Next Article in Journal
Comparative Study of Chemically Treated Sugarcane and Kevlar Fiber to Develop Brake Resistance Composites
Previous Article in Journal
The Introduction of Detergents in Thermal Proteome Profiling Requires Lowering the Applied Temperatures for Efficient Target Protein Identification
Previous Article in Special Issue
Anti-Inflammatory Effects of Cycloheterophyllin on Dinitrochlorobenzene-Induced Atopic Dermatitis in HaCaT Cells and BALB/c Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Comprehensive Update of Various Attempts by Medicinal Chemists to Combat COVID-19 through Natural Products

1
Department of Chemistry, Government College University, Faisalabad 38000, Pakistan
2
Department of Chemistry, Government College Women University, Faisalabad 38000, Pakistan
3
Department of Bioinformatics and Biotechnology, Government College University, Faisalabad 38000, Pakistan
4
Department of Pharmaceutical Chemistry, Faculty of Pharmaceutical Sciences, Government College University, Faisalabad 38000, Pakistan
5
Department of Chemistry, Faculty of Science, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11623, Saudi Arabia
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(12), 4860; https://doi.org/10.3390/molecules28124860
Submission received: 9 May 2023 / Revised: 5 June 2023 / Accepted: 9 June 2023 / Published: 20 June 2023
(This article belongs to the Special Issue Natural Product Chemistry in Drug Discovery)

Abstract

:
The ongoing COVID-19 pandemic has resulted in a global panic because of its continual evolution and recurring spikes. This serious malignancy is caused by the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). Since the outbreak, millions of people have been affected from December 2019 till now, which has led to a great surge in finding treatments. Despite trying to handle the pandemic with the repurposing of some drugs, such as chloroquine, hydroxychloroquine, remdesivir, lopinavir, ivermectin, etc., against COVID-19, the SARS-CoV-2 virus continues its out-of-control spread. There is a dire need to identify a new regimen of natural products to combat the deadly viral disease. This article deals with the literature reports to date of natural products showing inhibitory activity towards SARS-CoV-2 through different approaches, such as in vivo, in vitro, and in silico studies. Natural compounds targeting the proteins of SARS-CoV-2—the main protease (Mpro), papain-like protease (PLpro), spike proteins, RNA-dependent RNA polymerase (RdRp), endoribonuclease, exoribonuclease, helicase, nucleocapsid, methyltransferase, adeno diphosphate (ADP) phosphatase, other nonstructural proteins, and envelope proteins—were extracted mainly from plants, and some were isolated from bacteria, algae, fungi, and a few marine organisms.

1. Introduction

The “Coronavirus Disease 2019” (COVID-19) was first found as an endemic in Wuhan, China, and was declared a global pandemic by the World Health Organization (WHO) on 11 March 2020 [1,2]. COVID-19 has become a persistent threat to public health and an international concern in the scientific community because of its rapid spread. To date (7 January 2023), more than 667 million cases and 6.70 million associated deaths of the coronavirus disease 2019 (COVID-19) have been recorded worldwide (https://www.worldometers.info/coronavirus/ (accessed on 7 January 2023)). COVID-19 is caused by “Severe Acute Respiratory Syndrome Coronavirus 2” (SARS-CoV-2), which is more infectious and fatal, with a reproductive rate of 2.5 (range 1.8–3.6) compared to the hemaglutinin1-neuraminidase1 (H1N1) influenza A virus, the SARS-CoV reproductive rate, i.e., (range 2–3), and the Middle East respiratory syndrome coronavirus (MERS-CoV) with a 0.9 reproductive rate [3]. Due to its rapid transmission from person to person, it has been considered a highly contagious malignancy, leading from a common cold or mild flu to a fatal disease [4]. Its common symptoms include a throat infection, headache, fever, fatigue, dry cough, breath shortness, body aches, loss of taste and smell, and pneumonia [5]. The symptoms become more severe if the affected person is elderly, has a weak immune system, or has suffered from other illnesses, such as obesity, diabetes, and cardiovascular and pulmonary disorders [6,7].
The infection of SARS-CoV-2 starts with the entry of the virion through interaction with the human ACE2 (angiotensin-converting enzyme 2) receptor cells using spike glycoproteins present on its surface [8]. Open reading frames (ORF1a and ORF1b) use host cell ribosomes to generate polyproteins pp1a and pp1b, which are then processed by main proteases to generate 16 nonstructural proteins (NSPs). All these NSPs play their own significant roles in the replication and transcription of the viral genome [9]. After synthesizing structural proteins, viral envelope formation is carried out in the endoplasmic reticulum–Golgi intermediate complex and released from the host cell through budding [10,11].
Coronaviruses are categorized into four classes: Alpha, Beta, Gamma, and Delta. MERS-CoV, SARS-CoV, and SARS-CoV-2 belong to the Beta class of the Coronoviridae family [12]. The novel SARS-CoV-2 is an enveloped, positive-sense single-stranded RNA virus with a genome length in the range of 80–120 nm [13]. It is reported that, genetically, the virus possesses similarities with SARS-CoV or MERS-CoV [14]. Studies have shown that the genetic material encodes four structural proteins, spike protein (S), envelope protein (E), nucleocapsid protein (N), and matrix protein (M), sixteen nonstructural proteins, and nine accessory proteins [15]. The main protease (Mpro), papain-like protease (PLpro), RdRp, and spike glycoproteins are considered to play significant roles in the transcription and translation of the viral genome and ultimately lead to virus spread, so these are active targets for drug development against SARS-CoV-2 [16,17].
To combat the pandemic spanning 203 countries [18], different strategies were adopted, including lockdown, quarantine measures, social distancing, proper hygiene practice, face covering, and controlling sanitary conditions strictly [19]. All of these measures contributed to a reduction in normal social contacts [20]. Moreover, some already available antimalarial, immunomodulatory, and antiviral drugs (Figure 1)—chloroquine 1, hydroxychloroquine 2, ivermectin 3, azithromycin 4, remdesivir 5, lopinavir 6, ritonavir 7, favipiravir 8, galidesivir 9, dexamethasone 10, and ruxolitinib 11—have been repurposed for COVID-19 treatment [21,22,23,24]. Among the above-mentioned drugs, remdesivir, lopinavir/ritonavir, and chloroquine (or hydroxychloroquine) have received increased scientific attention, but only remdesivir has been approved by the Food and Drug Administration (FDA) for the treatment of patients with COVID-19 [25].
As the number of cases grew, understanding the immune response to SARS-CoV-2 became critical. Multiple vaccines were developed, including BNT162 vaccine by Pfizer and BioNTech; mRNA-1273 vaccine by Moderna; AZD1222 by AstraZeneca and the University of Oxford; CoronaVac by Sinovac; the COVID-19 vaccine by Sinopharm and the Wuhan Institute of Virology, China; Sputnik V by the Gamaleya Research Institute, Russia; BBIBP-CorV by Sinopharm and the Beijing Institute of Biological Products, China; and EpiVacCorona by the Federal Budgetary Research Institution State Research Center of Virology and Biotechnology, Russia [26]. In the beginning, vaccine hesitancy was constituted as a threat to tackling the COVID-19 pandemic because herd immunity depended on both the availability of vaccines and the population’s willingness to accept those vaccines [27]. But later on, the efficacies of the vaccines were shown with Pfizer at 95%, Moderna at 94.1%, and AstraZeneca at 70.4%, proving that these vaccines are effective at reducing the incidence and severity of SARS-CoV-2 infection among the study populations [28].
There was an emergence of SARS-CoV-2 variants due to the changes in the nucleotides that occur naturally during replication. The SARS-CoV-2 variants of concern that have emerged till today are the Alpha variant (B.1.1.7) that was detected in the UK in September 2020, the Beta variant (B.1.351) detected in South Africa in October 2020, the Gamma variant (P.1) detected in Brazil in November 2020, the Delta variant (B.1.617.2) detected in India in December 2020, and the Omicron variant (B.1.1.529) detected in South Africa in November 2021 [29]. These circulating SARS-CoV-2 variants are challenging therapeutic actions against COVID-19.
Natural products have effectively acted as lead compounds for various infectious diseases. On this ground, various research groups across the globe working in the field of medicinal chemistry have focused on known natural products for their ability to inhibit COVID-19. We reviewed the available literature on natural compounds that show some evidence of SARS-CoV-2 inhibitory potential: first, those for which in vivo and in vitro studies were performed, next those for which a combination of in vitro and in silico studies was carried out, and finally, a wide range of compounds for which only in silico studies were conducted. The latter is divided into sections for various SARS-CoV-2 proteins and the compounds that show evidence of interaction with them.

2. In Vivo Studies

Some marine natural products, including homofascaplysin A 12, (+)-aureol 13, and bromophycolide A 14 (Figure 2), have been reported as inhibitors of SARS-CoV-2 in human airway epithelial cells. SARS-CoV-2 inhibition was assessed by finding EC50 and CC50 values in human Calu-3 cells, i.e., (1.1 ± 0.4 μM, 4.0 ± 1.0 μM, and 6.9 ± 2.0 μM) and (~5 μM, >10 μM, and >10 μM), respectively, compared to those of remdesivir with EC50 (0.3 ± 0.0 μM) and CC50 (>5 μM) values [30].

3. In Vitro Studies

Baicalin 15 and baicalein 16 (isolated from S. baicalensis) were identified as inhibitors of 3-chymotrypsin-like protease (3CLpro), the main protease of SARS-CoV-2, through in vitro studies. The IC50 values were determined via a fluorescence resonance energy transfer (FRET)-based protease assay for baicalin 15 and baicalein 16, i.e., 6.41 μM and 0.94 μM, respectively. Half-maximal effective concentrations (EC50) of both compounds were found to be 10.27 μM and 1.69 μM, respectively [31]. Liu et al. (2021) also tested the ethanolic extract of S. baicalensis and baicalein 16 for their inhibitory action on 3CLpro of SARS-CoV-2. Both possessed inhibitory potentials with IC50 values of 8.52 μg/mL and 0.39 μM, respectively [32]. Zandi et al. (2021) reported their inhibitory potential against RdRp, the enzyme responsible for the replication of SARS-CoV-2, in Vero cells and in Calu-3 cells as well. Baicalein 16 was reported as a more potent compound with EC50 values of 4.5 ± 0.2 μM in Vero cells and 1.2 ± 0.03 μM in Calu-3 cells [33].
He et al. (2021) identified cepharanthine 17, a bis-benzylisoquinoline alkaloid, as the active drug candidate against SARS-CoV-2 with an EC50 value of 3.35 μM. Studies revealed that mechanistically, Compound 17 performed the pivotal role as a Ca-channel blocker and, hence, caused suppression in SARS-CoV-2 entry [34]. Sa-ngiamsuntorn et al. (2020) isolated andrographolide 18 from Andrographis paniculata and tested it in human lung epithelial cells (Calu-3) via an in vitro antiviral assay against SARS-CoV-2. The IC50, as determined by a plaque assay, was found to be 0.034 μM [35]. Saadh et al. (2021) demonstrated that zinc sulphate in combination with sauchinone, i.e., sauchinone/Zn-II 19, showed more additive and inhibitory effects against 3CLpro of SARS-CoV-2. Zinc, when combined with sauchinone, a well-known antiviral drug, displayed a 2.02-fold greater decrease in its inhibitory effect, i.e., the IC50 value of sauchinone was 4.325 μM [36]. Brown et al. (2021) identified the propylamylatinTM formula as an effective inhibitor of SARS-CoV-2 by performing plaque assays in Vero E6 cells. A 90% effective inactivation potential (EC90) of propylamylatinTM 20 was observed to be 4.28 μls, which was found to be better than the individual components (propionic acid; EC90 = 11.50 μls, isoamyl hexanoates; EC90 = 10-fold reduction in viral infection) of the mixture [37]. Liu et al. (2021) identified that the infection caused by SARS-CoV-2 was significantly suppressed by green tea beverages and their active catechin components. Among various tea catechins, epigallocatechin gallate (EGCG) 21 (Figure 3) effectively blocked viral entry into the host cell by interacting with the viral spike proteins and host ACE2 receptor (EC50 value = 43.48–107.6 ng/mL) with noncytotoxic doses [38].
Similarly, Tun et al. (2022) reported the inhibition of SARS-CoV-2 main protease by six types of Japanese green tea beverages and tea ingredients. These six types of tea beverages inhibited SARS-CoV-2 infection by 70–88% in a dilution-dependent way. The tea ingredients epigallocatechin gallate (EGCG) 21 and epicatechin gallate 22 (Figure 3) suppressed the antiviral activity with IC50 values of 12.5 and 6.5 μM, respectively. Moreover, these active compounds also interact with the viral entry into the host [39]. The antiviral potential of Spatholobus suberectus Dunn extract against SARS-CoV-2 infection was assessed using an in vitro analysis. The results have shown that the plant extract effectively blocks the viral spike proteins and host ACE2 cells interaction, and EC50 values were obtained in the range of 3.6 to 5.1 μg/mL. Further, in vivo studies showed that it could act as a potential inhibitor with no toxicity in long-term treatment [40]. Eggers et al. (2022) analyzed the inhibitory effect of various plant juices (black chokeberry, elderberry, and pomegranate) and green tea against SARS-CoV-2. Cell-based assays demonstrated a reduction in viral infection by ≥80% or ≥99%, while black chokeberry juice was found to be the most effective suppressor of SARS-CoV-2 with a percentage inhibition of 96% [41]. Baeshen et al. (2022) explored the anti-SARS-CoV-2 activity of natural extracts of the desert medicinal plant Rhazya stricta in a dose-dependent way. The highest activity was observed for the nonalkaloidal fraction (IC50: 0.0461 mg/mL; CC50: 0.18 mg/mL), and then weak base alkaloids had activity with an IC50 of 0.0474 mg/mL and a CC50 of 0.0464 mg/mL [42]. A natural nutraceutical, BEN815, contained extracts of green tea leaves (Camellia sinensis), guava leaves (Psidium guajava), and rose petals (Rosa hybrida) that displayed an antiviral effect against SARS-CoV-2 (IC50 value = 34.38 μg/mL). This study also showed that among the various components of BEN815, EGCG 21 demonstrated the highest inhibitory potential with an IC50 of 33.41 μM [43].

