Ublick, G.A.; Wang, J. Dual-functional plasmonic photothermal biosensors for highly accurate severe acute respiratory syndrome coronavirus 2 detection. ACS Nano 2020, 14, 5268–5277. [Google Scholar] [CrossRef] [Green Version]
  71. Seo, G.; Lee, G.; Kim, M.J.; Baek, S.H.; Choi, M.; Ku, K.B.; Lee, C.S.; Jun, S.; Park, D.; Kim, H.G.; et al. Rapid detection of COVID-19 causative virus (SARS-CoV-2) in human nasopharyngeal swab specimens using field-effect transistor-based biosensor. ACS Nano 2020, 14, 5135–5142. [Google Scholar] [CrossRef] [Green Version]
  72. Bhalla, N.; Pan, Y.; Yang, Z.; Payam, A.F. Opportunities and challenges for biosensors and nanoscale analytical tools for pandemics: COVID-19. ACS Nano 2020, 14, 7783–7807. [Google Scholar] [CrossRef]
  73. Murugan, D.; Bhatia, H.; Sai, V.V.; Satija, J. P-FAB: A fiber-optic biosensor device for rapid detection of COVID-19. Trans. Indian Natl. Acad. Eng. 2020, 5, 211–215. [Google Scholar] [CrossRef]
  74. Nag, P.; Sadani, K.; Mukherji, S. Optical fiber sensors for rapid screening of COVID-19. Trans. Indian Natl. Acad. Eng. 2020, 5, 233–236. [Google Scholar] [CrossRef]
  75. Ruiz-Vega, G.; Soler, M.; Lechuga, L.M. Nanophotonic biosensors for point-of-care COVID-19 diagnostics and coronavirus surveillance. J. Phys. Photonics. 2021, 3, 011002. [Google Scholar] [CrossRef]
  76. Van Tran, V.; Tran, N.H.; Hwang, H.S.; Chang, M. Development strategies of conducting polymer-based electrochemical biosensors for virus biomarkers: Potential for rapid COVID-19 detection. Biosens. Bioelectron. 2021, 182, 113192. [Google Scholar] [CrossRef] [PubMed]
  77. Choi, H.K.; Lee, M.J.; Lee, S.N.; Kim, T.H.; Oh, B.K. Noble metal nanomaterial-based biosensors for electrochemical and optical detection of viruses causing respiratory illnesses. Front. Chem. 2021, 9, 672739. [Google Scholar] [CrossRef]
  78. Fabiani, L.; Saroglia, M.; Galatà, G.; De Santis, R.; Fillo, S.; Lucam, V.; Faggioni, G.; D’Amore, N.; Regalbuto, E.; Salvatori, P.; et al. Magnetic beads combined with carbon black-based screen-printed electrodes for COVID-19: A reliable and miniaturized electrochemical immunosensor for SARS-CoV-2 detection in saliva. Biosens. Bioelectron. 2021, 171, 112686. [Google Scholar] [CrossRef]
  79. Kudo, E.; Israelow, B.; Vogels, C.B.; Lu, P.; Wyllie, A.L.; Tokuyama, M.; Venkataraman, A.; Brackney, D.E.; Ott, I.M.; Petrone, M.E.; et al. Detection of SARS-CoV-2 RNA by multiplex RT-qPCR. PLoS Biol. 2020, 18, e3000867. [Google Scholar] [CrossRef]
  80. 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]
  81. Ke, G.; Su, D.; Li, Y.; Zhao, Y.; Wang, H.; Liu, W.; Li, M.; Yang, Z.; Xiao, F.; Yuan, Y.; et al. An accurate, high-speed, portable bifunctional electrical detector for COVID-19. Sci. China Mater. 2021, 64, 739–747. [Google Scholar] [CrossRef]
  82. Chen, Z.; Zhang, Z.; Zhai, X.; Li, Y.; Lin, L.; Zhao, H.; Bian, L.; Li, P.; Yu, L.; Wu, Y.; et al. Rapid and sensitive detection of anti-SARS-CoV-2 IgG, using lanthanide-doped nanoparticles-based lateral flow immunoassay. Anal. Chem. 2020, 92, 7226–7231. [Google Scholar] [CrossRef]
  83. Damle, V.H.; Rajeswaran, B.; Tischler, Y.R. Low-Frequency Raman Spectroscopy as a Diagnostic Tool for COVID-19 and other Coronaviruses. R. Soc. Open Sci. 2020, 7, 1–28. [Google Scholar]
  84. Wu, Y.C.; Chen, C.S.; Chan, Y.J. The outbreak of COVID-19: An overview. J. Chin. Med. Assoc. 2020, 83, 217. [Google Scholar] [CrossRef]
  85. Argenziano, M.G.; Bruce, S.L.; Slater, C.L.; Tiao, J.R.; Baldwin, M.R.; Barr, R.G.; Chang, B.P.; Chau, K.H.; Choi, J.J.; Gavin, N.; et al. Characterization and clinical course of 1000 patients with coronavirus disease 2019 in New York: Retrospective case series. Bmj 2020, 1–9, 369. [Google Scholar] [CrossRef]
  86. Van Doremalen, N.; Bushmaker, T.; Morris, D.H.; Holbrook, M.G.; Gamble, A.; Williamson, B.N.; Tamin, A.; Harcourt, J.L.; Thornburg, N.J.; Gerber, S.I.; et al. Aerosol and surface stability of SARS-CoV-2 as compared with SARS-CoV-1. N. Engl. J. Med. 2020, 382, 1564–1567. [Google Scholar] [CrossRef]
  87. Wang, C.; Pan, R.; Wan, X.; Tan, Y.; Xu, L.; Ho, C.S.; Ho, R.C. Immediate psychological responses and associated factors during the initial stage of the 2019 coronavirus disease (COVID-19) epidemic among the general population in China. Int. J. Environ. Res. Public Health 2020, 17, 1729. [Google Scholar] [CrossRef] [Green Version]
