Next Article in Journal
Resistance of Tick Gut Microbiome to Anti-Tick Vaccines, Pathogen Infection and Antimicrobial Peptides
Previous Article in Journal
Genetic Diversity and Recombination in the Plant Pathogen Sclerotinia sclerotiorum Detected in Sri Lanka
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Origin, Potential Therapeutic Targets and Treatment for Coronavirus Disease (COVID-19)

1
Department of Biochemistry, Faculty of Science, King Abdulaziz University, Jeddah 21589, Saudi Arabia
2
Department of Microbiology, Abbottabad University of Science and Technology, Abbottabad 22010, Pakistan
3
Department of Genetics, Hazara University Garden Campus, Mansehra 21300, Pakistan
4
School of Biological Sciences, University of the Punjab, Lahore 54000, Pakistan
*
Author to whom correspondence should be addressed.
Pathogens 2020, 9(4), 307; https://doi.org/10.3390/pathogens9040307
Submission received: 5 April 2020 / Revised: 16 April 2020 / Accepted: 19 April 2020 / Published: 22 April 2020
(This article belongs to the Section Human Pathogens)

Abstract

:
The ongoing episode of coronavirus disease 19 (COVID-19) has imposed a serious threat to global health and the world economy. The disease has rapidly acquired a pandemic status affecting almost all populated areas of the planet. The causative agent of COVID-19 is a novel coronavirus known as SARS-CoV-2. The virus has an approximate 30 kb single-stranded positive-sense RNA genome, which is 74.5% to 99% identical to that of SARS-CoV, CoV-pangolin, and the coronavirus the from horseshoe bat. According to available information, SARS-CoV-2 is inferred to be a recombinant virus that originated from bats and was transmitted to humans, possibly using the pangolin as the intermediate host. The interaction of the SARS-CoV-2 spike protein with the human ACE2 (angiotensin-converting enzyme 2) receptor, and its subsequent cleavage by serine protease and fusion, are the main events in the pathophysiology. The serine protease inhibitors, spike protein-based vaccines, or ACE2 blockers may have therapeutic potential in the near future. At present, no vaccine is available against COVID-19. The disease is being treated with antiviral, antimalarial, anti-inflammatory, herbal medicines, and active plasma antibodies. In this context, the present review article provides a cumulative account of the recent information regarding the viral characteristics, potential therapeutic targets, treatment options, and prospective research questions.

1. Introduction

The outbreak of a novel respiratory syndrome, referred to as coronavirus disease 2019 (COVID-19), was first recognized in Wuhan, China, in December 2019. The causative agent for this deadly condition is a coronavirus known as SARS-CoV-2. COVID-19 is demonstrated by fever, dry cough, persistent pressure in the chest, and shortness of breath [1,2]. Sneezing, runny nose, and symptoms similar to the common cold are observed in only 5% of patients. About 2% to 10% of patients have shown diarrhea like symptoms [3,4]. The mortality rate of COVID-19 is 4.5% to 6%, which is less than that of SARS (severe acute respiratory syndrome), which has a mortality rate of 9.6%, and less than that of MERS (Middle East respiratory syndrome), up to 34.4% deaths. Individuals already suffering from cardiovascular disease, hypertension, respiratory disease, or diabetes are at a high risk of mortality. Age and gender-specific variations in the death rate have also been reported [5]. The disease became pandemic within a few months of its emergence, indicating a high transmission ability as compared with SARS and MERS [6,7,8]. COVID-19 lasts up to 6 weeks depending upon the individual’s immunity and the disease intensity. A variable incubation period has been reported for the infection to establish completely; a second exposure to the viral inoculum may decrease the incubation time [7]. An incubation time of 3 to 27 days (average 14 days) has been reported by different sources [6,9,10]. This incubation period is considerably longer than that required by SARS or MERS [11,12]. SARS-CoV-2 has an airborne route of transmission, whereby small aerosols spread in the surrounding air by the coughing and sneezing of infected individuals. These fine airborne droplets, harboring viral particles, can be directly inhaled by nearby healthy individuals [13]. The viral particles can stick to the fingertips and invade the healthy individuals by contact of contaminated hands with the nose, eyes, or mouth. Hence, hand hygiene is an expedient precaution to reduce SARS-CoV-2 transmission [14]. There is no evidence for the sexual transmission of SARS-CoV-2; however, the possibility of fecal transmission has been reported [15,16]. The infection mechanism of both the SARS and COVID-19 viruses involve an interaction with the angiotensin-converting enzyme 2 (ACE2) and cleavage of the viral spike protein by a serine protease [17,18]. Hence, a similar set of therapeutic targets can be the subject of prospective investigations. The present review article provides a cumulative account of recent information on the origin of the virus, its characteristics, and the potential therapeutics for COVID-19.

2. Origin, Transmission and Diagnosis of SARS-CoV-2

Coronaviruses, first discovered in the 1960s, are found in birds and mammals, especially in bats, cats, camels, and rats [19]. The causative agent of COVID-19 (SARS-CoV-2) belongs to the genus β-Coronavirus, family Coronaviridae, and order Nidovirales. A similar human coronavirus was found to be responsible for SARS in 2002 and 2003. The virus responsible for COVID-19 has a single-stranded positive-sense RNA genome of about 30 kb, which has 74% to 99% identity with that of the coronavirus from the pangolin (Manis javanica) and horseshoe bat (Rhinolophus sinicus) (Bat-CoVRaTG13), respectively [2,20,21,22,23]. Bats have been reported as being the rich source of coronaviruses [24,25], although only a few of these coronaviruses can infect humans [26,27]. According to the literature, the SARS and MERS viruses have zoonotic transmission, originating from bats using palm civets and camels, respectively, as the intermediate hosts [28,29,30,31,32]. The recent reports have suggested that SARS-CoV-2 is a modified coronavirus of bat origin [22,32], which came to humans as a result of zoonotic transmission [33,34]. A coronavirus identified from the Malayan pangolin has been shown to have a 99% similarity with SARS-CoV-2. The receptor-binding domain (RBD) of pangolin-CoV has only a one amino acid difference with that of SARS-CoV-2; the infected pangolins exhibit pathological symptoms similar to humans suffering from COVID-19, and their blood circulating antibodies can react with the spike protein of SARS-CoV-2 [35,36]. Although the RaTG13 coronavirus isolated from bat has about 96% identity with SARS-CoV-2, its RBD is different from that of the later, exhibiting a low binding ability to the human ACE2 [37]. However, the RBD of the S-protein from pangolin-CoV is highly similar to that of SARS-CoV-2, six residues critical for receptor binding being identical in both [38]. A comparative analysis of genetic data available to date has suggested that SARS-CoV-2 originated by the recombination of pangolin-CoV and the bat-CoV-RaTG13-like virus [35,39,40,41]. Based on this information, the pangolin is considered to be one of the possible intermediate hosts between bat and human. Snakes, minks, and turtles are also being investigated as the potential intermediate hosts [42,43] (Figure 1). Five out of the six critical amino acid residues comprising the RBD of the S-protein from SARS-CoV and SARS-CoV-2 are different, contradicting the theories about the laboratory origin of SARS-CoV-2 by the manipulation of SARS or MERS like viruses [43]. Studies based on the analysis of N, S, and ORF1a/1b genes have shown conserved sequences suggesting that SARS-CoV-2 is an animal virus, which was transmitted to humans by undergoing evolutionary adaptations [22,44]. The SARS infection from 2003, also involved zoonotic transmission of the virus to humans. Hence, further studies are required to confirm the intermediate hosts of coronaviruses to control zoonotic transmission and avoid the outbreak of such viral infections in the future [28].
On 2 March 2020, WHO published a PCR based detection method for SARS-CoV-2. The procedure could detect the virus in the blood, sputum, and nasopharyngeal swab [45,46]. Noncontrast chest CT (computed tomography) can also be used for the diagnosis of viral pneumonia. However, CT scans can be negative in the case of COVID-19 [47]. On the other hand, patients with negative RT-PCR test results can show pneumonia-like symptoms on a CT scan [48]. In a comparative study, the sensitivity of a chest CT was found to be 98%, whereas the sensitivity of the PCR test was only 71% [49,50]. RT-PCR based diagnosis also gave false-positive results [51]. Low viral load, inefficient sampling, poor sample storage or processing conditions, along with a lack of specific primers due to the high rate of mutations in RNA viruses, are some of the apparent factors for the poor sensitivity of PCR based diagnoses. Recently, some parallel procedures have also been reported for the diagnosis of COVID-19. One of these procedures is loop-mediated isothermal amplification (LAMP), which is a faster single-step procedure, having >95% sensitivity [52,53]. Further modifications of LAMP-based procedures have been reported, which can reduce the testing time with minimum equipment requirements [54,55]. To develop serological procedures, IgA and IgM have been evaluated against SARS-CoV-2 by immunofluorescence assays [56,57,58]. However, further refining of RT-PCR and the serological procedures are required to improve the sensitivity and specificity.

