Skip to content
Publicly Available Published by De Gruyter September 14, 2020

An evidence-based systematic review on emerging therapeutic and preventive strategies to treat novel coronavirus (SARS-CoV-2) during an outbreak scenario

  • Anupama M. Gudadappanavar EMAIL logo and Jyoti Benni

Abstract

A novel coronavirus infection coronavirus disease 2019 (COVID-19) emerged from Wuhan, Hubei Province of China, in December 2019 caused by SARS-CoV-2 is believed to be originated from bats in the local wet markets. Later, animal to human and human-to-human transmission of the virus began and resulting in widespread respiratory illness worldwide to around more than 180 countries. The World Health Organization declared this disease as a pandemic in March 2020. There is no clinically approved antiviral drug or vaccine available to be used against COVID-19. Nevertheless, few broad-spectrum antiviral drugs have been studied against COVID-19 in clinical trials with clinical recovery. In the current review, we summarize the morphology and pathogenesis of COVID-19 infection. A strong rational groundwork was made keeping the focus on current development of therapeutic agents and vaccines for SARS-CoV-2. Among the proposed therapeutic regimen, hydroxychloroquine, chloroquine, remdisevir, azithromycin, toclizumab and cromostat mesylate have shown promising results, and limited benefit was seen with lopinavir–ritonavir treatment in hospitalized adult patients with severe COVID-19. Early development of SARS-CoV-2 vaccine started based on the full-length genome analysis of severe acute respiratory syndrome coronavirus. Several subunit vaccines, peptides, nucleic acids, plant-derived, recombinant vaccines are under pipeline. This article concludes and highlights ongoing advances in drug repurposing, therapeutics and vaccines to counter COVID-19, which collectively could enable efforts to halt the pandemic virus infection.

Introduction

In December 2019, a novel coronavirus (CoV) infection emerged from Wuhan, Hubei Province of China, has spread to many countries worldwide [1], [2]. It is believed to be originated from bats, linked initially to animal-to-human transmission in local wet markets. Subsequently, human-to-human transmission of the virus commenced, resulting in widespread respiratory illness worldwide to around 183 countries [1], [2]. On February 11, 2020, the World Health Organization (WHO) named the virus 2019-nCoV (SARS-CoV-2), and the syndrome was named coronavirus disease 2019 (COVID-19), and on March 11, 2020, the WHO declared this disease pandemic as a global health emergency [3], [4]. The ongoing spread of this virus could potentially bring major challenges to worldwide health systems and consequences on the global economy and financial market if not controlled effectively [5].

As a part of the response to this outbreak, researchers from many nations are focusing their efforts on finding possible therapeutic options to save lives and to implement appropriate preventive and control strategies [6], [7]. Currently registered clinical trials under the WHO are evaluating a wide variety of interventions for SARS-CoV-2 and include mainly repurposing agents already known. A team of researchers reported recently to the Food and Drug Administration (FDA) that approximately 70 drugs and experimental compounds may be effective in treating the CoV [7].

In this review article, a literature search was performed using PubMed and Google Scholar to identify relevant English-language articles published on CoV infections. We have outlined different potential treatment options that could be pursued as a therapy for 2019-nCoV-2 virus, keeping the focus on agents that could be rapidly tested in patients today and broadly effective in spite of limited knowledge about virology of SARS-CoV-2. The information included in this report provides a strong intellectual groundwork for the ongoing development of therapeutic agents and vaccines.

Virology and protein targets of SARS-CoV-2

CoV is an enveloped virus with a positive sense single-stranded ribonucleic acid (RNA) genome (26e32 kb) which causes illness in humans and animals [8]. There are four classes of CoVs designated as alpha, beta, gamma and delta. The beta CoV is named after the crown-like spikes that its genome encodes several structural proteins, including the glycosylated spike (S) protein that functions as a major inducer of host immune responses [9], [10]. The viral genome also encodes nonstructural proteins which can be potential drug targets including RNA-dependent RNA polymerase (RdRp), CoV main protease (3CLpro) and papain-like protease (PLpro) [11], [12]. The beta CoV class includes severe acute respiratory syndrome (SARS) and severe acute respiratory syndrome coronavirus (SARS-CoV), Middle East respiratory syndrome (MERS) and Middle East respiratory syndrome coronavirus (MERS-CoV) and SARS-CoV-2, the COVID-19 causative agent [8], [9]. Similar to SARS-CoV and MERS-CoV, SARS-CoV-2 attacks the lower respiratory system to cause viral pneumonia, but it may also affect the kidney, gastrointestinal system, heart, liver and central nervous system leading to multiple organ failure [9].

Entering a vulnerable cell (fusion and endocytosis)

Several reports confirmed that the first step of the human COVID-19 pathogenesis is that the virus specifically recognizes the angiotensin converting enzyme 2 (ACE2) receptor by its spike protein, and it is well known that the ACE2 receptor is widely distributed on the human cell surface, especially the alveolar type II cells and capillary endothelium [13], [14]. Following receptor binding, a host type 2 transmembrane serine protease (TMPRSS2) facilitates cell entry of the virus via the S protein [14].

Releasing viral RNA and translation

After CoV enters the cell, it releases a snippet of genetic material called RNA which is read by the host cell. Infected cells begin to churn out new spikes and other proteins by viral proteinases 3CLpro and PLpro in making of more copies of the CoV [14], [15].

Assembling new copies and spreading the infection

The immune system acts as a natural harbor and helps assemble new copies of the virus which were carried to the outer edges of the cell. Before the cell finally breaks down and dies, each infected cell can release loads of copies of the virus. The viruses may infect nearby cells or end up in droplets that discharge from the lungs.

These viral lifecycle steps propose potential targets for novel drug therapy. ACE2 is an ectoenzyme that converts angiotensin II to angiotensin (1–7) at well-known sites of infection, including the lung and intestine. ACE2 acts as a potent negative regulator restraining, and overactivation of the renin-angiotensin system that may be involved in elicitation of inflammatory lung disease and further elevated pH can interfere with ACE2 glycosylation [12], [13]. Further promising drug targets which share homology with other CoVs include nonstructural proteins (3CLpro and RdRp) and a deubiquitinase, PLpro that deubiquinates certain host cell proteins, including interferon factor 3 and NF-κB, resulting in immune suppression in the host [14], [15], [16]. Table 1 summarizes the potential targets and major pharmacologic parameters of selected drugs with high target specificity and/or recognition of existing drugs that could be repurposed to treat SARS-CoV-2.

Table 1:

Stages of virus replication and key proteins involved in SARS-CoV-2 infection and potential drug candidates under research.

Stage of virus replication and key proteins involvedPossible classes of selective inhibitorsPotential drug candidate for SARS-CoV-2
Cell entry, attachment and penetration

S Protein: viral spike glycoprotein (a viral surface protein for binding to the host cell receptor ACE2)

TMPRSS2 (a host cell–produced protease that primes S protein to facilitate its binding to ACE2)

Endosome/ACE2
Soluble receptor decoys,

Antireceptor antibodies,

Fusion protein inhibitors
Umifenovir (brand name Arbidol) [22], [23]

Camostat mesylate [43], [44]

Hydroxychloroquine/chloroquine [45], [46]
Uncoating

Release of viral genome
Ion channel blockers,

Capsid stabilizers
-
Transcription of viral genome

Transcription of viral mRNA

Replication of viral genome

PLpro (papain-like protease)

3CLpro (coronavirus main protease)
Inhibitors of viral DNA polymerase, RNA polymerase, reverse transcriptase, helicase, primase or integraseLopinavir [17], [18], [19]

Ritonavir [17], [18], [19]

Darunavir
Translation of viral proteins

Regulatory proteins (early)

Structural proteins (late)

RdRp (RNA-dependent RNA polymerase)
Interferons, antisense oligonucleotides, ribozymes, inhibitors of regulatory proteinsRemdesivir [27], [28]

Ribavirin [31], [32]

Favipiravir [38], [39]
Posttranslational modifications proteolytic cleavage myristoylation, glycosylation aspartyl proteaseProtease inhibitorsNelfinavir [85] (anti-HIV drug)

Hydroxychloroquine and chloroquine
Assembly of virion componentsInterferons, assembly protein inhibitorsInterferons (α-1b, α-2b, β 1-a, β-1b) [71], [72], [73], [74], [75], [76]
Release

Budding, cell lysis
Neuraminidase inhibitors, antiviral antibodies, cytotoxic lymphocytes-
  1. TMPRSS2: type 2 transmembrane serine protease.

Investigational antiviral drugs

Human Immunodeficiency Virus (HIV) protease inhibitors

Lopinavir and ritonavir

These showed potential activity for other CoVs (SARS-CoV and MERS-CoV) in preclinical studies [17]. They may act by binding to M pro, a key enzyme for CoV replication, and suppress CoV activity [18], [19]. Various clinical studies done till now confirmed that there was no difference noted in the duration of viral shedding and no observed benefit with lopinavir–ritonavir treatment in hospitalized adult patients with severe COVID-19 [19], [20]. Certain adverse effects such as risk of cardiac arrhythmias (e.g., QT prolongation), risk of use in hepatic disease or hepatitis and significant drug interactions concern their use in critically ill adults with COVID-19 [19], [20], [21], [22].

Arbidol (ARB)/umifenovir

Arbidol (ARB)/umifenovir is an indole‐derivative molecule which functions as a virus–host cell fusion inhibitor targeting the S protein–ACE2 interaction to prevent viral entry into host cells [22], [23]. It is currently licensed in Russia and China for prophylaxis and treatment of influenza and other respiratory viral infections, as well as for hepatitis C virus [24]. The current dose of 200 mg orally every 8 h for influenza is being studied for COVID-19 treatment, but this lacks enough clinical experience [24], [25], [26].

Anti-Hepatitis C Virus (Anti-HCV), nucleotide inhibitors

Remdesivir is a monophosphoramidate prodrug of remdesivir triphosphate (RDV-TP), an adenosine analog that acts as an inhibitor of RdRps [27], [28]. RDV-TP competes with adenosine triphosphate for incorporation into nascent viral RNA chains [28], [29]. Once incorporated into the viral RNA, RDV-TP acts as a delayed RNA chain terminator by evading viral exoribonuclease’s proofreading process [30]. In both preclinical trials and clinical studies, remdesivir has demonstrated significant activity against CoV and a high genetic barrier to resistance [30], [31]. Sofosbuvir and Ribavirin can tightly bind to the newly emerged CoV RdRp and hence contradict the function of the protein leading to viral eradication hence under investigation for COVID-19 treatment. But the dose required to inhibit viral replication is high (e.g., 1.2–2.4 g orally every 8 h) hence, limit their use only in combination with other drugs [31], [32].