  88. Saw, L.H.; Leo, B.F.; Nor, N.S.M.; Yip, C.W.; Ibrahim, N.; Abd Hamid, H.H.; Latif, M.T.; Lin, C.Y.; Nadzir, M.S. Modeling aerosol transmission of SARS-CoV-2 from human-exhaled particles in a hospital ward. Environ. Sci. Pollut. Res. 2021, 28, 53478–53492. [Google Scholar] [CrossRef]
  89. Asadi, S.; Wexler, A.S.; Cappa, C.; Barreda, S.; Bouvier, N.M.; Ristenpart, W.D. Aerosol emission and superemission during human speech increase with voice loudness. Sci. Rep. 2019, 9, 1–10. [Google Scholar] [CrossRef] [Green Version]
  90. Qu, G.; Li, X.; Hu, L.; Jiang, G. An imperative need for research on the role of environmental factors in transmission of novel coronavirus (COVID-19). ACS Publ. 2020, 54, 3730–3732. [Google Scholar] [CrossRef]
  91. Medicine, The Lancet Respiratory. COVID-19 transmission—up in the air. Lancet. Respir. Med. 2020, 8, 1159. [Google Scholar] [CrossRef]
  92. Cheung, K.S.; Hung, I.F.; Chan, P.P.; Lung, K.C.; Tso, E.; Liu, R.; Ng, Y.Y.; Chu, M.Y.; Chung, T.W.; Tam, A.R.; et al. Gastrointestinal manifestations of SARS-CoV-2 infection and virus load in fecal samples from a Hong Kong cohort: Systematic review and meta-analysis. Gastroenterology 2020, 159, 81–95. [Google Scholar] [CrossRef]
  93. Dong, Y.; Mo, X.I.; Hu, Y.; Qi, X.; Jiang, F.; Jiang, Z.; Tong, S. Epidemiological characteristics of 2143 pediatric patients with 2019 coronavirus disease in China. Pediatrics 2020, 58, 712–713. [Google Scholar]
  94. Pan, L.; Mu, M.; Yang, P.; Sun, Y.; Wang, R.; Yan, J.; Li, P.; Hu, B.; Wang, J.; Hu, C.; et al. Clinical characteristics of COVID-19 patients with digestive symptoms in Hubei, China: A descriptive, cross-sectional, multicenter study. Am. J. Gastroenterol. 2020, 115. [Google Scholar] [CrossRef]
  95. Gao, Q.Y.; Chen, Y.X.; Fang, J.Y. 2019 Novel coronavirus infection and gastrointestinal tract. J. Dig. Dis. 2020, 21, 125. [Google Scholar] [CrossRef] [Green Version]
  96. Chan, J.F.; Kok, K.H.; Zhu, Z.; Chu, H.; To, K.K.; Yuan, S.; Yuen, K.Y. Genomic characterization of the 2019 novel human-pathogenic coronavirus isolated from a patient with atypical pneumonia after visiting Wuhan. Emerg. Microbes Infect. 2020, 9, 221–236. [Google Scholar] [CrossRef] [Green Version]
  97. De Graaf, M.; Beck, R.; Caccio, S.M.; Duim, B.; Fraaij, P.L.; Le Guyader, F.S.; Lecuit, M.; Le Pendu, J.D.W.E.; Schultsz, C. Sustained fecal-oral human-to-human transmission following a zoonotic event. Curr. Opin. Virol. 2017, 22, 1–6. [Google Scholar] [CrossRef]
  98. Du, M.; Cai, G.; Chen, F.; Christiani, D.C.; Zhang, Z.; Wang, M. Multiomics evaluation of gastrointestinal and other clinical characteristics of COVID-19. Gastroenterology 2020, 158, 2298–2301. [Google Scholar] [CrossRef]
  99. Zhang, H.; Kang, Z.; Gong, H.; Xu, D.; Wang, J.; Li, Z.; Cui, X.; Xiao, J.; Meng, T.; Zhou, W.; et al. The digestive system is a potential route of 2019-nCov infection: A bioinformatics analysis based on single-cell transcriptomes. BioRxiv. 2020, 1–26. [Google Scholar] [CrossRef]
  100. Xiao, F.; Tang, M.; Zheng, X.; Liu, Y.; Li, X.; Shan, H. Evidence for gastrointestinal infection of SARS-CoV-2. Gastroenterology 2020, 158, 1831–1833. [Google Scholar] [CrossRef]
  101. Masand, R.; Jadhao, A.; Jadhav, S. Review on Modes of Transmission of COVID-19. Int. J. Curr. Microbiol. Appl. Sci. 2021, 10, 1003–1014. [Google Scholar]
  102. Karimi, S.; Arabi, A.; Shahraki, T.; Safi, S. Detection of severe acute respiratory syndrome Coronavirus-2 in the tears of patients with Coronavirus disease 2019. Eye 2020, 34, 1220–1223. [Google Scholar] [CrossRef]
  103. Sungnak, W.; Huang, N.; Bécavin, C.; Berg, M.; Queen, R.; Litvinukova, M.; Talavera-López, C.; Maatz, H.; Reichart, D.; Sampaziotis, F.; et al. SARS-CoV-2 entry factors are highly expressed in nasal epithelial cells together with innate immune genes. Nat. Med. 2020, 26, 681–687. [Google Scholar] [CrossRef] [Green Version]
  104. Zhou, L.; Xu, Z.; Castiglione, G.M.; Soiberman, U.S.; Eberhart, C.G.; Duh, E.J. ACE2 and TMPRSS2 are expressed on the human ocular surface, suggesting susceptibility to SARS-CoV-2 infection. Ocul. Surf. 2020, 18, 537–544. [Google Scholar] [CrossRef] [PubMed]