2.1. SARS-CoV-2 vs. SARS-CoV—A Brief Comparison

SARS became epidemic in many countries around the world in 2002 and 2003. The disease had many symptoms similar to those of COVID-19. However, SARS-CoV-2 and SARS-CoV have shown differences, as well as similarities, in their genomic composition, incubation time, and infection mechanisms. A set of affinities has been tabulated that can help us to establish the correlation between the two viruses (Table 1).

2.2. Potential Therapeutics and Treatment for COVID-19

The intra- or inter-species transmission of β-coronaviruses (CoVs) requires a viral interaction with the host cell receptors, and the subsequent invasion of the host cells [85]. The genome of the coronavirus codes for a surface glycoprotein, known as a “spike” protein (S-protein), that specifically binds to the host cellular receptors to initiate the infection process. In fact, the spike protein performs a “key” like function to “unlock” the door and facilitate the cellular entry of a coronavirus. Studies based on 3D models of spike proteins from the SARS-CoV and SARS-CoV-2 viruses, have shown a considerable overall similarity [34,86]. The cryo-EM structure of the SARS-CoV-2 S-protein has also been reported [87] (Figure 2). The overall structure of the S-protein consists of several functional domains. The RBD, fusion domain (FD), and the S2 cleavage site could be critical for future studies to develop therapeutic strategies [87]. The protein exhibits a high binding affinity with ACE2, as represented by a low dissociation constant value (Kd ⁓ 15 nM). The receptor binding affinity of the SARS-CoV-2 S-protein is 10 times higher than that of the SARS-CoV S-protein [37,87,88,89,90]. Furthermore, studies using human, pig, and civet cell lines have allowed SARS-CoV-2 infection and replication, indicating that the virus makes use of the ACE2 receptor for infection [22,91,92,93]. ACE2 is cleaved by a protease (TMPRSS2) in order to activate virus entry. This can be inhibited by protease inhibitors such as camostat mesylate [92]. ACE2 is highly expressed in the lungs; a vast surface area makes the lung tissue highly susceptible to SARS-CoV-2 infection [94]. In addition to the lungs, the ACE2 receptor is also expressed in the endothelial cells of intestine, kidney, and heart cells [95].
According to recent information, the glutamine at the amino acid position 394 in the receptor-binding protein of SARS-CoV-2 that corresponds to the residue 479 in SARS-CoV, is recognized by lysine 31 residue in the human ACE2 receptor [90]. An interaction of polar residues in the ectodomain of ACE2 with the receptor binding domain of SARS-CoV-2 spike protein has been reported [96]. Downstream interaction and the invasion process include the cleavage of the spike by a serine protease at the “S2” domain (Figure 2) [97], followed by the interaction of the virus S-protein fusion domain with the host cell plasma membrane. Virus entry takes place either by fusion with the plasma membrane or by endocytosis, and the subsequent fusion of membranes in endosomes [98]. In addition to virus–plasma membrane fusion, SARS-CoV can adopt clathrin-dependent endocytosis [27]. Once inside the host cell, the viral RNA is translated in the cytoplasm producing polyproteins and structural proteins. After translation, the replication of the genome occurs [99]. New virus particles are formed in the membranes of the Golgi apparatus and the endoplasmic reticulum, after which the vesicles harboring the viral particles are fused with plasma membranes for the release of the virus [100,101]. Because SARS-CoV-2, in binding with ACE2, is dependent on an interaction and the cleavage of the spike protein by a serine protease, a spike protein-based vaccine or serine protease specific inhibitors could be potential therapies against SARS-CoV-2 infection [102]. The comparative homology studies of SARS-CoV-2 proteins (S, N, M, and E proteins) with those from SARS and MERS viruses, have suggested some targets for vaccine development [103]. Polyclonal antibodies raised against SARS-CoV are found to prevent a spike-mediated host cell invasion of SARS-CoV-2 [104]. Recently, 1.3 billion potential protease blockers have been investigated by molecular docking studies. Many of these molecules can be evaluated in wet labs in the near future [105]. ACE2 blockers can be another option to avoid the infection [106]. Similarly, there are some molecules including GSK1838705A, KT203, KT185, and BMS195614 that have strong binding affinities with RBD of the viral S-protein [107]. These molecules can help to control rapid infections by engaging the virus at entry points [107].
Currently, a tremendous amount of research is in progress to develop a vaccine against COVID-19. However, vaccine development is time consuming process, and the newly introduced vaccine will require several safety evaluations [4]. According to estimates, a vaccine against COVID-19 may take more than a year to become available [108]. Even after the preparation of an effective vaccine, under the present pandemic situation, human trials will be a big challenge for researchers. At present COVID-19 is being treated with some broad-spectrum antiviral drugs including remdesivir, favipiravir, and Chinese herbal medicine [101,109]. In vitro studies have shown that chloroquine and remdesivir are effective against SARS-CoV-2 [73,74,75]. Chloroquine phosphate has shown treatment efficacy and safety against SARS-CoV-2 associated pneumonia; these findings are based on multiple trials in hospitals in China [76]. Application of an anti-inflammatory drug such as baricitinib, together with an antiviral drug, has also been recommended to treat COVID-19 [110]. High doses of ascorbic acid (vitamin C) are suggested for the prevention of the COVID-19 disease [111]. Type I interferon can inhibit viral replication. According to studies, interferon β could inhibit the replication of SARS-CoV [95,112]. However, its efficacy against SARS-CoV-2 needs further investigation. The use of convalescent plasma for the treatment of COVID-19 has been suggested. However, the absence of multiple trials on a large scale, and the concern that antibodies may demonstrate donor dependent titters and specificities, are the main deficits of convalescent plasma therapy [113].