Other investigational drugs with potential antiviral effect

Several other antiviral drugs are being investigated, predominately those with activity against various influenza subtypes and other RNA viruses such as rintatolimod (Ampligen; toll-like receptor 3 agonist) [33], beta-D-N4-hydroxycytidine (NHC, EIDD-2801) [34], azvudine (nucleoside reverse transcriptase inhibitor) [35], danoprevir (NS3/4A HCV protease inhibitor) [36], plitidepsin (targets EF1A) [37] and favipiravir (viral RNA polymerase inhibitor) [38], [39]. Chymotrypsin‐like (3C‐like) inhibitor, Cinanserin a serotonin receptor antagonist shown promising results as inhibitor of replication of SARS-CoV [40]. Promazine, an antipsychotic drug, has been found to exhibit a significant effect in inhibiting the replication of SARS-CoV by blocking the interaction of S protein and ACE2 [41]. Niclosamide is an anthelmintic agent that targets the viral reservoir in the gut to decrease prolonged infection and transmission. It is under phase 2a/2b study for its promising antiviral actions [42]. Camostat mesylate, an approved drug for medical use in Japan for the treatment of chronic pancreatitis and postsurgery reflux esophagitis, has shown to prevent CoV cell entry in vitro through inhibition of the host serine protease, TMPRSS2 [43], [44]. This novel mechanism provides an additional drug target for future research.

Hydroxychloroquine and chloroquine

Hydroxychloroquine and chloroquine are commonly used antimalarial drugs and are also used to treat autoimmune conditions because of their immunomodulatory effects [45]. As inhibitors of hemepolymerase, they are recently believed to have additional antiviral activity via alkalinization of the phagolysosome, which inhibits the pH-dependent steps of viral replication [46].

The pharmacological activity of chloroquine and hydroxychloroquine was tested using SARS-CoV-2–infected Vero cells [47]. Hydroxychloroquine was found to be more potent than chloroquine in vitro; possible mechanisms may include inhibition of viral enzymes or processes such as viral DNA and RNA polymerase, viral protein glycosylation, virus assembly, new virus particle transport and virus release [48]. Other mechanisms may also involve ACE2 cellular receptor inhibition, acidification at the surface of the cell membrane inhibiting fusion of the virus and immunomodulation of cytokine release [49], [50], [51]. Though the studies on hydroxychloroquine for postexposure prophylaxis in healthcare workers or household contacts are underway, however, several countries have already adopted this as a measure to combat COVID-19 (Table 2) [52], [53]. Despite these promising results, this drug has several major limitations like risk of cardiac arrhythmias (e.g., QT prolongation) and retinal damage, especially with long-term use [54]. Patients with G6PD deficiency and diabetics should be cautious as it may cause significant drug interactions [54].

Table 2:

Medical management protocol for health workers and household contact of suspected/confirmed COVID-19 cases. (Indian Council of Medical Research, ICMR guidelines) [52], [53].

CategoryFor prophylaxisFor treatment
HydroxychloroquineHydroxychloroquine + azithromycinLopinavir (200 mg)/ritonavir (50 mg)
DescriptionDosageDosageDosage
Category 1Asymptomatic healthcare workers involved in the care of suspected or confirmed cases of COVID-19400 mg BD on day 1, f/by

400 mg once weekly for seven weeks
Hydroxychloroquine

200mg TID for 10 days

+

azithromycin

500 mg OD foe five days
2 tablets BD for 14 days or 7 days after becoming asymptomatic.

(whichever is earlier)
Category 2Asymptomatic household contact of laboratory confirmed cases of COVID-19400 mg BD on day 1, f/by

400 mg once weekly for three weeks
  1. COVID-19: coronavirus disease 2019.

Hydroxychloroquine plus azithromycin

Azithromycin is a macrolide antibacterial having immunomodulatory properties in pulmonary inflammatory disorders [55]. In one study, the patients receiving combination therapy had initially lower viral loads, when compared with patients receiving hydroxychloroquine alone with similar viral burden [56]. This may be explained by their ability to downregulate inflammatory responses and reduce the excessive cytokine production associated with respiratory viral infections [57]. Other antimicrobial actions include reducing chemotaxis of Polymorphonuclear Leucocytes (PMNs) to the lungs by inhibiting cytokines (i.e., IL-8), accelerating neutrophil apoptosis, inhibition of mucus hypersecretion, decreased production of reactive oxygen species and blocking the activation of nuclear transcription factors [58], [59], [60]. However, their direct effects on viral clearance are uncertain and awaits further research.

Ivermectin is an antiparasitic agent that has shown antiviral activity against a broad range of viruses, mainly HIV1 and dengue by various mechanisms [61]. Based on the hypothesis, ivermectin may interfere with viral replication process by acting as a specific inhibitor of importin-α/β-mediated nuclear import in RNA virus, ivermectin shows synergistic activity when combined with HCQ [62], [63].

Nitric oxide (NO) is a gas with diverse biological activities and is produced from arginine by NO synthases [64]. A phase 2 study of iNO is underway in patients with COVID-19 with the goal of preventing disease progression in those with severe Acute Respiratory Distress Syndrome (ARDS), and also, NO was also found to inhibit the synthesis of viral protein and RNA synthesis [65], [66].

Renin–angiotensin–aldosterone system (RAAS) blockade and COVID-19

As per literature search, SARS-CoV-2 is known to utilize angiotensin-converting enzyme 2 (ACE2) receptors for entry into target cells [13], [14]. It was therefore hypothesized that any agent that increases expression of ACE2 could potentially increase susceptibility to severe COVID-19 by improving viral cellular entry. However, physiologically, ACE2 converts angiotensin-II to angiotensin(1–7), which may protect against lung injury by lowering angiotensin 2 receptor binding and vasodilation [67]. There are also conflicting data regarding whether to continue or discontinue drugs that inhibit the renin–angiotensin–aldosterone system (RAAS), namely angiotensin-converting enzyme inhibitors (ACEIs) and angiotensin receptor blockers (ARBs) in patients with COVID-19 and comorbidities such as hypertension, cardiovascular disease and diabetes which are often treated with ACEIs and ARBs [67], [68], [69]. A trial is currently under evaluation for the use of losartan in patients with COVID-19 [70].

Biologics for CoV-associated disease (immunomodulators)

Most CoV infections cause fever as the immune system response to clear the virus. In severe cases, the immune system can overreact and start attacking lung cells, making it difficult to breathe as the lungs become obstructed with fluid and dying cells which can lead to acute respiratory distress syndrome and possibly death [71].

Interferons (IFNs)

During a viral infection, the most prominent cytokines produced are interferons (IFNs) that have the ability to interfere with viral replication [72]. A vivid correlation between the innate immune response threshold and the fatality rates in COVID-19 can be found. The higher threshold of IFN-mediated immune responses can increase mortality rates in the elderly [73]. The authors concluded that the delayed IFN-related antiviral response is a possible strategy implemented by CoV to evade the immune response [71], [74]. The virus also circumvents the immune system by hiding its double-stranded RNA in vesicles, causing less IFN induction [75]. The key for success in reducing the disease fatality might be stimulation of the innate immune responses to trigger IFN production at the very early stages of the disease, which might be done through administration of agents that are able to augment IFN production. Interferon-γ combination with an interferon-I might induce synergistic benefits [76]. However, in-depth research is needed to validate it and determine the optimum dosage and timing to prevent unwanted results. Interferon–human serum albumin fusion proteins (HSA-IFNs) have significantly lengthened the plasma half-life of IFNs (e.g., from 10 h to 12 days for HSA-IFN-α2b) due to slower free IFN release into the plasma and thus may prolong the duration of action of IFN for each injection [77].

Numerous studies have indicated a “cytokine storm” with release of interleukin-6 (IL-6), IL-1, IL-12 and IL-18, along with tumor necrosis factor alpha, and other inflammatory mediators may cause severe disease in patients with COVID-19 [71], [78], [79]. IL-6 is a proinflammatory cytokine that is involved in diverse physiological processes such as T-cell activation, immunoglobulin secretion induction, hepatic acute-phase protein synthesis initiation and hematopoietic precursor cell proliferation and differentiation stimulation. The increased inflammatory response from lungs may result in increased alveolar–capillary gas exchange, making oxygenation difficult in patients with severe illness [71], [78], [79]. IL-6 inhibitors may remediate severe damage to lung tissue caused by cytokine release in patients with serious coronaviral infections [80]. Initial data of few studies suggest that tocilizumab and sarilumab, IL-6 inhibitors, may have clinical benefit as adjunctive therapy which inhibits IL-6–mediated signaling by competitively binding to both soluble and membrane-bound IL-6 receptors [81], [82]. However, the drug increases the risk of GI perforation and hepatotoxicity in patients with thrombocytopenia and neutropenia and infusion-related reactions. Leronlimab is an investigational humanized monoclonal antibody to the chemokine receptor CCR5 [83]. This drug is currently under study for its ability to enhance immune response while mitigating cytokine storm in patients with COVID-19.

RNA therapies

Virus’ RNA genome first enters a cell; it interacts with the host’s protein-making machinery, using it to make proteins that can copy RNA molecules, and these RNA-copying proteins are called as “polymerases” [84]. By blocking these RNA polymerases, the virus can no longer reproduce that may reduce infection and can make a promising target for therapies. One such drug, remdesivir, was originally developed in the hope that it would limit Ebola virus and its relatives [27], [28]. CoV RNA encodes certain proteases that help RNA polymerase to get fully functional in order to adopt their mature configuration. Scientists have now found that HIV-1 protease inhibitor, nelfinavir, strongly inhibits the replication of the SARS-CoV, despite the fact that these viruses are unrelated [85].

Antisense oligonucleotides have also been developed to reduce the severity of SARS virus infections and to prevent or treat SARS virus–associated disease [86], [87]. Antisense oligonucleotides may also target disease-related proteins involved in the inflammatory process.

Antibody-directed therapy

The data indicate that the S protein is a putative target for SARS-CoV-2 antibody development as in viral infection but not the other structural proteins, M, E, and N in SARS-CoV, elicits an immune response [10]. A neutralizing antibody targeting the S protein on the surface of COVID-19 is likely the first therapy contemplated by biomedical researchers in the academia and industry, providing passive immunity to disease [88]. Certain monoclonal antibodies, for example, 80R, CR3014, F26G18, F26G19, m396, 1A9 and 4D4, which are used to target spike protein in SARS-CoV and MERS-CoV showed promising results in vitro and in vivo that could be potentially effective against SARS-CoV-2 [88], [89], [90].

Corticosteroids

Corticosteroid therapy is not recommended for viral pneumonia; however, use may be considered for patients with refractory shock or acute respiratory distress syndrome and there is no enough evidence-based data supporting its role in patients with COVID-19 [91], [92].

Convalescent plasma or immunoglobulins

Convalescent plasma is an antibody-rich product that is collected from eligible donors who have recovered from COVID-19 and FDA states its use only to confirmed patients having severe or immediately life-threatening COVID-19. It is important to determine its safety and efficacy via clinical trials before routinely administering to patients with SARS-CoV-2 [93], [94]. One possible explanation for the efficacy of convalescent plasma therapy is that the antibodies from convalescent plasma might suppress viremia [94]. However, convalescent plasma has not yet been shown to be effective in COVID-19 and needs further evaluation.

High-dose intravenous vitamin C treatment for COVID-19

The results of several meta-analyses have been established that intravenous (IV) high-dose vitamin C has significant benefit in the treatment of sepsis and septic shock, which may be because of its role as a prooxidant for immune cells but as an antioxidant for lung epithelial cells improving alveolar fluid clearance. Furthermore, vitamin C treatment known to inhibit lactate secretion produced by the activated immune cells may protect innate immunity [95], [96]. This effect may benefit patients with COVID‐19 as they SARS-CoV-2 mainly affects the lower respiratory tract.