  105. Roehrich, H.; Yuan, C.; Hou, J.H. Immunohistochemical study of SARS-CoV-2 viral entry factors in the cornea and ocular surface. Cornea 2020, 39, 1556–1562. [Google Scholar] [CrossRef] [PubMed]
  106. Chen, L.; Liu, M.; Zhang, Z.; Qiao, K.; Huang, T.; Chen, M.; Xin, N.; Huang, Z.; Liu, L.; Zhang, G.; et al. Ocular manifestations of a hospitalised patient with confirmed 2019 novel coronavirus disease. Br. J. Ophthalmol. 2020, 104, 748–751. [Google Scholar] [CrossRef] [PubMed]
  107. Ocansey, S.; Abu, E.K.; Abraham, C.H.; Owusu-Ansah, A.; Boadi-Kusi, S.B.; Ilechie, A.A.; Acheampong, D.O. Ocular Symptoms of SARS-CoV-2: Indication of Possible Ocular Transmission or Viral Shedding. Ocul. Im Munology Inflamm. 2020, 28, 1269–1279. [Google Scholar] [CrossRef]
  108. Lu, C.-W.; Liu, X.-F.; Jia, Z.-F. 2019-nCoV transmission through the ocular surface must not be ignored. Lancet 2020, 395, e39. [Google Scholar] [CrossRef] [Green Version]
  109. Luo, Y.; Yin, K. Management of pregnant women infected with COVID-19. Lancet Infect. Dis. 2020, 20, 513–514. [Google Scholar] [CrossRef] [Green Version]
  110. Cornish, E.F.; Filipovic, I.; Åsenius, F.; Williams, D.J.; McDonnell, T. Innate immune responses to acute viral in-fection during pregnancy. Front. Immunol. 2020, 11, 2404. [Google Scholar] [CrossRef]
  111. Lu-Culligan, A.; Chavan, A.R.; Vijayakumar, P.; Irshaid, L.; Courchaine, E.M.; Milano, K.M.; Tang, Z.; Pope, S.D.; Song, E.; Vogels, C.B.; et al. SARS-CoV-2 infection in pregnancy is associated with robust inflam-matory response at the maternal-fetal interface. medRxiv 2021, 1–45. [Google Scholar] [CrossRef]
  112. Li, M.; Chen, L.; Zhang, J.; Xiong, C.; Li, X. The SARS-CoV-2 receptor ACE2 expression of maternal-fetal inter-face and fetal organs by single-cell transcriptome study. PLoS ONE 2020, 15, e0230295. [Google Scholar] [CrossRef] [Green Version]
  113. Wang, S.; Guo, L.; Chen, L.; Liu, W.; Cao, Y.; Zhang, J.; Feng, L. A case report of neonatal COVID-19 infection in China. Clin. Infect. Dis. 2020, 71, 853–857. [Google Scholar] [CrossRef]
  114. Mor, G.; Aldo, P.; Alvero, A.B. The unique immunological and microbial aspects of pregnancy. Nat. Rev. Immunol. 2017, 17, 469–482. [Google Scholar] [CrossRef]
  115. Karimi-Zarchi, M.; Neamatzadeh, H.; Dastgheib, S.A.; Abbasi, H.; Mirjalili, S.R.; Behforouz, A.; Ferdosian, F.; Bahrami, R. Vertical transmission of coronavirus disease 19 (COVID-19) from infected pregnant mothers to neonates: A review. Fetal Pediatric Pathol. 2020, 39, 246–250. [Google Scholar] [CrossRef] [Green Version]
  116. Liu, H.; Wang, L.-L.; Zhao, S.-J.; Kwak-Kim, J.; Mor, G.; Liao, A.-H. Why are pregnant women susceptible to COVID-19? An immunological viewpoint. J. Reprod. Immunol. 2020, 139, 103122. [Google Scholar] [CrossRef]
  117. Massart, S.; Chiumenti, M.; De Jonghe, K.; Glover, R.; Haegeman, A.; Koloniuk, I.; Komínek, P.; Kreuze, J.; Kutnjak, D.; Lotos, L.; et al. Virus Detection by High-Throughput Sequencing of Small RNAs: Large-Scale Performance Testing of Sequence Analysis Strategies. Phytopathology 2019, 109, 488–497. [Google Scholar] [CrossRef] [Green Version]
  118. John, G.; Sahajpal, N.S.; Mondal, A.K.; Ananth, S.; Williams, C.; Chaubey, A.; Rojiani, A.M.; Kolhe, R. Next-Generation Sequencing (NGS) in COVID-19: A Tool for SARS-CoV-2 Diagnosis, Monitoring New Strains and Phylo-dynamic Modeling in Molecular Epidemiology. Curr. Issues Mol. Biol. 2021, 43, 845–867. [Google Scholar] [CrossRef]
  119. Drmanac, R.; Sparks, A.B.; Callow, M.J.; Halpern, A.L.; Burns, N.L.; Kermani, B.G.; Carnevali, P.; Nazarenko, I.; Nilsen, G.B.; Yeung, G.; et al. Human genome sequencing using unchained base reads on self-assembling DNA nanoar-rays. Science 2010, 327, 78–81. [Google Scholar] [CrossRef] [Green Version]
  120. Chakraborty, D.; Kumar, S.; Chandrasekaran, N.; Mukherjee, F.R.S.B. Viral diagnostics and preventive techniques in the era of COVID-19: Role of nanotechnology. Front. Nanotechnol. 2020, 2, 7. [Google Scholar] [CrossRef]
  121. Wan, Z.; Zhang, Y.N.; He, Z.; Liu, J.; Lan, K.; Hu, Y.; Zhang, C. A melting curve-based multiplex RT-qPCR assay for simultaneous detection of four human coronaviruses. Int. J. Mol. Sci. 2016, 17, 1880. [Google Scholar] [CrossRef] [Green Version]