3. Prospective Challenges and Research Questions

There are several challenges in the management and control of coronaviruses. A wide range of coronaviruses with a highly mutable single-stranded RNA genome are found [20,21] in many mammalian and avian sources [19] that closely interact with each other, as well as with humans. The long-term viability of coronaviruses in airborne aerosols, and on daily utensils composed of plastics, stainless steel, and cardboard, enhance the chances of transmission of infection between individuals and species [8,81,114]. The recurrence of SARS-CoV-2 has been reported in convalescence times [115], which can make the treatment more difficult and increase the chance of complications. The interventions into the ACE2 binding abilities of SARS-CoV-2 and similar viruses, by the inhibition of the corresponding serine protease [106], can be an area of investigation. Easy, highly reliable, and early-stage diagnosis procedures are still required as a challenge for biomedical researchers [52,53]. The presence of SARS-CoV-2 in stool samples of infected individuals raises the question about the fecal-oral transmission of the disease [10]. Recently, the viability of SARS-CoV and SARS-CoV-2 has been described (115); however, the genetic factors behind the long-term survival of these coronaviruses need further investigation [81]. SARS-CoV has shown low stability at higher temperatures and at specific air humidity levels [9,116]; the effect of temperature and other environmental factors on the viability of SARS-CoV-2 are unclear. Inactivation of coronaviruses by disinfectants, such as 60% to 70% ethanol or 0.1% sodium hypochlorite, is well established [117]. The efficacy and specificity of antiviral and antimalarial drugs being used in the treatment of COVID-19 need further clinical trials.

4. Conclusions

The ongoing COVID-19 outbreak that emerged from Wuhan, China, has acquired pandemic status. The causative agent of COVID-19 is a modified coronavirus, known as SARS-CoV-2 that has similarities with the coronaviruses responsible for SARS, MERS, and those identified as coming from various animals including the pangolin and bat. SARS-CoV-2 specific vaccine development, and the application of highly specific antiviral medicines, require time and investigation. The health management authorities have a major focus on known preventive measures for viral infections. The situation demands keen surveillance, and the development of early diagnostic and better treatment options. The present report provides an insight into the characteristics of the pathogen, its mode of infection, its potential targets, along with future research questions.

Author Contributions

M.S.N., M.A.Z., and H.C. conceived the idea and explored the literature; M.S.N., I.K., and B.N.M. explored the recent literature and prepared the manuscript; H.C., A.R.S., H.A., and M.A.Z. also contributed to the data analysis and the preparation of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

No funds were provided for this article.

Acknowledgments

The authors are grateful to the doctors, medical staff, and volunteers contributing to overcome the coronavirus disease (COVID-19).

Conflicts of Interest

The authors have no conflict of interest.

