Next Article in Journal
Responses of Issatchenkia terricola WJL-G4 upon Citric Acid Stress
Next Article in Special Issue
Fungal Proteins from Sargassum spp. Using Solid-State Fermentation as a Green Bioprocess Strategy
Previous Article in Journal
Density Functional Theory Approach to the Vibrational Properties and Magnetic Specific Heat of the Covalent Chain Antiferromagnet KFeS2
Previous Article in Special Issue
Novel Bio-Functional Aloe vera Beverages Fermented by Probiotic Enterococcus faecium and Lactobacillus lactis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Docking Analysis of Some Bioactive Compounds from Traditional Plants against SARS-CoV-2 Target Proteins

1
Department of Food Technology, Arid Lands Cultivation Research Institute (ALCRI), City of Scientific Research and Technological Applications (SRTA-City), Alexandria 21934, Egypt
2
Food Technology Department, Faculty of Agriculture, Benha University, Moshtohor, Benha 13736, Egypt
3
Department of Food Toxicology and Contaminants, National Research Centre, Dokki, Cairo 12622, Egypt
4
Department of Food Science and Nutrition, College of Sciences, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
5
Department of Plant Protection and Biomolecular Diagnosis, Arid Lands Cultivation Research Institute (ALCRI), City of Scientific Research and Technological Applications (SRTA-City), Alexandria 21934, Egypt
6
Food Research Section, R&D Division, Abu Dhabi Agriculture and Food Safety Authority (ADAFSA), Abu Dhabi P.O. Box 52150, United Arab Emirates
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(9), 2662; https://doi.org/10.3390/molecules27092662
Submission received: 4 January 2022 / Revised: 25 February 2022 / Accepted: 26 February 2022 / Published: 20 April 2022

Abstract

:
COVID-19 is still a global pandemic that has not been stopped. Many traditional medicines have been demonstrated to be incredibly helpful for treating COVID-19 patients while fighting the disease worldwide. We introduced 10 bioactive compounds derived from traditional medicinal plants and assessed their potential for inhibiting viral spike protein (S-protein), Papain-like protease (PLpro), and RNA dependent RNA polymerase (RdRp) using molecular docking protocols where we simulate the inhibitors bound to target proteins in various poses and at different known binding sites using Autodock version 4.0 and Chimera 1.8.1 software. Results found that the chicoric acid, quinine, and withaferin A ligand strongly inhibited CoV-2 S -protein with a binding energy of −8.63, −7.85, and −7.85 kcal/mol, respectively. Our modeling work also suggested that curcumin, quinine, and demothoxycurcumin exhibited high binding affinity toward RdRp with a binding energy of −7.80, −7.80, and −7.64 kcal/mol, respectively. The other ligands, namely chicoric acid, demothoxycurcumin, and curcumin express high binding energy than the other tested ligands docked to PLpro with −7.62, −6.81, and −6.70 kcal/mol, respectively. Prediction of drug-likeness properties revealed that all tested ligands have no violations to Lipinski’s Rule of Five except cepharanthine, chicoric acid, and theaflavin. Regarding the pharmacokinetic behavior, all ligand predicted to have high GI-absorption except chicoric acid and theaflavin. At the same way chicoric acid, withaferin A, and withanolide D predicted to be substrate for multidrug resistance protein (P-gp substrate). Caffeic acid, cepharanthine, chicoric acid, withaferin A, and withanolide D also have no inhibitory effect on any cytochrome P450 enzymes. Promisingly, chicoric acid, quinine, curcumin, and demothoxycurcumin exhibited high binding affinity on SARS-CoV-2 target proteins and expressed good drug-likeness and pharmacokinetic properties. Further research is required to investigate the potential uses of these compounds in the treatment of SARS-CoV-2.

1. Introduction

Coronavirus disease 2019 (COVID-19), the highly infectious viral illness caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), has had a disastrous effect on the world’s demographics, resulting in more than 4.5 million deaths worldwide. SARS-CoV-2 is a positive-sense RNA (30 kb) virus. Two types of proteins characterize the human coronaviruses HCoVs, structural proteins (Spike (S), Nucleocapsid (N), Matrix (M), and Envelope (E)), and non-structural proteins including the RdRp [1,2]. The membrane and envelope proteins are associated with virus assembly, while the spike protein plays the main role in facilitating virus entry via mediating its interaction with the transmembrane surface receptor on the host cells [3].
This virus contains S- protein, PLpro, and RdRp involved in viral entry, replication, and immune response evasion. Drugs targeting these proteins therefore have great potential for inhibiting the virus. The spike protein directly interacts with the peptidase domain (PD) of the angiotensin-converting enzyme 2 (ACE2) receptor [4], which technically marks the virus entry inside the cells [5]. ACE2 interaction with spike protein could aid in producing antivirals or a vaccine that can block CoV infection by targeting ACE2 [6]. The binding to the SARSCoV-2 spike protein at the same place where the original ACE2 PD domain interacts, which shows that 23-amino acids sequence independently has the potential to inhibit the interaction of the SARS-CoV-2 S- protein and ACE2 complex. Hence, the viral entrance. Interacting residues at the interface of the SARS-CoV-2 and ACE2 PD domain are ARG403 LYS417 GLY446 TYR449 TYR453 LEU455 PHEF456 TYR473 ALA475 GLY476 GLY485 PHE486 ASN487 TYR489 GLN493 TYR495 GLY496 GLNQ498 THR500 ASN501 GLY502 VAL503 TYR505 in SARS-CoV-2 S- protein [7].
Another viral protein, PLpro, which is vital for viral replication [8], is responsible for the proteolytic processing of the product of open reading frame 1a (ORF1a) in the replicase gene of CoV2, a large viral polyprotein containing non-structural proteins which form the replicase complex [9]. The peptide bond cleavage in the active site is catalyzed by a conserved catalytic triad comprised of residues CYS111, HIS272, and ASP286 [10]. In addition, PLpro possesses deubiquitinating and deISGylating capabilities [11] which interfere with critical signaling pathways leading to the expression of type I interferons, resulting in antagonistic effect on host innate immune response [12]. Therefore, inhibition of PLpro activity can halt viral replication and disrupt its role in host immune response evasion, making it an excellent anti-viral drug target. RdRp is also a vital enzyme for the life cycle of RNA viruses. It has been targeted in various viral infections (Hepatitis C virus (HCV), Zika virus (ZIKV), and human coronaviruses (HCoV) [13]. SARS-CoV-2 polymerase residues ARG 553, ARG 555, LYS 545, and ASN 691 interaction with some bioactive compounds play an important role in virus replication inhibition [14]. Thus, many research working in computer-aided drug design in generating new small molecules that might interact with SARS-CoV-2 S -protein, RdRp, and 3CL or Mpro protease proteins to combat the COVID-19 disease.
In the U.S., the Food and Drug Administration issued an emergency use authorization (EUA) for Pfizer’s Paxlovid (nirmatrelvir tablets and ritonavir tablets, co-packaged for oral use) for the treatment of mild-to-moderate coronavirus disease (COVID-19) in adults and pediatric patients (12 years of age and older weighing at least 40 kg or about 88 pounds) with positive results of direct SARS-CoV-2 testing, and who are at high risk for progression to severe COVID-19, including hospitalization or death. Paxlovid consists of nirmatrelvir, which inhibits a SARS-CoV-2 protein to stop the virus from replicating, and ritonavir, which slows down nirmatrelvir’s breakdown to help it remain in the body for a longer period at higher concentrations. Paxlovid is available by prescription only and should be initiated as soon as possible after diagnosis of COVID-19 and within five days of symptom onset [15]. Another antiviral drug use for the same purpose is Lagevrio (molnupiravir). Molnupiravir is a nucleoside analogue that inhibits SARS-CoV-2 replication by viral mutagenesis. The U.S. FDA has issued an EUA to make Molnupiravir available during the COVID-19 pandemic (Molnupiravir is not FDA-approved for any uses, including use as treatment for COVID-19. This medicine is still being studied, Molnupiravir may cause serious side effects [16].
The curative effects of traditional herbal formula sometimes are not necessarily by directly inhibiting or killing the virus, but through the integration of various aspects such as relieving cytokines storm syndrome, protecting human tissues and organs, relieving immunological injury, and enhancing the body’s ability [17].
Ayurveda, known as “The Science of Life,” an ancient traditional medicinal system that originated and is practiced in India, has been utilized for reducing SARS-CoV-2 infection and treating COVID-19-associated patients [18,19,20]. It describes many medicinal plants and herbs possessing a broad range of therapeutic usefulness in curing various kinds of ailments, diseases, and disorders, such as Allium sativum (Garlic), W. somnifera known commonly as ashwagandha, Zingiber officinale Roscoe (Ginger), Tinospora cordifolia (Giloy), Ocimum sanctum (Tulsi), Curcuma longa (Turmeric, Haldi), Glycyrrhiza glabra (Licorice, mulethi), and others [20,21,22].
In this study we aim to investigate docking characteristics of 10 bioactive compounds identified from traditional medicines (caffeic and chicoric acids from Echinacea purpurea; xanthorrhizol from Java turmeric; curcumin and demothoxycurcumin from Curcuma longa L.; Quinine from Chinchona bark; Theaflavin from Black tea; withaferin A and withanolide D from Withania somnifera L.; cepharanthine from Stephania cepharantha) on S-protein, RdRp and SARS-CoV-2 PLpro to explore the potential mode of the inhibitory activity toward SARS-CoV-2. The molecules that showed better hydrogen bonding and binding energy were selected further for drug likeness and pharmacokinetics prediction.