Stem cells, a therapeutic strategy

One of the concerns based on preclinical science is that the virus can infect the stem cells rendering them ineffective. It was recently brought to the light that ACE2 and umbilical cord mesenchymal stem cells (UC-MSCs) may benefit the patients with COVID-19 via immunoregulatory function. The first CoV case, treated with umbilical cord cells, was reported from China [97], leading to further to speculation and further investigations. It was known that UC-MSCs have been reported responsible of the bacterial clearance in preclinical models of sepsis, ARDS and cystic fibrosis infection. (ref) In patients with COVID-19, mesenchymal stem cell (MSC) therapy can inhibit the overactivation of the immune system and promote endogenous repair by improving the microenvironment. After entering the human body through IV infusion, part of the MSCs accumulate in the lung, which could improve the pulmonary microenvironment, protect alveolar epithelial cells, prevent pulmonary fibrosis and improve lung function [98], [99]. Therefore, the fact that the transplantation of MSCs improved the outcome of patients with COVID-19 may be due to regulating inflammatory response and promoting tissue repair and regeneration which needs further evaluation and research.

Progress and prospects on vaccine development against SARS-CoV-2

Presently, the epidemic is still spreading, and there is no effective means to prevent the infection. Vaccines are the most helpful and cost-effective means for prophylaxis and treatment of infectious diseases. Previously, lot of research has been done to develop vaccines against human CoV infections such as MERS and SARS. Conversely, till date, no licensed antiviral treatment or vaccine is available for MERS and SARS CoV infections [100], [101].

Lot of research work is directed toward the design and development of vaccines for SARS-CoV-2. Currently, those at the highest risk of acquiring COVID-19 like frontline healthcare workers, individuals aged 60 years and older or those with underlying diabetes and hypertension need vaccination [2]. Therefore, such populations might be prioritized for vaccine clinical trials. Important points to be considered while producing a vaccine against human COVID-19 are they should minimize immunopotentiation, must be suitable for stockpiling and can be given to healthcare workers, elderly and adults with underlying diabetes or hypertension [102]. Herewith, we provide a brief overview of the major candidates or vaccines under development.

COVID-19 and target product profile

Numerous approaches are adopted for the development of CoV vaccines; most of these target the surface-exposed spike (S) glycoprotein or S protein as it is the key inducer of neutralizing antibodies. The S protein molecule contains two subunits, S1 and S2. The S1 subunit has a receptor-binding domain (RBD) that interacts with its host cell receptor, angiotensin-converting enzyme 2 (ACE2), whereas the S2 subunit causes fusion between the virus and host cell membrane and then releases its viral RNA into the cytoplasm for further replication [102], [103]. Therefore, S protein–based vaccines can induce antibodies which block viral receptor binding and also virus genome uncoating in the cytoplasm. The S protein has a major role in the induction of protective immunity during infection with SARS-CoV by eliciting neutralizing-antibodies and T-cell responses [103].

The SARS-CoV-2 shares high genetic similarity with the SARS-CoV, based on the full-length genome phylogenetic analysis approximately 89% nucleotide similar to SARS-like CoVs (genus beta-CoV) found in Chinese bats [104], [105], [106]. On this basis, the early development of SARS-CoV-2 vaccine started and also suggested that the receptor of SARS-CoV-2 might be the same as that of SARS-CoV receptor (ACE2) [107].

Major COVID-19 vaccine development programs (Table 3)

Table 3:

List of various vaccine candidates for COVID-19 under development.

Type of vaccineVaccines under development [reference]
Whole virus vaccines: live-attenuated/inactive whole virus vaccine
  • – Johnson & Johnson – using adenovirus-vectored vaccine [110]

  • – University of Hong Kong – live influenza vaccine that expresses SARS-CoV-2 protein [111]

  • – Codagenix, Inc. in collaboration with the Serum Institute of India – live-attenuated vaccine by codon deoptimization technology. [112], [113]

Subunit vaccines
  • – University of Queensland – based on the “molecular clamp” technology developing a subunit vaccine against spike protein. [114]

  • – Novavax – using recombinant protein nanoparticle technology [115], [116].

  • – Clover biopharmaceuticals Inc. – using their patented Trimer-Tag® technology developed a subunit vaccine. [117]

  • – Texas Children’s Hospital Center – developed SARS-CoV-2 RBD vaccine. [118], [119]

  • – MNA-embedded SARS-CoV-2 vaccine [123]

Nucleic acid vaccines
  • – Inovio Pharmaceuticals /Beijing Advaccine Biotechnology Co developed a DNA vaccine “INO-4800” [125]

  • – LineaRx and takis biotech – analyzing DNA vaccine for SARS-CoV-2 [126]

  • – Zydus Cadila – recombinant DNA vaccine for COVID-19.[127]

  • – Moderna Therapeutics (mRNA-1273, encoding S protein) is in phase I clinical trials. [128]

  • – Fudan University/Shanghai Jiaotong University and Bluebird Biopharmaceutical Company – mRNA vaccine. [129]

  • – CureVac AG – mRNA vaccine [131]

  • – Stermirna Therapeutics [132]

Live vector vaccines
  • – Houston-based Greffex Inc. developed SARS-CoV-2 adenovirus vector vaccine with Greffex vector [133]

  • – Tonix Pharmaceuticals – (TNX-1800) vaccine [134]

  • – Johnson & Johnson – (AdVac®) vaccine [135]

  • – University of Oxford – (ChAdOx1 nCoV-19) [136]

  • – CanSino Biologics – (Ad5-nCoV) [137]

Synthetic peptide or epitope vaccine
  • – Hong Kong University of Science and Technology – peptide vaccine for COVID-19 138

Plant biopharmaceuticals
  • – Rapid response vaccines [140], [141]

  1. COVID-19: coronavirus disease 2019; RBD: receptor-binding domain; MNA: microneedle array.

  • a) Whole virus vaccines or live-attenuated or inactive whole virus vaccines: Whole virus vaccines represent a standard approach for viral vaccinations. The whole cell antigens include all the elements of the virus, i.e. proteins, lipids, polysaccharide and nucleic acids [108]. A major advantage of whole virus vaccines is their inherent immunogenicity and ability to stimulate toll-like receptors. But live virus vaccines often require extensive testing to validate their safety. This is especially an issue for CoV vaccines, given the findings of increased infectivity following immunization with live or killed whole virus SARS CoV vaccines [109]. Following are the live attenuated vaccines under development.

    • The US Healthcare giant Johnson & Johnson has started developing a vaccine against the Wuhan novel CoV SARS-CoV-2. The vaccine program is using Janssen’s adenovirus-vectored vaccine and PER.C6 technologies to rapidly upscale production of the optimal candidate [110] and has planned to initiate phase 1 clinical trials by September.

    • The researchers at the University of Hong Kong have developed a live influenza vaccine that expresses SARS-CoV-2 proteins [111].

    • The US Codagenix, Inc. is collaborating with the Serum Institute of India, Ltd. to develop a live-attenuated vaccine against SARS-CoV-2. They have developed a “codon deoptimization” technology to synthesize “rationally designed” live-attenuated vaccines. This technology helps for the rapid generation of multiple vaccine candidates against the virus [112], [113].

    • The US Healthcare giant Johnson & Johnson has started developing a vaccine against the Wuhan novel CoV SARS-CoV-2. The vaccine program is using Janssen’s adenovirus-vectored vaccine and PER.C6 technologies to rapidly upscale production of the optimal candidate [110] and has planned to initiate phase 1 clinical trials by September.

    • The researchers at the University of Hong Kong have developed a live influenza vaccine that expresses SARS-CoV-2 proteins [111].

    • The US Codagenix, Inc. is collaborating with the Serum Institute of India, Ltd. to develop a live-attenuated vaccine against SARS-CoV-2. They have developed a “codon deoptimization” technology to synthesize “rationally designed” live-attenuated vaccines. This technology helps for the rapid generation of multiple vaccine candidates against the virus [112], [113].

  • b) Subunit vaccines: The subunit vaccines with one or more antigens induce strong immunogenicity and efficiently stimulate the host immune system. These vaccines are safer and easier to manufacture and require the addition of adjuvants to elicit a strong protective immune response. Subunit vaccines for SARS-CoV-2 are directed against the S-spike protein to prevent its docking with the host ACE2 receptor [109].

    • The University of Queensland along with Viroclinics is synthesizing a subunit vaccine based on the “molecular clamp” technology, which locks the ‘spike’ protein into a shape which allows the immune system to be able to recognize and then neutralize the virus. It uses rapid response vaccine technology to develop a vaccine, and within three weeks, the candidate vaccine will be produced. This project is funded by the Coalition for Epidemic Preparedness Innovation (CEPI) [114].

    • Novavax, Inc. created the COVID-19 vaccine candidates using its proprietary recombinant protein nanoparticle technology to produce antigens derived from the CoV spike (S) protein. Novavax is awarded funding from the CEPI [115]. Novavax is using its proprietary Matrix-M™ adjuvant with its COVID-19 vaccine candidate which stimulates the entry of antigen-presenting cells into the injection site and enhances antigen presentation in local lymph nodes, boosting immune responses [116]. It is assessing the efficacy in animal models to identify an optimal vaccine candidate for human testing. Phase 1 clinical testing is expected to begin in late spring of 2020 [115].

    • Clover Biopharmaceuticals Inc. successfully produced 2019-nCoV subunit vaccine candidate against spike protein (“S-Trimer”) via mammalian cell expression system. They detected cross-reacting antibodies from sera of multiple infected patients. Clover is the first company in the world to disclose a 2019-nCoV vaccine candidate, announced on 10th February, 2020. The vaccine development was carried out with the support of leadership teams from Chengdu Hi-Tech Park and Chengdu Clinical Center for public health in China [117].

    • Texas Children’s Hospital Center for Vaccine Development at Baylor College of Medicine (including University of Texas Medical Branch and New York Blood Center) has developed and tested a subunit vaccine comprising the RBD of the SARS-CoV S-protein [109], [118], [119]. The RBD in spike protein located on the outer membrane of CoV mediates receptor binding and plays a major role in virus entry into the host cell [120]. This protein is used as a potential vaccine candidate as it is the major target for neutralizing antibodies [121]. When formulated on alum, the SARS-CoV RBD vaccine elicits high levels of protective immunity on the homologous virus challenge. The RBD-based vaccine minimizes host immunopotentiation [109].

    • Pasteur Institute, Johnson & Johnson and Chongqing Zhifei Biological Products Co., Ltd. also started subunit vaccine development against SARS-CoV-2 [122].

    • Microneedle array (MNA)–embedded SARS-CoV-2 S1 subunit vaccine [123]: The MNA technique is a minimally invasive subcutaneous approach, with low dose requirement, low cost and less toxicity, when compared to traditional needle injection [124]. In this preclinical study, C57BL/6 female mice (five animals per group) were inoculated intracutaneously with MNAs loaded with of SARS-CoV-2-S1 or SARS-CoV-2-S1fRS09 protein or phosphate buffered saline (PBS) as a negative control. In this study, by 2 weeks after immunization with the MNA delivered SARS-CoV-2, S1 subunit vaccine showed potent antigen-specific antibody responses [123].