References

  1. Huang, C.; Wang, Y.; Li, X.; Ren, L.; Zhao, J.; Hu, Y.; Zhang, L.; Fan, G.; Xu, J.; Gu, X.; et al. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. Lancet 2020, 395, 497–506. [Google Scholar] [CrossRef] [Green Version]
  2. Chinese Centre for Disease Control and Prevention (CCDC). The Epidemiological Characteristics of an Outbreak of 2019 Novel Coronavirus Diseases (COVID-19)–China. 2020. Available online: http://weekly.chinacdc.cn/en/article/id/e53946e2-c6c4-41e9-9a9b-fea8db1a8f51 (accessed on 17 February 2020).
  3. Wang, C.; Horby, P.W.; Hayden, F.G.; Gao, G.F. A novel coronavirus outbreak of global health concern. Lancet 2020, 395, 470–473. [Google Scholar] [CrossRef] [Green Version]
  4. Chen, D.; Xu, W.; Lei, Z.; Huang, Z.; Liu, J.; Gao, Z.; Peng, L. Recurrence of positive SARS-CoV-2 RNA in COVID-19: A case report. Int. J. Infect. Dis. 2020, 93, 297–299. [Google Scholar] [CrossRef] [PubMed]
  5. WHO. Director-General’s Opening Remarks at the Media Briefing on COVID-19—3 March 2020; World Health Organization: Geneva, Switzerland, 3 March 2020. [Google Scholar]
  6. WHO. Statement on the Second Meeting of the International Health Regulations Emergency Committee Regarding the Outbreak of Novel Coronavirus (2019-nCoV); World Health Organization: Geneva, Switzerland, 30 January 2020; Available online: https://www.who.int/news-room/detail/30-01-2020-statement-on-the-second-meeting-of-the-international-health-regulations-(2005)-emergency-committee-regarding-the-outbreak-of-novel-coronavirus-(2019-ncov) (accessed on 3 March 2020).
  7. Munster, V.J.; Koopmans, M.; van Doremalen, N.; van Riel, D.; de Wit, E. A novel coronavirus emerging in China—key questions for impact assessment. N. Engl. J. Med. 2020, 382, 692–694. [Google Scholar] [CrossRef] [PubMed]
  8. Guan, W.J.; Ni, Z.Y.; Hu, Y.; Liang, W.H.; Ou, C.Q.; He, J.X.; Liu, L.; Shan, H.; Lei, C.L.; Hui, D.S.; et al. Clinical characteristics of 2019 novel coronavirus infection in China. N. Engl. J. Med. 2020. [Google Scholar] [CrossRef] [PubMed]
  9. Bai, Y.; Yao, L.; Wei, T.; Tian, F.; Jin, D.Y.; Chen, L.; Wang, M. Presumed asymptomatic carrier transmission of COVID-19. JAMA 2020, 323, 1406–1407. [Google Scholar] [CrossRef] [Green Version]
  10. Lauer, S.A.; Grantz, K.H.; Bi, Q.; Jones, F.K.; Zheng, Q.; Meredith, H.R.; Azman, A.S.; Reich, N.G.; Lessler, J. The incubation period of coronavirus disease 2019 (COVID-19) from publicly reported confirmed cases: Estimation and application. Ann. Intern. Med. 2020. [Google Scholar] [CrossRef] [Green Version]
  11. Lessler, J.; Reich, N.G.; Brookmeyer, R.; Perl, T.M.; Nelson, K.E.; Cummings, D.A. Incubation periods of acute respiratory viral infections: A systematic review. Lancet Infec. Dis. 2009, 9, 291–300. [Google Scholar] [CrossRef]
  12. Backer, J.A.; Klinkenberg, D.; Wallinga, J. Incubation period of 2019 novel coronavirus (2019-nCoV) infections among travellers from Wuhan, China, 20–28 January 2020. Eur. Surveill. 2020, 25, 2000062. [Google Scholar] [CrossRef] [Green Version]
  13. Liu, Y.; Ning, Z.; Chen, Y.; Guo, M.; Liu, Y.; Gali, N.K.; Sun, L.; Duan, Y.; Cai, J.; Westerdahl, D.; et al. Aerodynamic Characteristics and RNA Concentration of SARS-CoV-2 Aerosol in Wuhan Hospitals during COVID-19 Outbreak. BioRxiv 2020. [Google Scholar] [CrossRef]
  14. Yang, C. Does hand hygiene reduce SARS-CoV-2 transmission? Graefe’s Arch. Clin. Exp. Ophthalmol. 2020, 1–2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Cui, P.; Chen, Z.; Wang, T.; Dai, J.; Zhang, J.; Ding, T.; Jiang, J.; Liu, J.; Zhang, C.; Shan, W.; et al. Clinical features and sexual transmission potential of SARS-CoV-2 infected female patients: A descriptive study in Wuhan, China. MedRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  16. Yeo, D. Enteric involvement of coronaviruses: Is faecal–oral transmission of SARS-CoV-2 possible? Lancet Gastroenterol. Hepatol. 2020, 5, 335–337. [Google Scholar] [CrossRef] [Green Version]
  17. Zhao, Y.; Zhao, Z.; Wang, Y.; Zhou, Y.; Ma, Y.; Zuo, W. Single-cell RNA expression profiling of ACE2, the putative receptor of Wuhan 2019-nCov. BioRxiv 2020. [Google Scholar] [CrossRef]
  18. Prabakaran, P.; Xiao, X.; Dimitrov, D.S. A model of the ACE2 structure and function as a SARS-CoV receptor. Biochem. Biophys. Res. Commun. 2004, 314, 235–241. [Google Scholar] [CrossRef] [PubMed]
  19. Woo, P.C.; Lau, S.K.; Lam, C.S.; Lau, C.C.; Tsang, A.K.; Lau, J.H.; Zheng, B.J. Discovery of seven novel Mammalian and avian coronaviruses in the genus delta-coronavirus supports bat coronaviruses as the gene source of alpha-coronavirus and betacoronavirus and avian coronaviruses as the gene source of gamma-coronavirus and delta-coronavirus. J. Virol. 2012, 86, 3995–4008. [Google Scholar] [CrossRef] [Green Version]
  20. Zhu, N.; Zhang, D.; Wang, W.; Li, X.; Yang, B.; Song, J.; Zhao, X.; Huang, B.; Shi, W.; Lu, R.; et al. A novel coronavirus from patients with pneumonia in China, 2019. N. Engl. J. Med. 2020, 382, 727–733. [Google Scholar] [CrossRef]
  21. Lai, C.C.; Shih, T.P.; Ko, W.C.; Tang, H.J.; Hsueh, P.R. Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) and corona virus disease-2019 (COVID-19): The epidemic and the challenges. Int. J. Antimicrob. Agents 2020, 105924. [Google Scholar] [CrossRef]
  22. Zhou, P.; Yang, X.L.; Wang, X.G.; Hu, B.; Zhang, L.; Zhang, W.; Si, H.R.; Zhu, Y.; Li, B.; Huang, C.L.; et al. A pneumonia outbreak associated with a new coronavirus of probable bat origin. Nature 2020, 579, 270–273. [Google Scholar] [CrossRef] [Green Version]
  23. Chen, L.; Liu, W.; Zhang, Q.; Xu, K.; Ye, G.; Wu, W.; Sun, Z.; Liu, F.; Wu, K.; Zhong, B.; et al. RNA based mNGS approach identifies a novel human coronavirus from two individual pneumonia cases in 2019 Wuhan outbreak. Emerg. Microb. Infect. 2020, 313–319. [Google Scholar] [CrossRef]
  24. Yang, L.; Wu, Z.; Ren, X.; Yang, F.; He, G.; Zhang, J.; Dong, J.; Sun, L.; Zhu, Y.; Du, J.; et al. Novel SARS-like betacoronaviruses in bats, China, 2011. Emerg. Infect. Dis. 2013, 19, 989–991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Hu, B.; Zeng, L.P.; Yang, X.L.; Ge, X.Y.; Zhang, W.; Li, B.; Xie, J.Z.; Shen, X.R.; Zhang, Y.Z.; Wang, N.; et al. Discovery of a rich gene pool of bat SARS-related coronaviruses provides new insights into the origin of SARS coronavirus. PLoS Pathogen. 2017, 11, e1006698. [Google Scholar] [CrossRef] [PubMed]
  26. Menachery, V.D.; Yount, B.L.; Sims, A.C.; Debbink, K.; Agnihothram, S.S.; Gralinskim, L.E.; Graham, R.L.; Scobey, T.; Plantem, J.A.; Royalm, S.R.; et al. SARS-like WIV1-CoV poised for human emergence. Proc. Natl. Acad. Sci. USA 2016, 113, 3048–3053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Wang, N.; Li, S.Y.; Yang, X.L.; Huang, H.M.; Zhang, Y.J.; Guo, H.; Luo, C.M.; Miller, M.; Zhu, G.; Chmura, A.A.; et al. Serological evidence of bat SARS-related coronavirus infection in humans, China. Virol. Sin. 2018, 33, 104–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Li, Q.; Guan, X.; Wu, P.; Wang, X.; Zhou, L.; Tong, Y.; Ren, R.; Leung, K.S.; Lau, E.H.; Wong, J.Y.; et al. Early transmission dynamics in Wuhan, China, of novel coronavirus–infected pneumonia. N. Engl. J. Med. 2020, 382, 1199–1207. [Google Scholar] [CrossRef] [PubMed]