4. In Vitro and In Silico Studies

Khan et al. (2021) confirmed that kaempferol 23 exhibited good inhibitory activity against 3CLpro of SARS-CoV-2 in vitro as well as through docking analysis. The inhibitory role played by kaempferol 23 on SARS-CoV-2 was found to be 34.46 μM in Vero E6 cells [44]. Alhadrami et al. (2021) demonstrated cnicin 24 as the potential inhibitor of the nonstructural proteins, RdRp (NSP12), ADPRP (NSP3), and endoribonuclease (NSP15) of SARS-CoV-2. The inhibition potency was validated by an IC50 value of 1.18 μg/mL and a CC50 value of 59.66 μg/mL. The selectivity index (CC50/IC50) was found to be 70.3. In vitro testing confirmed the results of the molecular docking simulations [45]. Li et al. (2021) evaluated some aromatic sesquiterpenoids as SARS-CoV-2 spike–ACE2 interaction inhibitors through an in vitro analysis, and their binding interactions were studied through molecular docking studies. Among the studied compounds, candinone sesquiterpenoids 25 showed the lowest IC50 value, i.e., 64.5 ± 1.8 μM. Their structure–activity relationship (SAR) has shown that the pharmacophore of candinone sesquiterpenoids 25 might be the hydroxyl and carboxyl groups [46]. Hijikata et al. (2021) evaluated cepharanthine 17 through in vitro as well as in silico studies. In vitro analysis revealed that the inhibitory concentration (IC50) of cepharanthine 17 against SARS-CoV-2 by interacting with the human lysosomal membrane protein Niemann–Pick type C intracellular cholesterol transporter 1 (NPC1) was found to be 1.90 μM. The diphenyl ester moiety was suggested as the pharmacophore of cepharanthine 17. Molecular docking was also carried out on AutoDock 4 (AD4) and Autodock Vina (ADV) software, and the root mean square deviation (RMSD) value obtained between the best docking poses was 1.3 in the S1 pocket of NPC1 [47]. Dogan et al. (2021) isolated artemisinin 26 from Artemisia annua L. and evaluated it against the Mpro and spike proteins of SARS-CoV-2 through in vitro and in silico evaluations. Artemisinin 26 showed maximum inhibitory potency, i.e., >200 μM, against SARS-CoV-2 in the HEK293T cell line [48]. Owis et al. (2021) isolated a flavonoid, mauritianin 27, from Salvadorapersica L. and checked its inhibition (IC50 and CC50) towards 3CLpro of SARS-CoV-2, which were found to be 8.59 ± 0.3 μg/mL and 24.5 ± 1.9 μg/mL, respectively [49]. Morita et al. (2021) reported that all-trans retinoic acid 28 exhibited anti-SARS-CoV-2 activity against 3CLpro through Alpha screening and in vitro approaches. AlphaScreen software provided high-throughput screening of the active compounds. Further, a FRET assay was carried out, and it was noticed that the IC50 value obtained was 24.7 ± 1.65 μM. It (28) inhibited the replication of Vero E6 cells and Calu-3 cells with IC50 values of 2.69 ± 0.09 μM and 0.82 ± 0.01 μM, respectively [50].
Gizawy et al. (2021) reported some bioactive compounds extracted from Pimentadioica (L.) as anti-SARS-CoV-2 Mpro agents. Rutin 29 appeared to be the best active compound, as evident from both in silico and in vitro studies. The antiviral and cytotoxic activities were determined by IC50 and CC50 values, i.e., 31 μg/mL and 8017 μg/mL, respectively [51]. In silico observations of Compounds 2328 and 30 (Figure 4) are depicted in (Table 1). Hou et al. (2022) identified phloroglucinols terpenoids as the target inhibitors of 3CLpro of SARS-CoV-2 through virtual screening and in vitro approaches. The terpenoid 30, isolated from Dryopteris wallichiana, showed the best inhibitory activity in Vero E6 cells with an effective inhibitory concentration of 4.5 μM. Compound 30 was virtually screened on AutoDock 4 to see the binding interactions between the ligand and amino acid residues of 3CLpro. It was noted that Compound 30 formed H-bonds with MET165, ASN142, GLU166, and GLN192 and hydrophobic interactions with PHE140 [52]. Kim et al. (2022) identified the antiviral potential of mulberrin (kuwanon C) 31 (Figure 5), isolated from the commonly known mulberry plant. In vitro studies showed that this natural compound inhibited the SARS-CoV-2 infection by interacting with both the viral spike proteins and the host ACE2 receptor cells with an IC50 value of 91.4 μM at a concentration range of 25 μM to 100 μM. Furthermore, in silico studies demonstrated the more stable binding interactions of ligands with spike proteins compared to ACE2 receptors [53]. Some naturally occurring flavonoids (myricetin, quercetin, kaempferol, flavanone, and licoflavone C) were found to inhibit the enzymatic activities of SARS-CoV-2 by selectively antagonizing the action of NSP13 at nanomolar and micromolar concentrations. However, computational analysis has shown that among these active compounds, licoflavone C 32 (Figure 5) displayed the highest affinity and strong binding interactions with amino acid residues, such as LYS569, GLY538, HIS290, and ARG442, followed by myricetin, baicalein, kaempferol, and quercetin [54].
Xiao et al. (2021) investigated the inhibitory activity of 15 natural compounds (flavonoids, coumarins, terpenoids, phenolics, aldehydes, and steroids) towards the SARS-CoV-2 main protease using enzymatic and virtual screening analysis. Among them, myricetin 33 had the most potent activity with an IC50 of 3.684 ± 0.076 μM. Molecular docking showed interactions with the Mpro binding pocket via the chromone ring and its 3′-, 4′-, and 7-hydroxyl groups [55]. Compounds 34 and 35 and extracts of R. japonica and Reynoutriasachalinensis species were evaluated against the main protease of SARS-CoV-2 through in silico and in vitro studies. The compounds vanicoside A 34 and vanicoside B 35 showed the best results towards inhibition of the targeted enzyme (Mpro), having an IC50 of 23.10 μM and 43.59 μM, respectively. Plant butanol fractions demonstrated the greatest suppression of SARS-CoV-2’s Mpro (the IC50 values for R. sachalinensis and R. japonica were 4.031 g/mL and 7.877 g/mL, respectively [56]. Chaves et al. (2022) analyzed the anti-SARS-CoV-2 potential of some classes of flavonoids: flavonol (fisetin, kaempferol, myricetin, and quercetin), flavone (apigenin and luteolin), and isoflavone (genistein). It was found that flavonols had better inhibitory potential than flavones and isoflavones, and this was observed due to the presence of a greater number of hydroxyl groups in the B ring of flavonols. Among flavonols, fisetin 36 and myricetin 33 mainly targeted the main protease with EC50 values of 2.03 ± 0.10 μM and 0.91 ± 0.05 μM, respectively. Furthermore, in silico analysis confirmed the better inhibitory potential of these flavonols compared to other flavonoids [57]. Hafez et al. (2022) isolated the natural compound 37 from Ophiocoma dentata (brittle star) and proposed it as an active inhibitor of SARS-CoV-2’s Mpro, NSP10, and RdRp through in vitro and in silico studies. An effective dose of 12.48 μM demonstrated 95% inhibition, and Compound 37 displayed an IC50 of 11,350 ± 1500 ng/mL against normal fibroblast cells. Moreover, in silico analysis provided strong binding interactions between the studied compounds and targeted viral enzymes [58]. Silibinin 38, a naturally occurring flavonoid, suppressed SARS-CoV-2 infection by inhibiting the main protease, S-proteins, and RdRp with IC50 values of 0.021, 0.029, and 0.042 μM, respectively. It showed >90% inhibition of viral activity at a concentration of 0.031 μM. Moreover, in silico studies showed that this potent compound showed good binding affinity with both spike proteins and the main protease [59]. Elhusseiny et al. (2022) analyzed the antiviral potential of aqueous extracts of Agaricus (A.) bisporus, Lentinula (L.) edodes, and Pleurotus (P.) ostreatus against SARS-CoV-2’s main protease, with IC50 values of 10.3.3, 26.17, and 39.19 μg/mL, respectively. Furthermore, docking analysis showed that chlorogenic acid 39, kaempferol 23, quercetin 40, and catechin 41 were the most active compounds that were bound effectively with binding interactions ranging between −22.8 and −37.61 kcal/mol [60]. Lopes et al. (2022) isolated 55 bioactive plant compounds and evaluated their potential against the Mpro, RdRp, PLpro, NSP15, endoribonuclease, spike protein, and ACE2 using molecular docking and in vitro analysis. The docking results demonstrated that 7-O-galloylquercetin, amentoflavone, gallagic acid, and kaempferitrin displayed good binding interactions with the targeted enzymes of SARS-CoV-2. Among these, amentoflavone 42 suppressed the activity of 3CLpro with an IC50 value of 8.3 μM [61]. Nine withanolides from Ashwagandha were found to cause inhibition of transmembrane serine protease 2 (TMPRSS2) (a human gene) and the Mpro of SARS-CoV-2. Among these nine tested compounds, the best binding affinities were observed for withanoside V 43 and withanoside IV 44, which formed hydrogen bonds and other interactions with targeted proteins. Additionally, in vitro analysis through a 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl-2H-tetrazolium bromide (MTT) assay confirmed the antiviral potential of selected compounds at a concentration >10 mM [62]. Kim et al. (2022) identified the antiviral potential of mulberrofuran G 45 (Figure 5) isolated from the commonly known mulberry plant. In vitro studies showed that this natural compound inhibited the SARS-CoV-2 infection by interacting with both the viral spike proteins and the host ACE2 receptor cells with an IC50 of 10.23 μM at a concentration of less than 50 μM. Furthermore, in silico studies demonstrated the more stable binding interactions of ligands with spike proteins compared to ACE2 receptors [63] (Table 1).
Remdesivir was observed to be the most frequently used control group in many in vitro studies reported in the literature. Significant IC50 and EC50 values observed for remdesivir fell in the range of 0.086–13.1 μM [31,33,34] and 0.77–5 μM [35,43,47,59], respectively. Naturally occurring Compounds 16 (EC50 = 1.69 μM [31]), 17 (IC50 = 1.90 μM [47] and EC50 = 3.35 μM [34]), 18 (IC50 = 0.034 μM [35]), and 38 (IC50 = 0.021 μM [59]) showed the most potent inhibitory activity. Moreover, Compounds 15 (EC50 = 10.27 μM [31]) and 21 (IC50 = 33.41 μM [43]) showed moderate potential to inhibit SARS-CoV-2.
Table 1. In silico results for compounds for which in vitro studies were also performed.
Table 1. In silico results for compounds for which in vitro studies were also performed.
Serial No.CompoundBinding PotentialInteractions at Enzyme Active Site *References
Docking Score
(kcal/mol)
Binding Free Energy (kcal/mol)H-Bond InteractionsHydrophobic
Interactions
Main Protease (Mpro/3CLpro)
1Kaempferol (23)−6.4−26.81PHE140, LEU141, ASN142, CYS145, ARG188-[44]
2Artemisinin (26)−5.21−25.84ASN142LEU141, CYS145, VAL42, CYS44, MET49, LEU27, MET106[48]
3Rutin (29)−9.19-HIS163, GLU166(2), PHE140, CYS145HIS41[51]
4Myricetin (33)<−8.0-PHE140, GLU166, ASP187HIS41[55]
55α-cholesta-4 (27), 24-dien-3β, 23 β-diol (37)-−24.68HIS41MET49, MET165[58]
6Chlorogenic acid (39)-−24.9GLY143, CYS145MET165[60]
7Kaempferol (23)-−36.5GLU166, MET49HIS41, CYS145[60]
8Quercetin (40)-−39.66GLU166, GLY143, CYS145MET49[60]
9Catechin (41)-−39.66GLU166, ASN142, HIS163, PHE140MET165, LEU141[60]
10Amentoflavone (42)−8.7--CYS145[61]
Papain-Like Protease (PLpro)
11Amentoflavone (42)−7.7--TYR264[61]
Spike Proteins
12Candinone sesquiterpenoids (25)--ARG403, ARG405, ARG408, ARG393, HIS34GLU406, LYS417, ARG403, GLN409[46]
13Artemisinin (26)−5.06−30.61GLU23, LYS26, THR27, ASP30, LYS417,
TYR421, PHE456, ARG457, TYR473
[48]
14Mauritianin (27)-−9.4799ASP405, GLY496, LYS403, GLU37, ASP30HIS34, ALA387[49]
15Kuwanon C (31)−7.1-ARG403TYR453, GLN493, SER494, TYR495, GLY496, PHE497, GLN498, ASN501, GLY502, TYR505[53]
16Silibinin (38)−7.78-ASN907, LYS1038, ILE909, GLU1092, TYR904, ASN907, GLY904, TYR904TYR904, ASN907, GLY910, GLY908, LYS1038, GLU1092, TYR904[59]
17Amentoflavone (42)-8.7--TYR449, GLN493[61]
18Mulberrofuran G (45)−8.4-GLN493, TYR453 GLY496GLU406, LYS417, LEU455, SER494, TYR495, GLY496, GLN498, ASN501, TYR505[63]
(a) ACE2
19Amentoflavone (42)−9.1--ASP30, ASP32, GLN42, TYR83, LYS353[61]
20Mulberrofuran G (45)−7.4-GLU23, THR27, ASN33, GLN96GLU23, GLN24, LYS26, THR27, LEU29, ASP30, VAL93, AND PRO389[63]
(b) TMPRSS2
21Withanoside V (43)−7.96-GLU299, TYR337, SER339HIS296, LYS300, TYR337, ASP338, LYS340, THR341, LYS342, ASP435, SER436, CYS437, GLN438, GLY439, SER441, THR459, SER460, TRP461, GLY462, GLY464, CYS465, GLY472, VAL473[62]
22Withanoside IV (44)−6.92-ASP338, LYS342, GLU389, SER436, SER441HIS296, GLU299, TYR337, ASP435, CYS437, GLN438, GLY439, ASP440, THR459, SER460, TRP461, GLY462, GLY464, CYS465, ALA466, GLY472, VAL473[62]
RDRP
23Cnicin (24)−9.7−10.3LYS41, LEU49, LYS50, THR51, ASN52PHE35, ASP36, ILE37, TYR38, ASN39, VAL42, PHE48, HIS75, ASP208, ASP218, ASP221[45]
245α-cholesta-4(27), 24-dien-3β, 23 β-diol (37)-−29.86TR680, CYS622URD20, ADE11, LYS545, URD10, VAL557, ALA547[58]
25Silibinin (38)−7.15-U20, ASP618, ILE548, ASP618ARG836, ASP618, SER814, GLU811, LYS545, LYS551, ALA547, ASP760, ILE-548[59]
26Amentoflavone (42)−9.4--ASP618[61]
Endoribonuclease
27Cnicin (24)−9.8−9.3GLN245, HIS250, LYS290ASN278, VAL292, SER294, THR341, TYR343, PRO344, LYS345, LEU346[45]
Adeno Diphosphate (ADP)-Ribose Phosphatase/ADPRP
28Cnicin (24)−9.2−10.1ALA38, VAL49, ALA50, LEU126, ALA129, ILE131, PHE132, ALA154, PHE156, ASP157ALA21, ILE23, GLY47, PRO125, SER128, LEU160[45]
Nonstructural proteins
295α-cholesta-4(27),24-dien-3β,23-β-diol (37)-−23.47ASP6912LEU6898, PRO6932[58]
* Catalytic site residues are shown in blue, and binding site residues are shown in black in Table 1.