2. Results and Discussion

Herbal medicines and medicinal plant-based natural compounds provide a rich resource for novel antiviral drug development. Some natural medicines have been shown to possess antiviral activities against various virus strains including coronavirus, herpes simplex virus [23,24,25], influenza virus [25], human immunodeficiency virus [26], hepatitis B and C viruses [27], SARS, and MERS [28,29]. These compounds antiviral action mechanisms caused by influencing of the viral life cycle, such as viral entry, replication, assembly, and release, as well as virus-host-specific interactions. In this study we analyze the binding mode of 10 ligands Figure 1 that previously identified in some traditional plants Table 1 using computational docking analysis.
All tested ligands showed hydrogen interaction to papain-like protease (6W9C) except cepharanthine and theaflavin. The ligand docking on CYS111, HIS272, and ASP286 catalytic residues (Figure 2A) ranked by binding energy ΔG in kcal/mol were as follows: chicoric acid > demothoxycurcumin > curcumin > withanolide D > xanthorrhizol > withaferin A > caffeic acid > quinine. It is known that the high binding affinity of the drug compounds depends on the type and amount of bonds occurring with the target protein [53]. The binding energies were −7.62, −6.81, −6.70, −6.58, −6.16, −6.13, −5.70, and −5.67 kcal/mol, respectively (Table 2; Figure 3).
The first reported bioactive effect of chicoric acid is its ability to inhibit infection with human immunodeficiency virus 1 (HIV-1) [54]. Several studies reported that chicoric acid inhibits infection with HIV-1 by deactivating the HIV-1 integrase. HIV-1 integrase is a multidomain enzyme required for integration of viral DNA into the host genome, a critical step in viral replication [55]. Inhibition of HIV-1 integration by chicoric acid results in stopping virus replication, leading to increased T-lymphoblastoid cell viability [56]. Chicoric acid found in many Egyptian source as Chicory and Echinacea purpurea (Table 1). Despite that cepharanthine (CEP) showed no binding behavior toward papain-like protease, it has unique anti-inflammatory, antioxidative, immunomodulating, antiparasitic, and antiviral properties. It can suppress nuclear factor-kappa B (NF-κB) activation, lipid peroxidation, nitric oxide (NO) production, cytokine production, and expression of cyclooxygenase; all of which are crucial to viral replication and inflammatory response [33]. Against SARS-CoV-2 and homologous viruses, CEP predominantly inhibits viral entry and replication at low doses [57].
Moreover, all tested ligands showed H-interaction to spike protein (6M17). The ligand docking on GLY 502, TYR 489 and TYR 505 (Figure 2B) ranked by binding energy in kcal/mol were as follows: chicoric acid > quinine = withaferin A > withanolide D > theaflavin > demothoxycurcumin > curcumin > xanthorrhizol > cepharanthine > caffeic acid. The binding energies were −8.63, −7.85, −7.85, −7.78, −7.43, −7.23, −6.34, −6.00, −5.26 and −5.06 kcal/mol respectively (Table 2; Figure 4). Cepharanthine (CEP) is a natural alkaloid, which has been widely used to treat many of the acute and chronic diseases. It is usually used in Japan in the form of Stephania cepharantha and Stephania rotunda (Table 1) for its anti-inflammatory, antiparasitic, anti-oxidative, antiviral, and anti-HIV activities [32,33,34,35]. In 2021 a Japanese research team reported the anti-COVID-19 activity of the combining cepharanthine/nelfinavir providing synergistic antiviral effects that CEP and nelfinavir inhibit SARS-CoV-2 entry and RNA replication, respectively [58]. In 2021, a Japanese research team reported that the combination of cepharanthine and nelfinavir inhibited SARS-CoV-2 entrance and RNA replication, offering synergistic antiviral effects [58].
This previous combination based on in silico docking simulation that confirms that CEP molecule can bind to SARS-CoV-2 S-protein and interfere with the spike engagement to its receptor, angiotensin-converting enzyme 2 (ACE2) [59,60]. Our results strongly supported this mechanism of CEP toward SARS-CoV-2 S-protein. However, chicoric acid, quinine, withaferin A, withanolide D, theaflavin, demothoxycurcumin, curcumin, and xanthorrhizol expressed higher binding energy than cepharanthine in case SARS-CoV-2 S-protein binding. From all of this we supposed that chicoric acid and quinine may play the same role in this combination (cepharanthine/nelfinavir) beside their acting mechanism toward SARS-CoV-2, PLpro and RdRp that may be promising to limit SARS-CoV-2 proliferation than CEP.
Regarding to RdRp (6M71), all tested ligands showed H-interaction except Cepharanthine. The ligand docking on ARG 553, ARG 555, LYS 545, and ASN 691 (Figure 2C) ranked by binding energy ΔG in kcal/mol were as follows: curcumin = quinine > demothoxycurcumin > chicoric acid > caffeic acid > withaferin A > withanolide D > xanthorrhizol > theaflavin. The binding energies were −7.80, −7.80, −7.64, −7.50, −6.75, −6.27, −6.10, −5.33, and −2.55 kcal/mol, respectively (Table 2; Figure 5).
Curcumin (CC) and its analogues are the main phytonutrients of turmeric (Curcuma longa L.) and other Curcuma spp., which are widely used around the world as culinary spices, traditional medicine as well as a popular dietary supplement ingredient due to its wide range of health benefits including anti-inflammation [61], anti-cancer [62], cardiovascular regulation [63], respiratory [64], and immune system benefits [65]. In addition, the suppression of multiple cytokines by curcumin suggested that it may be a useful approach in treating Ebola patients against cytokine storm [66]. As severe cases of COVID-19 are often associated with cytokine release syndrome, the use of anti-inflammatory molecules may reduce the proinflammatory cytokines involved. Quinine, is an extract of the bark of the Chinchona tree (native to the Andes of South America), that was used to treat feverish infections, particularly malaria, for hundreds of years almost worldwide [67]. Curcuma longa L. and Chinchona bark are rich source of curcumin and quinine respectively (Table 1).
Table 3 presents the drug likeness and pharmacokinetic properties of the tested ligands. Lipinski’s Rule of Five is generally used as an indicator of the drug likeness and pharmacological activities. In humans, this would make them orally active medications [68]. All the tested compounds are in the molecular weight range of 180.16 to 474.37 Da (<500 Da), except for cepharanthine and theaflavin are 606.71 and 564.49 Da. It is well known that drug molecules typically have low molecular weight (<500 kDa) are transported, diffused, and absorbed easily compared with large molecules [69]. All the tested compounds also have fewer than 15 rotatable bonds, and all have less than 5 hydrogen bond donors (NH and OH), except for quinine and theaflavin. In addition, the numbers of hydrogen bond acceptors (O and N atoms) predicted in all compounds are less than 10, except for chicoric acid and theaflavin (Table 3). At the same time, permeability (logP) of these ligands has also been investigated, and it was found that these ligands exhibited logP values of less than 5 except cepharanthine. Moreover, the topological polar surface area (TPSA) values of all ligands are less than 140 Å, except for chicoric acid and theaflavin. TPSA is closely related to the H-bonding ability of a compound [70]. The TPSA and the logP values are the two essential characteristics in the analysis of the bioavailability of drug molecules and permeability toward bio-membranes [71]. Compounds with 10 rotatable bonds and TPSA of ≤140 Å can be predicted to have good bioavailability [72]. As a result, all of the ligands examined have good bioavailability, with the exception of chicoric acid and theaflavin, which have TPSA values of 208.12 and 217.60, respectively. Pharmacokinetic prediction provides information to evaluate the time course of drugs and their effects on the body to design an appropriate drug regimen for any patient [73].
Our results of pharmacokinetic prediction (Table 3) showed that all tested ligands demonstrate a high absorption rate in the GI tract, except for chicoric acid and theaflavin. Moreover, none of the tested ligands were predicted to be able to pass through the blood–brain barrier, except for quinine and xanthorrhizol. Additionally, chicoric acid, withaferin A and withanolide D are the only compounds predicted to act as a substrate for P-gp, which decrease their further clinical application [74]. The predicted pharmacokinetic behavior indicates that only caffeic acid, cepharanthine, chicoric acid, withaferin A, and withanolide D would have no inhibitory effect on any cytochrome P450 enzymes (CYP1A2, CYP2C19, CYP2D6, and CYP3A4). Furthermore, CYP2C9 activity is only known to be affected by curcumin, demothoxycurcumin, theaflavin and xanthorrhizol. Cytochrome P450 enzymes are essential for the metabolism of many chemicals. Although this class contains more than 50 enzymes, six of them metabolize 90% of compounds. Cytochrome P450 enzyme inhibition causes unwanted adverse effects or therapeutic failures [75]. The proposed mechanism of SARS-CoV-2 inhibition by tested compounds is summarized in Figure 6; chicoric acid, quinine and withaferin A, interacting with S protein, inhibits viral ability to attach to human ACE2. Additionally, chicoric acid, demothoxycurcumin, and curcumin interacting with PL pro, inhibits viral replication, moreover, curcumin, demothoxycurcumin, and quinine interacting with RdRp, inhibits viral replication. Furthermore, interferon (IFN) production is triggered by virus infections that cause accumulation of RNA species in the cytoplasm, initiate the oligomerization of an essential mitochondrial antiviral signaling protein, MAVS, that serves as a scaffold for the activation of TNF receptor-associated factor (TRAF) family proteins, TRAF2, TRAF3, TRAF5, and TRAF6. TRAF proteins catalyze the assembly of K63-linked ubiquitin chains that are required for the activation of serine kinases, IKKa, IKKb, IKKc, IKKe, and TBK1. Kinase activation triggers the phosphorylation and nuclear import of IRF3, driving the production of primary antiviral effectors including IFNs [76]. On the other hand, PLpro of SARS-CoV-2 interfere with innate immune response by directly cleaving IRF3, inhibits IFNβ production. In general, in silico studies are a great starting point when looking for new therapeutic targets. In the event of an unpredictable pandemic, such as the recent COVID-19 epidemic, in silico research are critical. Drug targets cannot be discovered in a short period of time using existing methods. To reach bulk targets, it is unavoidable to have extremely rapid procedures. In silico approaches can be used to speed up the search for drug targets in natural medicinal plants. This article’s results warrant further evaluation of the potential anti-SARS-CoV-2 activity of these ligands in vitro.
Steps from one to six describe the replication cycle of SARS-CoV-2; attachment (1), viral entry (2), translation of viral proteins (3), viral replication (4), assembly (5), and release (6). In red: chicoric acid, quinine, and withaferin A, interacting with S protein, inhibiting its viral ability to attach to human ACE2 (1). Additionally, chicoric acid, demothoxycurcumin, and curcumin interacting with PL pro inhibits viral replication (2), moreover, curcumin, demothoxycurcumin, and quinine interacting with RdRp inhibits viral replication (3). Interferon (IFN) production is triggered by virus infections that cause accumulation of RNA species in the cytoplasm, initiating the oligomerization of an essential mitochondrial antiviral signaling protein, MAVS, which serves as a scaffold for the activation of TNF receptor-associated factor (TRAF) family proteins, TRAF2, TRAF3, TRAF5, and TRAF6. TRAF proteins catalyze the assembly of K63-linked ubiquitin chains that are required for the activation of serine kinases, IKKa, IKKb, IKKc, IKKe, and TBK1. Kinase activation triggers the phosphorylation and nuclear import of IRF3, driving the production of primary antiviral effectors including IFNs [74]. In blue: PLpro of SARS-CoV-2 interfere with innate immune response by directly cleaving IRF3, inhibits IFNβ production.

3. Experimental Section

3.1. Docked SARS-CoV-2 Protein Structures

The structure of the SARS-CoV-2 papain-like protease (6W9C), RNA-dependent RNA polymerase (6M71), and spike protein (6M17) used as a target for ligands binding was downloaded from RCSB website [77]. PDB (Protein Data Bank) has enabled breakthroughs in research, such as this study, and education worldwide [78].

3.2. Ligands and Drug Scan

The three-dimensional (3D) structures of all tested ligands were drawn in ACD/ChemSketch and then docked into the rigid binding pocket of 6W9C, 6M71, and 6M17 of SARS CoV-2. The compounds used in the present study were caffeic acid, cepharanthine, chicoric acid, curcumin, demothoxycurcumin, quinine, theaflavin, withaferin a, withanolide d and xanthorrhizol. The drug likeness and pharmacokinetic properties were calculated using the SWISSADME prediction website (http://www.swissadme.ch/) (accessed on 15 December 2021) [68,79]. Figure 1 shows the structure of the 10 docked ligands we used herein.