  122. Noh, J.Y.; Yoon, S.W.; Kim, D.J.; Lee, M.S.; Kim, J.H.; Na, W.; Song, D.; Jeong, D.G.; Kim, H.K. Simultaneous detection of se-vere acute respiratory syndrome, Middle East respiratory syndrome, and related bat coronaviruses by real-time reverse transcription PCR. Arch. Virol. 2017, 162, 1617–1623. [Google Scholar] [CrossRef] [Green Version]
  123. Sood, S.; Aggarwal, V.; Aggarwal, D.; Upadhyay, S.K.; Sak, K.; Tuli, H.S.; Kumar, M.; Kumar, J.; Talwar, S. COVID-19 Pandemic: From Molecular Biology, Pathogenesis, Detection, and Treatment to Global Societal Impact. Curr. Pharmacol. Rep. 2020, 6, 212–227. [Google Scholar] [CrossRef]
  124. Carter, L.J.; Garner, L.V.; Smoot, J.W.; Li, Y.; Zhou, Q.; Saveson, C.J. Janet, 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]
  125. Orooji, Y.; Sohrabi, H.; Hemmat, N.; Oroojalian, F.; Baradaran, B.; Mokhtarzadeh, A.; Karimi-Maleh, H. An overview on SARS-CoV-2 (COVID-19) and other human coronaviruses and their detection capability via am-plification assay, chemical sensing, biosensing, immunosensing, and clinical assays. Nano-Micro Lett. 2021, 13, 1–30. [Google Scholar] [CrossRef]
  126. Yang, W.; Dang, X.; Wang, Q.; Xu, M.; Zhao, Q.; Zhou, Y.; Zhao, H.; Wang, L.; Xu, Y.; Wang, J.; et al. Rapid detection of SARS-CoV-2 using reverse transcription RT-LAMP method. MedRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  127. Gill, P.; Ghaemi, A. Nucleic Acid Isothermal Amplification Technologies—A Review. Nucleosides Nucleotides Nucleic Acids 2008, 27, 224–243. [Google Scholar] [CrossRef] [PubMed]
  128. Tasrip, N.A.; Khairil Mokhtar, N.F.; Hanapi, U.K.; Abdul Manaf, Y.N.; Ali, M.E.; Cheah, Y.K.; Mustafa, S.; Desa, M. Loop mediated isothermal amplification; A review on its application and strategy in animal species authentication of meat based food products. Int. Food Res. J. 2019, 26, 1–10. [Google Scholar]
  129. Kumar, R.; Nagpal, S.; Kaushik, S.; Mendiratta, S. COVID-19 diagnostic approaches: Different roads to the same destination. VirusDisease 2020, 31, 97–105. [Google Scholar] [CrossRef] [PubMed]
  130. 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] [PubMed]
  131. Mukhopadhyay, C.D.; 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]
  132. Fozouni, P.; Son, S.; de León Derby, M.D.; Knott, G.J.; Gray, C.N.; D’Ambrosio, M.V.; Zhao, C.; Switz, N.A.; Kumar, G.R.; Stephens, S.I.; et al. Direct detection of SARS-CoV-2 using CRISPR-Cas13a and a mobile phone. MedRxiv 2020. [Google Scholar] [CrossRef]
  133. Hang, W.; Liu, K.; Zhang, P.; Cheng, W.; Li, L.; Zhang, F.; Yu, Z.; Li, L.; Zhang, X. CRISPR-Based approaches for efficient and accurate detection of SARS-CoV-2. Lab. Med. 2021, 52, 116–121. [Google Scholar]
  134. Lam, W.Y.; Yeung, A.C.M.; Tang, J.W.; Ip, M.; Chan, E.W.C.; Hui, M.; Chan, P.K.S. Rapid Multiplex Nested PCR for Detection of Respiratory Viruses. J. Clin. Microbiol. 2007, 45, 3631–3640. [Google Scholar] [CrossRef] [Green Version]
  135. Shi, L.; Sun, Q.; He, J.; Xu, H.; Liu, C.; Zhao, C.; Xu, Y.; Wu, C.; Xiang, J.; Gu, D.; et al. Development of SPR biosensor for simultaneous detection of multiplex respiratory viruses. Bio-Medical Mater. Eng. 2015, 26, S2207–S2216. [Google Scholar] [CrossRef] [Green Version]
  136. Lu, X.; Whitaker, B.; Sakthivel, S.K.K.; Kamili, S.; Rose, L.E.; Lowe, L.; Mohareb, E.; Elassal, E.M.; Al-sanouri, T.; Haddadin, A.; et al. Real-time reverse transcription-PCR assay panel for Middle East respiratory syndrome corona-virus. J. Clin. Microbiol. 2014, 52, 67–75. [Google Scholar] [CrossRef] [Green Version]
  137. Pfefferle, S.; Reucher, S.; Nörz, D.; Lütgehetmann, M. Evaluation of a quantitative RT-PCR assay for the detection of the emerging coronavirus SARS-CoV-2 using a high throughput system. Eurosurveillance 2020, 25, 2000152. [Google Scholar] [CrossRef] [Green Version]
  138. Ju, B.; Zhang, Q.; Ge, J.; Wang, R.; Sun, J.; Ge, X.; Yu, J.; Shan, S.; Zhou, B.; Song, S.; et al. Human neutralizing antibodies elicited by SARS-CoV-2 infection. Nature 2020, 584, 115–119. [Google Scholar] [CrossRef]
  139. Huang, P.; Wang, H.; Cao, Z.; Jin, H.; Chi, H.; Zhao, J.; Yu, B.; Yan, F.; Hu, X.; Wu, F.; et al. A rapid and specific assay for the detection of MERS-CoV. Front. Microbiol. 2018, 9, 1101. [Google Scholar] [CrossRef]
  140. Kim, J.H.; Kang, M.; Park, E.; Chung, D.R.; Kim, J.; Hwang, E.S. A simple and multiplex loop-mediated isothermal am-plification (LAMP) assay for rapid detection of SARS-CoV. Biochip J. 2019, 13, 341–351. [Google Scholar] [CrossRef] [Green Version]
