Next Article in Journal
Mitoxantrone-Loaded Nanoferritin Slows Tumor Growth and Improves the Overall Survival Rate in a Subcutaneous Pancreatic Cancer Mouse Model
Previous Article in Journal
T-Cell Acute Lymphoblastic Leukemia—Current Concepts in Molecular Biology and Management
Previous Article in Special Issue
OM-85 Broncho-Vaxom®, a Bacterial Lysate, Reduces SARS-CoV-2 Binding Proteins on Human Bronchial Epithelial Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Potential Therapeutic Approaches against SARS-CoV-2: A Review

1
Gachon Institute of Pharmaceutical Science and Department of Pharmacy, College of Pharmacy, Gachon University, Hambakmoeiro 191, Yeonsu-gu, Incheon 21924, Korea
2
Amity Institute of Biotechnology, Amity University Rajasthan, Jaipur 303002, India
3
Plasma Bioscience Research Center, Applied Plasma Medicine Center, Department of Electrical & Biological Physics, Kwangwoon University, Seoul 01897, Korea
4
Department of General, Visceral and Thoracic Surgery, University Medical Center Hamburg-Eppendorf, Martinistrasse 52/Gebäude N27, 20246 Hamburg, Germany
*
Authors to whom correspondence should be addressed.
Biomedicines 2021, 9(11), 1620; https://doi.org/10.3390/biomedicines9111620
Submission received: 15 September 2021 / Revised: 23 October 2021 / Accepted: 30 October 2021 / Published: 4 November 2021
(This article belongs to the Special Issue Drug Development for COVID-19)

Abstract

:
The ongoing SARS-CoV-2 pandemic is a serious threat to public health worldwide and, to date, no effective treatment is available. Thus, we herein review the pharmaceutical approaches to SARS-CoV-2 infection treatment. Numerous candidate medicines that can prevent SARS-CoV-2 infection and replication have been proposed. These medicines include inhibitors of serine protease TMPRSS2 and angiotensin converting enzyme 2 (ACE2). The S protein of SARS-CoV-2 binds to the receptor in host cells. ACE2 inhibitors block TMPRSS2 and S protein priming, thus preventing SARS-CoV-2 entry to host cells. Moreover, antiviral medicines (including the nucleotide analogue remdesivir, the HIV protease inhibitors lopinavir and ritonavir, and wide-spectrum antiviral antibiotics arbidol and favipiravir) have been shown to reduce the dissemination of SARS-CoV-2 as well as morbidity and mortality associated with COVID-19.

Graphical Abstract

1. Introduction

Coronaviruses contain positive-sense, single-stranded RNA, with a genome size ranging from 26–32 kb, and five structural proteins, and are classified into four categories: alpha, beta, gamma, and delta [1,2]. Human coronaviruses are alpha and beta coronaviruses which can cause respiratory and gastrointestinal tract infections [2]. The severe acute respiratory syndrome (SARS) outbreak between November 2002 and July 2003 (nine months) resulted in more than 8000 total cases and 774 deaths, with a fatality rate of 9.6% [3]. Middle East respiratory syndrome (MERS) was reported in 2012 resulting in more than 2400 cases and 858 deaths, with a fatality rate of 34.4%. Subsequently, in late December 2019, an unspecified case of pneumonia was reported in Wuhan, Hubei Province, the People’s Republic of China [1,2,3]. COVID-19 is the official name given by the WHO to the disease caused by SARS-CoV-2 infection. It has since been observed that the virus could spread from human to human [4]. Its incubation period is 2 to 14 days with various clinical presentations: asymptomatic, mild to severe illness, and mortality [5]. Symptoms include fever, cough, difficulty breathing, malaise and fatigue, gastrointestinal symptoms (decreased appetite, vomiting, watery diarrhea, and dehydration), loss of taste and smell, sore throat, rhinorrhoea, severe pneumonia, and acute respiratory distress, which can lead to multiple organ failure and death. The SARS-CoV-2 virus is mainly spread via airborne/aerosol particles; the virus has been observed to remain viable and infective for over 3 h in the air [6,7]. SARS-CoV-2 infection is a highly communicable disease, and this pandemic has been designated a world public health emergency by the World Health Organization (WHO) [7]. However, SARS-CoV-2 has many potential natural, intermediate, and final hosts, as do other viruses; thus, major problems in the prevention and diagnosis of viral infection are raised [8]. In this paper we discuss the genetic structure of SARS-CoV-2 and its mechanism of pathogenesis. We include consideration of the phylogenetic analysis of the SARS-CoV-2 genome, multiple sequence alignment analysis, and therapeutic approaches to SAR-Co-V-2 infection.

2. SARS-CoV-2 Genetic Structure and Pathogenic Mechanism

The SARS-CoV-2 genome codes for more than 20 distinct proteins. At least four structural proteins are present in coronaviruses, namely spike (S), envelope (E), membrane (M), and nucleocapsid (N) proteins (Figure 1). S proteins, which are involved in host attachment and virus-cell membrane fusion, determine the host range for viral infection (Figure 2) [9].
The SARS-CoV-2 main protease (Mpro) is recognised as one of the most essential viral proteins. SARS-CoV-2 Mpro is more than 96% similar to SARS-CoV Mpro. During viral translation, SARS-CoV-2 Mpro cleaves 11 polyproteins to polypeptides that are required for transcription and replication [10]. Some of the candidate drugs that can prevent SARS-CoV-2 viral replication target Mpro, such as remdesivir, griffithsin, nafamostat, disulfiram, lopinavir/ritonavir, nelfinavir, danoprevir and favipiravir [11].

3. Phylogenetic Analysis of SARS-CoV-2 Genome

A sequence alignment and phylogenetic analysis of SARS-CoV-2 genome is shown in Figure 3. The phylogenetic tree is primarily divided into three clades [12]. Clade I consist of SARS-CoV and Bat-SL-CoV genomes which share a sequence identity ranging from 88% to 99%. Clade II consist of 13 complete genomes of coronavirus and MERS-CoV genomes which share a sequence identity from 78% to 89%. Clade III consist of 23 SARS-CoV-2 and Bat-SL-CoV complete genomes which share a sequence identity ranging from 89% to 100%; the SARS-CoV-2 genomes isolated from human samples show a sequence identity ranging from 98% to 100% [13]. A particularly interesting observation from the analysis was that there is no major divergence in the SARS-CoV-2 genome sequence of different SARS-CoV-2 virus genomes isolated from different countries, as shown in Figure 3. The sequence alignment of the SARS-CoV-1 (Bat, PDB ID: 3TNT) and the SARS-CoV-2 (human, PDB ID: 7MBI) main proteases reveals that the amino acid sequence is conserved with a sequence identity of 96%; differences between these genomes are shown in Figure 4 at specific positions [13,14].

4. Therapeutic Approaches to SAR-COV-2 Infection

To identify therapeutic agents that are effective against SARS-CoV-2 infection, extensive research on the structure and pathogenesis of COVID-19 is in progress [15]. Therapeutic approaches to COVID-19 can be categorized into virus-based therapy and host-based therapy, as shown in Figure 5.

4.1. Virus-Based Therapy

Viral nucleic acids consist of nucleosides and nucleotides. Drugs capable of attacking nucleotides, nucleosides, or viral nucleic acids can affect the activity of a broad range of coronaviruses and other viruses, as shown in Table 1 [16]. Possible targets for antiviral therapy include major enzymes and proteins involved in SARS-CoV-2 viral replication. The PLpro enzymes and papain-like protease of SARS-CoV and MERS-CoV have been shown to exert proteolytic, deubiquitylating, and deISGylating activities. Studies have shown that lopinavir-ritonavir is the most potent protease inhibitor as shown in Table 1 [17]. The main SARS-CoV-2 immunogen antigen is the Spike glycoprotein with membrane anchor, which plays an important role in the interaction between host cells and viruses. Studies have shown that certain monoclonal antibodies can target the receptor binding domain (RBD) subunit epitopes and inhibit viral cell receptor binding, whereas other monoclonal antibodies bind to the S2 subunit and disrupt viral cell fusion [18]. A study using the CR3022 neutralising antibody of SARS-CoV shown in Table 1 [19]. Earlier trials also showed that adoptive transfer of plasma containing anti-MERS-COV-S antibodies had the ability to prevent infection and accelerate viral clearance.