3.3. Determination of SARS-CoV-2 Proteins Binding Hits

Table 4 contains the amino acids binding hits of the SARS-CoV-2 polymerase as described by Afonine et al. [80], papain-like protease (6W9C) as described by Baez-Santos et al., 2015 and spike protein (6M17) as described by Baig et al. [7].

3.4. Molecular Docking

Ligand optimization was performed using Open Babel, converting ligands from mol into the PDB format. Autodock version 4.0 was used for protein optimization through the removal of water and other atoms and then addition of a polar hydrogen group. Ligand tethering of the protein was performed by regulating the genetic algorithm (GA) parameters using 10 runs of the GA criteria. Docking analyses and determination of hydrogen bonds (H-bonds) were conducted using Chimera 1.8.1 [81].

4. Conclusions

Antiviral treatments targeting the coronavirus disease 2019 are urgently required. We screened a panel of ligands from traditional plant sources already used in medicinal fields and some of them described for severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). Chicoric acid, quinine, curcumin, and demothoxycurcumin exhibited high binding affinity on SARS-CoV-2 S- protein, PLpro, and RdRp. Binding of these proteins interfere with the viral entry, replication, and immune response evasion. Therefore, these compounds may have a great potential for inhibiting the virus. Our work provides data about the ligand mechanism toward SARS-CoV-2 target proteins and allows the comparison between them based on scientific information. Further in vitro cell-based investigations will be needed for chicoric acid, quinine, curcumin, and demothoxycurcumin compounds to confirm their previous results and determine their antiviral mechanism at cell level.