  141. Zhang, J.; Wang, X.; Jia, X.; Li, J.; Hu, K.; Chen, G.; Wei, J.; Gong, Z.; Zhou, C.; Yu, H.; et al. Risk factors for disease severity, unimprovement, and mortality in COVID-19 patients in Wuhan, China. Clin. Microbiol. Infect. 2020, 26, 767–772. [Google Scholar] [CrossRef]
  142. 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]
  143. Taha, B.; Al Mashhadany, Y.; Mokhtar, M.H.; Bin Zan, M.D.; Arsad, N. An Analysis Review of Detection Coronavirus Disease 2019 (COVID-19) Based on Biosensor Application. Sensors 2020, 20, 6764. [Google Scholar] [CrossRef]
  144. Dai, Y.; Liu, C.C. Recent advances on electrochemical biosensing strategies toward universal point-of-care systems. Angew. Chemie 2019, 131, 12483–12496. [Google Scholar] [CrossRef]
  145. Ibrahim, N.; Jamaluddin, N.; Tan, L.; Yusof, N.M. A Review on the Development of Gold and Silver Nanoparticles-Based Biosensor as a Detection Strategy of Emerging and Pathogenic RNA Virus. Sensors 2021, 21, 5114. [Google Scholar] [CrossRef] [PubMed]
  146. Russo, L.; Bueno, J.L.; Bergua, J.F.; Costantini, M.; Giannetto, M.; Puntes, V.; De La Escosura-Muñiz, A.; Merkoçi, A. Low-Cost Strategy for the Development of a Rapid Electrochemical Assay for Bacteria Detection Based on AuAg Nanoshells. ACS Omega 2018, 3, 18849–18856. [Google Scholar] [CrossRef]
  147. Wang, Y.; Xu, H.; Zhang, J.; Li, G. Electrochemical Sensors for Clinic Analysis. Sensors 2008, 8, 2043–2081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Song, Z.; Ma, Y.; Chen, M.; Ambrosi, A.; Ding, C.; Luo, X. Electrochemical Biosensor with Enhanced Antifouling Capability for COVID-19 Nucleic Acid Detection in Complex Biological Media. Anal. Chem. 2021, 93, 5963–5971. [Google Scholar] [CrossRef] [PubMed]
  149. Nehra, A.; Singh, K.P. Current trends in nanomaterial embedded field effect transistor-based biosensor. Biosens. Bioelectron. 2015, 74, 731–743. [Google Scholar] [CrossRef] [PubMed]
  150. Janissen, R.; Sahoo, P.K.; Santos, C.A.; da Silva, A.M.; von Zuben, A.A.G.; Souto, D.E.P.; Costa, A.D.T.; Celedon, P.; Zanchin, N.I.T.; Almeida, D.B.; et al. InP Nanowire Biosensor with Tailored Biofunctionalization: Ultrasensitive and Highly Selective Disease Biomarker Detection. Nano Lett. 2017, 17, 5938–5949. [Google Scholar] [CrossRef] [Green Version]
  151. Shabani, E.; Dowlatshahi, S.; Abdekhodaie, M.J. Laboratory detection methods for the human coronaviruses. Eur. J. Clin. Microbiol. Infect. Dis. 2021, 40, 225–246. [Google Scholar] [CrossRef]
  152. Zhang, X.; Qi, Q.; Jing, Q.; Ao, S.; Zhang, Z.; Ding, M.; Wu, M.; Liu, K.; Wang, W.; Ling, Y.; et al. Electrical probing of COVID-19 spike protein receptor binding domain via a graphene field-effect transistor. arXiv 2020, arXiv:2003.12529. [Google Scholar]
  153. Pohanka, M. Overview of Piezoelectric Biosensors, Immunosensors and DNA Sensors and Their Applications. Materials 2018, 11, 448. [Google Scholar] [CrossRef] [Green Version]
  154. Shons, A.; Dorman, F.; Najarian, J. An immunospecific microbalance. J. Biomed. Mater. Res. 1972, 6, 565–570. [Google Scholar] [CrossRef]
  155. Meyers, F.N.; Loh, K.J.; Dodds, J.S.; Baltazar, A. Active sensing and damage detection using piezoelectric zinc ox-ide-based nanocomposites. Nanotechnology 2013, 24, 185501. [Google Scholar] [CrossRef]
  156. Zuo, B.; Li, S.; Guo, Z.; Zhang, J.; Chen, C. Piezoelectric immunosensor for SARS-associated coronavirus in spu-tum. Anal. Chem. 2004, 76, 3536–3540. [Google Scholar] [CrossRef]
  157. Kabir, H.; Merati, M.; Abdekhodaie, M.J. Design of an effective piezoelectric microcantilever biosensor for rap-id detection of COVID-19. J. Med. Eng. Technol. 2021, 14, 1. [Google Scholar]
  158. Zhang, K.; Zhang, L.; Fu, L.; Li, S.; Chen, H.; Cheng, Z.-Y. Magnetostrictive resonators as sensors and actuators. Sens. Actuators A Phys. 2013, 200, 2–10. [Google Scholar] [CrossRef]
  159. Chen, Y.; Wang, Z.; Liu, Y.; Wang, X.; Li, Y.; Ma, P.; Gu, B.; Li, H. Recent advances in rapid pathogen detection method based on biosensors. Eur. J. Clin. Microbiol. Infect. Dis. 2018, 37, 1021–1037. [Google Scholar] [CrossRef]
  160. Narita, F.; Wang, Z.; Kurita, H.; Li, Z.; Shi, Y.; Jia, Y.; Soutis, C. A Review of Piezoelectric and Magnetostrictive Biosensor Materials for Detection of COVID-19 and Other Viruses. Adv. Mater. 2020, 33, e2005448. [Google Scholar] [CrossRef]