  29. Song, H.D.; Tu, C.C.; Zhang, G.W.; Wang, S.Y.; Zheng, K.; Lei, L.C.; Chen, Q.X.; Gao, Y.W.; Zhou, H.Q.; Xiang, H.; et al. Cross-host evolution of severe acute respiratory syndrome coronavirus in palm civet and human. Proc. Natl. Acad. Sci. USA 2005, 102, 2430–2435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Lau, S.K.; Li, K.S.; Huang, Y.; Shek, C.T.; Tse, H.; Wang, M.; Choi, G.K.; Xu, H.; Lam, C.S.; Guo, R.; et al. Ecoepidemiology and complete genome comparison of different strains of severe acute respiratory syndrome-related Rhinolophus bat coronavirus in China reveal bats as a reservoir for acute, self-limiting infection that allows recombination events. J. Virol. 2010, 84, 2808–2819. [Google Scholar] [CrossRef] [Green Version]
  31. Chu, D.K.; Poon, L.L.; Gomaa, M.M.; Shehata, M.M.; Perera, R.A.; Zeid, D.A.; El Rifay, A.S.; Siu, L.Y.; Guan, Y.; Webby, R.J.; et al. MERS coronaviruses in dromedary camels, Egypt. Emerg. Infect. Dis. 2014, 20, 1049–1053. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, L.; Shen, F.M.; Chen, F.; Lin, Z. Origin and evolution of the 2019 novel coronavirus. Clin. Infect. Dis. 2020. [Google Scholar] [CrossRef] [Green Version]
  33. Lu, R.; Zhao, X.; Li, J.; Niu, P.; Yang, B.; Wu, H.; Wang, W.; Song, H.; Huang, B.; Zhu, N.; et al. Genomic characterisation and epidemiology of 2019 novel coronavirus: Implications for virus origins and receptor binding. Lancet 2020, 395, 565–574. [Google Scholar] [CrossRef] [Green Version]
  34. Xu, X.; Chen, P.; Wang, J.; Feng, J.; Zhou, H.; Li, X.; Zhong, W.; Hao, P. Evolution of the novel coronavirus from the ongoing Wuhan outbreak and modelling of its Spike protein for risk of human transmission. Sci. China Life Sci. 2020, 63, 457–460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Xiao, K.; Zhai, J.; Feng, Y.; Zhou, N.; Zhang, X.; Zou, J.J.; Li, N.; Guo, Y.; Li, X.; Shen, X.; et al. Isolation and Characterization of 2019-nCoV-like Coronavirus from Malayan Pangolins. BioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, H.; Penninger, J.M.; Li, Y.; Zhong, N.; Slutsky, A.S. Angiotensin-converting enzyme 2 (ACE2) as a SARS-CoV-2 receptor: Molecular mechanisms and potential therapeutic target. Intensive Care Med. 2020, 3, 1–5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Wan, Y.; Shang, J.; Graham, R.; Baric, R.S.; Li, F. Receptor recognition by novel coronavirus from Wuhan: An analysis based on decade-long structural studies of SARS. J. Virol. 2020. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, Z.; Wu, Q.; Zhang, T. Pangolin homology associated with 2019-nCoV. bioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  39. Liu, P.; Chen, W.; Chen, J.P. Viral metagenomics revealed sendai virus and coronavirus infection of Malayan pangolins (Manis javanica). Viruses 2019, 11, 979. [Google Scholar] [CrossRef] [Green Version]
  40. Cui, J.; Li, F.; Shi, Z.L. Origin and evolution of pathogenic coronaviruses. Nat. Rev. Microbiol. 2019, 17, 181–192. [Google Scholar] [CrossRef] [Green Version]
  41. Li, X.; Zai, J.; Zhao, Q.; Nie, Q.; Li, Y.; Foley, B.T.; Chaillon, A. Evolutionary history, potential intermediate animal host, and cross-species analyses of SARS-CoV-2. J. Med. Virol. 2020. [Google Scholar] [CrossRef]
  42. Liu, Z.; Xiao, X.; Wei, X.; Li, J.; Yang, J.; Tan, H.; Zhu, J.; Zhang, Q.; Wu, J.; Liu, L. Composition and divergence of coronavirus spike proteins and host ACE2 receptors predict potential intermediate hosts of SARS-CoV-2. J. Med. Virol. 2020. [Google Scholar] [CrossRef] [Green Version]
  43. Andersen, K.G.; Rambaut, A.; Lipkin, W.I.; Holmes, E.C.; Garry, R.F. The proximal origin of SARS-CoV-2. Nat. Med. 2020, 26, 450–452. [Google Scholar] [CrossRef] [Green Version]
  44. Chen, N.; Zhou, M.; Dong, X.; Qu, J.; Gong, F.; Han, Y.; Qiu, Y.; Wang, J.; Liu, Y.; Wei, Y.; et al. Epidemiological and clinical characteristics of 99 cases of 2019 novel coronavirus pneumonia in Wuhan, China: A descriptive study. Lancet 2020, 395, 507–513. [Google Scholar] [CrossRef] [Green Version]
  45. WHO. Detection of 2019 novel coronavirus (2019-nCoV) in suspected human cases by RT-PCR. 2020. Available online: https://www.who.int/health-topics/coronavirus/laboratory-diagnostics-for-novel-coronavirus (accessed on 20 March 2020).
  46. Zou, L.; Ruan, F.; Huang, M.; Liang, L.; Huang, H.; Hong, Z.; Yu, J.; Kang, M.; Song, Y.; Xia, J.; et al. SARS-CoV-2 viral load in upper respiratory specimens of infected patients. N. Engl. J. Med. 2020, 382, 1177–1179. [Google Scholar] [CrossRef] [PubMed]
  47. Chung, M.; Bernheim, A.; Mei, X.; Zhang, N.; Huang, M.; Zeng, X.; Cui, J.; Xu, W.; Yang, Y.; Fayad, Z.A.; et al. CT imaging features of 2019 novel coronavirus (2019-nCoV). Radiology 2020, 295, 202–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Xie, X.; Zhong, Z.; Zhao, W.; Zheng, C.; Wang, F.; Liu, J. Chest CT for typical 2019-nCoV pneumonia: Relationship to negative RT-PCR testing. Radiology 2020, 200343. [Google Scholar] [CrossRef] [Green Version]
  49. Fang, Y.; Zhang, H.; Xie, J.; Lin, M.; Ying, L.; Pang, P.; Ji, W. Sensitivity of chest CT for COVID-19: Comparison to RT-PCR. Radiology 2020, 200432. [Google Scholar] [CrossRef]
  50. Ai, T.; Yang, Z.; Hou, H.; Zhan, C.; Chen, C.; Lv, W.; Tao, Q.; Sun, Z.; Xia, L. Correlation of chest CT and RT-PCR testing in coronavirus disease 2019 (COVID-19) in China: A report of 1014 cases. Radiology 2020, 200642. [Google Scholar] [CrossRef] [Green Version]
  51. Wang, Y.; Kang, H.; Liu, X.; Tong, Z. Combination of RT-qPCR Testing and Clinical Features for Diagnosis of COVID-19 facilitates management of SARS-CoV-2 Outbreak. J. Med. Virol. 2020. [Google Scholar] [CrossRef] [Green Version]
  52. Nguyen, T.; Duong Bang, D.; Wolff, A. 2019 Novel Coronavirus Disease (COVID-19): Paving the Road for Rapid Detection and Point-of-Care Diagnostics. Micromachines 2020, 11, 306. [Google Scholar] [CrossRef] [Green Version]
  53. Yu, L.; Wu, S.; Hao, X.; Li, X.; Liu, X.; Ye, S.; Han, H.; Dong, X.; Li, X.; Li, J.; et al. Rapid colorimetric detection of COVID-19 coronavirus using a reverse tran-scriptional loop-mediated isothermal amplification (RT-LAMP) diagnostic plat-form: iLACO. medRxiv 2020. [Google Scholar] [CrossRef]
  54. El-Tholoth, M.; Bau, H.H.; Song, J. A Single and Two-Stage, Closed-Tube, Molecular Test for the 2019 Novel Coronavirus (COVID-19) at Home, Clinic, and Points of Entry. ChemRxiv 2020. [Google Scholar] [CrossRef]