4.2. Host-Based Therapy

Viral entry of SARS-CoV-2 depends on the priming of its spike protein and on transmembrane protease 2 (TMPRSS2). Further studies have shown that camostat mesylate, a serine protease inhibitor, can block TMPRSS2 activity and is thus considered as a therapeutic candidate as shown in Table 2 [33]. Other research indicates a pH- and receptor-dependent endocytosis when coronavirus is introduced into the host cell. AP-2-associated protein kinase 1 (AAK1), a host kinase, controls clathrin-mediated endocytosis [34]. Since the virus structure is now established, various inhibitors have been tested in cell-based systems for their ability to prevent viral entry and replication within the host body, as shown in Table 2 [35]. These include spike (S) protein inhibitors, S-cleavage inhibitors, helicase and protease inhibitors, fusion core blockers, HCB monoclonal antibodies, RBD–ACE2 blockers, antiviral peptides, siRNAs, and antifreeze eutralizati antibodies [35,36]. The following section concentrates on the possible therapeutic treatment options based on our limited knowledge of SARS-CoV-2.

4.2.1. Neutralizing Antibodies

In general, coronavirus infection begins with the entry of the viral S protein, which binds to the cell surface. This S protein fuses with the cell membrane and facilitates the syncytial development and transmission of viral nucleocapsids into the cell for further replication [35]. Studies have shown that neutralization of the S protein RBD of SARS-CoV [36] and MERS-CoV [38,39,40] by antibodies can be effective against these diseases. Neutralisation of antigens can be highly useful in COVID-19 treatment, given that the S protein RBD sequence of SARS-CoV-2 is similar to those of SARS-CoV and MERS-CoV [37]. Critical COVID-19 patients are currently treated with immunoglobulin G [35,36]. FcR plays a role in inflammation in the lung; therefore, inflammation in COVID-19 can be reduced by blocking FcR activation. Thus, intravenous administration of immunoglobulins can be effective in the treatment of pulmonary inflammation, as shown in Table 3 and Figure 6 [41].

4.2.2. Antiviral Proteases

Lopinavir has been suggested as a treatment choice for persistent COVID-19 infection [41], as it inhibits the activity of coronavirus proteases. In addition, an inhibitor of vinyl sulfone protease has recently been proposed for treating patients with COVID-19 and can potentially be used widely as an anti-coronaviral agent [42]. Recently, the elucidation of the SARS-CoV-2 Mpro structure has brought about a breakthrough in antiviral studies, and numerous new drug candidates have been developed since. The Deep Docking platform, a virtual screening platform for structural proteins, containing approximately 1.2 billion molecules and 1000 suspected SARS-CoV-2 Mpro proteins, was recently established [42].

4.2.3. Combination Therapy

Combinations of wide-spectrum antivirals, such as lopinavir-ritonavir, neuraminidase inhibitors, peptides (EK1), and RNA synthesis inhibitors, have been tested as anti-viral treatments. The combination of antiviral medicines, however, is controversial [43]. Patients with COVID-19 have been treated with oseltamivir, lopinavir, ritonavir, and ganciclovir [44]. Additional medications have included antibiotics (e.g., cephalosporins, quinolones, carbapenems, methicillin-resistant Staphylococcus aureus tigecycline, linezolid, moxifloxacin, levofloxacin, azithromycin, or amoxicillin), antifungals, and corticosteroids (prednisolone or dexamethasone) [45].
Traditional Indian medicine, which includes Ayurveda, Siddha, Unani and Yoga, naturopathy, and homeopathy, is one of the oldest therapies in human history and has been used to treat various diseases [46]. However, numerous other drugs for COVID-19 patients are still being tested. Teicoplanin is a staphylococcus glycopeptide [45] recommended for the treatment of complex COVID-19 cases and S. aureus infections [45], and has been shown to be effective in treating complications in these patients. In a clinical trial, danoprevir, a viral protease inhibitor, led to substantial improvement in patients with COVID-19 [47]. A summary of the development of therapeutics against COVID-19 is shown in Figure 5.

4.2.4. SARS-CoV-2 Cell Entry Inhibitors

Inhibitors of TMPRSS2 Serine Protease

Transmembrane protease serine subfamily 2 (TMPRSS2), an airway and alveolar cell serine protease, is regulated by the cleavage and activation of the SARS-CoV spike protein (S protein), which is necessary for membrane fusion and host cell entry [48]. Camostat mesylate is a clinically proven serine defence inhibitor that partially prevents infection of HeLa cells expressing the ACE2 and TMPRSS2 by HCoV-NL63 and SARS-CoV-2 [49].

Nafamostat Mesylate

Nafamostat mesylate (FUT-175; CAS number: 81525-10-2) is an artificial serine protease inhibitor clinically approved in Japan for the treatment of acute pancreatitis, intravascular coagulation dissemination, and extracorporeal circulation antioxidation [49,50]. Through screening of approximately 1100 drugs for those that can inhibit the fusion of TMPRSS2, the FDA determined that nafamostat mesylate inhibited the fusion of TMPRSS2-expressing Calu-3 host cells with the MERS-CoV S protein [49,50].

Camostat Mesylate (Foipan™)

In a clinical trial, Foipan™ [(N,N-dimethylcarbamoylmethyl (4-(4-guanidinobenzoyloxy)-phenylacetate)) methanesulfate] [51], also known as camostat mesylate (NI-03; CAS number: 59721-28-7), was administered at a dose of 200 mg three times a day for 2 weeks to 95 patients to investigate its effect against dyspepsia associated with non-alcoholic mild pancreatic disease; the drug showed mild, but no serious, adverse effects, which indicated that camostat mesylate was a well-tolerated medicine [51,52].

Angiotensin Transforming Enzyme 2 (ACE2) Inhibitors and Antimalarial Drugs

SARS-CoV and corresponding coronaviruses interact directly with the host cell exoprotease and metallocarboxypeptidase (ACE2, S) via their S proteins, with angiotensin converting enzyme 2 (ACE2) catalysing the conversion of angiotensin I into nonapeptide angiotensin and angiotensin II into angiotensin 1–7 [53,54]. Unfortunately, the ACE inhibitors commonly used against hypertension and chronic heart failure cannot inhibit ACE2 [53,54]; however, other drugs and compounds have been shown to inhibit ACE2 [53].

Cepharanthine/Selamectin/Mefloquine Hydrochloride

The triple combination of mefloquine hydrochloride (Lariam™, used for the prophylaxis and treatment of malaria), cepharanthine (an anti-inflammatory alkaloid from Stephania cepharantha Hayata; and selamectin (an avermectin isolated from Streptomyces avermitilis and used as an anthelminthic and parasiticide drug in veterinary medicine; CAS number. 220119-17-5) [55,56] have recently been shown to inhibit the infection of simian Vero E6 cells with pangolin coronavirus GX_P2V/2017/Guangxi (GX_P2V), whose S protein shares 92.2% amino acid identity with that of SARS-CoV-2 [55,56].

Chloroquine Phosphate and Hydroxychloroquine

Chloroquine has been used for decades in malaria prophylaxis and for the treatment of chronic Q fever and autoimmune illnesses [57]. Hydroxychloroquine is transformed into chloroquine in vivo. These drugs have recently been shown as possible broad-spectrum antiviral medications [58]. Chloroquine phosphate inhibits ACE2 terminal phosphorylation and increases pH, both of which constitute the antiviral mechanisms of chloroquine and hydroxychloroquine [57,58].

4.2.5. Inhibitors of the Replication, Membrane Fusion, and Assembly of SARS-CoV-2

Lopinavir/Ritonavir (Kaletra™)

Lopinavir (ABT-378) was developed in 1998 to overcome HIV resistance to the protease inhibitor ritonavir (ABT-538), which is caused by a valine mutation at position 82 (Val 82) on an active site of HIV protease in response to ritonavir treatment [38]. As the metabolism of lopinavir is strongly inhibited by ritonavir, lopinavir has been administered at oral dosages exceeding ritonavir by >50-fold in rat, dog, and monkey-plasma for an in vitro EC50 for lopinavir after 8 h [39]. Lopinavir-ritonavir treatment was associated with inconsistencies in clinical improvement, but showed comparable mortality with the standard-care treatment after 28 days of treatment [38,39].