  • c) Nucleic acid vaccines

    DNA vaccines are characteristically composed of plasmid DNA molecules that encode one or more antigens.

    • Inovio Pharmaceuticals, Inc. in collaboration with Beijing Advaccine Biotechnology Co. has developed a DNA vaccine “INO-4800”, which is through the preclinical studies, and also, phase 1 human testing is done to evaluate the safety and immunogenicity, and the project is being funded by the CEPI [125].

    • Another DNA vaccine for SARS-CoV-2 is in preclinical studies being tested by collaboration of Applied DNA Sciences Subsidiary, LineaRx and Takis Biotech. They are working on preclinical development of a linear DNA vaccine based on polymerase chain reaction, and the vaccine constructs were produced at a scale using plasmid-based vaccine templates [126]. The initial data from preclinical animal tests showed immunogenicity and a strong antibody generation across all five vaccine candidates.

    • Zydus Cadila is developing a DNA vaccine against the major viral membrane protein responsible for the cell entry of the novel CoV. Also, it is planning to develop a live-attenuated recombinant measles virus–vectored vaccine against COVID-19 [127].

      mRNA vaccines have high potency, short production cycles, low-cost manufacturing and are safe for administration in comparison to the conventional vaccines [128].

    • Moderna Therapeutics has developed a SARS-CoV-2 mRNA vaccine (mRNA-1273, encoding S protein) which is in phase I clinical trials on healthy volunteers [129]. This phase 1 study is being conducted by the National Institutes of Health under its own Investigational New Drug application. The safety and immunogenicity of the vaccine is being evaluated by using three dose levels of mRNA-1273 (25, 100, 250 μg) with 2 doses per schedule, given 28 days apart. In this study, 45 healthy adults will be included. Participants will be followed during 12 months after the second dose vaccination. The primary objective of this study is to evaluate the safety and reactogenicity, and the secondary objective is to estimate the immunogenicity to the SARS-CoV-2 S protein [129]. Moderna is working to begin its phase 2 clinical trials.

    • Fudan University in collaboration with Shanghai Jiaotong University and Bluebird Biopharmaceutical Company is developing SARS-CoV-2 mRNA vaccine using two different strategies. The first is to use mRNA to express the SARS-CoV-2 S protein and RBD domain; the efficacy of this vaccine is now under evaluation in mice [130].

    • In addition, German biopharmaceutical company CureVac AG [131], Stermirna Therapeutics [132], BDGENE Therapeutics, Guanhao Biotech, ZY Therapeutics Inc., CanSino Vaccines Biologics Inc., Baylor College of Medicine, University of Texas, Tongji university also announced their progress on mRNA vaccine development against SARS-CoV-2 [123].

  • d) Live vector vaccines: Live vector vaccines are the live viruses (the vector) that express a heterologous antigen(s), and they are broadly used to induce cellular immunity. They are characterized by combining the strong immunogenicity of live-attenuated vaccines and the safety of subunit vaccines [123]. Below mentioned are few live vector vaccines.

    • Houston-based Greffex Inc. has completed the construction of SARS-CoV-2 adenovirus vector vaccine with genetically engineered Greffex Vector Platform and is planning for animal testing on approval by the USFDA [133].

    • Tonix Pharmaceuticals announced research to develop a potential SARS-CoV-2 vaccine TNX-1800 (live modified horsepox virus vaccine for percutaneous administration), based on Tonix’s proprietary horsepox vaccine platform [134].

    • Johnson & Johnson in collaboration with the U.S. Biomedical Advanced Research and Development Authority to develop a vaccine against CoV has adopted the AdVac® adenoviral vector platform for vaccine development [135].

    • The University of Oxford is testing their vaccine candidate (ChAdOx1 nCoV-19), an adenovirus vaccine vector against the COVID-19 spike protein. It has initiated their phase I/II randomized, placebo-controlled, multicentre clinical trial in healthy volunteers aged 18–55 years to determine the efficacy, safety and immunogenicity of the vaccine against COVID-19. The vaccine is administered via intramuscular route [136].

    • CanSino Biologics has developed a adenovirus type-5 vector-based recombinant COVID-19 vaccine, Ad5-nCoV, which has completed phase I studies, and sooner, they are initiating phase 2 clinical trials in China. The phase 2 randomized study has three primary endpoints: firstly, to evaluate adverse reactions within the first 14 days of vaccination; secondly, on day 28 to check for serum levels of anti-SARS-CoV-2 neutralizing antibody and antibody against the CoV’s spike protein and lastly, to follow up participants for 6 months [137].

  • e) Synthetic Or peptide vaccine: the Hong Kong University of Science and Technology has screened a set of B- and T-cell epitopes from S and N proteins of SARS-CoV; these epitopes are highly conserved in SARS-CoV-2 which guides toward the development of SARS-CoV-2 vaccines [138]. These peptide vaccines generally are prepared by chemical synthesis techniques and contain only certain fragments of intact antigens. Because of their low immunogenicity, they need additional structural modifications, delivery systems and adjuvants in the formulation [139].

  • f) Plant biopharmaceuticals: Plant expression systems for production of candidate vaccines offer advantage of affordable cost and rapid production, which aids in global vaccination programs. They produce “rapid response vaccines” as it produces more protein in a short time [140]. The plant-based biopharmaceutical production against 2019-nCoV will include the identification of potential epitopes and production of full-length viral surface proteins present in the envelope region or production of subunit vaccines expressing the immunogenic region or chimeric proteins [141].

Animal models for evaluation of SARS-CoV-2 virus

In the process of preclinical evaluation, animal models are vital to study the efficacy of vaccine candidates. Previously, by introducing hACE2 gene into the mouse genome, a SARS-CoV transgenic mouse model was developed [142]. Recently, in an animal study, they used the hACE2 transgenic mice infected with SARS-CoV-2 to evaluate the pathogenicity of the virus. In these transgenic mice infected with SARS-CoV-2, weight loss and virus replication in the lung tissue was seen. On histopathology, typical interstitial pneumonia findings were observed, and also, viral antigens were detected in the bronchial epithelial cells, alveolar epithelia and alveolar macrophages. This mouse model may facilitate the development of therapeutic drugs and vaccines against SARS-CoV-2 [143].

Cell culture systems for SARS-CoV-2

SARS-CoV-2 isolation has been attempted in Vero and the Huh-7 cells (human liver cancer cells) [144]. The monoclonal antibodies (fully human/humanized) that target both the S1-RBD and non-RBD, as well as the S2 domain of CoV, have been developed and tested in cell cultures for virus neutralizing capability, as well as in animal models for prophylactic and postexposure efficacies [145].

Bacille Calmette-Guérin vaccine (BCG) and COVID-19

Bacille Calmette-Guérin vaccine (BCG) is a live attenuated strain derived from an isolate of Mycobacterium bovis used broadly across the world as a vaccine for tuberculosis, with many countries, including India, Japan and China, having a universal BCG vaccination policy in infants. BCG vaccination has been shown to produce positive “heterologous” or nonspecific immune effects (trained immunity) leading to improved response against other nonmycobacterial pathogens and is proposed to be caused by metabolic and epigenetic changes leading to promotion of genetic regions encoding for proinflammatory cytokines specifically IL-1B, which has been shown to play a vital role in antiviral immunity [146], [147]. One study data suggested that BCG vaccination seems to significantly reduce mortality associated with COVID-19 [148] as this vaccination has been shown to produce broad protection against viral infections and sepsis [148]. The correlation between the beginning of universal BCG vaccination and the protection against COVID-19 needs further exploration.

Conclusion

This article concludes that, at present, neither vaccines nor direct-acting antiviral drugs are available for the treatment of human or animal CoV infections. The drug repurposing effort summarized in this review is applicable only to adult patients and focused primarily on agents currently known to be effective against other RNA viruses including SARS-CoV, MERS-CoV, influenza, HCV and Ebola, as well as antiinflammatory drugs. The biologics for treatment of COVID-19 have potential impacts as they have shown promising results, including bioengineered and vectored antibodies, cytokines and nucleic acid–based therapies targeting virus gene expression, as well as various types of vaccines. Finally, a concerted effort to develop effective drugs and vaccines against existing and potential future CoV infections and other highly pathogenic virus outbreaks is necessary to reduce overwhelming impacts on human life and worldwide healthcare systems.


Corresponding author: Dr. Anupama M. Gudadappanavar, Assistant professor, J N Medical College, KLE Academy of Higher Education and Research (KAHER), Belagavi, Karnataka, India, E-mail:

  1. Research funding: None declared.

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Competing interests: Authors state no conflict of interest.

  4. Informed consent: Informed consent was obtained from all individuals included in this study.

References

1. Li, Q, Guan, X, Wu, P, Wang, X, Zhou, L, Tong, Y, et al. Early transmission dynamics in Wuhan, China, of novel coronavirus-infected pneumonia. N Engl J Med 2020. Available from: https://doi.org/10.1056/NEJMoa2001316.Search in Google Scholar

2. Huang, C, Wang, Y, Li, X, Ren, L, Zhao, J, Hu, Y, et al. Clinical features of patients infected with 2019 novel coronavirus in wuhan, China. Lancet 2020;395:469–70. https://doi.org/10.1016/S0140-6736(20)30183-5.Search in Google Scholar

3. WHO. WHO Director-General’s opening remarks at the media briefing on COVID-19; 11 March 2020. Available from: https://www.who.int/dg/speeches/detail/who-director-general-s-opening-remarks-at-themedia-briefing-on-covid-19—11-march-2020 [Accessed 24 Mar 2020].Search in Google Scholar

4. Zhu, N, Zhang, D, Wang, W, Li, X, Yang, B, Song, J, et al. A novel coronavirus from patients with pneumonia in China, 2019. New Engl J Med 2020;382:727–33. https://doi.org/10.1056/nejmoa2001017.Search in Google Scholar PubMed PubMed Central

5. Coronavirus is now expected to curb global economic growth by 0.3% in; 2020. Available from: https://www.forbes.com/sites/sergeiklebnikov/2020/02/11/coronavirus-is-now-expected-to-curb-global-economic-growth by-03-in-2020/#5de149ad16da.Search in Google Scholar

6. Lu, H. Drug treatment options for the 2019-new coronavirus (2019-nCoV). Biosci Trends 2020. https://doi.org/10.5582/bst.2020.01020.Search in Google Scholar PubMed

7. Race to find COVID-19 treatments accelerates. Kai Kupferschmidt and Jon Cohen Science;367:1412–13. https://doi.org/10.1126/science.367.6485.1412.Search in Google Scholar PubMed

8. Perlman, S, Netland, J. Coronaviruses post-SARS: update on replication and pathogenesis, Nat Rev Microbiol 2009;7:439e450, https://doi.org/10.1038/nrmicro2147.Search in Google Scholar PubMed PubMed Central

9. Chen, Y, Liu, Q, Guo, D. Emerging coronaviruses: genome structure, replication, and pathogenesis. J Med Virol 2020;92:418–23. https://doi.org/10.1002/jmv.25681.Search in Google Scholar PubMed PubMed Central