  161. Ansari, A.A.; Alhoshan, M.; Alsalhi, M.S.; Aldwayyan, A.S. Prospects of nanotechnology in clinical immunodi-agnostics. Sensors 2010, 10, 6535–6581. [Google Scholar] [CrossRef]
  162. Sciacca, B.; François, A.; Hoffmann, P.; Monro, T.M. Multiplexing of radiative-surface plasmon resonance for the detection of gastric cancer biomarkers in a single optical fiber. Sens. Actuators B Chem. 2013, 183, 454–458. [Google Scholar] [CrossRef]
  163. Taha, B.; Ali, N.; Sapiee, N.; Fadhel, M.; Yeh, R.M.; Bachok, N.; Al Mashhadany, Y.; Arsad, N. Comprehensive Review Tapered Optical Fiber Configurations for Sensing Application: Trend and Challenges. Biosensors 2021, 11, 253. [Google Scholar] [CrossRef]
  164. Cotruvo, J.A., Jr.; Aron, A.T.; Ramos-Torres, K.M.; Chang, C.J. Synthetic fluorescent probes for studying copper in biological systems. Chem. Soc. Rev. 2015, 44, 4400–4414. [Google Scholar] [CrossRef] [Green Version]
  165. Sharma, A.; Khan, R.; Catanante, G.; Sherazi, T.A.; Bhand, S.; Hayat, A.; Marty, J.L. Designed strategies for flu-orescence-based biosensors for the detection of mycotoxins. Toxins 2018, 10, 197. [Google Scholar] [CrossRef] [Green Version]
  166. Diao, B.; Wen, K.; Chen, J.; Liu, Y.; Yuan, Z.; Han, C.; Chen, J.; Pan, Y.; Chen, L.; Dan, Y.; et al. Diagnosis of acute respiratory syndrome coronavirus 2 infection by detection of nucleocapsid protein. MedRxiv 2020. [Google Scholar] [CrossRef]
  167. Song, D.; Liu, J.; Xu, W.; Han, X.; Wang, H.; Cheng, Y.; Zhuo, Y.; Long, F. Rapid and quantitative detection of SARS-CoV-2 IgG antibody in serum using optofluidic point-of-care testing fluorescence biosensor. Talanta 2021, 235, 122800. [Google Scholar] [CrossRef] [PubMed]
  168. Chen, R.; Kan, L.; Duan, F.; He, L.; Wang, M.; Cui, J.; Zhang, Z.; Zhang, Z. Surface plasmon resonance aptasensor based on niobium carbide MXene quantum dots for nucleocapsid of SARS-CoV-2 detection. Microchim. Acta. 2021, 188, 1–10. [Google Scholar] [CrossRef] [PubMed]
  169. Huang, J.C.; Chang, Y.F.; Chen, K.H.; Su, L.C.; Lee, C.W.; Chen, C.C.; Chen, Y.M.; Chou, C. Detection of severe acute respiratory syndrome (SARS) coronavirus nucleocapsid protein in human serum using a localized sur-face plasmon coupled fluorescence fiber-optic biosensor. Biosens. Bioelectron. 2009, 25, 320–325. [Google Scholar] [CrossRef] [PubMed]
  170. Divagar, M.; Gayathri, R.; Rasool, R.; Shamlee, J.K.; Bhatia, H.; Satija, J.; Sai, V.V. Plasmonic Fiberoptic Absorb-ance Biosensor (P-FAB) for Rapid Detection of SARS-CoV-2 Nucleocapsid Protein. IEEE Sens. J. 2021, 21, 22758–22766. [Google Scholar] [CrossRef]
  171. Mahari, S.; Roberts, A.; Shahdeo, D.; Gandhi, S. eCovSens-ultrasensitive novel in-house built printed circuit board based electrochemical device for rapid detection of nCovid-19 antigen, a spike protein domain 1 of SARS-CoV-2. BioRxiv 2020, 172, 1–20. [Google Scholar] [CrossRef] [Green Version]
  172. Layqah, L.A.; Eissa, S. An electrochemical immunosensor for the corona virus associated with the Middle East respiratory syndrome using an array of gold nanoparticle-modified carbon electrodes. Microchim. Acta 2019, 186, 224. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Diverse types of coronaviruses within Nidovirales, Coronaviridae family, Orthocoronavirinae subfamily and the respective genera, (α), (β), (γ) and (δ).
Figure 1. Diverse types of coronaviruses within Nidovirales, Coronaviridae family, Orthocoronavirinae subfamily and the respective genera, (α), (β), (γ) and (δ).
Sensors 21 08362 g001
Figure 2. SARS-CoV and SARS-CoV-2 genomic structure and proteins.
Figure 2. SARS-CoV and SARS-CoV-2 genomic structure and proteins.
Sensors 21 08362 g002
Figure 3. Taxonomy of literature reviews on COVID-19 viral virus.
Figure 3. Taxonomy of literature reviews on COVID-19 viral virus.
Sensors 21 08362 g003
Figure 4. COVID-19 respiratory transmission Human-to-Human.
Figure 4. COVID-19 respiratory transmission Human-to-Human.
Sensors 21 08362 g004
Figure 5. Fecal–oral COVID-19 transmission.
Figure 5. Fecal–oral COVID-19 transmission.
Sensors 21 08362 g005
Table 1. Comparative Coronaviruses that infect the human body respiratory system.
Table 1. Comparative Coronaviruses that infect the human body respiratory system.