Author Contributions

Conceptualization: N.M.A.E.-A., M.G.S. and I.K.; Data curation: N.M.A.E.-A., M.G.S. and I.K.; Formal analysis: N.M.A.E.-A., M.G.S., A.N.B., A.M.G.D. and H.A.; Investigation: N.M.A.E.-A., M.G.S., I.K., A.N.B., E.-S.H., A.M.G.D. and H.A.; Methodology: N.M.A.E.-A., M.G.S. and I.K.; Validation: N.M.A.E.-A., M.G.S., I.K., A.N.B., E.-S.H., A.M.G.D. and H.A.; Visualization: N.M.A.E.-A., M.G.S., I.K., A.N.B. and E.-S.H.; Writing—original draft: N.M.A.E.-A., M.G.S. and I.K.; Writing—review and editing: N.M.A.E.-A., M.G.S., I.K., A.N.B., E.-S.H., A.M.G.D. and H.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Elfiky, A.A. Zika Virus: Novel Guanosine Derivatives revealed strong binding and possible inhibition of the polymerase. Future Virol. 2017, 12, 721–728. [Google Scholar] [CrossRef]
  2. Hasan, A.; Paray, B.A.; Hussain, A.; Qadir, F.A.; Attar, F.; Aziz, F.M.; Sharifi, M.; Derakhshankhah, H.; Rasti, B.; Mehrabi, M.; et al. A review on the cleavage priming of the spike protein on coronavirus by angiotensin-converting enzyme-2 and furin. J. Biomol. Struct. Dyn. 2020, 39, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Li, F. Structure, function, and evolution of coronavirus spike proteins. Annu. Rev. Virol. 2016, 3, 237–261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Yan, R.; Zhang, Y.; Li, Y.; Xia, L.; Guo, Y.; Zhou, Q. Structural basis for the recognition of SARS-CoV-2 by full-length human ACE2. Science 2020, 367, 1444–1448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Hoffmann, M.; Kleine-Weber, H.; Schroeder, S.; Kruger, N.; Herrler, T.; Erichsen, S. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and is blocked by a clinically proven protease inhibitor. Cell 2020, 181, 271–280. [Google Scholar] [CrossRef]
  6. Jando, J.; Camargo, S.M.R.; Herzog, B.; Verrey, F. Expression and regulation of the neutral amino acid transporter B0AT1 in rat small intestine. PLoS ONE 2017, 12, e0184845. [Google Scholar] [CrossRef]
  7. Baig, M.S.; Alagumuthu, M.; Rajpoot, S.; Saqib, U. Identification of a Potential Peptide Inhibitor of SARS-CoV-2 Targeting its Entry into the Host Cells. Drugs R&D 2020, 20, 161–169. [Google Scholar] [CrossRef]
  8. Harcourt, B.H.; Jukneliene, D.; Kanjanahaluethai, A.; Bechill, J.; Severson, K.M.; Smith, C.M. Identification of severe acute respiratory syndrome coronavirus replicase products and characterization of papain-like protease activity. J. Virol. 2004, 8, 13600–13612. [Google Scholar] [CrossRef] [Green Version]
  9. Wertz, I.E.; Murray, J.M. Structurally-defined deubiquitinase inhibitors provide opportunities to investigate disease mechanisms. Drug Discov. Today Technol. 2019, 31, 109–123. [Google Scholar]
  10. Baez-Santos, Y.M.; St John, S.E.; Mesecar, A.D. The SARS coronavirus papain-like protease: Structure, function and inhibition by designed antiviral compounds. Antivir. Res. 2015, 115, 21–38. [Google Scholar] [CrossRef] [PubMed]
  11. Sulea, T.; Lindner, H.A.; Purisima, E.O.; Ménard, R. Deubiquitination, a new function of the severe acute respiratory syndrome coronavirus papainlike protease. J. Virol. 2005, 79, 4550–4551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Bekes, M.; van Noort, G.J.V.; Ekkebus, R.; Ovaa, H.; Huang, T.T.; Lima, C.D. Recognition of Lys48-linked di-ubiquitin and deubiquitinating activities of the SARS coronavirus papain-like protease. Mol. Cell 2016, 62, 572–585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Elfiky, A.A.; Ismail, A. Molecular dynamics and docking reveal the potency of novel GTP derivatives against RNA dependent RNA polymerase of genotype 4a HCV. Life Sci. 2019, 238, 116958. [Google Scholar] [CrossRef] [PubMed]
  14. Abd El-Aziz, N.M.; Awad, O.M.; Shehata, M.S.; El-Sohaimy, S.A. Inhibition of the SARS-CoV-2 RNA-Dependent RNA Polymerase by Natural Bioactive Compounds: Molecular Docking Analysis. Egypt. J. Chem. 2021, 64, 1989–2001. [Google Scholar]
  15. Fact Sheet for Healthcare Providers: Emergency Use Authorization for Paxlovid™. Available online: https://www.fda.gov/media/155050/download (accessed on 22 December 2021).
  16. Mahase, E. COVID-19: UK becomes first country to authorise antiviral molnupiravir. BMJ 2021, 375, n2697. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, L. Traditional Chinese medicine contributes to the treatment of COVID-19 patients. Chin. Herb. Med. 2020, 12, 95–96. [Google Scholar] [CrossRef]
  18. Gautam, K.; Adhikari, R.P.; Gupta, A.S.; Shrestha, R.K.; Koirala, P.; Suraj, K.S. Self-reported psychological distress during the COVID-19 outbreak in Nepal: Fndings from an online survey. BMC Psychol. 2020, 8, 127. [Google Scholar] [CrossRef]
  19. Rastogi, A.; Bhansali, A.; Khare, N.; Suri, V.; Yaddanapudi, N.; Sachdeva, N.; Puri, G.D.; Malhotra, P. Short term, high-dose vitamin D supplementation for COVID-19 disease: A randomised, placebo-controlled, study (SHADE study). Postgrad. Med. J. 2022, 98, 87–90. [Google Scholar] [CrossRef]
  20. Singh, P.; Guleri, R.; Singh, V.; Kaur, G.; Kataria, H.; Singh, B.; Kaur, G.; Kaul, S.C.; Wadhwa, R.; Pati, P.K. Biotechnological interventions in Withania somnifera (L.) dunal. Biotechnol. Genet. Eng. Rev. 2015, 31, 1–20. [Google Scholar] [CrossRef]
  21. Dhama, K.; Sharun, K.; Tiwari, R.; Dadar, M.; Malik, Y.S.; Singh, K.P.; Chaicumpa, W. COVID-19, an emerging coronavirus infection: Advances and prospects in designing and developing vaccines, immunotherapeutics, and therapeutics. Hum. Vaccin Immunother. 2020, 16, 1232–1238. [Google Scholar] [CrossRef] [Green Version]