Favipiravir

Favipiravir (Avigan™; T-705) is an oral pyrazinecarboxamide derivative developed by Toyama Chemical, China. It is a potent, selective inhibitor of the RNA-dependent RNA polymerase (RdRp) of RNA viruses, inducing lethal RNA transversion mutations and consequently, producing a nonviable virus phenotype [40]. Favipiravir is effective against various RNA viruses, including influenza A virus, flavivirus, alphavirus, filovirus, bunyavirus, arenavirus, norovirus, West Nile virus, yellow fever virus, foot-and-mouth disease viruses, Ebola virus, and Lassa virus [59].

Umifenovir

Umifenovir(ethyl-6-bromo-4-[(dimethylamino)methyl]-5-hydroxy-1-methyl-2 [(phenylthio) methyl]-indole-3-carboxylate hydrochloride monohydrate; CAS number: 131707-25-0) is a small indole-derivate molecule manufactured by JSC Pharmstandard, Russia [59]. Umifenovir prevents viral host cell entry by inhibiting fusion of the viral envelope and host cell cytoplasmic membrane through inhibition of clathrin-mediated endocytosis, thereby preventing virus infection [60].

Inhibitors of SARS-CoV-2 3CLpro Protease

3CLpro (also known as Mpro) is the major betacoronavirus protease necessary for viral RNA translation in polyprotein synthesis [10,30]. Recently, the x-ray structures and α-ketoamide complex of SARS-CoV-2 3CLpro have been reported [10,30,61]. Two compound pulmonary tropism pyridine-containing α-ketoamides, called 13a and 13b, exerted pharmacokinetic properties at sufficient concentration in the lungs and bronchial-alveolar liquid lavage of mice within 4 to 24 h of administration [47]. N3, formerly known as Michael acceptor inhibitor, was developed using a computer-aided method. N3 can specifically inhibit the Mpro of multiple coronaviruses, including SARS-CoV and MERS-CoV, and has shown potent antiviral activity against infectious bronchitis virus in an animal model [51,52,61]. N3 was shown to form a covalent bond with SARS-CoV-2 3CLpro, thereby acting as an irreversible inhibitor of SARS-CoV and MERS-CoV 3CLpro [51,52].

SARS-CoV-2 Mpro Inhibitor

The SARS-CoV-2 Mpro structure in the apo form and α-ketoamide bound with the protein form a dimer crystallographic composed of two monomers of identical conformations [62]. Each protomer is made up of three domains, as shown in Figure 7A,B, Domain I (residues 8–101) and II (residues 102–184) interface to form the active site of the protein, where α-ketoamide derivative bound complex and composed of a Cys145-His41 dyad with an antiparallel β-barrel structure. Domain III (residues 201–303) contains five α-helices, connected to domain II by a long loop region (residues 185–200), with the substrate-binding site located in a cleft between domain I and domain II. It has shown several highly conserved interactions and forms a catalytic bond with the His41 and Cys145 and adjacent residues in the substrate binding cleft such as Gly143 and Ser144 dyad. SARS-CoV uses a chymotrypsin-like protease (3CLpro) and a papain-like protease (PLpro) to process and cleave its long polyprotein precursor into individually functional non-structural proteins [63,64].
In the active site region of SARS-CoV-2 Mpro, the S1 residues contributed by Cys145, Gly143, and Ser144 also serve as the oxyanion hole. Bulky Gln189 and Pro168 residues form the S4 site, the S1 residue is His163, while Glu166 and Gln189 are located at the S2 position (Figure 8A). The Mpro recognizes and binds specific residues at each subsite of the peptide substrate to determine the initiation of proteolysis and the production of non-structural proteins for the formation of the replication-transcription complex. Crystal structures of Mpro have revealed that the most variable regions are the helical domain III and surface loops, and that the substrate-binding pocket (located in a cleft between domain I and domain II) is highly conserved among Mpro in all coronaviruses; this suggests that antiviral inhibitors targeting this pocket should have wide-spectrum activity against coronaviruses, as shown in Figure 8B. A previous study proposed that Mpro has a substrate-recognition pocket that is highly conserved among all coronaviruses and that this pocket could serve as a drug target for the design of broad-spectrum inhibitors. The recent discovery of new coronaviruses, and the accumulation of structural data for Mpro from coronaviruses of various species, has provided the opportunity to further examine this hypothesis.

5. Nanomaterials Approach for Anti-Coronavirus

Nanomaterials, such as silver colloid, titanium dioxide, and diphyllin nanoparticles, are considered promising antiviral agents and drug-delivery platforms for the effective management of coronavirus infection (Table 4, Table 5 and Table 6) [65,66,67,68,69,70,71,72]. Small interfering RNAs (siRNAs) are highly efficient in reducing the replication of RNA viruses, such as coronaviruses. The efficacy of siRNA-based treatments strictly depends on specific targeting of the viral sequence of interest and targeted cellular delivery of therapeutic siRNA [73]. In this context, nontoxic, biocompatible nanocarriers composed of polymers, lipids, polymer/lipid hybrid nanoparticles, nanohydrogels, silica, dendrimers, iron oxide nanoparticles, or gold nanoparticles are considered as promising siRNA-delivery platforms [74]. These nanocarriers can improve siRNA stability by preventing enzymatic degradation. For inhalable antiviral siRNA loading and aerosol-based delivery of antiviral siRNA in the lungs, polymer/lipid nanocarriers have shown promising outcomes [73,74].
Antiviral nanoparticles have been used as potential immunostimulatory agents for vaccine development. For example, gold nanoparticles conjugated with swine transmissible gastroenteritis virus have been used to activate macrophages, induce interferon production, and increase anti-coronavirus neutralizing antibody levels in vaccinated animals [75]. Similarly, conjugates of ribonucleic acid and ferritin-based nanoparticles have been used as molecular chaperons to develop a vaccine against MERS-CoV [74,75]. The vaccine has been shown to induce a strong T cell response and promote interferon production. Currently, nanotechnology is playing an increasingly important role in antiviral therapy for coronaviruses [73,75]. Nanomaterials have been developed specifically to improve the delivery of biotherapeutics across physiological barriers [73,74,75]. A broad range of potential nanodevices, such as nanosensors, nano-based vaccines, and smart nanomedicines, offer great hope for combating current and future mutated versions of coronaviruses [72].
Table 4. List of Nanomaterial’s approach against coronavirus.
Table 4. List of Nanomaterial’s approach against coronavirus.
NanomaterialChitosan NanospheresCQDsc) Nanocrystals4–Gold NanorodsPLGAg) Hollow NanoparticlesTiO2 NanoparticlesRef
Average size10 nm–10 µm9 nm18–54 nm114 nmNot reported[65,66,67,68,69,70,71,72,73,74,75]
ShapeSphericalSphericalRodSphericalPredominantly spherical
StrategyGenipin-crosslinked chitosanBoronic acid-functionalized CQDs
(Carbon quantum dots)
Gold nanorods-based peptidesViral antigens and STING agonists-loaded hollow nanoparticlesTNPsi)-coated glass coverslips UVC radiation
CoronavirusHCoV NL63 (Human coronavirus NL63) HCoV-229E-LucMERS-CoVMERS-CoVHCoV NL63a)
Potential applicationAdsorbentsAntiviral drugsAntiviral drugsVaccineSelf-cleaning surfaces
Table 5. Nanomaterial’s approach against SARS-CoV-2.
Table 5. Nanomaterial’s approach against SARS-CoV-2.
NanomaterialIron Oxide NanoparticlesPoly Silica NanoparticlesGold NanoparticlesNano-Sized FormazansPolylactic-Co-Glycolic Acid (PLGA)-NanoparticlesSilver NanoparticlesRef.
SizeN/ra210 ± 40 nmN/ra23.75 ± 7.16 nmN/ra [65,66,67,68,69,70,71,72,73,74,75,76]
StrategyNano-mineral structure of Fe2O3 and Fe3O4Optimized polyP encapsulated by SiNPsPeptide-functionalized gold nanoparticlesFormazan analogs by dithizone and α-haloketones reactionOptimized Remdesivir-loaded L-PLGA NPsArtemisinin, Artemether, and Artesunate delivery by silver nanoparticles
Ligand-receptor binding resultsInteractions with S1-RBD of SARS-CoV-2Inhibition of binding of ACE2 to S-protein SARS-CoV-2, at a physiological solutionMore stable complex with RBD of SARS-CoV-2 than ACE2Inhibition of SARS-CoV-2 chymotrypsin-like protease, at a physiological solutionInteractions Lisinopril-ACE1 and remdesivir-intracellular targeting protein RdRpInteractions between negative charges of oxygen atoms of drugs with Ag surface
Potential applicationRepurposing medicationImmunologic agentsAntiviral agentsAntiviral agentsAntiviral therapyAntiviral drugs
Table 6. Nanomaterials based approach to manage SARS-CoV-2.
Table 6. Nanomaterials based approach to manage SARS-CoV-2.
NanomaterialShape, SizeStrategyPotential ApplicationRef.
Graphene sheetsLayersModified graphene sheets conjugated with spike antibody of SARS-CoV-2Immunodetection[77]
Gold nanoparticles2.4 nm Spherical in sizeSulfonated gold nanomaterialsAntiviral agents[66]
Polymeric nanoparticlesSphericalBioinspired DNase-coated melanin-like nanospheresSepsis or acute respiratory distress syndrome (ARDS) in sever COVID-19 Patients[78]
Polymeric nanoparticles10 nm–1 μm colloidal particlesIvermectin-delivery by (Poly (lactic-co-glycolic acid)-b-Maleimide PEG (polyethylene glycol) Copolymers Antiviral drug[79]
Nanostructured lipid carriersSphericalPulmonary delivery of Salinomycin by nanostructured lipid carriersDrug delivery[66]
Polymeric nanoparticlesSpherical 22.0 nmDNase-I-coated polydopamine-PEG poly (ethylene glycol) Sepsis or acute respiratory distress syndrome (ARDS) in sever COVID-19 Patients[80]
Decoy nanoparticles(Not reported)Fusing genetically engineered cell membrane nanovesicles (293T/ACE2 and THP-1cells)Therapeutic vaccines[81]