5. In Silico Studies

5.1. Natural Compounds as SARS-CoV-2 Main Protease Inhibitors

Recent studies have shown that during their life cycle, coronaviruses typically accumulate a few polypeptides and then develop proteolytic breakdown to produce an additional 20 proteins [11]. Among these proteins, it is highlighted that the main protease (Mpro/3CLpro) and papain-like protease (PLpro) play a crucial role in the transcription and replication of viruses. These proteases have been the subject of significant research to find specific COVID-19 inhibitors. Computational studies using various in silico approaches made it possible to find potential inhibitors against viral proteases.
Bernardi et al. (2021) used a computational virtual screening approach to investigate phenylethanoid glycosides as inhibitors of SARS-CoV-2’s main protease (Mpro). Three major phenylethanoid glycosides, forsythiaside A 46 (Figure 6), isoacteoside, and verbacoside, were studied due to their good docking scores and strong binding interactions at the active site [64]. (The docking scores and amino acid residues are listed in Table 2).
Some potential phytochemicals from Cameroonian plants and bioactive lactones from Saussureacostus were investigated to combat SARS-CoV-2’s Mpro. ADMET (absorption, distribution, metabolism, excretion, and toxicity) analysis was also carried out to check the pharmacological properties of pycnanthuquinone C 47 (Figure 6) and pycnanthuquinone A 48, which were extracted from Pycnanthusangolensis [65]. According to the visualization results, the interaction between the Mpro and cyanropicrin 49 was the best one, as proved by the molecular docking (MD) simulation study. The drug-likeness of cyanropicrin 49 was confirmed using ADMET analysis and Lipinski’s rule [66]. The phytocompounds from Indian medicinal plants and some polyphenols were examined as potential inhibitors of SARS-CoV-2’s main protease. The studies revealed that withanolide R 50 had the lowest relative free binding energy value and was declared to be the most potent among the studied compounds [67]. Eighty polyphenols were initially tested, and four of them—hesperidin 51, rutin 29, diosmin 52, and apiin 53—showed active inhibitory activity against the Mpro [68]. Phytochemicals from Jordanian hawksbeard, jaceidin 54, pachypodol 55, and chrysosplenetin 56 showed good binding affinities to the main protease (Mpro) of SARS-CoV-2 [70]. Imidazoline-4-one-2-imino-1-(4-methoxy-6-dimethylamino-1,3,5-triazin-2-yl) 57, spiro[4,5]dec-6-en-1-ol, 2,6,10,10-tetramethyl 58, and 3-hydroxy-5-cholen-24-oic acid 59 extracted from Tinosporacrispa showed the best binding affinities against the Mpro [71]. Withacoagulin H 60, ajugin E 61, withacoagulin 62 [72], crocin 63 (Figure 6) [73], rhamnocitrin 64 (Figure 7) from Artemisia annua [74], and sterenin M 65, a fungal metabolite [75], were shown to be the best active compounds against the Mpro. Rhamnocitrin 64 also possessed a favorable ADMET profile with no hepatotoxicity, carcinogenicity, mutagenicity, cytotoxicity, and immunotoxicity [74]. The inhibitory activity of honeybee natural products and some natural flavonoids and peonidin was studied against SARS-CoV-2’s main protease (Mpro) through an in silico evaluation. It was found that 3,4,5-tricaffeoylquinic acid 66 possessed the highest binding affinity value [76]. Quercetin 40 and the peonidin 67 moiety were shown to have inhibitory potency against the Mpro. Djiboutian medicinal plants and some natural products have potential as inhibitors of SARS-CoV-2’s Mpro. Nine molecules have been studied, out of which rutin 29, catechin 41, and kaempferol 23 showed the best binding affinities with the Mpro compared to the reference drug, remdesivir (−7.194 kcal/mol) [69], whereas amentoflavone 42 was found to be the best active compound compared to the reference drug hydroxychloroquine with a binding free energy value of −6.3 kcal/mol. Lipinski’s rule of five also showed its drug-like properties. Amentoflavone 42 can be found in many medicinal plants, such as Selaginellaceae, Cupressaceae, and Euphorbiaceae family plants [78]. (The binding affinities and amino acid residue interactions are depicted in Table 2).
Out of 10 active compounds found in Aloe vera, ferolide (68) (Figure 7) was demonstrated as the most potent compound against a viral protein, i.e., 3CLpro, an enzyme that plays a key role in post-translational protein regulation, particularly the cleavage of viral polyproteins into functional protein units. According to the results of virtual screening, it was observed that ferolide 68 also followed Lipinski’s rule of five to be used as a drug [80]. The terpenoid ginkgolide A 69, extracted from Ginkgo biloba [81], showed the highest S-score. Gracillin 70 extracted from Paris vietnamensis and proanthocyanidin 71 extracted from Cinnamomum sp. showed the lowest docking score compared to the reference drug, boceprevir (−7.7 kcal/mol) [82]. Ginkgolide M 72, mezerein 73, and tubocuraine 74 showed the best binding affinities against 3CLpro compared to the reference drugs nelfinavir and lopinavir, with −9.1 kcal/mol and −8.4 kcal/mol binding affinities, respectively [83]. Choline 75 exhibited drug-like properties approved via the Lipinski, Veber, and Egon rules. The pharmacokinetic study revealed that it also showed gastrointestinal absorption [84]. The biflavonoids amentoflavone 42 and volkensiflavone 76 displayed the highest binding affinity to 3CLpro.The biflavonoid amentoflavone 42 also exhibited the highest binding affinity to PLpro [79]. Jamhour et al. (2021) tested thirty-six phytochemicals under an in silico perspective; six (rutin, quercetin, catechin gallate, rhamnetin, campesterol, and stigmasterol) out of 36 were found to be bioactive. Stigmasterol 77 had the lowest binding energy value of −6.30 kcal/mol against 3CLpro [85]. Hesperidin 51 showed the highest binding affinity to 3CLpro compared to the standard drugs nelfinavir, hydroxychloroquine sulphate, and chloroquine [86]. Natural compounds extracted from Amphimedon sp. were investigated through a computational study. Amphimedoside C 78 was found to be the most active ligand against 3CLpro of SARS-CoV-2 [87]. The marine-derived bioactive compound fasciospongide A 79 was found to be the most active compound against 3CLpro, compared to the reference drugs lopinavir and ritonavir, which had molecular mechanics Poisson–Boltzmann surface area (MM/PBSA) scores of −101.13 kJ/mol and −97.40 kJ/mol, respectively. Constanolactone B 80 was revealed to be the most bioactive compound with a lower MM/PBSA score against PLpro compared to the reference drugs lopinavir and ritonavir, which had binding free energy values of −60.84 kJ/mol and −70.74 kJ/mol, respectively [88]. Glaucogenin D 81 was found to be the best active compound against 3CLpro, whereas glaucogenin D 81 and glaucogenin A 82 were observed to have the best binding affinity towards papain-like protease (PLpro) [89]. Mehmood, A. et al. (2021) identified the Quranic and prophetic medicinal plants capable of inhibiting SARS-CoV-2’s essential enzymatic functions of 3CLpro, the viral main proteinase, which was inhibited by calcium elenolate 83 [90]. (E)-7-(4-hydroxy-3-methoxyphenyl)-1-phenylhept-4-en-3-one 84, extracted from Alpinia officinarium, showed the lowest binding energies and the highest ligand efficiency values in closed (−47 kJ/mol, 0.49) and open (−28 kJ/mol, 0.27) conformations, respectively. The 8-gingerol 85 (Figure 7) from ginger showed the best binding energies and ligand efficiency values in closed (−43 kJ/mol, 0.45) and open (−15 kJ/mol, 0.16) conformations, respectively, [124] (as depicted in Table 2).
Abodunrin et al. (2022) examined the therapeutic functions of the active chemicals found in ten common African medicinal herbs. Five active compounds, including curcumin 86 (Figure 8), kolaviron 87, bisdemethoxycurcumin 88, 6-gingerol 89, and artemisinin 26, were chosen and docked against the main protease. The results of the pharmacokinetic prediction showed that none of these five active substances exhibited any affinity for cytotoxicity, hepatotoxicity, mutagenicity, or carcinogenicity, maintaining their excellent ADMET profile [94]. A total of 29 compounds were isolated from the medicinal plant Passiflora, and all showed comparable binding affinity with various amino acid residues of the main protease. Compounds, such as luteolin 90, lucenin 91, olealonic acid 92, isoorientin 93, isochaphoside 94, saponarin 95, and schaftoside 96, were bound above −8.0 kcal/mol binding energy (Figure 8). In addition to ADMET, Lipinski, Veber, and Ghose criteria were used to investigate the drug-like characteristics of Passiflora chemical compounds [95]. The molecular docking results showed that Compound 97 (CID 11170714) (C31H30Br6N4O11), which belongs to the family Aplysinidae, showed a good docking score and better binding interactions. MD simulation (RMSD and RMSF) studies have also been performed, which reflect a stable binding contact between the ligand and the target enzyme, Mpro [149]. Later on, 1018 natural brown algal compounds were taken from the COVID-19 major protease-screening database MarinLit, which is devoted to marine natural materials. The interactions between the top seven chemicals (7,2″-bieckol 98, 7-hydroxyeckol-hepta-acetate 99, 5-hydroxy-cystofurano-quinol 100, sargaquinoic acid 101, triacetoxy-18-hydroxy-2,7-dolabelladiene 102, fallahydroquinone 103, and methoxybifurcarenone 104 (Figure 8)) and the active site of the Mpro were investigated in more detail. Compound 98 displayed the lowest binding energy (high binding affinity) among all the compounds under investigation [96] (Table 2).
Ibrahim et al. (2021) screened 360 metabolites (cembranoid diterpenes) from the genus Sarcophyton (soft coral) against the Mpro of SARS-CoV-2. Almost 59 compounds showed better docking scores compared to the docking score of the reference (darunavir = −8.2 kcal/mol). Sarelengan B 105 possessed the highest docking score, followed by bislatumlide A 106 (Figure 8). Further, from the MD simulation studies and molecular mechanics with generalized Born and surface area salvation (MM/GBSA) binding energy calculation results, it was found that Compound 105 showed favorable binding affinity with ∆G binding of <−44.0 kcal/mol against the target protein. Drug-likeness studies also demonstrated convenient physicochemical properties for Compound 106 (Figure 8) [97]. The molecular interactions of bioactive metabolites from the oils of Eucalyptus and Corymbia against SARS-CoV-2 were assessed. The compounds that showed excellent ADMET profiles and properties of drug-likeness included citronellol, α-terpineol, o-cymene, d-limonene, eucalyptol, α-pinene, and 3-carene. In addition to this, the primary component in the leaves of Eucalyptus globulus essential oil, eucalyptol 107 (Figure 9), had good bioavailability and blood–brain barrier (BBB) penetration and was an inhibitor of CYP 2C9 and CYP 3A4, as well as being noncarcinogenic in rat and mouse studies. Eucalyptol 107 showed less binding energy than α-pinene 108, α-terpineol 109, 3-carene 110, d-limonene 111, o-cymene 112, and citronellol 113 (Figure 9). Based upon these results, these bioactive substances may work as potential inhibitors of virus replication and transcription by inhibiting the Mpro [98]. Compounds from 10 different species of medicinal plants were isolated and investigated for their therapeutic potential against the SARS-CoV-2 main proteases. The results of biological activity, pharmacological behavior, and binding affinities showed high absorption and bioavailability for harsingar, aloe vera, and giloy plants. Other active SARS-CoV-2 protease inhibitors included turmeric, neem, and ginger. All these plants showed more inhibition potential compared to chloroquine and hydroxychloroquine. The docking results demonstrated the highest inhibition potential for the extracts of harsingar and aloe vera, namely, nictoflorin 114 and aloenin 115 (Figure 9) [99]. Vijayakumar et al. (2022) investigated the therapeutic potential of phytochemicals present in a medicinal herb named Andrographis panniculata (A. panniculata). Five diterpenoid molecules, andrographolide, neoandrographolide, 14-deoxyandrographolide, 14-deoxy-11, 12-didehydroandrographolide, and andrograpanin, extracted from A. panniculata were screened. The results from pharmacokinetics and molecular dynamic simulation showed that all the selected diterpenoids possessed a significant inhibitory potential against the Mpro of SARS-CoV-2. Among these bioactive compounds, andrographolide 116 showed a high potential for Mpro inhibition [100]. (Binding affinities, binding free energies, and amino acid residue interactions are depicted in Table 2).