  22. Tiwari, R.; Latheef, S.K.; Ahmed, I.; Iqbal, H.M.N.; Bule, M.H.; Dhama, K.; Samad, H.A.; Karthik, K.; Alagawany, M.; El-Hack, M.E.A.; et al. Herbal Immunomodulators—A remedial panacea for designing and developing effective drugs and medicines: Current scenario and future prospects. Curr. Drug Metab. 2018, 19, 264–301. [Google Scholar] [CrossRef] [PubMed]
  23. Calland, N.; Dubuisson, J.; Rouille, Y.; Seron, K. Hepatitis C virus and natural compounds: A new antiviral approach? Viruses 2012, 4, 2197–2217. [Google Scholar] [CrossRef] [Green Version]
  24. Du, J.; He, Z.D.; Jiang, R.W.; Ye, W.C.; Xu, H.X.; But, P.P.H. Antiviral flavonoids from the root bark of Morus alba L. Phytochemistry 2003, 62, 1235–1238. [Google Scholar] [CrossRef]
  25. Luganini, M.E.; Terlizzi, G.; Catucci, G.; Gilardi, M.E.; Maffei, G.G. The cranberry extract oximacro exerts in vitro virucidal activity against influenza virus by interfering with hemagglutinin. Front. Microbiol. 2018, 9, 1826. [Google Scholar]
  26. Xu, H.X.; Ming, D.S.; Dong, H.; But, P.P. A new anti-HIV triterpene from Geum japonicum. Chem. Pharm. Bull. 2000, 48, 1367–1369. [Google Scholar] [CrossRef] [Green Version]
  27. Sahuc, M.E.; Sahli, R.; Riviere, C.; Pene, V.; Lavie, M.; Vandeputte, A. Dehydrojuncusol, a natural phenanthrene compound extracted from Juncus maritimus, is a new inhibitor of hepatitis C virus RNA replication. J. Virol. 2019, 93, e02009–e02018. [Google Scholar] [CrossRef] [Green Version]
  28. Cinatl, J.; Morgenstern, B.; Bauer, G.; Chandra, P.; Rabenau, H.; Doerr, H.W. Glycyrrhizin, an active component of liquorice roots, and replication of SARS-associated coronavirus. Lancet 2003, 361, 2045–2046. [Google Scholar] [CrossRef] [Green Version]
  29. Lin, S.C.; Ho, C.T.; Chuo, W.H.; Li, S.; Wang, T.T.; Lin, C.C. Effective inhibition of MERS-CoV infection by resveratrol. BMC Infect. Dis. 2017, 17, 144. [Google Scholar] [CrossRef] [Green Version]
  30. Elansary, H.O.; Szopa, A.; Kubica, P.; Ekiert, H.; Ali, H.M.; Elshikh, M.S.; Abdel-Salam, E.M.; El-Esawi, M.; El-Ansary, D.O. Bioactivities of Traditional Medicinal Plants in Alexandria. Evid.-Based Complement. Altern. Med. 2018, 1463579, 13. [Google Scholar]
  31. Wang, Z.; Song, Q.; Su, J.; Tang, W.; Song, J.; Huang, X.; An, J.; Li, Y.; Ye, W.; Wang, Y. Caffeic acid oligomers from Mesona chinensis and their in vitro antiviral activities. Fitoterapia 2020, 144, 104603. [Google Scholar] [CrossRef]
  32. Liu, T.; Liu, X.; Li, W. Tetrandrine, a Chinese plant-derived alkaloid, is a potential candidate for cancer chemotherapy. Oncotarget 2016, 7, 40800–40815. [Google Scholar] [CrossRef] [Green Version]
  33. Bailly, C. Cepharanthine: An update of its mode of action, pharmacological properties and medical applications. Phytomedicine 2019, 62, 152956. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, C.H.; Wang, Y.F.; Liu, X.J.; Lu, J.H.; Qian, C.W.; Wan, Z.Y.; Yan, X.G.; Zheng, H.Y.; Zhang, M.Y.; Xiong, S.; et al. Antiviral activity of cepharanthine against severe acute respiratory syndrome coronavirus in vitro. Chin. Med. J. 2005, 20, 493–496. [Google Scholar]
  35. Okamoto, M.; Ono, M.; Baba, M. Potent inhibition of HIVtype 1 replication by an antiinflammatory alkaloid cepharanthinein chronically infected monocytic cells. AIDS Res. Hum. 1998, 14, 1239–1245. [Google Scholar] [CrossRef] [PubMed]
  36. Mulinacci, N.; Innocenti, M.; Gallori, S.; Romani, A.; LaMarca, G.; Vincieri, F.F. Optimization of the chromatographic determination of polyphenols in the aerial parts of Cichoriumintybus L. Chromatographia 2001, 24, 455–461. [Google Scholar] [CrossRef]
  37. Ziamajidi, N.; Khagjani, S.; Hassanzadeh, G.; Vardasbi, S.; Ahmadian, S.; Nowrouzi, A. Amelioration by chicory seed extract of diabetes-and oleic acid-inducednon-alcoholicfattyliverdisease(NAFLD)/non-alcoholicsteato- hepatitis (NASH) via modulkation of PPARα and SREBP-1. Food Chem. Toxicol. 2013, 58, 198–209. [Google Scholar] [CrossRef]
  38. Peng, Y.; Sun, Q.; Park, Y. The Bioactive Effects of Chicoric Acid As a Functional Food Ingredient. J. Med. Food. 2019, 22, 645–652. [Google Scholar] [CrossRef] [PubMed]
  39. Araújo, C.C.; Leon, L.L. Biological activities of Curcuma longa L. Mem Inst. Oswaldo Cruz. 2001, 96, 723–728. [Google Scholar] [CrossRef] [PubMed]
  40. Soni, V.K.; Mehta, A.; Ratre, Y.K.; Tiwari, A.K.; Amit, A.; Singh, R.P.; Sonkar, S.C.; Chaturvedi, N.; Shukla, D.; Vishvakarma, N.K. Curcumin, a traditional spice component, can hold the promise against COVID-19? Eur. J. Pharmacol. 2020, 886, 173551. [Google Scholar] [CrossRef]
  41. Siviero, A.; Gallo, E.; Maggini, V.; Gori, L.; Mugelli, A.; Firenzuoli, F.; Vannacci, A. Curcumin, a golden spice with a low bioavailability. J. Herb. Med. 2015, 5, 57–70. [Google Scholar] [CrossRef]
  42. Prakash, D.; Suri, S.; Upadhyay, G.; Singh, B.N. Total phenol, antioxidant and free radical scavenging activities of some medicinal plants. Int. J. Food Sci. Nutr. 2007, 58, 18–28. [Google Scholar] [CrossRef]
  43. Sivaprabha, J.; Dharani, B.; Padma, P.R.; Sumathi, S. Induction of DNA damage by the leaves and rhizomes of Curcuma amada Roxb in breast cancer cell lines. J. Acute Dis. 2015, 4, 12–17. [Google Scholar] [CrossRef] [Green Version]
  44. Mahadevi, R.; Kavitha, R. Phytochemical and pharmacological properties of Curcuma amada: A Review. Int. J. Res. Pharm. Sci. 2020, 11, 3546–3555. [Google Scholar] [CrossRef]