CoronavirusesYear of FindingEmergenceTypeHOSTCellular ReceptorIncubation PeriodRespiratory System InfectionSymptomsMortality RateReference
HCoV-HKU12005Hong KongBetaHuman9-O-Acetylated sialic acid2–4 daysCommon cold, Bronchitis, and pneumonia.N.A.[33]
HCoV-Nl632004HollandAlphaHumanACE22–4 daysCommon cold, sore throat, bronchiolitis/croup in children, high temperature, malaise, coughing and rhinitisN.A.[34]
HCoV-229E1966N.A.AlphaHumanHuman aminopeptidase N (CD13)2–5 daysCommon cold, Headache, Fever, Running nose, Pneumonia (in neonates), malaise, Bronchiolitis,N.A.[35]
HCoV-OC431967N.A.BetaHuman9-O-Acetylated sialic acid2–5 daysRunning nose, Common cold, Fever, Headache, Malaise, Bronchiolitis, Pneumonia (in neonates)N.A.[35]
SARS-CoV2002Guangdong, southern ChinaBetaCivets, HumanACE22–11 daysHeadache, Diarrhea, Fever, Shivering, Dyspnea, Cough, Pneumonia, Myalgia10%[33]
MERS-CoV2012Jeddah, Saudi ArabiaBetaCamel, HumanDPP42–14 daysFever, sore throat, dyspnea, dry cough, Chills, Pneumonia, Myalgia, Diarrhea, Hemoptysis, Headache, Rhinorrhoea35%[19]
SARS-CoV-22019Hubei province, Wuhan city, ChinaBetaBat, HumanACE23–6 daysHigh body temperature, Short of breath, Headache, Sore throat, myalgia, Dry coughing, Anosmia, rarely pneumonia, Diarrhea, Generalized weakness, Nasal congestion, Rhinorrhea, Sneezing2.8%[36]
Delta Variant *2020IndiaDeltaHumanACE25–6 daysHeadaches, Sore Throat, Runny Nose, Replacing Cough and Loss of Taste, Loss of SmellN.A.[37]
NA = Not Available, * Delta Variant is a mutation of SARS-CoV-2.
Table 2. Human exchange transmission comparison.
Table 2. Human exchange transmission comparison.
ComparisonTransmission Human Exchange
Respiratory TransmissionFecal–Oral TransmissionOcular TransmissionVertical Transmission
ProbabilityMore likelyMore likelyLess likelyLess likely
PossibilityHighHighRareRarer
Confirmed CasesConfirmedConfirmedConfirmedUnconfirmed
Virus entry organMouth, NoseMouthEyeUterus
Transmissibility approachDirect, IndirectIndirectDirect, IndirectDirect
GenusMale and FemaleMale and FemaleMale and FemaleFemale
Table 3. RNA Coronavirus detection methods analysis based on RT-PCR, LAMP, NGS and CRISPR.
Table 3. RNA Coronavirus detection methods analysis based on RT-PCR, LAMP, NGS and CRISPR.
ReferenceCoronavirusAnalyteTarget GenesDetection MethodsLimit of DetectionConcentration
Range
Detection TimeTested Sample
[134]SARS-CoVRNAPolymeraseRT-PCR10 copies/reactionN.A.(5) hNasal aspirate
[135]SARS-CoVRNANART-PCR2 nMN.A.(~2) hThroat swab samples
[136]MERS-CoVRNA(N) generRT-PCR10 copies/reaction or 0.0013 TCID50/ml10–108 copies/-reaction(~2) hSerum, nasopharyngeal/- oropharyngeal swab, and sputum samples
[137]COVID-19RNA(E)-generRT-PCR275.7 copies/reactionN.A.(~1) hSwab samples
[138]COVID-19RNA(N) generRT-PCR10 copies/reactionN.A.(~30) minPlasmids containing the complete N gene
[139]MERS-CoVRNA(N) geneRT-LAMP10 copies/μL5 × 101–5 × 108 copies/-reaction(35) minThroat swab specimens
[140]SARS-CoVRNA(ORF1b) and (N) geneLAMP104 copies/reactionN.A.(20–25) minSynthetic RNA solutions
[141]COVID-19RNA(ORF1b) and (N) geneRT-LAMP20 copies/reactionN.A.(20–30) minNasopharyngeal swab and bronchoalveolar lavage fluid samples
[126]COVID-19RNA(ORF1ab), (N) and (E) geneRT-LAMP5 copies/reactionN.A.(30) minNasopharyngeal swab specimens
[142]COVID-19RNA(S) geneNGS125 GCE/mLN.A.N.A.Nasopharyngeal swab
[132]COVID-19RNA(N) and (E) geneCRISPR/Cas13a~100 copies/µL3.2 × 105 –1.65 × 103 copies/µL(30) minNasal swab
NA = NOT available, N = nucleocapsid, E = envelope, ORF1b = open reading frame 1b, S = s pike protein.
Table 4. Coronavirus detection methods analysis based on biosensor application.
Table 4. Coronavirus detection methods analysis based on biosensor application.
ReferencePublication YearCoronavirusBiosensor Detection TechniqueMaterialTargetDetection TimeLinear RangeTested SampleLimit of DetectionTemperatures
[148]2 April 2021COVID-19Electrochemical(PANI)N gene1 h10−14 to 10−9 MNR3.5 fM37 °C
[171]11 May 2020COVID-19ElectrochemicalGoldS protein10–30 s1 fM to 1 μMSpiked saliva samples90 fM4 °C
[172]27 February 2019MERS-CoVElectrochemicalGoldS protein20 min1 pg·mL−1 to 10 μg·mL−1Spiked nasal samples0.4 and 1.0 pg·mL−1RT
[71]15 April 2020COVID-19Electrical (FET)GrapheneS protein4 hNRnasopharyngeal swab1.6 × 101 pfu/mLNR
[152]2020COVID-19Electrical (FET)GrapheneS protein2 minNRSpiked spike protein solutions0.2 pMNR
[156]1 July 2004SARS-CoVPiezoelectricCrystal with quartz waferAntigen sputum1 h1–4 µg/µLNR0.60 mg/mLRT
[166]13 March 2020COVID-19Optical (fluorescence)Not SpecifiedN protein10 minNRNasopharyngeal aspirate swabs and urineNot SpecifiedNR
[167]14 August 2021COVID-19Optical (fluorescence)Not SpecifiedIgG25 minNRHuman serum12.5 ng/mLNR
[168]11August 2021SARS-CoV-2Optical (SPR)Nb2C-SH QDN geneNA0.05 to 100 ng·mL−1Human serum4.9 pg·mL−1NR
[169]17 July 2009SARS-CoVOptical (LSPCF)polymethyl methacrylateN protein2 h0.1 pg/mL to 1 ng/mLHuman serum∼1 pg/mL37 °C
[170]1 September 2021COVID-19Optical (P-FAB)Gold nanoparticlesN protein10 min0.1 ng/mL and 100 ng/mLPBS Buffer~2.5 ng/mLNR
Electropolymerized polyaniline (PANI) nanowires, (G/SPCE) = A graphene-modified screen-printed carbon electrode, RT = Room Temperature, (P-FAB) = plasmonic fiber optic absorbance biosensor, NR = Not Reported, N = nucleocapsid, S = spike protein.