Flavonoids are naturally occurring phytochemical compounds that possess various biological applications, including antiviral activities. Bora et al. (2021) assessed the inhibitory potential of four naturally occurring flavonoids (quercetin, luteolin, galangin 117, and naringenin) against the SARS-CoV-2 main protease. All four compounds—quercetin, luteolin, galangin, and naringenin—exhibited satisfactory docking scores. Among these four, galangin 117 displayed the highest number of interactions with a large number of amino acids. Furthermore, the pharmacokinetic results revealed that galangin could be more potent against SARS-CoV-2’s Mpro [101]. Moreover, the flavonoids, including amentoflavone 42, scutellarin 118, morusin 119, apigenin 120, wogonin 121, kaempferol 23, fisetin 36, kuwanon C 31, and morin 122, showed strong binding interactions with the Mpro at several amino acid residues. In addition to these, prenylated flavonoids, flavanones, bioflavonoids, flavones, and flavan-3-ols demonstrated good interactions and binding affinities with the COVID-19 main protease [102]. Among the 2360 natural compounds, 12 compounds from different natural sources (microbes, fungi, and plants) showed better docking results below −12 kcal/mol. Kazinol T 123 (discovered from Broussonetiakazinoki) exhibited the highest score, followed by butyrolactone 1,3 sulphate 124 (Figure 9) (extracted from Aspergillus terreus). ADMET analysis predicted no considerable toxicity for the active lead compounds [103] (Table 2).
Baildya et al. (2021) predicted the inhibitory potential of neem extracts containing 19 natural compounds on the PLpro of SARS-CoV-2. All the extracted compounds showed satisfactory inhibitory potential against the target enzyme. Among them, desacetyl gedunin 125, which was extracted from neem seed, exhibited the highest binding affinity towards PLpro. However, the ADMET analysis and pharmacokinetic studies confirmed high blood–brain barrier permeability, bioavailability, and low toxicity of selected compounds compared to the standard drugs [104]. The binding affinities of palmatine 126, sauchinone 127, and tabersonine 128 were recorded from Autodock Vina. Protein–ligand interaction results have shown that palmatine 126, sauchinone 127, and tabersonine 128 (Figure 9) were bound to the active site of the Mpro. Further, MD simulation studies were performed, and the MM/PBSA results reflected that two compounds, palmatine 126 and sauchinone 127, formed very stable complexes with the Mpro and showed free energy values of −71.47 kJ/mol and −71.68 kJ/mol, respectively, compared to the reference (−69.58 kJ/mol) [105]. Bharadwaj et al. (2021) extracted some natural compounds from Echinacea angustifolia and docked them against the main protease of a novel coronavirus. Almost 50 natural compounds showed binding affinities ranging from −12.93 to 0.0897 kcal/mol, and top five compounds (echinacoside 129, quercetagetin 7-glucoside 130, levan N 131, inulin from chicory 132 (Figure 9), and 1,3-dicaffeoylquinic acid 133) were selected for further analysis. Redocking, intermolecular interaction, and MD simulation studies showed that all five selected compounds exhibited a binding affinity of >10 kcal/mol [106]. Papia Chowdhury in 2021 screened the chemical constituents of Tinospora cordifolia (an Indian medicinal plant) and showed that berberine 134 (Figure 10), choline 75, and tetrahydropalmatine satisfied all the required screening attributes. Molecular docking and MD simulation studies showed that among all the tested compounds, berberine 134 exhibited strong binding affinity and better inhibition towards the 3CLpro. All the inhibitors possessed drug-like properties and better pharmacokinetics [107] (Table 2).
Shree et al. (2022) investigated the potency of various natural compounds from three different plant species, including Withania somnifera, Tinospora cordifolia, and Ocimum sanctum, against the Mpro of SARS-CoV-2. Six phytochemicals showed better binding results and stable complexes with the target, i.e., withanoside V 43: somniferine 135, tinocordiside 136, vicenin 137, 4′-O-glucoside 2″-O-p-hydroxybenzoate 138, and ursolic acid 139 (Figure 10). Furthermore, ADMET and drug-likeness analysis showed decent results [108]. Four active compounds that included three flavonoids (podocarpus flavone A, methoxyquercitrin, and proanthocyanidin) and a lignoid, chimarrhinin 140, displayed the best inhibitory potential against the main protease. Among these four, the lignoid (140) showed the highest binding affinity (−9.0 kcal/mol) compared to the reference (−8.9 kcal/mol) [109]. The binding capacity of five naturally occurring alkaloids—berberine 134, lycorine 141, hemanthamine 142, aloperin, and dendrobine—to SARS-CoV-2’s main protease was monitored. The molecular docking results revealed that Compounds 141 and 142 showed the best docking scores. In addition, pharmacokinetics and ADMET analysis confirmed the biocompatibility and drug-like properties of the tested compounds, especially 141 and 142 [110]. Laxman Durgam and Lalitha Guruprasad recently performed a virtual screening of natural compounds from the NCI database. Using AutoDock Vina and CDOCKER, eight active compounds that showed the best docking scores in the range of –7.3 kcal/mol to –8.1 kcal/mol were selected, and after assessing their drug-likeness properties, molecular dynamic stimulations were performed. The amino acid residues that highly contributed to the binding free energies of all the compounds were Cys145, Met165, and Glu166. The binding affinities of Compounds 143, 144, 145, and 146 (Figure 10) were recorded. The complex of Compound 143 with the targeted enzyme was found to possess high conformational changes [111]. (The docking scores and amino acid residue interactions are depicted in Table 2).
Hawary et al. (2022) identified the active metabolites of Citrus nobilis L. and Citrus deliciosa Tenora. It was found that out of 21 compounds that belong to coumarins, phenolic acids, and flavonoids, quercetin-7-O-glucoside-3-O-rutinoside 147 showed the best docking score towards the active site of the main protease, followed by luteoline-7-rutinoside 148, quercetin-3-O-rutinoside 149, and apigenin-8-C-glucoside 150 (Figure 10). MD simulation studies revealed that Compounds 148 and 149 showed better binding with the active site of the Mpro [112]. Out of these 100 screened metabolites, pyranonigrin A 151 exhibited similar behavior as the reference compound and showed the best docking score of −7.3 kcal/mol. The MD simulation studies showed similar behaviors of both ligands (N3 and pyranonigrin A 151) towards binding with the Mpro, and the ADMET analysis results were found to be acceptable [113]. The phytochemicals of Indian medicinal plants showed 26 active compounds against SARS-CoV-2. Among these, the compounds 4,8-dihydroxysesamin 152 and arboreal 153 showed the best activities against PLpro and 3CLpro, respectively. Molecular docking studies revealed that 4,8-dihydroxysesamin 152 showed binding affinity with PLpro, and arboreal 153 (Figure 10) showed binding affinity with 3CLpro [114] (Table 2).
Moharana et al. (2022) performed an in silico analysis on 12 biologically active compounds out of 424 that were isolated from the extract of medicinal plants. It was shown that acacetin 154 (Figure 10) showed the best docking results and binding interactions with various amino acid residues of the SARS-CoV-2 main protease. MD simulation studies and free energy calculation analyses confirmed the flexibility and stability of the ligand–receptor protein complex [115]. Epoxy-linalool oxide 155 that was found in Cymbopogon citrates oil showed good binding interactions with the main protease. MD stimulation and pharmacokinetic studies revealed that epoxy-linalool oxide 155 possessed a more stable complex with the target protein and had more drug-likeness properties [116]. The potential of glycyrrhizin 156 towards Mpro and PLpro inhibition showed good docking scores [117]. An active compound, rutin 29, was isolated from a library of phytochemicals from Peruvian plants. MM/GBSA analysis of Compound 29 showed favorable interactions (−40.293 kcal/mol and 21.713 kcal/mol) with the Mpro and PLpro, respectively [118]. Africa et al. (2022) found some potent antitubercular phytochemicals and analyzed their potential for novel coronaviruses. Vobtusine lactone 157 showed greater binding affinity with 3CLpro, and deoxyvobtusine lactone 158 showed higher binding affinity with PLpro. ADMET analysis confirmed the drug-likeness attributes of these active compounds [119]. Some natural compounds with antiviral activities were selected and docked against the main protease. Among the tested compounds, seven compounds were found to be more potent based on their high binding affinities with target proteins. The highest binding affinity was reported for sotetsuflavone 159 (Figure 10) with hydrogen bonds and other alkyl interactions with various residues. ADMET analysis and pharmacokinetic studies showed good results for this compound, confirming its ability to act as an inhibitor of SARS-CoV-2 [120]. (Binding affinities and amino acid residue interactions are mentioned in Table 2).
Shaldam et al. (2021) screened many phenolic and terpene compounds from honeybee products against SARS-CoV-2 protein targets using in silico techniques. Through molecular docking analysis, it was found that Compounds 160, 23, 161, 162, and 163 (Figure 10) were more potent against the main protease of virus [121]. Maackin A 164 [91], anthracene dione 165 [92], and fortunellin 166 [93] were investigated as the Mpro inhibitors, and jezonofol 167 was identified as a PLpro inhibitor.
The antiviral potential of the secondary metabolites of Streptomyces sp. GMR22 was evaluated by Melinda et al. (2021). Two active compounds, echoside A 168 and echoside B 169, displayed a higher docking score than remdesivir towards 3CLpro [122]. Suleimen et al. (2022) reported the anti-SARS-CoV-2 PLpro inhibition potential of the two active compounds 6-demethoxy-4′-O-capillarsine 170 and tenuflorin C 171 from Artemisia commutata and A. glauca, respectively. The results from molecular docking analysis demonstrated that Compounds 170 and 171 (Figure 10) showed a good binding score and had interactions with various residues [125]. Similarly, Dutta et al. (2021) analyzed the antiviral potential of a medicinal plant, Calotropis gigantean, towards SARS-CoV-2 Mpro inhibition via in silico approaches. Among various bioactive compounds, juniper camphor 172 (Figure 10) displayed the best docking score [123] (Table 2).
The active site of an enzyme further comprised two sites: the binding site and the catalytic site [150]. Catalytic site residues reported in the literature are for the Mpro, HIS41 and CYS145 [75]; for PLpro, HIS272, ASP28 and TRP106 [89]; for RdRp, SER759, ASP760, and ASP761 [77,89]; and for endoribonuclease, HIS235, HIS250, LYS290 [145], and THR341 [85]. The catalytic residues are shown in blue, and the binding site residues are shown in black in the following table.