10. Su, S, Wong, G, Shi, W, Liu, J, Lai, ACK, Zhou, J, et al. Epidemiology, genetic recombination, and pathogenesis of coronaviruses. Trends Microbiol 2016;24:490−502. https://doi.org/10.1016/j.tim.2016.03.003.Search in Google Scholar PubMed PubMed Central

11. Yang, H, Bartlam, M, Rao, Z. Drug design targeting the main protease, the Achilles’ heel of coronaviruses. Curr Pharm Des 2006;12:4573–90. https://doi.org/10.2174/138161206779010369.Search in Google Scholar PubMed

12. Anand, K, Ziebuhr, J, Wadhwani, P, Mesters, JR, Hilgenfeld, R. Coronavirus main proteinase (3CLpro) structure: basis for design of anti-SARS drugs. Science 2003;300:1763–7. https://doi.org/10.1126/science.1085658.Search in Google Scholar PubMed

13. Kuba, K, Imai, Y, Rao, S, Gao, H, Guo, F, Guan, B, et al. A crucial role of angiotensin converting enzyme 2 (ACE2) in SARS coronavirus-induced lung injury. Nat Med 2005 Aug;11:875–9. https://doi.org/10.1038/nm1267.Search in Google Scholar PubMed PubMed Central

14. Hoffmann, M, Kleine-Weber, H, Schroeder, S, Krüger, N, Herrler, T, Erichsen, S, et al. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and is blocked by a clinically proven protease inhibitor. Cell 2020 Mar 4. https://doi.org/10.1016/j.cell.2020.02.052 [Epub ahead of print].Search in Google Scholar PubMed PubMed Central

15. Baez-Santos, YM, St John, SE, Mesecar, AD. The SARS coronavirus papain-like protease: structure, function and inhibition by designed antiviral compounds. Antiviral Res 2015;115:21−38. https://doi.org/10.1016/j.antiviral.2014.12.015.Search in Google Scholar PubMed PubMed Central

16. Gordon, DE, Gwendolyn, MJ, Mehdi, B, Jiewei, X, Kirsten, O, White, KM, et al. A SARS-CoV-2-human protein-protein interaction map reveals drug targets and potential drug-repurposing. bioRxiv 2020. https://doi.org/10.1101/2020.03.22.002386.Search in Google Scholar PubMed PubMed Central

17. Chu, CM, Cheng, VC, Hung, IF, Wong, MML, Chan, KH, Chan, KS, et al. Role of lopinavir/ritonavir in the treatment of SARS: initial virological and clinical findings. Thorax 2004;59:252–6. https://doi.org/10.1136/thorax.2003.012658.Search in Google Scholar PubMed PubMed Central

18. Huang, J, Song, W, Huang, H, Sun, Q. Pharmacological therapeutics targeting RNA-dependent RNA polymerase, proteinase and spike protein: from mechanistic studies to clinical trials for COVID-19. J Clin Med 2020;9:1131. https://doi.org/10.3390/jcm9041131.Search in Google Scholar PubMed PubMed Central

19. Jin, Z, Du, X, Xu, Y, Deng, Y, Liu, M, Zhao, Y, et al. Structure of Mpro from COVID-19 virus and discovery of its inhibitors. bioRxiv 2020 Jan 1. https://doi.org/10.1101/2020.02.26.964882.Search in Google Scholar

20. Cao, B, Wang, Y, Wen, D, Liu, W, Wang, J, Fan, G, et al. A trial of lopinavir–ritonavir in adults hospitalized with severe Covid-19. N Engl J Med 2020. https://doi.org/10.1056/NEJMoa2001282.Search in Google Scholar PubMed PubMed Central

21. Li, Y, Xie, Z, Lin, W, Cai, W, Wen, C, Guan, Y, et al. An exploratory randomized, controlled study on the efficacy and safety of lopinavir/ritonavir or arbidol treating adult patients hospitalized with mild/moderate COVID-19 (ELACOI). medRxiv 2020. https://doi.org/10.1016/j.medj.2020.04.001.Search in Google Scholar PubMed PubMed Central

22. A new antiviral drug heading into clinical trials offers hope for COVID-19 treatment – in part because it can be taken as a pill. Emery University; 2020 Apr 06. Available from: https://news.emory.edu/stories/2020/04/covid_eidd_2801_lung/index.html.Search in Google Scholar

23. Sheahan, TP, Sims, AC, Zhou, S, Graham, RL, Pruijssers, AJ, Agostini, ML, et al. An orally bioavailable broad-spectrum antiviral inhibits SARS-CoV-2 and multiple endemic, epidemic and bat coronavirus. Biorxiv 2020 Mar 20. https://doi.org/10.1101/2020.03.19.997890.Search in Google Scholar

24. Kadam, RU, Wilson, IA. Structural basis of influenza virus fusion inhibition by the antiviral drug Arbidol. Proc Natl Acad Sci USA 2017;114:206214. https://doi.org/10.1073/pnas.1617020114.Search in Google Scholar PubMed PubMed Central

25. Khamitov, RA, SIa, L, Shchukina, VN, Borisevich, SV, Maksimov, VA, Shuster, AM. Antiviral activity of arbidol and its derivatives against the pathogen of severe acute respiratory syndrome in the cell cultures [in Russian]. Vopr Virusol 2008;53:9–13.Search in Google Scholar

26. Yousefifard, M, Zali, A, Ali, KM, Neishaboori, AM, Zarghi, A, Hosseini, M, et al. Antiviral therapy in management of COVID-19: a systematic review on current evidence. Archives of Academic Emergency Medicine 2020;8.10.1111/ijcp.13557Search in Google Scholar PubMed PubMed Central

27. Elfiky, AA. Anti-HCV, nucleotide inhibitors, repurposing against COVID-19. Life Sci 2020 Feb 28:117477. https://doi.org/10.1016/j.lfs.2020.117477.Search in Google Scholar PubMed PubMed Central

28. Al-Tawfiq, JA, Al-Homoud, AH, Memish, ZA. Remdesivir as a possible therapeutic option for the COVID-19. Travel Med Infect Dis Published online March 5, 2020. https://doi.org/10.1016/j.tmaid.2020.101615.Search in Google Scholar PubMed PubMed Central

29. Yethindra, V. Role of GS-5734 (Remdesivir) in inhibiting SARS-CoV and MERS-CoV: the expected role of GS-5734 (Remdesivir) in COVID-19 (2019-nCoV)-VYTR hypothesis. Int J Res Pharm Sci 2020 Mar 6;11:1–6. https://doi.org/10.26452/ijrps.v11iSPL1.1973.Search in Google Scholar

30. Ko, WC, Rolain, JM, Lee, NY, Chen, PL, Huang, CT, Lee, PI, et al. Arguments in favour of remdesivir for treating SARS-CoV-2 infections. Int J Antimicrob Agents 2020 Jan 1. https://doi.org/10.1016/j.ijantimicag.2020.105933.Search in Google Scholar PubMed PubMed Central

31. Sheahan, TP, Sims, AC, Leist, SR, Schäfer, A, Won, J, Brown, AJ, et al. Comparative therapeutic efficacy of remdesivir and combination lopinavir, ritonavir, and interferon-beta against MERS-CoV. Nat Commun 2020;11:222. https://doi.org/10.1038/s41467-019-13940-6.Search in Google Scholar PubMed PubMed Central

32. Khalili, JS, Zhu, H, Mak, A, Yan, Y, Zhu, Y. Novel coronavirus treatment with ribavirin: groundwork for evaluation concerning COVID‐19. J Med Virol 2020 Mar 30. https://doi.org/10.1002/jmv.25798.Search in Google Scholar PubMed PubMed Central

33. Overton, ET, Goepfert, PA, Cunningham, P, Carter, WA, Horvath, J, Young, D, et al. Intranasal seasonal influenza vaccine and a TLR-3 agonist, rintatolimod, induced cross-reactive IgA antibody formation against avian H5N1 and H7N9 influenza HA in humans. Vaccine 2014 Sep 22;32:5490–5. https://doi.org/10.1016/j.vaccine.2014.07.078.Search in Google Scholar PubMed

34. Pruijssers, AJ, Denison, MR. Nucleoside analogues for the treatment of coronavirus infections. Curr Opin Virol 2019 Apr 1;35:57–62. https://doi.org/10.1016/j.coviro.2019.04.002.Search in Google Scholar PubMed PubMed Central

35. Sheahan, TP, Sims, AC, Zhou, S, Graham, RL, Pruijssers, AJ, Agostini, ML, et al. An orally bioavailable broad-spectrum antiviral inhibits SARS-CoV-2 and multiple endemic, epidemic and bat coronavirus. bioRxiv 2020 Mar 20. https://doi.org/10.1126/scitranslmed.abb5883.Search in Google Scholar PubMed PubMed Central

36. Lythgoe, MP, Middleton, P. Ongoing clinical trials for the management of the COVID-19 pandemic. Trends Pharmacol Sci 2020 Apr 9. https://doi.org/10.1016/j.tips.2020.03.006.Search in Google Scholar PubMed PubMed Central

37. Sorbera, LA, Graul, AI, Dulsat, C. Taking aim at a fast-moving target: targets to watch for SARS-CoV-2 and COVID-19. Drugs Future 2020;45. https://doi.org/10.1358/dof.2020.45.4.3150676.Search in Google Scholar

38. Vafaei, S, Razmi, M, Mansoori, M, Asadi-Lari, M, Madjd, Z. Spotlight of remdesivir in comparison with ribavirin, favipiravir, oseltamivir and umifenovir in coronavirus disease 2019 (COVID-19) pandemic. Favipiravir, oseltamivir and umifenovir in coronavirus disease. 2019. https://doi.org/10.2139/ssrn.3569866.Search in Google Scholar

39. Qingxian, C, Yang, M, Liu, D, Chen, J, Shu, D, Xia, J, et al. Experimental treatment with Favipiravir for COVID-19: an open-label control study. Engineering 2020. https://doi.org/10.1016/j.eng.2020.03.007.Search in Google Scholar PubMed PubMed Central

40. Basha, SH. Corona virus drugs–a brief overview of past, present and future. J Peer Scientist 2020;2: e1000013.Search in Google Scholar

41. Mani, D, Wadhwani, A, Krishnamurthy, PT. Drug repurposing in antiviral research: a current Scenario. J Young Pharm 2019 Apr 1;11:117. https://doi.org/10.5530/jyp.2019.11.26.Search in Google Scholar

42. First Wave Bio to Initiate Phase 2a/2b Study of FW-1022. A proprietary form of Niclosamide, to treat COVID-19. First wave bio 2020 Apr 09. Available from: https://www.firstwavebio.com/firstwave-bio-to-initiate-phase-2a-2b-study-of-fw-1022-a-proprietary-form-of-niclosamide-to-treat-covid-19/.Search in Google Scholar

43. Bittmann, S, Luchter, E, Moschüring-Alieva, E, Villalon, G, Weissenstein, AC. 19: Camostat and the role of serine protease entry inhibitor TMPRSS2. J Regen Biol Med 2020;2:1–2. https://doi.org/10.14302/issn.2691-5014.jphn-20-3295.Search in Google Scholar