Table 5. COVID-19 detection techniques advantages and limitations comparison.
Table 5. COVID-19 detection techniques advantages and limitations comparison.
Categories of Coronavirus DetectionTechniquesAdvantagesLimitation
Indirect detection RNART-PCR
  • ❖ Reliable/detects current viral infection
  • ❖ Rapid results
  • ❖ Higher sensitivity
  • ❖ Needs a slight amount of DNA
  • ❖ It can be performed in a single step
  • ❖ Well established methodology in viral diagnostics
  • ❖ Inability to detect already recovered patients.
  • ❖ Depends on Spiked-in material.
  • ❖ Requires sophisticated instruments.
  • ❖ Low sensitivity.
  • ❖ The need to extract RNA.
  • ❖ Higher costs due to expensive consumables usage.
  • ❖ Complex detection.
Indirect detection RNALAMP
  • ❖ Extremely repeatable
  • ❖ Precise
  • ❖ One working temperature (60–65°)
  • ❖ Rapid reaction
  • ❖ Accurate
  • ❖ Fast amplification
  • ❖ Operation Simplicity
  • ❖ Detection Simplicity
  • ❖ Low versatility
  • ❖ Possibility of primer–primer interactions
  • ❖ Low detection sensitivity.
  • ❖ Long detection time.
  • ❖ The need to extract RNA from clinical samples.
  • ❖ Cannot detect recovered patients.
  • ❖ Too sensitive, highly susceptible to false
  • ❖ positives due to carry-over or cross-contamination
Indirect detection RNANGS
  • ❖ Highly sensitive
  • ❖ Specific
  • ❖ Can identify the novel strain
  • ❖ Slow detection.
  • ❖ The need to extract RNA from clinical samples.
  • ❖ High expertise required
  • ❖ Equipment dependency
  • ❖ Expensive
  • ❖ Requires highly sophisticated Lab
  • ❖ Unable to detect recovered patients.
Indirect detection RNACRISPR
  • ❖ Affordable
  • ❖ High detection speed
  • ❖ High sensitivity
  • ❖ Several CRISPR-based kits are still in development phase
  • ❖ Clinical validation is required
  • ❖ Not yet widespread and in clinical trials
Indirect detection: Spike (S) and Nucleocapsid (N) proteinsElectrochemical sensors
  • ❖ Structure simplicity
  • ❖ High-level sensitivity
  • ❖ Economical
  • ❖ Instant response
  • ❖ Label-Free
  • ❖ High need for kits.
  • ❖ Tiresome sample collection.
  • ❖ Time-consuming process
  • ❖ Requires trained personnel
  • ❖ Requires adequate laboratory infrastructure
Indirect detection: Spike (S) proteinElectronic sensors(FET)
  • ❖ Low cost
  • ❖ High speed
  • ❖ Small size
  • ❖ Excellent compatibility with integrated circuits (ICs).
  • ❖ The signal transduction process found is not always apparent.
  • ❖ Heterogeneous interface structures.
  • ❖ Long time result.
Indirect detection: Antigen sputumPiezoelectric sensor
  • ❖ Rapid
  • ❖ Highly sensitive
  • ❖ Specific
  • ❖ Label-free
  • ❖ Depends on sample preparation
  • ❖ Complex pretreatment steps
Indirect detection: IgM antibody and Nucleocapsid (N) ProteinOptical (fluorescence, SPR, LSPCF, and P-FAB)
  • ❖ Rapid
  • ❖ Sensitive
  • ❖ Specific
  • ❖ Requirement for point of care remains difficult.
  • ❖ High cost.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alathari, M.J.A.; Al Mashhadany, Y.; Mokhtar, M.H.H.; Burham, N.; Bin Zan, M.S.D.; A Bakar, A.A.; Arsad, N. Human Body Performance with COVID-19 Affectation According to Virus Specification Based on Biosensor Techniques. Sensors 2021, 21, 8362. https://doi.org/10.3390/s21248362

AMA Style

Alathari MJA, Al Mashhadany Y, Mokhtar MHH, Burham N, Bin Zan MSD, A Bakar AA, Arsad N. Human Body Performance with COVID-19 Affectation According to Virus Specification Based on Biosensor Techniques. Sensors. 2021; 21(24):8362. https://doi.org/10.3390/s21248362

Chicago/Turabian Style

Alathari, Mohammed Jawad Ahmed, Yousif Al Mashhadany, Mohd Hadri Hafiz Mokhtar, Norhafizah Burham, Mohd Saiful Dzulkefly Bin Zan, Ahmad Ashrif A Bakar, and Norhana Arsad. 2021. "Human Body Performance with COVID-19 Affectation According to Virus Specification Based on Biosensor Techniques" Sensors 21, no. 24: 8362. https://doi.org/10.3390/s21248362

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