  55. Zhu, X.; Wang, X.; Han, L.; Chen, T.; Wang, L.; Li, H.; Li, S.; He, L.; Fu, X.; Chen, S.; et al. Reverse transcription loop-mediated isothermal amplification combined with nanoparticles-based biosensor for diagnosis of COVID-19. MedRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  56. Haveri, A.; Smura, T.; Kuivanen, S.; Österlund, P.; Hepojoki, J.; Ikonen, N.; Pitkäpaasi, M.; Blomqvist, S.; Rönkkö, E.; Kantele, A.; et al. Serological and molecular findings during SARS-CoV-2 infection: The first case study in Finland, January to February 2020. Eur. Surveill. 2020, 25, 2000266. [Google Scholar] [CrossRef] [PubMed]
  57. Li, X.; Geng, M.; Peng, Y.; Meng, L.; Lu, S. Molecular immune pathogenesis and diagnosis of COVID-19. J. Pharm. Anal. 2020. [Google Scholar] [CrossRef] [PubMed]
  58. Lee, N.Y.; Li, C.W.; Tsai, H.P.; Chen, P.L.; Syue, L.S.; Li, M.C.; Tsai, C.S.; Lo, C.L.; Hsueh, P.R.; Ko, W.C. A case of COVID-19 and pneumonia returning from Macau in Taiwan: Clinical course and anti-SARS-CoV-2 IgG dynamic. J. Microbiol. Immunol. Infect. 2020. [Google Scholar] [CrossRef] [PubMed]
  59. Cheng, V.C.; Lau, S.K.; Woo, P.C.; Yuen, K.Y. Severe acute respiratory syndrome coronavirus as an agent of emerging and reemerging infection. Clin. Microbiol. Rev. 2007, 20, 660–694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Wang, D.; Hu, B.; Hu, C.; Zhu, F.; Liu, X.; Zhang, J.; Wang, B.; Xiang, H.; Cheng, Z.; Xiong, Y.; et al. Clinical characteristics of 138 hospitalized patients with 2019 novel coronavirus–infected pneumonia in Wuhan, China. JAMA 2020. [Google Scholar] [CrossRef] [PubMed]
  61. Yang, X.; Yu, Y.; Xu, J.; Shu, H.; Xia, J.; Liu, H.; Wu, Y.; Zhang, L.; Yu, Z.; Fang, M.; et al. Clinical course and outcomes of critically ill patients with SARS-CoV-2 pneumonia in Wuhan, China: A single-centered, retrospective, observational study. Lancet Resp. Med. 2020. [Google Scholar] [CrossRef] [Green Version]
  62. Nuttall, I.; Dye, C. The SARS wake-up call. Science 2013, 339, 1287–1288. [Google Scholar] [CrossRef]
  63. Jiang, X.; Rayner, S.; Luo, M.H. Does SARS-CoV-2 has a longer incubation period than SARS and MERS? J. Med. Virol. 2020, 92, 476–478. [Google Scholar] [CrossRef]
  64. Chan-Yeung, M.; Xu, R.H. SARS: Epidemiology. Respirology 2003, 8, S9–S14. [Google Scholar] [CrossRef]
  65. Wang, M.; Yan, M.; Xu, H.; Liang, W.; Kan, B.; Zheng, B.; Chen, H.; Zheng, H.; Xu, Y.; Zhang, E.; et al. SARS-CoV infection in a restaurant from palm civet. Emerg. Infect. Dis. 2005, 11, 1860–1865. [Google Scholar] [CrossRef] [PubMed]
  66. Yuan, J.; Hon, C.C.; Li, Y.; Wang, D.; Xu, G.; Zhang, H.; Zhou, P.; Poon, L.L.; Lam, T.T.; Leung, F.C.; et al. Intraspecies diversity of SARS-like coronaviruses in Rhinolophus sinicus and its implications for the origin of SARS coronaviruses in humans. J. Gen. Virol. 2010, 91, 1058–1062. [Google Scholar] [CrossRef] [Green Version]
  67. Ji, W.; Wang, W.; Zhao, X.; Zai, J. Homologous recombination within the spike glycoprotein of the newly identified coronavirus may boost cross-species transmission from snake to human. J. Med. Virol. 2020, 92, 433–440. [Google Scholar] [CrossRef]
  68. Guo, Q.; Li, M.; Wang, C.; Fang, Z.; Tan, J.; Wu, S.; Xiao, Y.; Zhu, H. Host and infectivity prediction of Wuhan 2019 novel coronavirus using deep learning algorithm. BioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  69. Bhadra, S.; Jiang, Y.S.; Kumar, M.R.; Johnson, R.F.; Hensley, L.E.; Ellington, A.D. Real-time sequence-validated loop-mediated isothermal amplification assays for detection of Middle East respiratory syndrome coronavirus (MERS-CoV). PLoS ONE 2015, 10, e0123126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Chan, J.F.W.; Choi, G.K.Y.; Tsang, A.K.L.; Tee, K.M.; Lam, H.Y.; Yip, C.C.; To, K.K.; Cheng, V.C.; Yeung, M.L.; Lau, S.K.; et al. Development and evaluation of novel real-time reverse transcription-PCR assays with locked nucleic acid probes targeting leader sequences of human-pathogenic coronaviruses. J. Clin. Microbiol. 2015, 53, 2722–2726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Rodriguez-Morales, A.; Tiwari, R.; Sah, R.; Dhama, K. COVID-19, an Emerging Coronavirus Infection: Current Scenario and Recent Developments-An Overview. J. Pure Appl. Microbiol. 2020, 14, 6150. [Google Scholar] [CrossRef] [Green Version]
  72. Tai, D.Y. Pharmacologic treatment of SARS: Current knowledge and recommendations. Ann. Acad. Med. Singapore 2007, 36, 438. [Google Scholar]
  73. Wang, M.; Cao, R.; Zhang, L.; Yang, X.; Liu, J.; Xu, M.; Shi, Z.; Hu, Z.; Zhong, W.; Xiao, G. Remdesivir and chloroquine effectively inhibit the recently emerged novel coronavirus (2019-nCoV) in vitro. Cell Res. 2020, 30, 269–271. [Google Scholar] [CrossRef]
  74. Touret, F.; de Lamballerie, X. Of chloroquine and COVID-19. Antiv. Res. 2020, 104762. [Google Scholar] [CrossRef]
  75. Colson, P.; Rolain, J.M.; Lagier, J.C.; Brouqui, P.; Raoult, D. Chloroquine and hydroxychloroquine as available weapons to fight COVID-19. Int. J. Antimicrob. Agents 2020, 105932. [Google Scholar] [CrossRef] [PubMed]
  76. Gao, J.; Tian, Z.; Yang, X. Breakthrough: Chloroquine phosphate has shown apparent efficacy in treatment of COVID-19 associated pneumonia in clinical studies. Biosci. Trends 2020, 14, 72–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Wang, M.; Jiang, A.; Gong, L.; Luo, L.; Guo, W.; Li, C.; Zheng, J.; Li, C.; Yang, B.; Zeng, J.; et al. Temperature significant change COVID-19 Transmission in 429 cities. medRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  78. Battegay, M.; Kuehl, R.; Tschudin-Sutter, S.; Hirsch, H.H.; Widmer, A.F.; Neher, R.A. 2019-Novel coronavirus (2019-nCoV): Estimating the case fatality rate: A word of caution. Swiss Med. Wkly. 2020, 150, w20203. [Google Scholar] [CrossRef] [PubMed]
  79. Wu, Z.; McGoogan, J.M. Characteristics of and important lessons from the coronavirus disease 2019 (COVID-19) outbreak in China: Summary of a report of 72 314 cases from the Chinese Center for Disease Control and Prevention. JAMA 2020. [Google Scholar] [CrossRef] [PubMed]
  80. Colizza, V.; Barrat, A.; Barthélemy, M.; Vespignani, A. Predictability and epidemic pathways in global outbreaks of infectious diseases: The SARS case study. BMC Med. 2007, 5, 34. [Google Scholar] [CrossRef] [PubMed]
  81. 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]
  82. Peiris, J.S.M.; Chu, C.M.; Cheng, V.C.; Chan, K.S.; Hung, I.F.; Poon, L.L.; Law, K.I.; Tang, B.S.; Hon, T.Y.; Chan, C.S.; et al. SARS Study Group Clinical progression and viral load in a community outbreak of coronavirus-associated SARS pneumonia: A prospective study. Lancet 2003, 361, 1767–1772. [Google Scholar] [CrossRef] [Green Version]
  83. Tsang, T.K.; Cowling, B.J.; Fang, V.J.; Chan, K.H.; Ip, D.K.; Leung, G.M.; Peiris, J.S.; Cauchemez, S. Influenza A virus shedding and infectivity in households. J. Infect. Dis. 2015, 212, 1420–1428. [Google Scholar] [CrossRef] [Green Version]
  84. Rothe, C.; Schunk, M.; Sothmann, P.; Bretzel, G.; Froeschl, G.; Wallrauch, C.; Zimmer, T.; Thiel, V.; Janke, C.; Guggemos, W.; et al. Transmission of 2019-nCoV infection from an asymptomatic contact in Germany. N. Engl. J. Med. 2020, 382, 970–971. [Google Scholar] [CrossRef] [Green Version]