44. Hoffmann, M, Kleine-Weber, H, Schroeder, S, Krüger, N, Herrler, T, Erichsen, S, et al. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and is blocked by a clinically proven protease inhibitor. Cell 2020 Mar 5. https://doi.org/10.1016/j.cell.2020.02.052.Search in Google Scholar PubMed PubMed Central

45. Pastick, KA, Okafor, EC, Wang, F, Lofgren, SM, Skipper, CP, Nicol, MR, et al. Review: hydroxychloroquine and chloroquine for treatment of SARS-CoV-2 (COVID-19). Open Forum Infect Dis, ofaa130 https://doi.org/10.1093/ofid/ofaa130.Search in Google Scholar PubMed PubMed Central

46. Chen, Z, Hu, J, Zhang, Z, Jiang, S, Han, S, Yan, D, et al. Efficacy of hydroxychloroquine in patients with COVID-19: results of a randomized clinical trial. MedRxiv 2020 Jan 1. https://doi.org/10.1101/2020.03.22.20040758.Search in Google Scholar

47. Liu, J, Cao, R, Xu, M, Wang, X, Zhang, H, Hu, H, et al. Hydroxychloroquine, a less toxic derivative of chloroquine, is effective in inhibiting SARS-CoV-2 infection in vitro. Cell discovery 2020 Mar 18;6:1–4. https://doi.org/10.1038/s41421-020-0156-0.Search in Google Scholar

48. Fantini, J, Di Scala, C, Chahinian, H, Yahi, N. Structural and molecular modeling studies reveal a new mechanism of action of chloroquine and hydroxychloroquine against SARS-CoV-2 infection. Int J Antimicrob Agents 2020 Apr 3:105960. https://doi.org/10.1016/j.ijantimicag.2020.105960.Search in Google Scholar

49. Colson, P, Rolain, JM, Lagier, JC, Brouqui, P, Raoult, D. Chloroquine and hydroxychloroquine as available weapons to fight COVID-19. Int J Antimicrob Agents 2020 Mar 4:105932–1016. https://doi.org/10.1016/j.ijantimicag.2020.105932.Search in Google Scholar

50. Gao, J, Tian, Z, Breakthrough, YX. Chloroquine phosphate has shown apparent efficacy in treatment of COVID-19 associated pneumonia in clinical studies. Biosci Trends 2020 Mar 16;14:72–3. https://doi.org/10.5582/bst.2020.01047.Search in Google Scholar

51. Gendrot, M, Javelle, E, Le Dault, E, Clerc, A, Savini, H, Pradines, B. Chloroquine as prophylactic agent against COVID-19?. Int J Antimicrob Agents 2020 Apr 12. https://doi.org/10.1016/j.ijantimicag.2020.105980.Search in Google Scholar

52. What’s Up? @St John’s Hospital SPECIAL ISSUE ON COVID-19 PANDEMIC Issue 41, Supplement 1. Release Date; 01/04/2020.Search in Google Scholar

53. Post-exposure prophylaxis for SARS-coronavirus-2. ClinicalTrials.gov 2020 Mar 23. Available from: https://clinicaltrials.gov/ct2/show/NCT0430866 [Accessed 24 March 2020].Search in Google Scholar

54. Brunton, LL, et al. editors. Goodman & Gilman’s: The Pharmacological Basis of Therapeutics, 13e New York, NY: McGraw-Hill; 2018. p. 969–86.Search in Google Scholar

55. Ohe, M, Shida, H, Jodo, S, Kusunoki, Y, Seki, M, Furuya, K, et al. Macrolide treatment for COVID-19: will this be the way forward?. BioScience Trends 2020. https://doi.org/10.5582/bst.2020.03058.Search in Google Scholar

56. Dahly, D, Gates, S, Morris, T. Statistical review of hydroxychloroquine and azithromycin as a treatment of COVID-19: results of an open-label nonrandomized clinical trial. Preprint. Posted online 2020 Mar 23:23.Search in Google Scholar

57. Lane, JC, Weaver, J, Kostka, K, Duarte-Salles, T, Abrahao, MT, Alghoul, H, et al. Safety of hydroxychloroquine, alone and in combination with azithromycin, in light of rapid wide-spread use for COVID-19: a multinational, network cohort and self-controlled case series study. medRxiv 2020 Jan 1. https://doi.org/10.1016/s2665-9913(20)30276-9.Search in Google Scholar

58. Gabriels, J, Saleh, M, Chang, D, Epstein, LM. Inpatient use of mobile continuous telemetry for COVID-19 patients treated with hydroxychloroquine and azithromycin. HeartRhythm Case Reports 2020 Apr 1. https://doi.org/10.1016/j.hrcr.2020.03.017.Search in Google Scholar PubMed PubMed Central

59. Gautret, P, Lagier, J, Parola, P, Hoang, VT, Meddeb, L, Mailhe, M, et al. Hydroxychloroquine and azithromycin as a treatment of COVID-19: results of an open-label non-randomized clinical trial. Int J Antimicrob Agents 2020 Mar 20. https://doi.org/10.1016/j.ijantimicag.2020.105949.Search in Google Scholar PubMed PubMed Central

60. Lane, JWJ. Safety of hydroxychloroquine, alone and in combination with azithromycin, in light of rapid wide-spread use for COVID-19: a multinational, network cohort and self-controlled case series study. Available from: https://data.ohdsi.org/Covid19EstimationHydroxychloroquine/ [Accessed 17 April 2020].Search in Google Scholar

61. Wagstaff, KM, SivakumaranH, HeatonSM, HarrichD, JansDA. Ivermectin is a specific inhibitor of importin alpha/beta169 mediated nuclear import able to inhibit replication of HIV-1 and dengue virus. Biochem J 2012;443:851–6. https://doi.org/10.1042/bj20120150.Search in Google Scholar

62. Patrì, A, Fabbrocini, G. Hydroxychloroquine and ivermectin: a synergistic combination for COVID-19 chemoprophylaxis and/or treatment?. J Am Acad Dermatol 2020 Apr 10. https://doi.org/10.1016/j.jaad.2020.04.017. Epub ahead of print. PMCID: PMC7146719.10.1016/j.jaad.2020.04.017Search in Google Scholar PubMed PubMed Central

63. Caly, L, Druce, JD, Catton, MG, Jans, DA, Wagstaff, KM. The FDA-approved Drug Ivermectin inhibits the replication of SARS-CoV-2 in vitro. Antivir Res 2020 Apr 3:104787. https://doi.org/10.1016/j.antiviral.2020.104787.Search in Google Scholar PubMed PubMed Central

64. Regev-Shoshani, G, Vimalanathan, S, McMullin, B, Road, J, Av-Gay, Y, Miller, C, et al. Gaseous nitric oxide reduces influenza infectivity in vitro. Nitric Oxide 2013;31:48–53. https://doi.org/10.1016/j.niox.2013.03.007.Search in Google Scholar PubMed PubMed Central

65. Chen, L, Liu, P, Gao, H, Sun, B, Chao, D, Wang, F, et al. Inhalation of nitric oxide in the treatment of severe acute respiratory syndrome: a rescue trial in Beijing. Clin Infect Dis 2004;39:1531–5. https://doi.org/10.1086/425357.Search in Google Scholar PubMed PubMed Central

66. Nitric oxide gas inhalation for severe acute respiratory syndrome in COVID-19. ClinicalTrials.gov 2020 Mar 06. Available from: https://clinicaltrials.gov/ct2/show/NCT04290871 [Accessed 17 April 2020].Search in Google Scholar

67. Kreutz, R, Algharably, EA, Azizi, M, Dobrowolski, P, Guzik, T, Januszewicz, A, et al. Hypertension, the renin–angiotensin system, and the risk of lower respiratory tract infections and lung injury: implications for COVID-19European Society of Hypertension COVID-19 Task Force Review of Evidence. Cardiovasc Res 2020.10.1093/cvr/cvab224Search in Google Scholar PubMed PubMed Central

68. Vaduganathan, M, Vardeny, O, Michel, T, McMurray, JJV, Pfeffer, MA, Solomon, SD. Renin-angiotensin-aldosterone system inhibitors in patients with Covid-19. N Engl J Med 2020 Mar 30.10.1056/NEJMsr2005760Search in Google Scholar PubMed PubMed Central

69. Ferrario, CM, Jessup, J, Chappell, MC, Averill, DB, Brosnihan, KB, Tallant, EA, et al. Effect of angiotensin-converting enzyme inhibition and angiotensin II receptor blockers on cardiac angiotensin-converting enzyme 2. Circulation 2005 May 24;111:2605–10. https://doi.org/10.1161/circulationaha.104.510461.Search in Google Scholar PubMed

70. University of Minnesota. Losartan for patients with COVID-19 requiring hospitalization (NCT04312009). ClinicalTrials.gov 2020 Mar 17. Available from: https://www.clinicaltrials.gov/ct2/show/NCT04312009?term=NCT04312009&draw=2&rank=1 [Accessed 27 March 2020].Search in Google Scholar

71. Li, G, Fan, Y, Lai, Y, Han, T, Li, Z, Zhou, P, et al. Coronavirus infections and immune responses. J Med Virol 2020 Apr;92:424–32. https://doi.org/10.1002/jmv.25685.Search in Google Scholar PubMed PubMed Central

72. Mantlo, EK, Bukreyeva, N, Maruyama, J, Paessler, S, Huang, C. Potent antiviral activities of type I interferons to SARS-CoV-2 infection. bioRxiv 2020 Jan 1.10.1101/2020.04.02.022764Search in Google Scholar PubMed PubMed Central

73. Coronavirus age, sex, demographics (COVID-19) – worldometer. [Internet]. Available from: https://www.worldometers.info/coronavirus/coronavirus-age-sexdemographics/ [Accessed on 20 April 2020].Search in Google Scholar

74. Mosaddeghi, P, Negahdaripour, M, Dehghani, Z, Farahmandnejad, M, Moghadami, M, Nezafat, N, et al. Therapeutic approaches for COVID-19 based on the dynamics of interferon-mediated immune responses 2020. https://doi.org/10.20944/preprints202003.0206.v1.Search in Google Scholar

75. Deng, X, Yu, X, Pei, J. Regulation of interferon production as a potential strategy for COVID-19 treatment. arXivpreprint arXiv:2003.00751. 2020 Mar 2.Search in Google Scholar

76. Prokunina-Olsson, L, Alphonse, N, Dickenson, RE, Durbin, JE, Glenn, JS, Hartmann, R, et al. COVID-19 and emerging viral infections: the case for interferon lambda. J Exp Med 2020 May 4:217. https://doi.org/10.1084/jem.20200653.Search in Google Scholar

77. Liu, C, Zhou, Q, Li, Y, Garner, LV, Watkins, SP, Carter, LJ. Research and development on therapeutic agents and vaccines for COVID-19 and related human coronavirus diseases. Acs Central Sci 2020. https://doi.org/10.1021/acscentsci.0c00272.Search in Google Scholar