6. Conclusions

As vaccines will not be available for the general population until the end of 2020, approved off-label and experimental drugs that are effective against COVID-19 have been identified, including TMPRSS2 and ACE2 inhibitors, antifungal and antimalarial products, antiviral medicines inhibiting viral RdRp, proteases, virus/host cell membrane fusion inhibitors, and antiviral phytochemicals. In addition to infection preventive strategies that have been implemented in many countries, such as quarantine of infected persons, mobile monitoring of contacts, protection of persons at high risk of infection, restriction of national curfew, and urgent vaccine production and supply, trials of these drugs have been conducted by the WHO and the European Union. This review of therapeutic approaches to COVID-19 aims to facilitate the production of drugs and vaccines against this disease. In future studies, furin inhibitors may be investigated as a strategy for developing SARS-CoV-2 vaccines.

Author Contributions

D.K.Y. and D.D.S. conceived and designed the project. D.D.S., I.H. and E.-H.C. collected data from the literature. I.H., E.-H.C., D.K.Y., D.D.S. and Y.K. analyzed the data and wrote the manuscript. All authors contributed to the interpretation and discussion of the results. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Foreign Research Institute Recruitment Program through the National Research Foundation of Korea (NRF) funded by the Korean government (MSIP) (NRF-2016K1A4A3914113).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

D.K.Y. is thankful to the Gachon Institute of Pharmaceutical Science and the Department of Pharmacy, College of Pharmacy, Gachon University of Medicine and Science, Incheon, Korea. D.D.S. is thankful to Amity University Rajasthan. We also thankful to Chandni Kumari have collect the data. Figure was created using BioRender (https://biorender.com/ accessed on 16 July 2021).