5.2. Natural Compounds as SARS-CoV-2 Spike Protein Inhibitors

Glycosylated spike proteins that are present on the outer surface of the viral membrane are responsible for the attachment and entry of viruses into the host cell. The host angiotensin-converting enzyme 2 (ACE2) receptor and GRP78 binding domain are responsible for viral attachment and entry. Inhibition of spike protein binding is an alternate route to achieve viral inhibition. Different researchers have performed in silico studies on natural compounds and the spike protein of SARS-CoV-2.
Muhseen et al. (2020) demonstrated terpenes are SARS-CoV-2 spike receptor binding domain attachment inhibitors to the human angiotensin-converting enzyme 2 (hACE2) receptor via molecular docking, ADMET screening, and MD simulations. A terpene, NPACT01552 173 (Figure 11), was discovered to be more potent than the others [126]. Potential phyto compounds of Brassica oleracea and naturally occurring biflavones were identified as the targets of the spike protein of SARS-CoV-2 through molecular dynamics, classical molecular dynamics simulations, and ADMET analysis. 3-p-coumaroylquinic acid 174 had a high affinity for the S2 domain of SARS-CoV-2 spike proteins [138]. Hinokiflavone 175 and robustaflavone 176 were found to be more active than nefamostat, which had a binding energy value of −8.40 kcal/mol [127]. Geraniin 177 was found to be an effective blocker of the interaction between the spike protein receptor binding domain and the human ACE2 receptor. The simulation demonstrated that geraniin 177 might bind more steadily to the spike protein than to the hACE2 receptor [132]. Abietatriene 178 (Figure 11) was demonstrated as the best active inhibitor of spike proteins [128]. Quercetin 40 was found to be potent enough to block interaction sites on spike proteins [77]. Hesperidin 51 and nabiximols 179 (Figure 11) performed inhibitory activity against spike proteins with binding free energy values compared to the standard drugs nelfinavir, hydroxychloroquine sulphate, and chloroquine [86]. Nine molecules from Djiboutian medicinal plants were studied, out of which rutin 29, catechin 41, and kaempferol 23 exhibited lower binding energy values than the reference drug, hydroxychloroquine (−4.828 kcal/mol) against the SARS-CoV-2 receptor binding domain [69]. Amentoflavone 42 was found to be the best active compound against the spike proteins compared to the reference drug hydroxychloroquine, with a binding free energy of −6.4 kcal/mol. Lipinski’s rule of five also showed its drug-likeness [78]. Crocin (63) was shown to be an active carotenoid against the spike protein of SARS-CoV-2 [73].
Mehmood, A. et al. (2021) examined Quranic and prophetic medicinal plants as inhibitors of the SARS-CoV-2 essential enzyme spike protein. These inhibitors showed a strong binding affinity for pelargonidin-3-galactoside 180 (Figure 11) [90]. Bhowmik et al. (2020) identified various phytochemicals that showed binding with the SARS-CoV-2 spike protein and with the GRP78 binding domain. The molecular docking results of the tested phytochemicals showed that orientin 181 (a flavonoid) showed the best binding affinity with the SARS-CoV-2 spike protein and with the GRP78 binding domain. It was also found that orientin 181 showed binding interactions with spike proteins in the same region where GRP78 interactions occurred. Further, MD simulation studies confirmed that orientin formed stable complexes with the GRP78 binding domain and inhibited the attachment of the SARS-CoV-2 spike protein to this receptor [129]. A total of 37 compounds of Kabasura Kudineer, Official Siddha Formulation, and JACOM were screened. It was shown that chrysoeriol 182, luteolin 183, quercetin 40, and scutellarein 118, possessed high binding affinities by forming hydrogen bonds with four amino acid residues of the target protein. Pharmacokinetic analysis showed that all the selected compounds possessed good bioavailability and no toxicity [130]. Acacetin 154 showed the best docking result with a binding energy of −7.75 kcal/mol by forming hydrogen bonds and hydrophobic interactions with various residues. MD simulation studies and free energy calculation analyses confirmed the flexibility and stability of the ligand–receptor protein complex [115]. Mhatre et al. (2021) investigated the therapeutic potential of different active catechins against SARS-CoV-2 spike proteins using computation techniques. Molecular docking studies revealed that all the tested catechins showed better docking scores in the range of −6.3 kcal/mol to −5.7 kcal/mol than the reference NGA (−5.0 kcal/mol). Among all the screened compounds, epigallocatechin gallate 21 exhibited the highest docking score, and MD simulation studies confirmed the stability of the protein–ligand complex. The pharmacokinetic analysis results revealed that the drug-likeness of these compounds needs to be improved [131]. The neem plant extract was evaluated for its inhibition of SARS-CoV-2 entry into the host cell. Out of the total 19 compounds that were extracted and screened, 3 compounds demonstrated the best binding scores towards the spike receptor binding domain–angiotensin-converting enzyme 2 complex (RBD–ACE2) compared to the reference, hesperidin 51 (−7 kcal/mol). Azadirachtin H 184 showed a higher binding affinity than quercetin 40 and margocin 185. Compound 184 also showed better binding interactions (−8 kcal/mol) with spike proteins and the ACE2 receptor. The MM/PBSA binding free energy calculations and pharmacokinetic studies confirmed the potential of the studied compounds as drug candidates towards spike SARS-CoV-2 inhibition [139]. (The binding energies and amino acid residue interactions are described in Table 2).
Cheke et al. (2021) analyzed the potential of various medicinally active phytochemicals towards spike protein–ACE2 inhibition. Molecular docking studies revealed that among various screened compounds, the higher binding affinity was shown by indigo blue 186, followed by glycyrrhizin 156, β-sitosterol 187, indirubin 188, bicyclogermacrene 189, curcumin 86, hesperetin 190, rhein 191, and berberine 134 [140]. Acetogenins isolated from A. muricate showed good docking scores with the target protein in the range of −5.3 to −7.7 kcal/mol. The highest binding affinity was reported for cis-annonacin 192 compared to the reference (−7.5 kcal/mol) [133]. Tuftsin 193, a naturally occurring peptide, possessed binding interactions with both ACE2 and neuropilin-1 (NRP1), with binding affinities of −6.9 kcal/mol and −8.1 kcal/mol, respectively. The surface plasmon resonance (SPR) analysis further proved the potency of tuftsin towards the inhibition of SARS-CoV-2 spread [141]. Thuy et al. (2021) evaluated 34 compounds that were found in Cymbopogon citratus oil against ACE2 receptor proteins using docking and MD simulation studies. It was found that five compounds showed the best binding affinities, of which the highest affinity was shown by epoxy-linalool oxide 155 through hydrogen bonding and van der Waals interactions with different residues. MD stimulation and pharmacokinetic studies revealed that Compound 155 possessed a more stable complex with the target protein and had a more drug-like attitude [116]. Bromelain 194 (Figure 11) (an enzyme, present in fruits) bound more effectively in the region of RBD-ACE2 binding sites. The molecular docking results of bromelain with the receptor binding domain (RBD) variants, WT, the United Kingdom, BR, SA, and the United States, possessed binding affinities of −14.9 kcal/mol, −15.0 kcal/mol, −15.6 kcal/mol, −15.4 kcal/mol, and −15.0 kcal/mol, respectively. MD simulation studies were also performed to confirm the stability of these complexes [151]. Dharmashekara et al. (2021) conducted in silico investigations on various phytochemicals to evaluate their potential towards the spike proteins of SARS-CoV-2. Among the tested compounds, trigoneoside IB 195 showed the highest binding affinity with target spike glycoproteins, with interactions with various amino acid residues [134]. Two active compounds, 196 and 197, extracted from Jinhua Qinggan granules (JQGs), were analyzed against ACE2. Arctiin 196 and linarin 197 (Figure 11) showed the best docking scores, having strong interactions with METB1007 and ALA1000 amino acid residues, respectively [135] (Table 2).
Kashyap et al. (2021) investigated the anti-COVID-19 activity of phytochemicals from some medicinal plants. The isolated compounds were docked against spike proteins of the virus and two host proteins (ACE2 and TMPRSS2). Out of 12 screened compounds that showed the best binding affinities with target proteins, the highest docking score was shown by withanolide D 198 (Figure 11) with the spike proteins, ACE2, and TMPRSS2, respectively. ADMET analysis confirmed the drug-likeness of these compounds [136]. Some natural compounds with antiviral activities were tested against spike glycoproteins and ACE2. Among the tested compounds, seven compounds were found to be more potent based on their high binding affinities with the target proteins. The highest binding affinities were reported by morellic acid 199 against spike proteins and human ACE2, respectively [120]. Roshni et al. (2022) investigated the phytochemicals of Indian medicinal plants and reported 26 active compounds against SARS-CoV-2. Among these, scutellarein 118 showed the best activity against the spike glycoproteins of SARS-CoV-2 [114]. Melinda et al. (2021) evaluated the antiviral activity of metabolites of Streptomyces sp. GMR22 against spike proteins and ACE2 receptors. Echoside A 168 and echoside B 169 displayed higher docking scores than remdesivir against the target proteins [122]. Siddiqui et al. (2022) analyzed the therapeutic potential of an ethanolic extract of Moringa oleifera fruits against SARS-CoV-2. Among all the extracted phytochemicals, 2-pyrrolidinone 200 (Figure 11) displayed good binding interactions with both the spike protein and the ACE2 receptor [137].