  85. Li, F. Structure, function, and evolution of coronavirus spike proteins. Ann. Rev. Virol. 2016, 3, 237–261. [Google Scholar] [CrossRef] [Green Version]
  86. Li, F.; Li, W.; Farzan, M.; Harrison, S.C. Structure of SARS coronavirus spike receptor-binding domain complexed with receptor. Science 2005, 309, 1864–1868. [Google Scholar] [CrossRef]
  87. Wrapp, D.; Wang, N.; Corbett, K.S.; Goldsmith, J.A.; Hsieh, C.L.; Abiona, O.; Graham, B.S.; McLellan, J.S. Cryo-EM structure of the 2019-nCoV spike in the prefusion conformation. Science 2020, 367, 1260–1263. [Google Scholar] [CrossRef] [Green Version]
  88. Peiris, J.; Guan, Y.; Yuen, K. Severe acute respiratory syndrome. Nat. Med. 2004, 10, S88–S97. [Google Scholar] [CrossRef]
  89. Kirchdoerfer, R.N.; Wang, N.; Pallesen, J.; Wrapp, D.; Turner, H.L.; Cottrell, C.A.; Corbett, K.S.; Graham, B.S.; McLellan, J.S.; Ward, A.B. Stabilized coronavirus spikes are resistant to conformational changes induced by receptor recognition or proteolysis. Sci. Rep. 2018, 8, 1–11. [Google Scholar] [CrossRef] [Green Version]
  90. Wu, K.L.; Peng, G.Q.; Wilken, M.; Geraghty, R.J.; Li, F. Mechanisms of host receptor adaptation by severe acute respiratory syndrome coronavirus. J. Biol. Chem. 2012, 287, 8904–8911. [Google Scholar] [CrossRef] [Green Version]
  91. Zhu, Y.; Yu, D.; Yan, H.; Chong, H.; He, Y. Design of potent membrane fusion inhibitors against SARS-CoV-2, an emerging coronavirus with high fusogenic activity. BioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  92. Hoffmann, M.; Kleine-Weber, H.; Krüger, N.; Mueller, M.A.; Drosten, C.; Pöhlmann, S. The novel coronavirus 2019 (COVID-19) uses the SARS-1 coronavirus receptor ACE2 and the cellular protease TMPRSS2 for entry into target cells. BioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  93. Letko, M.; Marzi, A.; Munster, V. Functional assessment of cell entry and receptor usage for SARS-CoV-2 and other lineage B betacoronaviruses. Nat. Microbiol. 2020, 5, 562–569. [Google Scholar] [CrossRef] [Green Version]
  94. Hamming, I.; Timens, W.; Bulthuis, M.L.C.; Lely, A.T.; Navis, G.; van Goor, H. Tissue distribution of ACE2 protein, the functional receptor for SARS coronavirus: A first step in understanding SARS pathogenesis. J. Pathol. 2004, 203, 631–663. [Google Scholar] [CrossRef]
  95. Zhang, L.; Liu, Y. Potential interventions for novel coronavirus in China: A systematic review. J. Med. Virol. 2020, 92, 479–490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Yan, R.; Zhang, Y.; Li, Y.; Xia, L.; Guo, Y.; Zhou, Q. Structural basis for the recognition of the SARS-CoV-2 by full-length human ACE2. Science 2020, 367, 1444–1448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Bertram, S.; Glowacka, I.; Müller, M.A.; Lavender, H.; Gnirss, K.; Nehlmeier, I.; Niemeyer, D.; He, Y.; Simmons, G.; Drosten, C.; et al. Cleavage and activation of the severe acute respiratory syndrome coronavirus spike protein by human airway trypsin-like protease. J. Virol. 2011, 85, 13363–13372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Millet, J.K.; Whittaker, G.R. Host cell entry of Middle East respiratory syndrome coronavirus after two-step, furin-mediated activation of the spike protein. Proc. Natl. Acad. Sci. USA 2014, 111, 15214–15219. [Google Scholar] [CrossRef] [Green Version]
  99. Perlman, S.; Netland, J. Coronaviruses post-SARS: Update on replication and pathogenesis. Nat. Rev. Microbiol. 2009, 7, 439–450. [Google Scholar] [CrossRef] [Green Version]
  100. De Wit, E.; van Doremalen, N.; Falzarano, D.; Munster, V.J. SARS and MERS: Recent insights into emerging coronaviruses. Nat. Rev. Microbiol. 2016, 14, 523–534. [Google Scholar] [CrossRef]
  101. Li, G.; De Clercq, E. Therapeutic options for the 2019 novel coronavirus (2019-nCoV). Nat. Rev. Drug Discov. 2020, 19, 149–150. [Google Scholar] [CrossRef] [Green Version]
  102. Gurwitz, D. Angiotensin receptor blockers as tentative SARS-CoV-2 therapeutics. Drug Dev. Res. 2020. [Google Scholar] [CrossRef] [Green Version]
  103. Ahmed, S.F.; Quadeer, A.A.; McKay, M.R. Preliminary identification of potential vaccine targets for the COVID-19 coronavirus (SARS-CoV-2) based on SARS-CoV immunological studies. Viruses 2020, 12, 254. [Google Scholar] [CrossRef] [Green Version]
  104. Walls, A.C.; Park, Y.J.; Tortorici, M.A.; Wall, A.; McGuire, A.T.; Veesler, D. Structure, function, and antigenicity of the SARS-CoV-2 spike glycoprotein. Cell 2020. [Google Scholar] [CrossRef]
  105. Beigel, J.H.; Nam, H.H.; Adams, P.L.; Krafft, A.; Ince, W.L.; El-Kamary, S.S.; Sims, A.C. Advances in respiratory virus therapeutics - A meeting report from the 6th isirv Antiviral Group conference. Antivir. Res. 2019, 167, 45–67. [Google Scholar] [CrossRef]
  106. Ton, A.T.; Gentile, F.; Hsing, M.; Ban, F.; Cherkasov, A. Rapid Identification of Potential Inhibitors of SARS-CoV-2 Main Protease by Deep Docking of 1.3 Billion Compounds. Mol. Inform. 2020, 10. [Google Scholar] [CrossRef] [Green Version]
  107. Choudhary, S.; Malik, Y.S.; Tomar, S.; Tomar, S. Identification of SARS-CoV-2 Cell Entry Inhibitors by Drug Repurposing Using in Silico Structure-Based Virtual Screening Approach. ChemRxiv 2020. [Google Scholar] [CrossRef]
  108. Huaxia. WHO says vaccines against novel coronavirus 18 months away, pushes global research. Xinhuanet. 2020. Available online: http://www.xinhuanet.com/english/2020-02/12/c_138777886.htm (accessed on 15 March 2020).
  109. Luo, C.H.; Tang, Q.L.; Shang, Y.X.; Liang, S.B.; Yang, M.; Robinson, N.; Liu, J.P. Can Chinese medicine be used for prevention of Corona virus disease 2019 (COVID-19)? A review of historical classics, research evidence and current prevention programs. Chin. J. Integr. Med. 2020. [Google Scholar] [CrossRef] [Green Version]
  110. Stebbing, J.; Phelan, A.; Griffin, I.; Tucker, C.; Oechsle, O.; Smith, D.; Richardson, P. COVID-19: Combining antiviral and anti-inflammatory treatments. Lancet Infect. Dis. 2020. [Google Scholar] [CrossRef]
  111. Matthay, M.A.; Aldrich, J.M.; Gotts, J.E. Treatment for severe acute respiratory distress syndrome from COVID-19. Lancet Respir. Med. 2020. [Google Scholar] [CrossRef] [Green Version]
  112. Morgenstern, B.; Michaelis, M.; Baer, P.C.; Doerr, H.W.; Cinatl, J., Jr. Ribavirin and interferon-beta synergistically inhibit SARS-associated coronavirus replication in animal and human cell lines. Biochem. Biophys. Res. Commun. 2005, 326, 905–908. [Google Scholar] [CrossRef]
  113. Roback, J.D.; Guarner, J. Convalescent Plasma to Treat COVID-19: Possibilities and Challenges. JAMA 2020. [Google Scholar] [CrossRef] [Green Version]
  114. Van Doremalen, N.; Bushmaker, T.; Munster, V.J. Stability of Middle East respiratory syndrome coronavirus (MERS-CoV) under different environmental conditions. Eur. Surveill. 2013, 18, 20590. [Google Scholar] [CrossRef] [Green Version]