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. 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. [Google Scholar] [CrossRef] [Green Version]
  2. Zhang, Y.Z.; Holmes, E.C. A genomic perspective on the origin and emergence of SARS-CoV-2. Cell 2020, 181, 223–227. [Google Scholar] [CrossRef] [PubMed]
  3. Lai, C.C.; Liu, Y.H.; Wang, C.Y.; Wang, Y.H.; Hsueh, S.C.; Yen, M.Y.; Ko, W.C.; Hsueh, P.R. Asymptomatic carrier state, acute respiratory disease, and pneumonia due to severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2): Facts and myths. J. Microbiol. Immunol. Infect. 2020, 53, 404–412. [Google Scholar] [CrossRef] [PubMed]
  4. Ruan, Q.; Yang, K.; Wang, W.; Jiang, L.; Song, J. Clinical predictors of mortality due to COVID-19 based on an analysis of data of 150 patients from Wuhan, China. Intensive Care Med. 2020, 46, 846–848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Tian, Y.; Rong, L.; Nian, W.; He, Y. Review article: Gastrointestinal features in COVID-19 and the possibility of faecal transmission. Aliment. Pharmacol. Ther. 2020, 51, 843–851. [Google Scholar] [CrossRef] [PubMed]
  6. Davies, N.G.; Klepac, P.; Liu, Y.; Prem, K.; Jit, M.; CMMID COVID-19 Working Group; Eggo, R.M. Age-dependent effects in the transmission and control of COVID-19 epidemics. Nat. Med. 2020, 26, 1205–1211. [Google Scholar] [CrossRef] [PubMed]
  7. Khan, S.; Siddique, R.; Shereen, M.A.; Ali, A.; Liu, J.; Bai, Q.; Bashir, N.; Xue, M. Emergence of a novel Coronavirus, severe acute respiratory syndrome Coronavirus 2: Biology and therapeutic options. J. Clin. Microbiol. 2020, 58, e00187-20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Singh, D.D.; Jain, A. Multipurpose Instantaneous Microarray Detection of Acute Encephalitis Causing Viruses and Their Expression Profiles. Curr. Microbiol. 2012, 65, 290–303. [Google Scholar] [CrossRef]
  9. Wan, Y.; Shang, J.; Graham, R.; Baric, R.S.; Li, F. Receptor recognition by the novel Coronavirus from Wuhan: An analysis based on decade-long structural studies of SARS Coronavirus. J. Virol. 2020, 94, e00127-20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. 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, 202000028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Pan, H.; Peto, R.; Henao-Restrepo, A.-M.; Karim, Q.A.; Alejandria, M.; García, C.H.; Kieny, M.-P.; Malekzadeh, R.; Murthy, S.; Preziosi, M.-P.; et al. Repurposed antiviral drugs for COVID-19—Interim WHO Solidarity Trial Results. N. Engl. J. Med. 2021, 384, 497–511. [Google Scholar]
  12. Lam, T.T.-Y.; Shum, M.H.-H.; Zhu, H.-C.; Tong, Y.-G.; Ni, X.-B.; Liao, Y.-S.; Wei, W.; Cheung, W.Y.-M.; Li, W.-J.; Li, L.-F.; et al. Identifying SARS-CoV-2 related coronaviruses in Malayan pangolins. Nature 2020, 583, 282–285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Khateeb, J.; Li, Y.; Zhang, H. Emerging SARS-CoV-2 variants of concern and potential intervention approaches. Crit. Care 2021, 25, 244. [Google Scholar] [CrossRef] [PubMed]
  14. Monteil, V.; Kwon, H.; Patricia, P.; Prado, P.; Hagelkrüys, A.; Wimmer, R.A.; Stahl, M.; Leopoldi, A.; Garreta, E.; Hurtado del Pozo, C.; et al. Inhibition of SARS-CoV-2 infections in engineered human tissues using clinical-grade soluble human ACE2. Cell Press 2020, 181, 905–913. [Google Scholar] [CrossRef]
  15. Zhou, Y.; Hou, Y.; Shen, J.; Huang, Y.; Martin, W.; Cheng, F. Network-based drug repurposing for novel coronavirus 2019-nCoV/SARS-CoV-2. Cell Discov. 2020, 6, 14–18. [Google Scholar] [CrossRef] [Green Version]
  16. Rabaan, A.A.; Al-Ahmed, S.H.; Haque, S.; Sah, R.; Tiwari, R.; Malik, Y.S.; Dhama, K.; Yatoo, M.I.; Bonilla-Aldana, D.K.; RodriguezMorales, A.J.; et al. SARS-CoV-2, SARS-CoV, and MERS-CoV: A comparative overview. Infez. Med. 2020, 28, 174–184. [Google Scholar] [PubMed]
  17. Ko, M.; Chang, S.; Byun, S.; Ianevski, A.; Choi, I.; D’Orengiani, A.-L.P.H.D.; Ravlo, E.; Wang, W.; Bjørås, M.; Kainov, D.; et al. Screening of FDA-Approved Drugs Using a MERS-CoV Clinical isolate from South korea identifies potential therapeutic options for COVID-19. Viruses 2021, 13, 651. [Google Scholar] [CrossRef]
  18. Yadav, R.; Chaudhary, J.; Jain, N.; Chaudhary, P.; Khanra, S.; Dhamija, P.; Sharma, A.; Kumar, A.; Handu, S. Role of structural and non-structural proteins and therapeutic targets of SARS-CoV-2 for COVID-19. Cells 2021, 10, 821. [Google Scholar] [CrossRef]
  19. Finkelstein, M.T.; Mermelstein, A.G.; Parker Miller, E.; Seth, P.C.; Stancofski, E.D.; Fera, D. Structural analysis of neutralizing epitopes of the SARS-CoV-2 spike to guide therapy and vaccine design strategies. Viruses 2021, 13, 134. [Google Scholar] [CrossRef]
  20. Lai, C.-C.; Chen, C.-H.; Wang, C.-Y.; Chen, K.-H.; Wang, Y.-H.; Hsueh, P.-R. Clinical efficacy and safety of remdesivir in patients with COVID-19: A systematic review and network meta-analysis of randomized controlled trials. J. Antimicrob. Chemother. 2021, 76, 1962–1968. [Google Scholar] [CrossRef]
  21. Kocayiğit, H.; Süner, K.Ö.; Tomak, Y.; Demir, G.; Yaylacı, S.; Dheir, H.; Güçlü, E.; Erdem, A.F. Observational study of the effects of Favipiravir vs Lopinavir/Ritonavir on clinical outcomes in critically Ill patients with COVID-19. J. Clin. Pharm. Ther. 2021, 46, 454–459. [Google Scholar] [CrossRef]
  22. Saw, P.E.; Song, E.-W. siRNA therapeutics: A clinical reality. Sci. China Life Sci. 2020, 63, 485–500. [Google Scholar] [CrossRef] [PubMed]
  23. Zanella, I.; Zizioli, D.; Castelli, F.; Quiros-Roldan, E. Tenofovir, another inexpensive, well-known and widely available old drug repurposed for SARS-CoV-2 infection. Pharmaceuticals 2021, 14, 454. [Google Scholar] [CrossRef]
  24. Dejmek, M.; Konkol’ová, E.; Eyer, L.; Straková, P.; Svoboda, P.; Šála, M.; Krejčová, K.; Růžek, D.; Boura, E.; Nencka, R. Non-Nucleotide RNA-Dependent RNA Polymerase Inhibitor That Blocks SARS-CoV-2 Replication. Viruses 2021, 13, 1585. [Google Scholar] [CrossRef]
  25. Singh, D.D.; Han, I.; Choi, E.H.; Yadav, D.K. Immunopathology, host-virus genome interactions, and effective vaccine development in SARS-CoV-2. Comput. Struct. Biotechnol. J. 2020, 18, 3774–3787. [Google Scholar] [CrossRef]
  26. Hurt, A.; Wheatley, A. Neutralizing antibody therapeutics for COVID-19. Viruses 2021, 13, 628. [Google Scholar] [CrossRef]
  27. Li, X.; Zhang, L.; Chen, S.; Ouyang, H.; Ren, L. Possible targets of Pan-coronavirus antiviral strategies for emerging or re-emerging Coronaviruses. Microorganisms 2021, 9, 1479. [Google Scholar] [CrossRef] [PubMed]
  28. Stoddard, S.V.; Wallace, F.E.; Stoddard, S.D.; Cheng, Q.; Acosta, D.; Barzani, S.; Bobay, M.; Briant, J.; Cisneros, C.; Feinstein, S.; et al. In silico design of peptide-based SARS-CoV-2 fusion inhibitors that target wt and mutant versions of SARS-CoV-2 HR1 Domains. Biophysica 2021, 1, 311–327. [Google Scholar] [CrossRef]
  29. Yamamoto, M.; Kiso, M.; Sakai-Tagawa, Y.; Iwatsuki-Horimoto, K.; Imai, M.; Takeda, M.; Kinoshita, N.; Ohmagari, N.; Gohda, J.; Semba, K.; et al. The anticoagulant nafamostat potently inhibits SARS-CoV-2 s protein-mediated fusion in a cell fusion assay system and viral infection in vitro in a cell-type-dependent manner. Viruses 2020, 12, 629. [Google Scholar] [CrossRef]
  30. Sardanelli, A.M.; Isgrò, C.; Palese, L.L. SARS-CoV-2 main protease active site ligands in the human metabolome. Molecules 2021, 26, 1409. [Google Scholar] [CrossRef]
  31. Citarella, A.; Scala, A.; Piperno, A.; Micale, N. SARS-CoV-2 Mpro: A potential target for peptidomimetics and small-molecule inhibitors. Biomolecules 2021, 11, 607. [Google Scholar] [CrossRef] [PubMed]