  45. Gurung, P.; De, P. Spectrum of biological properties of cinchona alkaloids: A brief review. J. Pharmacogn. Phytochem. 2017, 6, 162–166. [Google Scholar]
  46. O’Neill, E.J.; Termini, D.; Albano, A.; Tsiani, E. Anti-Cancer Properties of Theaflavins. Molecules 2021, 26, 987. [Google Scholar] [CrossRef]
  47. Tong, T.; Liu, Y.; Kang, J.; Zhang, C.; Kang, S. Antioxidant Activity and Main Chemical Components of a Novel Fermented Tea. Molecules 2019, 24, 2917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Behl, T.; Sharma, A.; Sharma, L.; Sehgal, A.; Zengin, G.; Brata, R.; Fratila, O.; Bungau, S. Exploring the Multifaceted Therapeutic Potential of Withaferin A and Its Derivatives. Biomedicines 2020, 8, 571. [Google Scholar] [CrossRef]
  49. Wang, G.; Xu, L.; Liu, W.; Xu, W.; Mu, Y.; Wang, Z.; Huang, X.; Li, L. New anti-inflammatory withanolides from Physalis pubescens fruit. Fitoterapia 2020, 146, 104692. [Google Scholar] [CrossRef] [PubMed]
  50. Tan, J.; Liu, Y.; Cheng, Y.; Sun, Y.; Pan, J.; Guan, W.; Li, X.; Huang, J.; Jiang, P.; Guo, S.; et al. New withanolides with anti-inflammatory activity from the leaves of Datura metel L. Bioorganic Chem. 2020, 95, 103541. [Google Scholar] [CrossRef] [PubMed]
  51. Jung, Y.; Lee, J.; Kim, H.K.; Moon, B.C.; Ji, Y.; Ryu, D.H.; Hwang, G.S. Metabolite profiling of Curcuma species grown in different regions using 1HNMR spectroscopy and multivariate analysis. Analyst 2012, 137, 5597–5606. [Google Scholar] [CrossRef]
  52. Kamazeri, T.S.A.T.; Samah, O.A.; Taher, M.; Susanti, D.; Qaralleh, H. Antimicrobial activity and essential oils of Curcuma aeruginosa, Curcuma mangga, and Zingiber cassumunar from Malaysia. Asian Pac. J. Trop. Med. 2012, 5, 202–209. [Google Scholar] [CrossRef] [Green Version]
  53. Khaerunnisa, S.; Kurniawan, H.; Awaluddin, R.; Suhartati, S.; Soetjipto, S. Potential Inhibitor of COVID-19 Main Protease (Mpro) From Several Medicinal Plant Compounds by Molecular Docking Study. Preprints 2020. [Google Scholar] [CrossRef] [Green Version]
  54. Crosby, D.C.; Lei, X.Y.; Gibbs, C.G.; McDougall, B.R.; Robinson, W.E.; Reinecke, M.G. Design, synthesis, and biological evaluation of novel hybrid dicaffeoyltartaric/diketo acid and tetrazole-substituted lchicoric acid analogue inhibitors of human immunodeficiency virus type 1 integrase. J. Med. Chem. 2010, 53, 8161–8175. [Google Scholar] [CrossRef] [PubMed]
  55. Hazuda, D.J.; Felock, P.; Witmer, M.; Wolfe, A.; Stillmock, K.; Grobler, J.A.; Espeseth, A.; Gabryelski, L.; Schleif, W.; Blau, C. Inhibitors of strand transfer that prevent integration and inhibit HIV-1 replication in cells. Science 2000, 287, 646–650. [Google Scholar] [CrossRef] [PubMed]
  56. Reinke, R.A.; Lee, D.J.; McDougall, B.R. l-Chicoric acid inhibits human immunodeficiency virus type 1 integration in vivo and is a noncompetitive but reversible inhibitor of HIV-1 integrase in vitro. Virology 2004, 326, 203–219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Moshe, R.; Paul, O.; Igor, K. Cepharanthine: A review of the antiviral potential of a Japanese-approved alopecia drug in COVID-19. Pharmacol. Rep. 2020, 72, 1509–1516. [Google Scholar]
  58. Ohashi, H.; Watashi, K.; Saso, W.; Shionoya, K.; Iwanami, S.; Hirokawa, T.; Shirai, T.; Kanaya, S.; Ito, Y.; Kim, K.S.; et al. Potential anti-COVID-19 agents, cepharanthine and nelfinavir, and their usage for combination treatment. iScience 2021, 23, 102367. [Google Scholar] [CrossRef] [PubMed]
  59. Lan, J.; Ge, J.; Yu, J.; Shan, S.; Zhou, H.; Fan, S.; Zhang, Q.; Shi, X.; Wang, Q.; Zhang, L. Structure of the SARS-CoV-2 spike receptor-binding domain bound to the ACE2 receptor. Nature 2020, 581, 215–220. [Google Scholar] [CrossRef] [Green Version]
  60. Walls, A.C.; Park, Y.J.; Tortorici, M.A.; Wall, A.; McGuire, A.T.; Veesler, D. Structure, function, and antigenicity of the SARS-CoV-2 spike glycoprotein. Cell 2020, 181, 281–292. [Google Scholar] [CrossRef]
  61. Deguchi, A. Curcumin targets in inflammation and cancer. Endocr. Metab. Immune Disord. Drug Targets 2015, 15, 88–96. [Google Scholar] [CrossRef]
  62. Pulido-Moran, M.; Moreno-Fernandez, J.; Ramirez-Tortosa, C.; Ramirez-Tortosa, M. Curcumin and Health. Molecules 2016, 21, 264. [Google Scholar] [CrossRef] [PubMed]
  63. Kim, Y.; Clifton, P. Curcumin, Cardiometabolic Health and Dementia. Int. J. Environ. Res. Public Health 2018, 15, 2093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Lelli, D.; Sahebkar, A.; Johnston, T.P.; Pedone, C. Curcumin use in pulmonary diseases: State of the art and future perspectives. Pharmacol. Res. 2017, 115, 133–148. [Google Scholar] [CrossRef] [PubMed]
  65. Bright, J.J. Curcumin and autoimmune disease. Adv. Exp. Med. Biol. 2007, 595, 425–451. [Google Scholar]
  66. Sordillo, P.P.; Helson, L. Curcumin Suppression of Cytokine Release and Cytokine Storm. A Potential Therapy for Patients with Ebola and Other Severe Viral Infections. In Vivo 2015, 29, 1–4. [Google Scholar]
  67. Achan, J.; Talisuna, A.O.; Erhart, A.; Yeka, A.; Tibenderana, J.K.; Baliraine, F.N.; Rosenthal, P.J.; D’Alessandro, U. Quinine, an old anti-malarial drug in a modern world: Role in the treatment of malaria. Malar. J. 2011, 10, 144. [Google Scholar] [CrossRef] [Green Version]
  68. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 1997, 23, 3–25, reprinted in Adv. Drug Deliv. Rev. 2001, 46, 3–26. [Google Scholar] [CrossRef]
  69. Lipinski, C.A. Lead- and drug-like compounds: The rule-of-five revolution. Drug Discov. Today Technol. 2004, 1, 337–341. [Google Scholar]