78. Mehta, P, McAuley, DF, Brown, M, Sanchez, E, Tattersall, RS, Manson, JJ. COVID-19: consider cytokine storm syndromes and immunosuppression. Lancet 2020 Mar 28;395:1033–4. https://doi.org/10.1016/s0140-6736(20)30628-0.Search in Google Scholar

79. Ye, Q, Wang, B, Mao, J. Cytokine storm in COVID-19 and treatment. J Infect 2020 Apr 10.Search in Google Scholar

80. Liu, B, Li, M, Zhou, Z, Guan, X, Xiang, Y. Can we use interleukin-6 (IL-6) blockade for coronavirus disease 2019 (COVID-19)-induced cytokine release syndrome (CRS)?. J Autoimmun 2020 Apr 10:102452. https://doi.org/10.1016/j.jaut.2020.102452.Search in Google Scholar PubMed PubMed Central

81. Michot, JM, Albiges, L, Chaput, N, Saada, V, Pommeret, F, Griscelli, F, et al. Tocilizumab, an anti-IL6 receptor antibody, to treat Covid-19-related respiratory failure: a case report. Ann Oncol 2020 Apr 2. https://doi.org/10.1016/j.annonc.2020.03.300.Search in Google Scholar PubMed PubMed Central

82. Zhang, C, Wu, Z, Li, JW, Zhao, H, Wang, GQ. The cytokine release syndrome (CRS) of severe COVID-19 and Interleukin-6 receptor (IL-6R) antagonist Tocilizumab may be the key to reduce the mortality. Int J Antimicrob Agents 2020 Mar 29:105954. https://doi.org/10.1016/j.ijantimicag.2020.105954.Search in Google Scholar PubMed PubMed Central

83. Velavan, TP, Meyer, CG. The COVID-19 epidemic. Trop Med Int Health 2020 Mar 1;25:278–80. https://doi.org/10.1111/tmi.13383.Search in Google Scholar PubMed PubMed Central

84. Lo, IL, Lio, CF, Cheong, HH, Lei, CI, Cheong, TH, Zhong, X, et al. Evaluation of SARS-CoV-2 RNA shedding in clinical specimens and clinical characteristics of 10 patients with COVID-19 in Macau. Int J Biol Sci 2020;16:1698. https://doi.org/10.7150/ijbs.45357.Search in Google Scholar PubMed PubMed Central

85. Yamamoto, N, Matsuyama, S, Hoshino, T, Yamamoto, N. Nelfinavir inhibits replication of severe acute respiratory syndrome coronavirus 2 in vitro. bioRxiv 2020 Jan 1. https://doi.org/10.1101/2020.04.06.026476.Search in Google Scholar

86. Rangan, R, Zheludev, IN, Das, R. RNA genome conservation and secondary structure in SARS-CoV-2 and SARS-related viruses. bioRxiv 2020 Jan 1. https://doi.org/10.1101/2020.03.27.012906.Search in Google Scholar PubMed PubMed Central

87. Rabaan, AA, Al-Ahmed, SH, Sah, R, Tiwari, R, Yatoo, MI, Patel, SK, et al. SARS-CoV-2/COVID-19 and advances in developing potential therapeutics and vaccines to counter this emerging pandemic virus–A review. https://doi.org/10.20944/preprints202004.0075.v1.Search in Google Scholar

88. Shanmugaraj, B, Siriwattananon, K, Wangkanont, K, Phoolcharoen, W. Perspectives on monoclonal antibody therapy as potential therapeutic intervention for Coronavirus disease-19 (COVID-19). Asian Pac J Allergy Immunol 2020 Mar 1;38:10–8. https://doi.org/10.12932/AP-200220-0773.Search in Google Scholar

89. Jiang, S, Hillyer, C, Du, L. Neutralizing antibodies against SARS-CoV-2 and other human coronaviruses. Trends Immunol 2020 Apr 2. https://doi.org/10.1016/j.it.2020.08.003.Search in Google Scholar

90. Tian, X, Li, C, Huang, A, Xia, S, Lu, S, Shi, Z, et al. Potent binding of 2019 novel coronavirus spike protein by a SARS coronavirus-specific human monoclonal antibody. Emerg Microb Infect 2020 Jan 1;9:382–5. https://doi.org/10.1080/22221751.2020.1729069.Search in Google Scholar

91. Yang, Z, Liu, J, Zhou, Y, Zhao, X, Zhao, Q, Liu, J. The effect of corticosteroid treatment on patients with coronavirus infection: a systematic review and meta-analysis. J Infect 2020 Apr 10. https://doi.org/10.1016/j.jinf.2020.03.062.Search in Google Scholar

92. Russell, CD, Millar, JE, Baillie, JK. Clinical evidence does not support corticosteroid treatment for 2019-nCoV lung injury. Lancet 2020 Feb 15;395:473–5. https://doi.org/10.1016/s0140-6736(20)30317-2.Search in Google Scholar

93. Chen, L, Xiong, J, Bao, L, Shi, Y. Convalescent plasma as a potential therapy for COVID-19. Lancet Infect Dis 2020 Apr 1;20:398–400. https://doi.org/10.1016/s1473-3099(20)30141-9.Search in Google Scholar

94. Yoo, JH. Convalescent plasma therapy for Corona virus disease 2019: a long way to go but worth trying. J Kor Med Sci 2020 Feb 17:35. https://doi.org/10.3346/jkms.2020.35.e150.Search in Google Scholar PubMed PubMed Central

95. Erol, A. High-dose intravenous vitamin C treatment for COVID-19. OSF Preprints 26 Feb. 2020. https://doi.org/10.31219/osf.io/p7ex8. Web.Search in Google Scholar

96. Grant, WB, Lahore, H, McDonnell, SL, Baggerly, CA, French, CB, Aliano, JL, et al. Evidence that vitamin D supplementation could reduce risk of influenza and COVID-19 infections and deaths. Nutrients 2020 Apr;12:988. https://doi.org/10.3390/nu12040988.10.3390/nu12040988Search in Google Scholar PubMed PubMed Central

97. Liang, B, Chen, J, Li, T, Wu, H, Yang, W, Li, Y, et al. Clinical remission of a critically ill COVID-19 patient treated by human umbilical cord mesenchymal stem cells. China Xiv; 2020. p. 202002. 00084.10.1097/MD.0000000000021429Search in Google Scholar

98. Leng, Z, Zhu, R, Hou, W. Transplantation of ACE2 Mesenchymal stem cells improves the outcomes of patients with COVID-19 pneumonia. Aging Dis 2020;11:216–28. https://doi.org/10.14336/ad.2020.0228.Search in Google Scholar

99. Shetty, AK. Mesenchymal stem cell infusion shows promise for combating coronavirus (COVID-19)-Induced pneumonia. Aging and disease 2020 Apr;11:462. https://doi.org/10.14336/ad.2020.0301.Search in Google Scholar

100. Dhama, K, Sharun, K, Tiwari, R, Dadar, M, Malik, YS, Singh, KP, et al. COVID-19, an emerging coronavirus infection: advances and prospects in designing and developing vaccines, immune therapeutics, and therapeutics. Human Vaccines & Immunotherapeutics. Available from: https://doi.org/10.1080/21645515.2020.1735227.Search in Google Scholar

101. Graham, RL, Donaldson, EF, Baric, RS. A decade after SARS: strategies for controlling emerging coronaviruses. Nat Rev Microbiol 2013;11:836–48. https://doi.org/10.1038/nrmicro3143.Search in Google Scholar

102. Chen, W-H, Ulrich, S, Hotez, PJ, Elena Bottazzi, M. The SARS-CoV-2 vaccine pipeline: an overview. Curr Trop Med Rep; March 2020;03. Available from: https://doi.org/10.1007/s40475-020-00201-6.Search in Google Scholar

103. Du, L, He, Y, Zhou, Y, Liu, S, Zheng, BJ, Jiang, S. The spike protein of SARS-CoV–a target for vaccine and therapeutic development. Nat Rev Microbiol 2009;7:226–36. https://doi.org/10.1038/nrmicro2090.Search in Google Scholar

104. Jiang, S, Du, L, Shi, Z. An emerging coronavirus causing pneumonia outbreak in Wuhan, China: calling for developing therapeutic and prophylactic strategies. Emerg Microb Infect 2020;9:275–77. https://doi.org/10.1080/22221751.2020.1723441.Search in Google Scholar

105. Wu, F, Zhao, S, Yu, B, Chen, Y-M, Wang, W, Hu, Y, et al. Complete genome characterisation of a novel coronavirus associated with severe human respiratory disease in Wuhan, China. bioRxiv 2020. https://doi.org/10.1101/2020.01.24.919183.Search in Google Scholar

106. Lu, R, Zhao, X, Li, J, Niu, P, Yang, B, Wu, H, et al. Genomic characterisation and epidemiology of 2019 novel coronavirus: implications for virus origins and receptor binding. Lancet 2020;6736:1–10.10.1016/S0140-6736(20)30251-8Search in Google Scholar

107. Veljkovic, V, Vergara-Alert, J, Segalés, J, Paessler, S. Use of the informational spectrum methodology for rapid biological analysis of the novel coronavirus 2019-nCoV: prediction of potential receptor, natural reservoir, tropism and therapeutic/vaccine target. F1000Research 2020;9:52. https://doi.org/10.12688/f1000research.22149.1.Search in Google Scholar

108. Minor, PD. Live attenuated vaccines: historical successes and current challenges. Virology 2015;479–480:379–92. https://doi.org/10.1016/j.virol.2015.03.032.Search in Google Scholar PubMed

109. Jiang, S, Bottazzi, ME, Du, L, Lustigman, S, Tseng, CT, Curti, E, et al. Roadmap to developing a recombinant coronavirus S protein receptor-binding domain vaccine for severe acute respiratory syndrome. Expert Rev Vaccines 2012;11:1405–13. https://doi.org/10.1586/erv.12.126.Search in Google Scholar PubMed PubMed Central

110. J&J working on coronavirus vaccine. The pharma letter. 2020. Available from https://www.thepharmaletter.com/article/j-j-working-on-coronavirusvaccine [Accessed 14 April 2020].Search in Google Scholar

111. Cheung, E. China coronavirus: Hong Kong researchers have already developed vaccine but need time to test it, expert reveals. South China Morning Post. Available from: https://www.scmp.com/news/hong-kong/healthenvironment/article/3047956/ china-coronavirus-hong kong-researchers-have [Accessed 14 April 2020].Search in Google Scholar

112. Shieber, J. Codagenix raises $20 million for a new flu vaccine and other therapies. Tech Crunch. Available from: https://techcrunch.com/2020/01/13/codagenix-raises-20-million-for-a-new-flu-vaccine-and-othertherapies/ [Accessed28 Feb 2020].Search in Google Scholar

113. Codagenix and serum Institute of India initiate Co-development of a scalable, live-attenuated vaccine against the 2019 Novel coronavirus, COVID-19. BioSpace 2020. Available from: https://www.biospace.com/article/releases/codagenix-and-serum-institute-of-indiainitiate-co development-of-a-scalable-live-attenuated-vaccineagainst-the-2019-novel-coronavirus-covid-9/ [Accessed 14 April 2020].Search in Google Scholar