  32. Riva, A.; Conti, F.; Bernacchia, D.; Pezzati, L.; Sollima, S.; Merli, S.; Siano, M.; Lupo, A.; Rusconi, S.; Cattaneo, D.; et al. Da-runavir does not prevent SARS-CoV-2 infection in HIV patients. Pharmacol. Res. 2020, 157, 104826. [Google Scholar] [CrossRef] [PubMed]
  33. Nersisyan, S.; Shkurnikov, M.; Turchinovich, A.; Knyazev, E.; Tonevitsky, A. Integrative analysis of miRNA and mRNA sequencing data reveals potential regulatory mechanisms of ACE2 and TMPRSS2. PLoS ONE 2020, 15, e0235987. [Google Scholar] [CrossRef] [PubMed]
  34. Gatti, M.; Turrini, E.; Raschi, E.; Sestili, P.; Fimognari, C. Janus kinase inhibitors and Coronavirus dDisease (COVID)-19: Rationale, clinical evidence and safety issues. Pharmaceuticals 2021, 14, 738. [Google Scholar] [CrossRef] [PubMed]
  35. Arisan, E.D.; Dart, A.; Grant, G.H.; Arisan, S.; Cuhadaroglu, S.; Lange, S.; Uysal-Onganer, P. The prediction of miRNAs in SARS-CoV-2 genomes: Hsa-miR databases identify 7 Key miRs linked to host responses and virus pathogenicity-related KEGG pathways significant for comorbidities. Viruses 2020, 12, 614. [Google Scholar] [CrossRef]
  36. Chien, M.; Anderson, T.K.; Jockusch, S.; Tao, C.; Li, X.; Kumar, S.; Russo, J.J.; Kirchdoerfer, R.N.; Ju, J. Nucleotide Analogues as Inhibitors of SARS-CoV-2 Polymerase, a Key Drug Target for COVID-19. J. Proteome Res. 2020, 19, 4690–4697. [Google Scholar] [CrossRef]
  37. Belouzard, S.; Millet, J.K.; Licitra, B.N.; Whittaker, G.R. Mechanisms of Coronavirus cell entry mediated by the viral spike protein. Viruses 2012, 4, 1011–1033. [Google Scholar] [CrossRef] [Green Version]
  38. Gil Martínez, V.; Avedillo Salas, A.; Santander Ballestín, S. Antiviral therapeutic approaches for SARS-CoV-2 infection: A systematic review. Pharmaceuticals 2021, 14, 736. [Google Scholar] [CrossRef] [PubMed]
  39. Janik, E.; Niemcewicz, M.; Podogrocki, M.; Saluk-Bijak, J.; Bijak, M. Existing drugs considered as promising in COVID-19 therapy. Int. J. Mol. Sci. 2021, 22, 5434. [Google Scholar] [CrossRef] [PubMed]
  40. Malhani, A.A.; Enani, M.A.; Sharif-Askari, F.S.; Alghareeb, M.R.; Bin-Brikan, R.T.; AlShahrani, S.A.; Halwani, R.; Tleyjeh, I.M. Combination of (interferon beta-1b, lopinavir/ritonavir and ribavirin) versus favipiravir in hospitalized patients with non-critical COVID-19: A cohort study. PLoS ONE 2021, 16, e0252984. [Google Scholar] [CrossRef] [PubMed]
  41. Jonsdottir, H.R.; Bielecki, M.; Siegrist, D.; Buehrer, T.W.; Züst, R.; Deuel, J.W. Titers of neutralizing antibodies against SARS-CoV-2 are independent of symptoms of non-severe COVID-19 in young adults. Viruses 2021, 13, 284. [Google Scholar] [CrossRef] [PubMed]
  42. Magro, P.; Zanella, I.; Pescarolo, M.; Castelli, F.; Quiros-Roldan, E. Lopinavir/ritonavir: Repurposing an old drug for HIV infection in COVID-19 treatment. Biomed. J. 2021, 44, 43–53. [Google Scholar] [CrossRef]
  43. Tampere, M.; Pettke, A.; Salata, C.; Wallner, O.; Koolmeister, T.; Cazares-Körner, A.; Visnes, T.; Hesselman, M.C.; Kunold, E.; Wiita, E.; et al. Novel broad-spectrum antiviral inhibitors targeting host factors essential for replication of pathogenic RNA viruses. Viruses 2020, 12, 1423. [Google Scholar] [CrossRef] [PubMed]
  44. Jeong, G.U.; Song, H.; Yoon, G.Y.; Kim, D.; Kwon, Y.-C. Therapeutic strategies against COVID-19 and structural characterization of SARS-CoV-2: A Review. Front. Microbiol. 2020, 11, 1723. [Google Scholar] [CrossRef] [PubMed]
  45. Singh, D.D.; Han, I.; Choi, E.H.; Yadav, D.K. Recent advances in pathophysiology, drug development and future perspectives of SARS-CoV-2. Front. Cell Dev. Biol. 2020, 8, 580202. [Google Scholar] [CrossRef] [PubMed]
  46. Singh, D.D. Assessment of antimicrobial activity of hundreds extract of twenty Indian medicinal plants. Biomed. Res. 2018, 29, 1797–1814. [Google Scholar] [CrossRef] [Green Version]
  47. Kumar, S.; Zhi, K.; Mukherji, A.; Gerth, K. Repurposing antiviral protease inhibitors using extracellular vesicles for potential therapy of COVID-19. Viruses 2020, 12, 486. [Google Scholar] [CrossRef]
  48. Xiao, L.; Sakagami, H.; Miwa, N. ACE2: The key molecule for understanding the pathophysiology of severe and critical conditions of COVID-19: Demon or Angel? Viruses 2020, 12, 491. [Google Scholar] [CrossRef] [PubMed]
  49. Amraei, R.; Rahimi, N. COVID-19, renin-angiotensin system and endothelial dysfunction. Cells 2020, 9, 1652. [Google Scholar] [CrossRef]
  50. Eze, P.; Mezue, K.N.; Nduka, C.U.; Obianyo, I.; Egbuche, O. Efficacy and safety of chloroquine and hydroxychloroquine for treatment of COVID-19 patients-a systematic review and meta-analysis of randomized controlled trials. Am. J. Cardiovasc. Dis. 2021, 11, 93–107. [Google Scholar]
  51. Huang, X.; Pearce, R.; Omenn, G.; Zhang, Y. Identification of 13 Guanidinobenzoyl- or Aminidinobenzoyl-containing drugs to potentially inhibit TMPRSS2 for COVID-19 treatment. Int. J. Mol. Sci. 2021, 22, 7060. [Google Scholar] [CrossRef]
  52. Hoffmann, M.; Hofmann-Winkler, H.; Smith, J.C.; Kruger, N.; Arora, P.; Sorensen, L.K.; Sogaard, O.S.; Hasselstrom, J.B.; Winkler, M.; Hempel, T.; et al. Camostat mesylate inhibits SARS-CoV-2 activation by TMPRSS2-related proteases and its metabolite GBPA exerts antiviral activity. EBioMedicine 2021, 65, 103255. [Google Scholar] [CrossRef] [PubMed]
  53. Hoffmann, M.; Kleine-Weber, H.; Schroeder, S.; Kruger, N.; Herrler, T.; Erichsen, S.; Schiergens, T.S.; Herrler, G.; Wu, N.H.; Nitsche, A.; et al. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and is blocked by a clinically proven protease inhibitor. Cell 2020, 181, 271–280.e8. [Google Scholar] [CrossRef] [PubMed]
  54. Shang, J.; Wan, Y.; Luo, C.; Ye, G.; Geng, Q.; Auerbach, A.; Li, F. Cell entry mechanisms of SARS-CoV-2. Proc. Natl. Acad. Sci. USA 2020, 117, 11727–11734. [Google Scholar] [CrossRef]
  55. Shah, B.; Modi, P.; Sagar, S.R. In silico studies on therapeutic agents for COVID-19: Drug repurposing approach. Life Sci. 2020, 252, 117652. [Google Scholar] [CrossRef]
  56. Zandi, K.; Amblard, F.; Musall, K.; Downs-Bowen, J.; Kleinbard, R.; Oo, A.; Cao, D.; Liang, B.; Russell, O.O.; McBrayer, T.; et al. Repurposing nucleoside analogs for human Coronaviruses. Antimicrob. Agents Chemother. 2020, 65, e01652-20. [Google Scholar] [CrossRef] [PubMed]
  57. Touret, F.; de Lamballerie, X. Of chloroquine and COVID-19. Antiv. Res. 2020, 177, 104762. [Google Scholar] [CrossRef]
  58. Ho, T.-C.; Wang, Y.-H.; Chen, Y.-L.; Tsai, W.-C.; Lee, C.-H.; Chuang, K.-P.; Chen, Y.-M.; Yuan, C.-H.; Ho, S.-Y.; Yang, M.-H.; et al. Chloroquine and hydroxychloroquine: Efficacy in the treatment of the COVID-19. Pathogens 2021, 10, 217. [Google Scholar] [CrossRef]
  59. Marcianò, G.; Roberti, R.; Palleria, C.; Mirra, D.; Rania, V.; Casarella, A.; De Sarro, G.; Gallelli, L. SARS-CoV-2 Treatment: Current therapeutic options and the pursuit of tailored therapy. Appl. Sci. 2021, 11, 7457. [Google Scholar] [CrossRef]
  60. Asselah, T.; Durantel, D.; Pasmant, E.; Lau, G.; Schinazi, R.F. COVID-19: Discovery, diagnostics and drug development. J. Hepatol. 2021, 74, 168–184. [Google Scholar] [CrossRef]
  61. Zhang, L.; Lin, D.; Sun, X.; Curth, U.; Drosten, C.; Sauerhering, L.; Becker, S.; Rox, K.; Hilgenfeld, R. Crystal structure of SARS-CoV-2 main protease provides a basis for design of improved α-ketoamide inhibitors. Science 2020, 368, 409–412. [Google Scholar] [CrossRef] [Green Version]