  70. Clark, D.E. Rapid calculation of polar molecular surface area and its application to the prediction of transport phenomena 1. Prediction of intestinal absorption. J. Pharm. Sci. 1999, 88, 807–814. [Google Scholar]
  71. Chang, L.C.W.; Spanjersberg, R.F.; von Frijtag Drabbe, K.; Unzel, J.K.; Mulder-Krieger, T.; van den Hout, G.; Beukers, M.W.; Brussee, J.; Ijzerman, A.P. 2,4,6-Trisubstituted pyrimidines as a new class of selective adenosine A, receptor antagonists. J. Med. Chem. 2004, 47, 6529–6540. [Google Scholar] [CrossRef]
  72. Veber, D.F.; Johnson, S.R.; Cheng, H.Y.; Smith, B.R.; Ward, K.W.; Kopple, K.D. Molecular properties that influence the oral bioavailability of drug candidates. J. Med. Chem. 2002, 45, 2615–2623. [Google Scholar] [CrossRef]
  73. Andersson, S.; Antonsson, M.; Elebring, M.; Jansson-Lofmark, R.; Weidolf, L. Drug metabolism and pharmacokinetic strategies for oligonucleotide- and mRNA-based drug development. Drug Discov. Today 2018, 23, 1733–1745. [Google Scholar] [CrossRef] [PubMed]
  74. Yu, J.; Zhou, Z.; Tay-Sontheime, J.; Levy, R.H.; Ragueneau-Majlessi, I. Intestinal Drug Interactions Mediated by OATPs: A Systematic Review of Preclinical and Clinical Findings. J. Pharm. Sci. 2017, 106, 2312–2325. [Google Scholar] [CrossRef] [Green Version]
  75. Lynch, T.; Price, A. The effect of cytochrome P450 metabolism on drug response, interactions, and adverse effects. Am. Fam. Physician 2007, 1, 391–396. [Google Scholar]
  76. Paz, S.; Vilasco, M.; Werden, S.J.; Arguello, M.; Joseph-Pillai, D.; Zhao, T.; Nguyen, T.L.; Sun, Q.; Meurs, E.F.; Lin, R. A functional C-terminal TRAF3- binding site in MAVS participates in positive and negative regulation of the IFN antiviral response. Cell Res. 2011, 21, 895–910. [Google Scholar] [CrossRef] [PubMed]
  77. Gao, Y.; Yan, L.; Huang, Y.; Liu, F.; Zhao, Y.; Cao, L.; Wang, T.; Sun, Q.; Ming, Z.; Zhang, L.; et al. Structure of the RNA-dependent RNA polymerase from COVID-19 virus. Science 2020, 368, 779–782. [Google Scholar] [CrossRef] [Green Version]
  78. Berman, H.M.; Battistuz, T.; Bhat, T. The protein data bank. Acta Crystallogr. Sect. D Biol. Crystallogr. 2002, 58, 899–907. [Google Scholar] [CrossRef]
  79. Giménez, B.G.; Santos, M.S.; Ferrarin, M.; Dos Santos Fernandes, J.P. Evaluation of blockbuster drugs under the rule-of-five. Pharmazie 2010, 65, 148–152. [Google Scholar]
  80. Afonine, P.V.; Klaholz, B.P.; Moriarty, N.W.; Poon, B.K.; Sobolev, O.V.; Terwilliger, T.C.; Adams, P.D.; Urzhumtsev, A. New tools for the analysis and validation of cryo-EM maps and atomic models. Acta Crystallogr. D Struct. Bio. 2018, 74, 814–840. [Google Scholar] [CrossRef] [Green Version]
  81. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couc, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structures of docking ligands from ZINC drug database.
Figure 1. Structures of docking ligands from ZINC drug database.
Molecules 27 02662 g001
Figure 2. Histogram showing molecular docking results between (A) 6W9C, (B) 6M17, and (C) 6M71 and several drug candidates compounds (the binding energy value ΔG is shown in minus kcal/mol).
Figure 2. Histogram showing molecular docking results between (A) 6W9C, (B) 6M17, and (C) 6M71 and several drug candidates compounds (the binding energy value ΔG is shown in minus kcal/mol).
Molecules 27 02662 g002
Figure 3. Chimera visualization of 6W9C docking with chicoric acid (1), curcumin (2), demothoxycurcumin, and (3) withanolide D (4). The yellow dots show H-bonds.
Figure 3. Chimera visualization of 6W9C docking with chicoric acid (1), curcumin (2), demothoxycurcumin, and (3) withanolide D (4). The yellow dots show H-bonds.
Molecules 27 02662 g003
Figure 4. Chimera visualization of 6M17 docking with chicoric acid (1), quinine (2), withaferin A (3) and withanolide D (4). The yellow dots show H-bonds.
Figure 4. Chimera visualization of 6M17 docking with chicoric acid (1), quinine (2), withaferin A (3) and withanolide D (4). The yellow dots show H-bonds.
Molecules 27 02662 g004
Figure 5. Chimera visualization of 6M71 docking with chicoric acid (1), curcumin (2), demothoxycurcumin (3) and quinine (4). The yellow dots show H-bonds.
Figure 5. Chimera visualization of 6M71 docking with chicoric acid (1), curcumin (2), demothoxycurcumin (3) and quinine (4). The yellow dots show H-bonds.
Molecules 27 02662 g005
Figure 6. The proposed mechanism of SARS-CoV-2 inhibition by tested compounds.
Figure 6. The proposed mechanism of SARS-CoV-2 inhibition by tested compounds.
Molecules 27 02662 g006
Table 1. Traditional medicinal source for bioactive compounds docking ligands toward COVID-19.
Table 1. Traditional medicinal source for bioactive compounds docking ligands toward COVID-19.
LigandsTraditional Medicine SourceRegionReported Biological ActivityReferences
1Citrullus colocynthis (fruit coat), C. colocynthis (Seeds),
Paraguariensis
Mesona chinensis
Echinacea purpurea
Egypt,
Poland,
Argentina,
China
Antiviral activities
Antioxidant activity
[30]
[31]
2Stephania cepharantha
Stephania rotunda
Japan Anti-inflammatory activity, Antiparasitic activities,
Anti-oxidative properties, Antiviral activities and
Anti-HIV activity
[32,33,34,35]
3Sickle seagrass, starflower, sweet basil, African basil, Chicory and Echinacea purpureaEgypt, China, India, and North AmericaAntivirus, anti- inflammation,
glucose and lipid homeostasis, Neuroprotection,
antioxidation effects.
Antimicrobial activity and
Antioxidant activity