114. UQ News. International partnership progresses UQ COVID-19 vaccine project. Published on 09-04-2020. Available from: https://www.uq.edu.au/news/article/2020/04/international-partnership-progresses-uq-covid-19-vaccine-project [Accessed 14 April 2020].Search in Google Scholar

115. NOVAVAX press release. Novavax advances development of Novel COVID-19 vaccine. Available from: https://ir.novavax.com/news-releases/news-release-details/novavax-advances-development-novel-covid-19-vaccine [Accessed 14 April 2020].Search in Google Scholar

116. Novavax advances development of Novel COVID-19 vaccine. Drug development and delivery. April 2020. Available from: https://drug-dev.com/novavax-advances-development-of-novel-covid-19-vaccine/ [Accessed 14 April 2020].Search in Google Scholar

117. Clover Biopharmaceuticals. Clover successfully produced 2019-nCoV subunit vaccine candidate. CHENGDU, China. Published on 10th February 2020 http://www.cloverbiopharma.com/index.php?m=content&c=index&a=show&catid=11&id=41 [Accessed 16 April 2020].Search in Google Scholar

118. World Health Organization. Public statement for collaboration on COVID-19 vaccine development. Published on 13th April 2020. Available from: https://www.who.int/news-room/detail/13-04-2020-public-statement-for-collaboration-on-covid-19-vaccine-development [Accessed 16 April 2020].Search in Google Scholar

119. Chen, WH, Chag, SM, Poongavanam, MV, Biter, AB, Ewere, EA, RezendeW, et al. Optimization of the production process and characterization of the yeast-expressed SARS-CoV recombinant receptor-binding domain (RBD219-N1), a SARS vaccine candidate. J Pharmacol Sci 2017;106:1961–70. https://doi.org/10.1016/j.xphs.2017.04.037.Search in Google Scholar PubMed PubMed Central

120. Song, Z, Xu, Y, Bao, L, Zhang, L, Yu, P, Qu, Y, et al. From SARS to MERS, thrusting coronaviruses into the spotlight. Viruses 2019;11:59. https://doi.org/10.3390/v11010059.Search in Google Scholar PubMed PubMed Central

121. Peiris, J.S., Guan, Y., Yuen, K.Y. Severe acute respiratory syndrome. Nat Med 2004;10:S88–97. https://doi.org/10.1038/nm1143.Search in Google Scholar PubMed PubMed Central

122. Zhang, J, Zeng, H, Gu, J, Li, H, Zheng, L, Zou, Q. Progress and prospects on vaccine development against SARS-CoV-2. Vaccines 2020;8:153. https://doi.org/10.3390/vaccines8020153.10.3390/vaccines8020153Search in Google Scholar PubMed PubMed Central

123. Kim, E., et al. Microneedle array delivered recombinant coronavirus vaccines: Immunogenicity and rapid translational development. EBioMedicine 2020 https://doi.org/10.1016/j.ebiom.2020.102743.Search in Google Scholar PubMed PubMed Central

124. Balmert, SC, Carey, CD, Falo, GD, et al. Dissolving undercut microneedle arrays for multicomponent cutaneous vaccination. J Contr Release 2020;317:336–46. https://doi.org/10.1016/j.jconrel.2019.11.023.Search in Google Scholar PubMed PubMed Central

125. Inovio Pharmaceuticals. Inovio collaborating with beijing advaccine to advance INO-4800 vaccine against New coronavirus in China; 2020 http://ir.inovio.com/news-and-media/news/press-release-details/2020/Inovio-Collaborating-With-Beijing-Advaccine-To-Advance-INO-4800-Vaccine-Against-New-Coronavirus-In-China/default.aspx. Accessed 28 Feb 2020 [Accessed 15 Apr 2020].Search in Google Scholar

126. Pharmaceutical Technology News. Applied DNA supplies Covid-19 vaccine candidates for preclinical tests. Published on April 13, 2020. https://www.pharmaceutical-technology.com/news/applied-dna-ships-covid-19-vaccine-candidates/ [Accessed 16 Apr 2020].Search in Google Scholar

127. BloombergQuint. Zydus Cadila launches research programme to develop coronavirus vaccine; 2020. Published on Feb. 15 https://www.bloombergquint.com/business/zydus-cadila-launches-research-programme-to-develop-coronavirus-vaccineLast [Accessed 17 Apr 2020].Search in Google Scholar

128. Pardi, N, Hogan, MJ, Porter, FW, Weissman, D. mRNA vaccines—a new era in vaccinology. Nat Rev Drug Discov 2018;17:261–79. https://doi.org/10.1038/nrd.2017.243.Search in Google Scholar PubMed PubMed Central

129. Moderna Press releases. Moderna announces first participant dosed in NIH-led phase 1 study of mRNA vaccine (mRNA-1273) against Novel coronavirus. Published on March 16, 2020 https://investors.modernatx.com/news-releases/news-release-details/moderna-announces-first-participant-dosed-nih-led-phase-1-study [Accessed 16 Apr 2020].Search in Google Scholar

130. Genetic Engineering & Biotechnology News. Vanquishing the virus: 160+ COVID-19 drug and vaccine candidates in development. Published on April 13, 2020. https://www.genengnews.com/a-lists/vanquishing-the-virus-160-covid-19-drug-and-vaccine-candidates-in-development/4/ [Accessed 16 Apr 2020].Search in Google Scholar

131. Smith, J. CureVac bids to develop first mRNA coronavirus vaccine; 2020. Available from: https://www.labiotech.eu/medical/curevac-coronavirusoutbreak-cepi/ [Accessed 28 Feb 2020].Search in Google Scholar

132. Xinhua. China fast-tracks novel coronavirus vaccine development Xinhua. Published on January 29 2020. Available from: http://www.xinhuanet.com/english/2020-01/28/c_138739378.htm [Accessed 16 Apr 2020].Search in Google Scholar

133. Innovationews. Greffex completes COVID-19 vaccine, prepares for FDA animal testing. Published on March 12, 2020. Available from: https://www.innovationews.com/ [Accessed 16 Apr 2020].Search in Google Scholar

134. Tonix Pharmaceuticals, Releases Press. Tonix Pharmaceuticals announces research collaboration with southern research to develop a potential vaccine to protect against New coronavirus disease 2019 (COVID-19) based on horsepox virus (TNX-1800). Published on february 26, 2020. Available from: https://www.tonixpharma.com/news-events/press-releases/detail/1191/tonix-pharmaceuticals-announces-research-collaboration-with [Accessed 16 Apr 2020].Search in Google Scholar

135. J&J News. Johnson & Johnson announces a lead vaccine candidate for COVID-19.Published on March 30, 2020. Available from: https://www.jnj.com/johnson-johnson-announces-a-lead-vaccine-candidate-for-covid-19-landmark-new-partnership-with-u-s-department-of-health-human-services-and-commitment-to-supply-one-billion-vaccines-worldwide-for-emergency-pandemic-use [Accessed 16 Apr 2020].Search in Google Scholar

136. Trial site news. University of Oxford Commences clinical trial for vaccine candidate (ChAdOx1 nCoV-19) targeting COVID-19. Published on March 31, 2020. Available from: https://www.trialsitenews.com/university-of-oxford-commences-clinical-trial-for-vaccine-candidate-chadox1-ncov-19-targeting-covid-19-2/ [Accessed 18 Apr 2020].Search in Google Scholar

137. FiercePharma, Vaccines. China’s CanSino Bio advances COVID-19 vaccine into phase 2 on preliminary safety data. Published on April 10; 2020. Available from: https://www.fiercepharma.com/vaccines/china-s-cansino-bio-advances-covid-19-vaccine-into-phase-2-preliminary-safety-data [Accessed 18 Apr 2020].Search in Google Scholar

138. Ahmed, SF, Quadeer, AA, McKay, MR. Preliminary identification of potential vaccine targets for 2019-nCoV based on SARS-CoV immunological studies. BioRxiv 2020;12:254. https://doi.org/10.3390/v12030254.Search in Google Scholar PubMed PubMed Central

139. Azmi, F, Ahmad Fuaad, AA, Skwarczynski, M, Toth, I. Recent progress in adjuvant discovery for peptide-based subunit vaccines. Hum Vaccines Immunother 2014;10:778–96. https://doi.org/10.4161/hv.27332.Search in Google Scholar PubMed PubMed Central

140. Shanmugaraj, BM, Ramalingam, S. Plant expression platform for the production of recombinant pharmaceutical proteins. Austin J Biotechnol Bioeng 2014;1:4.Search in Google Scholar

141. Emergence of Novel coronavirus 2019-nCoV: Need for rapid vaccine and biologics development. Balamurugan shanmugaraj 1,2, ashwini malla 1,2 and waranyoo phool charoen. Pathogens 2020;9:148. https://doi.org/10.3390/pathogens9020148.Search in Google Scholar PubMed PubMed Central

142. Yang, XH, Deng, W, Tong, Z, Liu, YX, Zhang, LF, Zhu, H, et al. Mice transgenic for human angiotensin-converting enzyme 2 provide a model for SARS coronavirus infection. Comp Med 2007;57:450–59.Search in Google Scholar

143. Bao, L, Deng, W, Huang, B, Gao, H, Ren, L, Wei, Q, et al. The pathogenicity of 2019 Novel coronavirus in hACE2 transgenic mice. BioRxiv 2020. https://doi.org/10.1101/2020.02.07.939389.Search in Google Scholar

144. Wang, M, Cao, R, Zhang, L, Yang, X, Liu, J, Xu, M, et al. Remdesivir and chloroquine effectively inhibit the recently emerged novel coronavirus (2019-nCoV) in vitro. Cell Res 2020a. https://doi.org/10.1038/s41422-020-0282-0.Search in Google Scholar PubMed PubMed Central

145. Xu, J, Jia, W, Wang, P, Zhang, S, Shi, X, Wang, X, et al. Antibodies and vaccines against Middle East respiratory syndrome coronavirus. Emerg Infect Dis 2019;8:841–56. https://doi.org/10.1080/22221751.2019.1624482.Search in Google Scholar PubMed PubMed Central

146. Kleinnijenhuis, J, Quintin, J, Preijers, F, Benn, CS, Joosten, LA, Jacobs, C, et al. Long-lasting effects of BCG vaccination on both heterologous Th1/Th17 responses and innate trained immunity. J Innate Immun 2014;6:152–8. https://doi.org/10.1159/000355628.Search in Google Scholar PubMed PubMed Central

147. Moorlag, SJ, Arts, RJ, van Crevel, R, Netea, MG. Non-specific effects of BCG vaccine on viral infections. Clin Microbiol Infect 2019 Dec 1;25:1473–8. https://doi.org/10.1016/j.cmi.2019.04.020.Search in Google Scholar PubMed

148. Miller, A, Reandelar, MJ, Fasciglione, K, Roumenova, V, Li, Y, Otazu, GH. Correlation between universal BCG vaccination policy and reduced morbidity and mortality for COVID-19: an epidemiological study. medRxiv 2020 Jan 1. https://doi.org/10.1101/2020.03.24.20042937.Search in Google Scholar

Received: 2020-04-23
Accepted: 2020-07-19
Published Online: 2020-09-14

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/jbcpp-2020-0113/html
Scroll to top button