5.3. Natural Compounds as SARS-CoV-2 RdRp Inhibitors

RNA-dependent RNA polymerase (RdRp) is involved in viral RNA replication and transcription. Various researchers have found some active natural phytochemicals against this enzyme.
Mir, S.A. et al. (2021) examined SARS-CoV-2 RdRp potential inhibitors extracted from Nigella sativa through an in silico approach. When compared to the standard inhibitor remdesivir, α-hederin 201 (Figure 12) displayed the highest binding affinity to RdRp (PDB ID: 6M71) [142]. Some Quranic and prophetic medicinal plants acted as potential inhibitors of the SARS-CoV-2 essential enzymatic functions of RdRp; kaempferol 23 inhibited the viral transcription machinery in the best way [90]. Polyphenolic anti-HIV reverse transcriptase natural compounds were reported as potential inhibitors of NSP12 (RdRp) of the SARS-CoV-2. RNA-dependent RNA polymerase (RdRp) polymerizes the nucleotide chain of the daughter strand, which helps in the inhibition of virus proteins. NSP12 (RdRp) demonstrated a good binding score with ellagitannin punicalin 202 [79]. Cyanidin 203 was found to be active with a −7.7 kcal/mol binding affinity value against RdRp [77]. Amphimedoside C 78 was a potent inhibitor of SARS-CoV-2 RdRp [87]. 14-debromoaraplysilin I 204 [88], glaucogenin D 81, glaucogenin C 205 (Figure 12) [89], crocin 63, a food-derived carotenoid [73], and jezonofol 167 [91] were demonstrated to be the best active compounds against the replication enzyme RdRp (as depicted in Table 2).
Shaldam et al. (2021) screened many phenolic and terpene compounds from honeybee products. Through molecular docking analysis, it was found that ellagic acid 206, kaempferol 23, quercetin 40, p-Coumaric acid 163, and naringenin 207 were more potent against the RdRp enzyme of SARS-CoV-2. Compound 206 showed hydrogen bond interactions with RdRp active site amino acid residues [121]. Some potent antitubercular phytochemicals were analyzed against the novel coronavirus. It was demonstrated that 10 compounds showed active interactions with various protein targets of the virus. Among these, vobtusine lactone 157, deoxyvobtusine lactone 158, deoxyvobtusine 208, and globospiramine 209 showed greater binding affinities with the RdRp of SARS-CoV-2. The most potent compound, 157, showed many interactions with active site residues, including hydrogen bond interactions with SER682 and ALA688 residues. ADMET analysis confirmed the drug-likeness attributes of these active compounds [119]. Roshni et al. (2022) investigated the phytochemicals of Indian medicinal plants and found 26 active compounds against SARS-CoV-2. Among these, the compound arboreol 153 showed the best activity against the RdRp proteins of the virus [114]. Melinda et al. (2021) displayed the antiviral potential of two active compounds named echoside A 168 and echoside B 169. Both displayed higher docking scores than remdesivir against the target proteins of SARS-CoV-2 [122] (Table 2).

5.4. Natural Compounds as SARS-CoV-2 Nucleocapsid Inhibitors

Nucleocapsids or N-proteins are important structural and functional proteins that have been found to play an important role in viral replication and translational properties. They have two domains, i.e., NTD (N-terminal domain) and CTD (C-terminal domain), both of which function to bind with viral RNA and translate it. So, the inhibition of N-proteins is considered an important target to control this viral disease. Roshni et al. (2022) investigated the phytochemicals of Indian medicinal plants and found 26 active compounds against SARS-CoV-2. Among these, 4,8-dihydroxysesamin 152 and arboreol 153 showed the best activities against the nucleocapsid proteins of SARS-CoV-2. Molecular docking studies revealed that Compounds 152 and 153 showed the highest binding affinities. Both of these compounds were found to be efficient compared to the reference drugs [114]. It was found that glycyrrhizin 156 showed the highest binding affinity with nucleocapsid proteins. The MM/GBSA and MM/PBSA estimations showed binding free energies of −30.05 kcal/mol and −25.95 kcal/mol, respectively, for the ligand–receptor complex [117]. Mamani et al. (2021) found active rutin 29 from a library of phytochemicals from Peruvian plants. MM/GBSA analysis of the compound showed favorable interactions (−34.342 kcal/mol) with the N-domain of the nucleocapsid, and further MD simulation studies confirmed the stability of the ligand–receptor complex [118]. It was shown that trigoneoside IB 195 exhibited the highest binding affinity, with target proteins having interactions with different amino acid residues. Other than this, two more compounds showed docking scores of −4.8 kcal/mol and −4.6 kcal/mol, respectively [134]. Husain et al. (2022) analyzed the antiviral effect of some bioactive compounds against the NTD and CTD of the nucleocapsid proteins of SARS-CoV-2. The compounds were isolated using HPLC, and then docking studies were performed, followed by MD simulation and pharmacokinetic analysis. From the molecular docking results, it was found that with the NTD, apigenin 120 showed the highest docking score, followed by catechin 41 and apiin 53. With the CTD of the nucleocapsid, apigenin 120 showed the highest score, followed by cinnamic acid 210 (Figure 13) and apiin 53. MD simulation and pharmacokinetics confirmed the therapeutic potential of these four compounds towards the inhibition of this viral disease [143]. The inhibitory potential of curcumin 86 towards SARS-CoV-2 nucleocapsid proteins was evaluated. Molecular docking and MD simulation analyses were carried out to find the binding affinity of the tested ligand with the target proteins. The study concluded that curcumin formed a stable complex with the target protein and would prove the best candidate for SARS-CoV-2 drug development [144]. (The docking scores and amino acid residue interactions are listed in Table 2).

5.5. Natural Compounds as SARS-CoV-2 Endoribonuclease Inhibitors

Endoribonucleases prevent host dsRNA sensors from recognizing dsRNA intermediates. The biflavonoid hinokiflavone 175 [79] and rutin 29 [85] exhibited a good binding score against NSP15.The diosmetin glucopyranoside derivative NPC198199 211 [145] and the NPASS compound with the ID NPC10737 212 (Figure 14) [146] both demonstrated the greatest anti-NSP15 potency.

5.6. Natural Compounds as SARS-CoV-2 Helicase Inhibitors

Crocin 63 [73], the natural product biflavonoid rhusflavanone 213, morelloflavone 214 [79], and scirpusin A 217 (Figure 14) [91] have the highest binding affinities towards helicase (essential for replication). NPASS compounds with the IDs NPC270578 215 and NPC52382 216 (Figure 14) both displayed a higher MMGBSA score [146].

5.7. Natural Compounds as SARS-CoV-2 Methyltransferase Inhibitors

Polyphenolic compounds, which were well known as anti-HIV reverse transcriptase inhibitors, also played a role as potential inhibitors of the nonstructural protein NSP16–NSP10 complex (S-adenosylmethionine complex) of SARS-CoV-2. In the case of SAM-dependent 2′-O-methyltransferase complex enzymes (NSP16–NSP10 complex), the biflavonoid robustaflavone 176 and the alkaloid michellamine B 218 (Figure 15) both demonstrated active binding affinities [79]. The NPASS compound with the ID NPC226294 219 (Figure 15) exhibited good binding free energies towards methyltransferase with an MM/GBSA score of −75.24 compared to the control, i.e., sinefungin, with an MM/GBSA score of −60.32 [146].

5.8. Natural Compounds as SARS-CoV-2 ADP-Ribose Phosphatase Inhibitors

The NSP–complex formation was potentially blocked by ADRP inhibitors. In this context, Mujwar, S. et al. (2022) investigated food-derived carotenoids against the ADP-ribose phosphatase (ADPRP) (PDB ID: 6W02) of SARS-CoV-2 through an in silico approach. Crocin 63 was also reported as a potential inhibitor of ADPRP [73].

5.9. Natural Compounds as SARS-CoV-2 Exoribonuclease Inhibitors

An exoribonuclease inhibitor was identified by Naik et al. (2020). Naik and co-workers tackled the bioactive compounds from the Natural Product Activity and Species Source database that might impede the activity of the essential enzyme of SARS-CoV-2, i.e., exoribonuclease, through a molecular docking study. It was observed that the NPASS compound with the ID NPC137813 220 (Figure 16) displayed the highest binding capability to exoribonuclease compared to the control MES, with a −32.10 binding score [146]. (The binding affinities and amino acid residue interactions are represented in Table 2).

5.10. Natural Compounds as Inhibitors of Other SARS-CoV-2 NSPs

Medicinal plant metabolites acted as potent inhibitors of the SARS-CoV-2 NSP9 RNA-binding protein. Those inhibitors were identified by Bandyopadhyay, S. et al. (2021) through a molecular dynamics evaluation. Hispaglabridin B 221, licoflavone B 222, and ochnaflavone 223 (Figure 17) were found to have the best binding with the SARS-CoV-2 NSP9 protein [147]. 2,3-dehydrosomnifericin 224 was also identified as an NSP3 inhibitor [148].

5.11. Natural Compounds as SARS-CoV-2 Envelope Protein Inhibitors

Coronavirus E proteins are incorporated into the virion lipidic envelope along with the spike protein (S) and the membrane protein (M). Roshni et al. (2022) investigated the phytochemicals of Indian medicinal plants and found 26 active compounds against the SARS-CoV-2 envelope protein. Among these, dicumarol 225 (Figure 18) showed the best activity against envelope protein. Molecular docking studies revealed that Compound 225 showed a binding affinity of −7.4 kcal/mol against the target protein [114]. Similarly, trigoneoside IB 195 showed the highest binding affinity against target E-proteins with interactions with various amino acid residues [134] (Table 2).
In this review, we have highlighted the potent anti-SARS-CoV-2 activities of natural compounds that constitute various drug classes, i.e., flavonoids, bioflavonoids, alkaloids, carotenoids, terpenes, steroids, quinones, polyphenols, and glycosides. Numerous docking simulations studies have recommended to use these compounds as COVID-19 therapy. These classes have shown their promising actions on multiple therapeutic targets of SARS-CoV-2. Among them, flavonoids and their subgroups have shown their potential in inhibiting the viral infection by targeting all the enzyme targets of SARS-CoV-2. For example, amentoflavone 42, one of the most abundant plant flavonoids, is proposed as a lead candidate with its ability to inhibit spike glycoproteins, viral proteases (Mpro and PLpro), and RdRp activities of the virus, as well as to inhibit the ACE2 activity of the host cell [61,78,79]. Rutin 29, which is a flavonoid glycoside, showed inhibitory effects on the main protease, PLpro, RBD-ACE2 complex, nucleocapsid, and endoribonuclease [68,69,85,118]. Apart from this, many other flavonoids (myricetin 33 [54,55], baicalein 16 [31,32], kaempferol 23 [54,57], quercetin 40 [54,60], and catechins 41 [54,60,69]) have been found to display encouraging in silico outcomes against the COVID-19 disease. It was also found that this class of compounds has shown interactions with the catalytic and binding site amino acid residues of targeted enzymes.
Moreover, carotenoids, such as crocin 63, showed potent anticoronavirus properties by inhibiting the coronavirus targets (the Mpro, spike glycoproteins, RdRp, helicases, and ADP-phosphates) [73]. The other drug classes, i.e., alkaloids (tubocuraine 74 [83] and palmatine 126 [105]), terpenes (cyanopyrin 49 [66], ferolide 68 [80], and abietatriene 178 [128]), steroids (withanolide R 50 [67], withacoagulin H 60 [72], withacoagulin 62 [72], ajugin E 61 [72], sterenin M 65 [75], and stigmasterol 77 [85]), quinones (pycnanthuquinone C 47 [65] and pycnanthuquinone B 48 [65]), polyphenols (geraniin 177 [132], ellagitannin punicalin 202 [79], and ellagic acid 206 [121]), and glycosides (cyanidin 203 [77], arctiin 196 [135], forsythiaside A 46 [64], and hesperidin 51 [86]) have also been found to show promising anti-SARS-CoV-2 activities. It is anticipated that the phytoconstituents discussed in this report will aid the development of an effective and safe anti-SARS-CoV-2 treatment option from naturally procured compounds.