[36,37,38]
4Curcuma longa L.Southern Asia,
China,
India,
Indonesia,
Indochina
Anti-inflammatory activity, Antioxidant activity,
Anti-bacteria activity Chemopreventive and Chemotherapeutic activity
Anti-HIV activity,
and Nematocidal activities
[39,40,41]
5Curcuma longa L.Southern Asia,
China,
India,
Indonesia,
Indochina,
Antimicrobial activity, Anti-inflammatory activity, Antioxidant activity, Platelet aggregation inhibitory activity, Antiallergy activity, Anticancer activity[42,43,44]
6Chinchona barkPeru, Bolivia, Colombia, Ecuador, India, and Sri Lanka.Antimalarial activity, Antioxidant activity, Anti-cancer agent, Anti-inflammatory, Antiparasitic activity and Antimicrobial property[45]
7Black teaAsia and EuropeAntioxidant activity and Anti-cancer activity[46,47]
8Withania somnifera L.Sri LankaAnti-cancer activity and Anti-COVID activity[48]
9Withania somnifera L.
Datura metel L. leaves
IndiaAnti-inflammatory and Antioxidant activity[49,50]
10Java turmeric
Curcuma species (C. zedoaria, C. xanthorrhiza, C. aeruginosa and C. mangga)
India and Southeast AsiaAnti-inflammatory, Antioxidant, and Anti-cancer activities[51,52]
1: Caffeic acid; 2: Cepharanthine; 3: Chicoric acid; 4: Curcumin; 5: Demothoxycurcumin; 6: Quinine; 7: Theaflavin; 8: withaferin A; 9: withanolide D; 10: Xanthorrhizol.
Table 2. Molecular docking analysis of several compounds against papain-like protease (6W9C), RNA-dependent RNA polymerase (6M71), and spike protein (6M17) of SARS CoV-2.
Table 2. Molecular docking analysis of several compounds against papain-like protease (6W9C), RNA-dependent RNA polymerase (6M71), and spike protein (6M17) of SARS CoV-2.
Ligand Name12345678910
Molecular formulaC9H8O4C37H38N2O6C22H18O12C21H20O6C20H18O5C20H24N2O2C29H24O12C28H38O6C28H38O6C15H22O
ClassificationPhenolics AlkaloidPhenylpropanoidPhenolicsPhenolicsAlkaloidPhenolicsPhenolicsPhenolicsSesquiterpenoid
6W9C
Binding energy ΔG−5.70-−7.62−6.70−6.81−5.67-−6.13−6.58−6.16
No. of H bonding6-4362-122
Binding sites CYS 270, GLN 269, TYR 268, ASN 267 and THR 265-CYS 270, TYR 268, ASN 267 and
ASP 286
THR 265 and
TYR 268
CYS 270, TYR 268, ASN 267, THR 265 and
GLU 263
ASP 286 and
LEU 290
-TYR 264ASN 267 and
TYR 268
ASN 267 and
TYR 268
6M71
Binding energy ΔG−6.75-−7.50−7.80−7.64−7.80−2.55−6.27−6.10−5.33
No. of H bonding6-411847472
Binding sites THR 680,
ARG 555,
ARG 553,
THR 556,
and SER 682
-ALA 554, ARG 553 and ARG 555ARG 555,
ARG 553,
THR 556,
and LYS 545
ARG 555,
and THR 556,
ARG 555,
THR 556,
SER 682
and ALA 554
ARG 553,
THR 556,
and SER 682
THR 556,
and SER 682
THR 556,
SER 682 and
ASP 623
TYR 455 and ALA 554
6M17
Binding energy ΔG−5.06−5.26−8.63−6.34−7.23−7.85−7.43−7.85−7.78−6.00
No. of H bonding21441471273
Binding sites ASN 487GLN 76GLU 484,
CYS 488, GLN 493 and GLN 76
ASP 355, THR 500
and THR 324
ASN 487CYS 488,
GLY 484 and ASN 487
THR 324, GLY 404, ARG 408 and THR 508THR 324, SER 375, TYR 376 and TYR 508THR 324, SER 375, and TYR 508THR 324 and VAL 503
1: Caffeic acid; 2: Cepharanthine; 3: Chicoric acid; 4: Curcumin; 5: Demothoxycurcumin; 6: Quinine; 7: Theaflavin; 8: withaferin A; 9: withanolide D; 10: Xanthorrhizol.
Table 3. Predicted drug likeness and pharmacokinetics of SARS-CoV-2 potential inhibitors.
Table 3. Predicted drug likeness and pharmacokinetics of SARS-CoV-2 potential inhibitors.
Ligands Name12345678910
Lipinski’s Rule of Five
Molecular weight (<500 Da)180.16 606.71474.37368.38338.35324.42564.49470.60470.60218.33
LogP (<5)0.935.351.013.033.002.811.312.292.364.34
No. rotatable bonds (<15) 22118742324
No. H-Bond donors (5)3062219441
No. H-bond acceptors (<10)481265412661
TPSA Å 77.7661.86208.1293.0683.8345.59217.60115.06115.0620.23
Violations0000003000
Pharmacokinetics
GI absorption HighHighLowHighHighHighLowHighHighHigh
BBBNoNoNoNoNoYesNoNoNoYes
P-gp substrateNoNoYesNoNoNoNoYesYesNo
CYP1A2 inhibitorNoNoNoNoYesNoNoNoNoNo
CYP2C19 inhibitorNoNoNoNoNoNoNoNoNoNo
CYP2C9 inhibitorNoNoNoYesYesNoYesNoNoYes
CYP2D6 inhibitorNoNoNoNoNoYesNoNoNoYes
CYP3A4 inhibitorNoNoNoYesYesNoYesNoNoNo
1: Caffeic acid; 2: Cepharanthine; 3: Chicoric acid; 4: Curcumin; 5: Demothoxycurcumin; 6: Quinine; 7: Theaflavin; 8: withaferin A; 9: withanolide D; 10: Xanthorrhizol.
Table 4. Protein target amino acids for molecular docking.
Table 4. Protein target amino acids for molecular docking.
Amino Acids HitsPapain-Like Protease (PLpro)
(6W9C)
RNA Dependent RNA Polymerase (RdRp)
(6M71)
Spike Protein (S Protein)
(6M17)
ASP 286, HIS 272, and CYS 111ARG 553, ARG 555, and LYS 545GLY 502, TYR 489, and TYR 505
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abd El-Aziz, N.M.; Khalifa, I.; Darwish, A.M.G.; Badr, A.N.; Aljumayi, H.; Hafez, E.-S.; Shehata, M.G. Docking Analysis of Some Bioactive Compounds from Traditional Plants against SARS-CoV-2 Target Proteins. Molecules 2022, 27, 2662. https://doi.org/10.3390/molecules27092662

AMA Style

Abd El-Aziz NM, Khalifa I, Darwish AMG, Badr AN, Aljumayi H, Hafez E-S, Shehata MG. Docking Analysis of Some Bioactive Compounds from Traditional Plants against SARS-CoV-2 Target Proteins. Molecules. 2022; 27(9):2662. https://doi.org/10.3390/molecules27092662

Chicago/Turabian Style

Abd El-Aziz, Nourhan M., Ibrahim Khalifa, Amira M. G. Darwish, Ahmed N. Badr, Huda Aljumayi, El-Sayed Hafez, and Mohamed G. Shehata. 2022. "Docking Analysis of Some Bioactive Compounds from Traditional Plants against SARS-CoV-2 Target Proteins" Molecules 27, no. 9: 2662. https://doi.org/10.3390/molecules27092662

Article Metrics

Back to TopTop