  115. Chen, W.H.; Strych, U.; Hotez, P.J.; Bottazzi, M.E. The SARS-CoV-2 Vaccine Pipeline: An Overview. Curr. Tropl. Med. Rep. 2020, 3, 1–4. [Google Scholar] [CrossRef] [Green Version]
  116. Casanova, L.M.; Jeon, S.; Rutala, W.A.; Weber, D.J.; Sobsey, M.D. Effects of air temperature and relative humidity on coronavirus survival on surfaces. Appl. Environ. Microbiol. 2010, 76, 2712–2717. [Google Scholar] [CrossRef] [Green Version]
  117. Kampf, G.; Todt, D.; Pfaender, S.; Steinmann, E. Persistence of coronaviruses on inanimate surfaces and its inactivation with biocidal agents. J. Hosp. Infect. 2020. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Intermediate hosts for the SARS virus (civet cat), the MERS virus (camel), and the possible intermediate hosts for SARS-CoV-2 (pangolin or snake). The dotted lines indicate intermediate hosts under investigation (adopted and modified from literature) [33,34,43].
Figure 1. Intermediate hosts for the SARS virus (civet cat), the MERS virus (camel), and the possible intermediate hosts for SARS-CoV-2 (pangolin or snake). The dotted lines indicate intermediate hosts under investigation (adopted and modified from literature) [33,34,43].
Pathogens 09 00307 g001
Figure 2. The interaction of the viral S-protein with ACE2, its subsequent activation by protease (TMPRSS2), and its viral entry into the cell. The schematic primary structure of the SARS-CoV-2 spike protein is elaborated indicating the major domains. SS—signal sequence, RBD—receptor binding domain, RBD subdomains 1 and 2—SD1 and SD2, S1/S2—the protease cleavage site, S2′—the protease restriction site indicated by the arrows, FP—fusion peptide, HR1 and HR2—heptad repeats 1 and 2, CH—central helix, CD—connector domain, TM—transmembrane domain, and CT—cytoplasmic tail. (The schematic was adopted and modified from [87,88].)
Figure 2. The interaction of the viral S-protein with ACE2, its subsequent activation by protease (TMPRSS2), and its viral entry into the cell. The schematic primary structure of the SARS-CoV-2 spike protein is elaborated indicating the major domains. SS—signal sequence, RBD—receptor binding domain, RBD subdomains 1 and 2—SD1 and SD2, S1/S2—the protease cleavage site, S2′—the protease restriction site indicated by the arrows, FP—fusion peptide, HR1 and HR2—heptad repeats 1 and 2, CH—central helix, CD—connector domain, TM—transmembrane domain, and CT—cytoplasmic tail. (The schematic was adopted and modified from [87,88].)
Pathogens 09 00307 g002
Table 1. Comparative analysis of COVID-19 and SARS with reference to their corresponding causative agents, symptoms, origins, and therapeutics.
Table 1. Comparative analysis of COVID-19 and SARS with reference to their corresponding causative agents, symptoms, origins, and therapeutics.
Sr. No.COVID-19
(SARS-CoV-2)
SARS
(SARS-CoV)
References
1COVID-19 is represented by pneumonia-like symptoms, fever, cough, or diarrhea. The outbreak of disease was recorded in December 2019, in China.SARS showed many symptoms similar to that of COVID-19. The outbreak was detected in November 2002 (winter), in China.[44,59,60,61]
2To date, the mortality rate of COVID-19 is 4.5% to 5.5%. There are more than 1 million reported infections and 50,000 deaths (as recorded on 3 April 2020).The mortality rate was between 9.6% to 21%. It was restricted to 8437 individuals and 813 deaths.[6,7,62]
3The virus needs a longer incubation time (average 14 days) to represent COVID-19 symptoms.The virus needed a relatively short incubation time (1–4 days) to exhibit symptoms.[11,63]
4In COVID-19, the infection ratio between males and females is 2.7:1, indicating that the disease is more prevalent among males. Old aged people also have a high mortality rate.The male to female ratio was 1:1.25; more prevalent in females. There was a higher death rate in old aged patients.[3,64]
5SARS-CoV-2 has a potential origin from bats, and it is suspected to have a zoonotic transmission involving an unclear intermediate host. The pangolin is considered as a probable intermediate host; snakes, minks, and turtles are also being investigated.SARS-CoV originated from bats. It has zoonotic transmission via the civet cat as the intermediate host.[65,66,67,68]
6Several diagnostic tools including RT-PCR, chest CT, LAMP, etc., have been applied to detect COVID-19. However, the efficacy and sensitivity of these methods is still under investigation.SARS was efficiently diagnosed by RT-PCR.[53,69,70,71]
7COVID-19 is being treated by antiviral, antimalarial, and anti-inflammatory medicine. It is also being treated by the transfusion of active plasma antibodies into the blood circulation of infected patients.SARS was treated by antiviral drugs including ribavirin and interferon.[72,73,74,75,76]
8Low temperature is more suitable for SARS-CoV-2 viability and pathogenicity.Infection ability and viability were temperature dependent.[11,77]
9COVID-19 infections have spread to over 99.8% of the global populated area.SARS was restricted to 29 countries in the world.[78,79,80]
10SARS-CoV-2 remains viable in aerosols, and on plastic and steel surfaces, for a considerable time. The virus is not viable on copper after 4 h, nor on cardboard after 24 h.SARS-CoV was found viable in aerosols for 3 h, on plastic for 72 h, and on steel for 48 h. It was not viable on copper after 8 h, nor on cardboard after 8 h.[81]
11The mechanism of the SARS-CoV-2 infection transmission is similar to that of the influenza virus.The transmission mode of SARS-CoV is not similar to that of the influenza virus.[46,82,83]
12Recent reports advocate the asymptomatic transmission of SARS-CoV-2.Asymptomatic transmission of SARS-CoV has also been reported.[9,46,64,84]
13Six amino acids, Leu455, Phe486, Gln493, Ser494, Asn501, and Tyr505 are critical in ACE2 binding to the domain of SARS-CoV-2.The corresponding amino acids in SAR-CoV are: Tyr442, Leu472, Asn479, Asp480, Thr487, and Tyr4911. This indicates that five out of six amino acids are different to that of SARS-CoV-2.[37,43]

Share and Cite

MDPI and ACS Style

Nadeem, M.S.; Zamzami, M.A.; Choudhry, H.; Murtaza, B.N.; Kazmi, I.; Ahmad, H.; Shakoori, A.R. Origin, Potential Therapeutic Targets and Treatment for Coronavirus Disease (COVID-19). Pathogens 2020, 9, 307. https://doi.org/10.3390/pathogens9040307

AMA Style

Nadeem MS, Zamzami MA, Choudhry H, Murtaza BN, Kazmi I, Ahmad H, Shakoori AR. Origin, Potential Therapeutic Targets and Treatment for Coronavirus Disease (COVID-19). Pathogens. 2020; 9(4):307. https://doi.org/10.3390/pathogens9040307

Chicago/Turabian Style

Nadeem, Muhammad Shahid, Mazin A. Zamzami, Hani Choudhry, Bibi Nazia Murtaza, Imran Kazmi, Habib Ahmad, and Abdul Rauf Shakoori. 2020. "Origin, Potential Therapeutic Targets and Treatment for Coronavirus Disease (COVID-19)" Pathogens 9, no. 4: 307. https://doi.org/10.3390/pathogens9040307

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