6. Conclusions

The pandemic of the novel coronavirus disease has become challenging because of the lack of specific treatment and the continuous resistance caused by mutant strains of the virus. Chloroquine, hydroxychloroquine, ivermectin, azithromycin, remdesivir, lopinavir, ritonavir, favipiravir, galidesivir, dexamethasone, and ruxolitinib are considered the alternative treatments for this viral pandemic, but they are not as effective as one would hope. Natural products have been proven to be the best source of treatment for various human illnesses. In this regard, several plant-based remedies have been applied to alleviate COVID-19. This review article mainly focuses on all the natural compound-based treatments that have been suggested for SARS-CoV-2 through in vivo, in vitro, and in silico analyses. Many important phytochemicals, such as flavonoids, bioflavonoids, catechins, alkaloids, chalcones, terpenes, sterols, quinones, glycosides, and polyphenols, extracted from medicinal plants, algae, fungi, bacteria, and marine natural sources are suggested to be active ingredients to combat the coronavirus disease. Among them, flavonoids and their subgroups have shown their potential in inhibiting viral infection by targeting all the enzyme targets of the SARS-CoV-2. The main targets of these active natural compounds are the main proteases (Mpro/3Clpro), papain-like proteases (PLpro), viral spike glycoproteins, human receptor cells (ACE2, TMPRSS2, and NRP1), RBD-ACE2, RNA-dependent RNA polymerase (RdRp), nucleocapsid, endoribonuclease, helicase, methyltransferase (NSP16–NSP10 complex), exoribonuclease, nonstructural proteins, and envelope proteins.
Computational approaches are emerging techniques to analyze the potential of biologically active compounds against targeted diseases. These techniques help researchers find effective potential treatments against the novel SARS-CoV-2. In recent years, various in silico analyses, such as molecular docking, MD simulations, MM/GBSA, MM/PBSA, ADMET, and Lipinski’s rule of five have been carried out to check the potential of active phyto-constituents towards targeted enzymes of this viral infection. Many active natural materials with greater binding affinities towards targeted areas are highlighted in this review article. Moreover, it is suggested that the antiviral benefits of these natural compounds should be studied on an experimental level, which will benefit the researchers in designing new anti-SARS-CoV-2 drugs.