  62. Kneller, D.W.; Phillips, G.; O’Neill, H.M.; Jedrzejczak, R.; Stols, L.; Langan, P.; Joachmiak, A.; Coates, L.; Kovalevsky, A. Structural plasticity of SARS-CoV-2 3CL Mpro active site cavity revealed by room temperature X-ray crystallography. Nat. Commun. 2020, 11, 3202. [Google Scholar] [CrossRef] [PubMed]
  63. Kumar, Y.; Singh, H.; Patel, C.N. In silico prediction of potential inhibitors for the main protease of SARS-CoV-2 using molecular docking and dynamics simulation-based drug-repurposing. J. Infect. Public Health 2020, 13, 1210–1223. [Google Scholar] [CrossRef] [PubMed]
  64. Ianevski, A.; Yao, R.; Biza, S.; Zusinaite, E.; Mannik, A.; Kivi, G.; Planken, A.; Kurg, K.; Tombak, E.-M.; Ustav, M.; et al. Identification and tracking of antiviral drug combinations. Viruses 2020, 12, 1178. [Google Scholar] [CrossRef] [PubMed]
  65. Stepniowski, W.J.; Misiolek, W.Z. Review of fabrication methods, physical properties, and applications of nanostructured copper oxides formed via electrochemical oxidation. Nanomaterials 2018, 8, 379. [Google Scholar] [CrossRef] [Green Version]
  66. Carvalho, J.P.F.; Silva, A.C.Q.; Silvestre, A.J.D.; Freire, C.S.R.; Vilela, C. Spherical Cellulose Micro and Nanoparticles: A Review of Recent Developments and Applications. Nanomaterials 2021, 11, 2744. [Google Scholar] [CrossRef] [PubMed]
  67. Bhardwaj, S.K.; Bhardwaj, N.; Kumar, V.; Bhatt, D.; Azzouz, A.; Bhaumik, J.; Kim, K.-H.; Deep, A. Recent progress in nanomaterial-based sensing of airborne viral and bacterial pathogens. Environ. Int. 2021, 146, 106183. [Google Scholar] [CrossRef] [PubMed]
  68. Zeng, M.; Zhang, Y. Colloidal nanoparticle inks for printing functional devices: Emerging trends and future prospects. J. Mater. Chem. A 2019, 7, 23301–23336. [Google Scholar] [CrossRef]
  69. Zeng, M.; Chen, M.; Huang, D.; Lei, S.; Zhang, X.; Wang, L.; Cheng, Z. Engineered two-dimensional nanomaterials: An emerging paradigm for water purification and monitoring. Mater. Horiz. 2021, 8, 758–802. [Google Scholar] [CrossRef]
  70. De la Sota, P.G.; Lorente, E.; Notario, L.; Mir, C.; Zaragoza, O.; López, D. Mitoxantrone Shows in vitro, but not in vivo antiviral activity against human respiratory syncytial virus. Biomedicines 2021, 9, 1176. [Google Scholar] [CrossRef]
  71. Mosselhy, D.; Virtanen, J.; Kant, R.; He, W.; Elbahri, M.; Sironen, T. COVID-19 pandemic: What about the safety of Anti-Coronavirus nanoparticles? Nanomaterials 2021, 11, 796. [Google Scholar] [CrossRef]
  72. Sivasankarapillai, V.S.; Pillai, A.M.; Rahdar, A.; Sobha, A.P.; Das, S.S.; Mitropoulos, A.C.; Mokarrar, M.H.; Kyzas, G.Z. On facing the SARS-CoV-2 (COVID-19) with combination of nanomaterials and medicine: Possible strategies and first challenges. Nanomaterials 2020, 10, 852. [Google Scholar] [CrossRef] [PubMed]
  73. Tolksdorf, B.; Nie, C.; Niemeyer, D.; Röhrs, V.; Berg, J.; Lauster, D.; Adler, J.M.; Haag, R.; Trimpert, J.; Kaufer, B.; et al. Inhibition of SARS-CoV-2 replication by a small interfering RNA targeting the leader sequence. Viruses 2021, 13, 2030. [Google Scholar] [CrossRef] [PubMed]
  74. Tang, Z.; Ma, Q.; Chen, X.; Chen, T.; Ying, Y.; Xi, X.; Wang, L.; Ma, C.; Shaw, C.; Zhou, M. Recent advances and challenges in nanodelivery systems for antimicrobial peptides (AMPs). Antibiotics 2021, 10, 990. [Google Scholar] [CrossRef] [PubMed]
  75. Nasrollahzadeh, M.; Sajjadi, M.; Soufi, G.J.; Iravani, S.; Varma, R.S. Nanomaterials and nanotechnology-associated innovations against viral infections with a focus on Coronaviruses. Nanomaterials 2020, 10, 1072. [Google Scholar] [CrossRef] [PubMed]
  76. Perrella, F.; Coppola, F.; Petrone, A.; Platella, C.; Montesarchio, D.; Stringaro, A.; Ravagnan, G.; Fuggetta, M.; Rega, N.; Musumeci, D. Interference of Polydatin/Resveratrol in the ACE2: Spike recognition during COVID-19 infection. A focus on their potential mechanism of action through computational and biochemical assays. Biomolecules 2021, 11, 1048. [Google Scholar] [CrossRef]
  77. Cordaro, A.; Neri, G.; Sciortino, M.T.; Scala, A.; Piperno, A. Graphene-based strategies in liquid biopsy and in viral diseases diagnosis. Nanomaterials 2020, 10, 1014. [Google Scholar] [CrossRef]
  78. Kassirian, S.; Taneja, R.; Mehta, S. Diagnosis and management of acute respiratory distress syndrome in a time of COVID-19. Diagnostics 2020, 10, 1053. [Google Scholar] [CrossRef]
  79. Begines, B.; Ortiz, T.; Pérez-Aranda, M.; Martínez, G.; Merinero, M.; Argüelles-Arias, F.; Alcudia, A. Polymeric nanoparticles for drug delivery: Recent developments and future prospects. Nanomaterials 2020, 10, 1403. [Google Scholar] [CrossRef]
  80. Rasmi, Y.; Saloua, K.; Nemati, M.; Choi, J. Recent progress in nanotechnology for COVID-19 prevention, diagnostics and treatment. Nanomaterials 2021, 11, 1788. [Google Scholar] [CrossRef] [PubMed]
  81. Verma, R.; Sahu, R.; Singh, D.D.; Egbo, T.E. A CRISPR/Cas9 based polymeric nanoparticles to treat/inhibit microbial infections. Semin. Cell Dev. Biol. 2019, 96, 44–52. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Genome structure of SARS-CoV-2. Figure was created by using BioRender (https://biorender.com, accessed on 15 September 2021).
Figure 1. Genome structure of SARS-CoV-2. Figure was created by using BioRender (https://biorender.com, accessed on 15 September 2021).
Biomedicines 09 01620 g001
Figure 2. Crystallographic structure SARS-CoV-2. Figure was created using by BioRender (https://biorender.com, accessed on 15 September 2021).
Figure 2. Crystallographic structure SARS-CoV-2. Figure was created using by BioRender (https://biorender.com, accessed on 15 September 2021).
Biomedicines 09 01620 g002
Figure 3. The phylogenetic tree was generated using the latest complete genome sequences of different neighbors, MERS-CoV, SARS-CoV, and Bat-SL-CoV. The tree is divided into three major clades according to the grouping of clusters: Clade I: Bat-SL-CoV-2 and SARS-CoV viruses showing a close evolutionary relationship with each other. Clade II: Human and bat coronaviruses, including MERS-CoV. Clade III: All of the SARS-CoV-2 genomes isolated from humans—it was observed that these genomes show a close evolutionary relationship with Bat-SL-CoV-2.
Figure 3. The phylogenetic tree was generated using the latest complete genome sequences of different neighbors, MERS-CoV, SARS-CoV, and Bat-SL-CoV. The tree is divided into three major clades according to the grouping of clusters: Clade I: Bat-SL-CoV-2 and SARS-CoV viruses showing a close evolutionary relationship with each other. Clade II: Human and bat coronaviruses, including MERS-CoV. Clade III: All of the SARS-CoV-2 genomes isolated from humans—it was observed that these genomes show a close evolutionary relationship with Bat-SL-CoV-2.
Biomedicines 09 01620 g003
Figure 4. Multiple sequence alignment analysis of the amino acid sequence of SARS-CoV-1 and SARS-CoV-2 Mpro. Amino acids marked underneath with an arrow represent catalytic residues; residues marked underneath with * represent substrate-binding residues of various subsites.
Figure 4. Multiple sequence alignment analysis of the amino acid sequence of SARS-CoV-1 and SARS-CoV-2 Mpro. Amino acids marked underneath with an arrow represent catalytic residues; residues marked underneath with * represent substrate-binding residues of various subsites.
Biomedicines 09 01620 g004
Figure 5. Therapeutic approaches to SARS-CoV-2 infection.
Figure 5. Therapeutic approaches to SARS-CoV-2 infection.
Biomedicines 09 01620 g005
Figure 6. Schematic of binding mechanism of SARS-CoV-2 spike protein to the receptor.
Figure 6. Schematic of binding mechanism of SARS-CoV-2 spike protein to the receptor.
Biomedicines 09 01620 g006
Figure 7. (A) Structural features of the main protease of SARS-CoV-2 dimer. A SARS-CoV-2 main protease. (B) Sphere representation of main protease monomer, in the active site groove.
Figure 7. (A) Structural features of the main protease of SARS-CoV-2 dimer. A SARS-CoV-2 main protease. (B) Sphere representation of main protease monomer, in the active site groove.
Biomedicines 09 01620 g007
Figure 8. (A) Structural features of the main protease of SARS-CoV-2 monomer and consists of three domains. A linker joins domain II to domain III, which is critical for the dimerization of protein. (B) Different S1′, S1, S2, S3 and S4 subsites groups in the substrate-binding subsites of SARS-CoV-2 MPro.
Figure 8. (A) Structural features of the main protease of SARS-CoV-2 monomer and consists of three domains. A linker joins domain II to domain III, which is critical for the dimerization of protein. (B) Different S1′, S1, S2, S3 and S4 subsites groups in the substrate-binding subsites of SARS-CoV-2 MPro.
Biomedicines 09 01620 g008
Table 1. Virus-based therapy: Drugs capable of attacking nucleotides, nucleosides, or viral nucleic acids of a broad range of coronaviruses and other viruses.
Table 1. Virus-based therapy: Drugs capable of attacking nucleotides, nucleosides, or viral nucleic acids of a broad range of coronaviruses and other viruses.
Antiviral AgentDrug TargetMechanism of ActionInfectious DiseaseReferences
RemdesivirRdRpTerminates the non-obligate chainSARS-CoV-2, MERS-CoV, SARS-CoV[20]
FavipiravirRdRpInhibits RdRpSARS-CoV-2, Influenza[21]
siRNARdRpShort chains of dsRNA that interfereSARS-CoV, MERS-CoVWu[22]
GalidesivirRdRpInhibits viral RNA polymerase function byGalidesivir SARS-CoV-2,[23]
RibavirinRdRpInhibits viral RNA synthesis and mRNA cappingSARS-CoV-2, MERS-CoV, SARS-CoV,[24]
LJ001 and JL103Lipid membraneMembrane-binding photosensitizers that induceEnveloped viruses (IAV, filoviruses, poxviruses, arenaviruses, bunyaviruses, paramyxoviruses, flaviviruses and HIV-1)[25]
CR3022Spike glycoproteinImmunogenic antigen against Spike proteinSARS-CoV-2, SARS-CoV[26]
GriffithsinSpike glycoproteinGriffithsin binds to the SARSCoV-2 spikeSARS-CoV-2[27]
Peptide (P9)Spike glycoproteinInhibits spike protein-mediated cell-cell entry orBroad-spectrum (SARS-CoV, MERS-CoV, influenza)[28]
NafamostatSpike glycoproteinInhibits spike-mediated membrane fusion ASARS-CoV-2, MERS-CoV[29]
Ritonavir3CLproInhibits 3CLproSARS-CoV-2, MERS-CoV[30]
Lopinavir3CLproInhibits 3CLproSARS-CoV-2, MERS-CoV, SARS-CoV, HCoV-229E, HIV, HPV[31]
Darunavir and cobicistat3CLproInhibits 3CLproSARS-CoV-2[32]
Table 2. Host-based therapy: Drug target and mechanism of action against infectious diseases.
Table 2. Host-based therapy: Drug target and mechanism of action against infectious diseases.
Antiviral AgentDrug TargetMechanism of ActionInfectious DiseaseReferences
BaricitinibClathrin-mediated endocytosisBaricitinibClathrin-mediated endocytosis[34]
ChloroquineEndosomal
acidification
A lysosomatropic base that appears to disrupt intracellular trafficking and viral fusion eventsSARS-CoV-2, SARS-CoV, MERS-CoV[33]
Convalescent plasma-Inhibits virus entry to the target cellsSARS-CoV, SARS-CoV-2, Influenza[35,36]
Camostat MesylateSurface proteasePotent serine protease inhibitorSARS-CoV, MERS-CoV, HcoV-229E[33]
CorticosteroidsPulsed methylprednisolonePatients with severe MERS who were treated with systemic corticosteroid with or without antivirals and interferons had no favorable responseSARS-CoV, MERS-CoVL[35]
NitazoxanideInterferon responseInduces the host innate immune response Coronaviruses, SARS-CoV-2[19]
Recombinant interferonsInterferon responseExogenous interferonsSARS-CoV-2, SARS-CoV, MERS-CoV[37]
Table 3. Neutralizing antibodies against SARS-CoV-2.
Table 3. Neutralizing antibodies against SARS-CoV-2.
S.N.Antibody NameAntibody TypeOriginPDB IDEpitopesNeutralizing MechanismCross Neutralizing ActivityProtective EfficacyRef
1CV30Human IgGInfected COVID-19 patients6XE1D420-Y421, Y453, L455-N460, Y473-S477, F486-N487, Y489, Q493, T500, G502, Y505Block hACE2-RBD interactionnoIC50 value of 0.03 µg/mL[35]
2REGN10933 Recombinantfull-human antibodiesHumanized mice and COVID-19-convalescent patients6XDGR403, K417, Y421, Y453, L455-F456, A475-G476, E484-Y489, Q493Block hACE2-RBD interaction, ADCC & ADCPnoIC50 value of 37.4 pM
3B38Human IgGCOVID-19-convalescent patient7BZ5R403, D405-E406, Q409, D420-Y421, Y452, L454-N460, Y473-S477, F486-N487, Y489-F490, Q493-G496, Q498, T500-V503, Y505Block hACE2-RBD interactionnoA single dose of B38 (25 mg/kg)[35]
4CC12.1Human IgGCOVID-19-convalescent patient6XC3R403, D405-E406, R408-Q409, D420-Y421, Y453, L455-N460, Y473-S477, F486-N487, Y489, Q493-G496, Q498, T500-V503, Y505Block hACE2-RBD interactionnoIC50 value of 0.019 µg/mL[36]
5CB6Human IgGCOVID-19-convalescent patient7C01R403, D405-E406, R408-Q409, D420-Y421, L455-N460, Y473-S477, F486-N487, Y489, Q493, Y495, N501-G502, G504-Y505Block hACE2-RBD interactionnoA single dose of CB6-LALA (50 mg/kg)[37]
6C105Human IgGCOVID-19-convalescent patient6XCN, 6XCMR403, D405, R408, D420-Y421, Y453, L455-N460, Y473, A475-G476, F486-N487, G502, Y505Block hACE2-RBD interactionnoIC50 value of 26.1 ng/mL[41]
7CC12.3Human IgGCOVID-19-convalescent patient6XC7R403, D405, D420-Y421, Y453, L455-N460, Y473-S477, F486-N487, Y489, Q493, G496, N501, Y505Block hACE2-RBD interactionnoIC50 value of 0.018 µg/mL[42]
8CR3022Human IgGSARS-convalescent patient6YOR, 6 W41Y369-N370, F374-K386, L390, F392, D428, T430, F515-L517Trapping RBD in the less stable up conformation while leading to destabilization of SSARS-CoV, SARS-CoV-2ND50 value of 0.114 µg/mL[19]
9EY6AHuman IgGLate-stage COVID-19 patient6ZDH, 6ZER, 6ZCZY369, F374-S375, F377-K386, N388, L390, P412-G413, D427-F429, L517destabilization of SSARS-CoV, SARS-CoV-2ND50 value of ~10.8 µg/mL[26]
10VHH-72Llama single domain antibodyllama immunized with prefusionstabilized betacoronavirus spikes6WAQY356-T359, F361-C366, A371-T372, G391-D392, R395, N424, I489, Y494Trapping RBD in the less stable up conformation while leading to destabilization of S, Block hACE2_RBD interactionSARS-CoV, SARS-Co-V-2IC50 values of 0.14 µg/mL and 0.2 mg/mL.[19]
11BD23Human IgGCOVID-19-convalescent patient7BYRG446, Y449, L452, T470, E484-F486, Y489-F490, L492-S494, G496, Q498, T500-N501, Y505Block hACE-RBD2 interactionnoIC50 value of 8.5 µg/mL[26]
12Fab 2–4Human IgGInfected COVID-19 patients6XEYY449, Y453, L455-F456, E484-F486, Y489-F490, L492-S494, G496Locking RBD in the down conformation while occluding access to ACE2noNeutralizing SARS-CoV-2 live virus with IC50 value of 0.057 µg/mL[41]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yadav, D.K.; Singh, D.D.; Han, I.; Kumar, Y.; Choi, E.-H. Current Potential Therapeutic Approaches against SARS-CoV-2: A Review. Biomedicines 2021, 9, 1620. https://doi.org/10.3390/biomedicines9111620

AMA Style

Yadav DK, Singh DD, Han I, Kumar Y, Choi E-H. Current Potential Therapeutic Approaches against SARS-CoV-2: A Review. Biomedicines. 2021; 9(11):1620. https://doi.org/10.3390/biomedicines9111620

Chicago/Turabian Style

Yadav, Dharmendra Kumar, Desh Deepak Singh, Ihn Han, Yogesh Kumar, and Eun-Ha Choi. 2021. "Current Potential Therapeutic Approaches against SARS-CoV-2: A Review" Biomedicines 9, no. 11: 1620. https://doi.org/10.3390/biomedicines9111620

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