Author Contributions

Conceptualization, M.A., U.A.A., N.u.A.M., M.E.A.Z. and S.A.A.-H.; methodology, A.R., T.J. and S.A.; resources, M.A., U.A.A., N.u.A.M., S.A.A.-H. and M.E.A.Z.; writing—original draft preparation, A.R., T.J., S.A. and M.A.; writing—review and editing, N.u.A.M., S.A., M.A., S.A.A.-H. and M.E.A.Z.; supervision, M.A. and S.A.A.-H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Deanship of Scientific Research, Imam Mohammad Ibn Saud Islamic University (IMSIU), Saudi Arabia, Grant No. (21-13-18-067) and Government College University Faisalabad Grant No. (HEC, NRPU-3715).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are thankful to the Deanship of Scientific Research, Imam Mohammad Ibn Saud Islamic University, and Government College University Faisalabad for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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]
  2. Dinda, B.; Dinda, M.; Dinda, S.; Chakraborty, M. Some Natural Compounds and Their Analogues Having Potent Anti- SARS-CoV-2 and Anti-Proteases Activities as Lead Molecules in Drug Discovery for COVID-19. Eur. J. Med. Chem. Rep. 2022, 6, 100079. [Google Scholar] [CrossRef] [PubMed]
  3. Petersen, E.; Koopmans, M.; Go, U.; Hamer, D.H.; Petrosillo, N.; Castelli, F.; Storgaard, M.; Al Khalili, S.; Simonsen, L. Comparing SARS-CoV-2 with SARS-CoV and Influenza Pandemics. Lancet Infect. Dis. 2020, 20, e238–e244. [Google Scholar] [CrossRef] [PubMed]
  4. Gorbalenya, A.E.; Baker, S.C.; Baric, R.S.; de Groot, R.J.; Drosten, C.; Gulyaeva, A.A.; Haagmans, B.L.; Lauber, C.; Leontovich, A.M.; Neuman, B.W.; et al. The Species Severe Acute Respiratory Syndrome-Related Coronavirus: Classifying 2019-NCoV and Naming It SARS-CoV-2. Nat. Microbiol. 2020, 5, 536–544. [Google Scholar]
  5. Souid, I.; Korchef, A.; Souid, S. In Silico Evaluation of Vitis Amurensis Rupr. Polyphenol Compounds for Their Inhibition Potency Against COVID-19 Main Enzymes Mpro and RdRp. Saudi Pharm. J. 2022, 30, 570–584. [Google Scholar] [CrossRef]
  6. Tirado-Kulieva, V.A.; Hernández-Martínez, E.; Choque-Rivera, T.J. Phenolic Compounds Versus SARS-CoV-2: An Update on the Main Findings Against COVID-19. Heliyon 2022, 8, e10702. [Google Scholar] [CrossRef]
  7. Yusuf, A.P.; Zhang, J.; Li, J.; Muhammad, A.; Abubakar, M.B. Herbal Medications and Natural Products for Patients with COVID-19 and Diabetes Mellitus: Potentials and Challenges. Phytomedicine Plus 2022, 2, 100280. [Google Scholar] [CrossRef]
  8. Prasansuklab, A.; Theerasri, A.; Rangsinth, P.; Sillapachaiyaporn, C.; Chuchawankul, S.; Tencomnao, T. Anti-COVID-19 Drug Candidates: A review on Potential Biological Activities of Natural Products in the Management of New Coronavirus Infection. J. Tradit. Complement. Med. 2021, 11, 144–157. [Google Scholar] [CrossRef] [PubMed]
  9. Stobart, C.C.; Sexton, N.R.; Munjal, H.; Lu, X.; Molland, K.L.; Tomar, S.; Mesecar, A.D.; Denison, M.R. Chimeric Exchange of Coronavirus NSP5 proteases (3CLpro) Identifies Common and Divergent Regulatory Determinants of Protease Activity. J. Virol. 2013, 87, 12611–12618. [Google Scholar] [CrossRef] [Green Version]
  10. De Wit, E.; Van Doremalen, N.; Falzarano, D. SARS and MERS: Recent insights into emerging coronaviruses. Nat. Rev. Microbiol. 2016, 14, 523–534. [Google Scholar] [CrossRef]
  11. Trougakos, I.P.; Stamatelopoulos, K.; Terpos, E.; Tsitsilonis, O.E.; Aivalioti, E.; Paraskevis, D.; Kastritis, E.; Pavlakis, G.N.; Dimopoulos, M.A. Insights to SARS-CoV-2 Life Cycle, Pathophysiology, and Rationalized Treatments that Target COVID-19 Clinical Complications. J. Biomed. Sci. 2021, 28, 154. [Google Scholar] [CrossRef] [PubMed]
  12. 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 a-Ketoamide Inhibitors. Science 2020, 368, 409–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Oladele, J.O.; Ajayi, E.I.; Oyeleke, O.M.; Oladele, O.T.; Olowookere, B.D.; Adeniyi, B.M.; Oyewole, O.I.; Oladiji, A.T. A Systematic Review on COVID-19 Pandemic with Special Emphasis on Curative Potentials of Nigeria Based Medicinal Plants. Heliyon 2020, 6, e04897. [Google Scholar] [CrossRef] [PubMed]
  14. Chojnacka, K.; Witek-Krowiak, A.; Skrzypczak, D.; Mikula, K.; Młynarz, P. Phytochemicals Containing Biologically Active Polyphenols as an Effective Agent Against COVID-19-Inducing Coronavirus. J. Funct. Foods 2020, 73, 104146. [Google Scholar] [CrossRef]
  15. Le Maréchal, M.; Morand, P.; Epaulard, O.; Némoz, B. COVID-19 in Clinical Practice: A Narrative Synthesis. Med. Mal. Infect. 2020, 50, 639–647. [Google Scholar] [CrossRef]
  16. Tiwari, V.; Beer, J.C.; Sankaranarayanan, N.V.; Swanson-Mungerson, M.; Desai, U.R. Discovering Small-Molecule Therapeutics Against SARS-CoV-2. Drug Discov. Today 2020, 25, 1535–1544. [Google Scholar] [CrossRef]
  17. Russo, M.; Moccia, S.; Spagnuolo, C.; Tedesco, I.; Russo, G.L. Roles of Flavonoids Against Coronavirus Infection. Chem. Biol. Interact. 2020, 328, 109211. [Google Scholar] [CrossRef]
  18. Chen, Y.; Liu, Q.; Guo, D. Emerging Coronaviruses: Genome Structure, Replication, and Pathogenesis. J. Med. Virol. 2020, 92, 418–423. [Google Scholar] [CrossRef] [Green Version]
  19. Dos Santos, O.V.; Pinaffi Langley, A.C.d.C.; Mota de Lima, A.J.; Vale Moraes, V.S.; Dias Soares, S.; Teixeira-Costa, B.E. Nutraceutical Potential of Amazonian Oilseeds in Modulating the Immune System Against COVID-19—A Narrative Review. J. Funct. Foods 2022, 94, 105123. [Google Scholar] [CrossRef]
  20. Benke, C.; Autenrieth, L.K.; Asselmann, E.; Pané-farré, C.A. Lockdown, Quarantine Measures, and Social Distancing: Associations with Depression, Anxiety and Distress at the Beginning of the COVID-19 Pandemic among Adults from Germany. Psychiatry Res. 2020, 293, 113462. [Google Scholar] [CrossRef]
  21. Fuzimoto, A.D.; Isidoro, C. The Antiviral and Coronavirus-Host Protein Pathways Inhibiting Properties of Herbs and Natural Compounds—Additional Weapons in the Fight Against the COVID-19 Pandemic? J. Tradit. Complement. Med. 2020, 10, 405–419. [Google Scholar] [CrossRef]
  22. Córdoba-Tovar, L.; Barón, P.A.R.; Marrugo-Negrete, J.; Roa-Fuentes, L.L.; Jonathan, M.P. Scientific Production on Medicinal Plants and Their Efficacy Against COVID-19: A Review and Scientometric Analysis Based on VOS viewer. Acta Ecol. Sin. 2022; in press. [Google Scholar] [CrossRef]
  23. Cai, Q.; Yang, M.; Liu, D.; Chen, J.; Shu, D.; Xia, J.; Liao, X.; Gu, Y.; Cai, Q.; Yang, Y.; et al. Experimental Treatment with Favipiravir for COVID-19: An Open-Label Control Study. Engineering 2020, 6, 1192–1198. [Google Scholar] [CrossRef]
  24. Sharma, S.; Suvalka, C.; Bharti; Joshi, M.; Bahuguna, A. Potential Inhibitors and Plant Based Technology: An Alternative Approach to Treat Corona Virus. Bioorg. Chem. 2021, 117, 105460. [Google Scholar] [CrossRef]
  25. Ceramella, J.; Iacopetta, D.; Sinicropi, M.S.; Andreu, I.; Mariconda, A.; Saturnino, C.; Giuzio, F.; Longo, P.; Aquaro, S.; Catalano, A. Drugs for COVID-19: An Update. Molecules 2022, 27, 8562. [Google Scholar] [CrossRef]
  26. Kashte, S.; Gulbake, A.; El-Amin, S.F.; Gupta, A. COVID-19 Vaccines: Rapid Development, Implications, Challenges and Future Prospects. Hum. Cell 2021, 34, 711–733. [Google Scholar] [CrossRef]
  27. de Albuquerque Veloso Machado, M.; Roberts, B.; Wong, B.L.H.; van Kessel, R.; Mossialos, E. The relationship Between the COVID-19 Pandemic and Vaccine Hesitancy: A Scoping Review of Literature until August 2021. Front. Public Health 2021, 9, 1370. [Google Scholar] [CrossRef] [PubMed]
  28. Francis, A.I.; Ghany, S.; Gilkes, T.; Umakanthan, S. Review of COVID-19 Vaccine Subtypes, Efficacy and Geographical Distributions. Postgrad. Med. J. 2022, 98, 389–394. [Google Scholar] [CrossRef]
  29. Mistry, P.; Barmania, F.; Mellet, J.; Peta, K.; Strydom, A.; Viljoen, I.M.; James, W.; Gordon, S.; Pepper, M.S. SARS-CoV-2 Variants, Vaccines, and Host Immunity. Front. Immunol. 2022, 12, 5400. [Google Scholar] [CrossRef] [PubMed]
  30. Chhetri, B.K.; Tedbury, P.R.; Sweeney-jones, A.M.; Mani, L.; Soapi, K.; Manfredi, C.; Sorscher, E.; Sara, S.G.; Kubanek, J. Marine Natural Products as Leads Against SARS-CoV-2 Infection. J. Nat. Prod. 2022, 85, 657–665. [Google Scholar] [CrossRef] [PubMed]
  31. Su, H.; Yao, S.; Zhao, W.; Li, M.; Liu, J.; Shang, W.J.; Xie, H.; Ke, C.; Gao, M.; Yu, K.; et al. Discovery of Baicalin and Baicalein as Novel, Natural Product Inhibitors of SARS-CoV-2 3CL Protease In Vitro. BioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  32. Liu, H.; Ye, F.; Sun, Q.; Liang, H.; Li, C.; Li, S.; Lu, R.; Huang, B.; Tan, W.; Lai, L. Scutellaria Baicalensis Extract and Baicalein Inhibit Replication of SARS-CoV-2 and Its 3C-like Protease In Vitro. J. Enzym. Inhib. Med. Chem. 2021, 36, 497–503. [Google Scholar] [CrossRef] [PubMed]
  33. Zandi, K.; Musall, K.; Oo, A.; Cao, D.; Liang, B.; Hassandarvish, P.; Lan, S.; Slack, R.L.; Kirby, K.A.; Bassit, L.; et al. Baicalein and Baicalin Inhibit SARS-CoV-2 RNA-dependent-RNA polymerase. Microorganisms 2021, 9, 893. [Google Scholar] [CrossRef] [PubMed]
  34. He, C.L.; Huang, L.Y.; Wang, K.; Gu, C.J.; Hu, J.; Zhang, G.J.; Xu, W.; Xie, Y.H.; Tang, N.; Huang, A.L. Identification of bis-Benzylisoquinoline Alkaloids as SARS-CoV-2 Entry Inhibitors from a Library of Natural Products. Signal Transduct. Target. Ther. 2021, 6, 131. [Google Scholar] [CrossRef]
  35. Sa-Ngiamsuntorn, K.; Suksatu, A.; Pewkliang, Y.; Thongsri, P.; Kanjanasirirat, P.; Manopwisedjaroen, S.; Charoensutthivarakul, S.; Wongtrakoongate, P.; Pitiporn, S.; Chaopreecha, J.; et al. Anti-SARS-CoV-2 Activity of Andrographis Paniculata Extract and Its Major Component Andrographolide in Human Lung Epithelial Cells and Cytotoxicity Evaluation in Major Organ Cell Representatives. J. Nat. Prod. 2021, 84, 1261–1270. [Google Scholar] [CrossRef]
  36. Saadh, M.J. Sauchinone with Zinc Sulphate Significantly Inhibits the Activity of SARS-CoV-2 3CL-Protease. Pharmacologyonline 2021, 2, 242–248. [Google Scholar]
  37. Brown, A.N.; Strobel, G.; Hanrahan, K.C.; Sears, J. Antiviral Activity of the Propylamylatintm Formula Against the Novel Coronavirus SARS-CoV-2 In Vitro Using Direct Injection and Gas Assays in Virus Suspensions. Viruses 2021, 13, 415. [Google Scholar] [CrossRef]
  38. Liu, J.; Bodnar, B.H.; Meng, F.; Khan, A.I.; Wang, X.; Saribas, S.; Wang, T.; Lohani, S.C.; Wang, P.; Wei, Z.; et al. Epigallocatechin Gallate from Green Tea Effectively Blocks Infection of SARS-CoV-2 and New Variants by Inhibiting Spike Binding to ACE2 Receptor. Cell Biosci. 2021, 11, 168. [Google Scholar] [CrossRef]
  39. Ngwe Tun, M.M.; Luvai, E.; Nwe, K.M.; Toume, K.; Mizukami, S.; Hirayama, K.; Komatsu, K.; Morita, K. Anti-SARS-CoV-2 Activity of Various PET-Bottled Japanese Green Teas and Tea Compounds In Vitro. Arch. Virol. 2022, 167, 1547–1557. [Google Scholar] [CrossRef]
  40. Liu, Q.; Kwan, K.Y.; Cao, T.; Yan, B.; Ganesan, K.; Jia, L.; Zhang, F.; Lim, C.; Wu, Y.; Feng, Y.; et al. Broad-Spectrum Antiviral Activity of Spatholobus Suberectus Dunn Against SARS-CoV-2, SARS-CoV-1, H5N1, and Other Enveloped Viruses. Phyther. Res. 2022, 36, 3232–3247. [Google Scholar] [CrossRef]
  41. Eggers, M.; Jungke, P.; Wolkinger, V.; Bauer, R.; Kessler, U.; Frank, B. Antiviral Activity of Plant Juices and Green Tea against SARS-CoV-2 and Influenza Virus. Phyther. Res. 2022, 36, 2109–2115. [Google Scholar] [CrossRef] [PubMed]
  42. Baeshen, M.N.; Attar, R.; Bouback, T.A.; Albeshri, A.O.; Baeshen, N.N.; Karkashan, A.; Abbas, B.; Aljaddawi, A.A.; Almulaiky, Y.Q.; Mahmoud, S.H.; et al. Assaying for Antiviral Activity of the Folkloric Medicinal Desert Plant Rhazya Stricta on Coronavirus SARS-CoV-2. Biotechnol. Biotechnol. Equip. 2022, 36, 67–73. [Google Scholar] [CrossRef]
  43. Shin, J.A.; Oh, S.; Jeong, J.M. The Potential of BEN815 as an Anti-Inflammatory, Antiviral and Antioxidant Agent for the Treatment of COVID-19. Phytomedicine Plus 2021, 1, 100058. [Google Scholar] [CrossRef] [PubMed]
  44. Khan, A.; Heng, W.; Wang, Y.; Qiu, J.; Wei, X.; Peng, S.; Saleem, S.; Khan, M.; Ali, S.S.; Wei, D.Q. In Silico and In Vitro Evaluation of Kaempferol as a Potential Inhibitor of the SARS-CoV-2 Main Protease (3CLpro). Phyther. Res. 2021, 35, 2841–2845. [Google Scholar] [CrossRef]
  45. Alhadrami, H.A.; Sayed, A.M.; Hassan, H.M.; Youssif, K.A.; Gaber, Y.; Moatasim, Y.; Kutkat, O.; Mostafa, A.; Ali, M.A.; Rateb, M.E.; et al. Cnicin as an Anti-SARS-CoV-2: An Integrated in Silico and In Vitro Approach for the Rapid Identification of Potential COVID-19 Therapeutics. Antibiotics 2021, 10, 542. [Google Scholar] [CrossRef]
  46. Li, X.; Gao, J.; Li, M.; Cui, H.; Jiang, W.; Tu, Z.C.; Yuan, T. Aromatic Cadinane Sesquiterpenoids from the Fruiting Bodies of Phellinus Pini Block SARS-CoV-2 Spike-ACE2 Interaction. J. Nat. Prod. 2021, 84, 2385–2389. [Google Scholar] [CrossRef]
  47. Hijikata, A.; Shionyu-Mitsuyama, C.; Nakae, S.; Shionyu, M.; Ota, M.; Kanaya, S.; Hirokawa, T.; Nakajima, S.; Watashi, K.; Shirai, T. Evaluating Cepharanthine Analogues as Natural Drugs against SARS-CoV-2. FEBS Open Bio 2022, 12, 285–294. [Google Scholar] [CrossRef]
  48. Dogan, K.; Erol, E.; Didem Orhan, M.; Degirmenci, Z.; Kan, T.; Gungor, A.; Yasa, B.; Avsar, T.; Cetin, Y.; Durdagi, S.; et al. Instant Determination of the Artemisinin from Various Artemisia annua L. Extracts by LC-ESI-MS/MS and Their In-Silico Modelling and In Vitro Antiviral Activity Studies Against SARS-CoV-2. Phytochem. Anal. 2022, 33, 303–319. [Google Scholar] [CrossRef]
  49. Owis, A.I.; El-Hawary, M.S.; El Amir, D.; Refaat, H.; Alaaeldin, E.; Aly, O.M.; Elrehany, M.A.; Kamel, M.S. Flavonoids of Salvadora persica L. (Meswak) and Its Liposomal Formulation as a Potential Inhibitor of SARS-CoV-2. RSC Adv. 2021, 11, 13537–13544. [Google Scholar] [CrossRef] [PubMed]
  50. Morita, T.; Miyakawa, K.; Jeremiah, S.S.; Yamaoka, Y.; Sada, M.; Kuniyoshi, T.; Yang, J.; Kimura, H.; Ryo, A. All-Trans Retinoic Acid Exhibits Antiviral Effect Against SARS-CoV-2 by Inhibiting 3CLpro Activity. Viruses 2021, 13, 1669. [Google Scholar] [CrossRef] [PubMed]
  51. Alsfouk, A.A.; Taher, A.T.; Al-karmalawy, A.A. Pimenta dioica (L.) Merr. Bioactive Constituents Exert Anti-SARS-CoV-2 and Anti-Inflammatory Activities: Molecular Docking and Dynamics, In Vitro, and In Vivo Studies. Molecules 2021, 26, 5844. [Google Scholar]
  52. Hou, B.; Zhang, Y.; Liao, H.; Fu, L.; Li, D.; Zhao, X.; Qi, J.; Yang, W.; Xiao, G.; Yang, L.; et al. Target-Based Virtual Screening and LC/MS-Guided Isolation Procedure for Identifying Phloroglucinol-Terpenoid Inhibitors of SARS-CoV-2. J. Nat. Prod. 2022, 85, 327–336. [Google Scholar] [CrossRef]
  53. Kim, Y.S.; Kwon, E.B.; Kim, B.; Chung, H.S.; Choi, G.; Kim, Y.H.; Choi, J.G. Mulberry Component Kuwanon C Exerts Potent Therapeutic Efficacy In Vitro against COVID-19 by Blocking the SARS-CoV-2 Spike S1 RBD: ACE2 Receptor Interaction. Int. J. Mol. Sci. 2022, 23, 12516. [Google Scholar] [CrossRef]
  54. Corona, A.; Wycisk, K.; Talarico, C.; Manelfi, C.; Milia, J.; Cannalire, R.; Esposito, F.; Gribbon, P.; Zaliani, A.; Iaconis, D.; et al. Natural Compounds Inhibit SARS-CoV-2 Nsp13 Unwinding and ATPase Enzyme Activities. ACS Pharmacol. Transl. Sci. 2022, 5, 226–239. [Google Scholar] [CrossRef]
  55. Xiao, T.; Cui, M.; Zheng, C.; Wang, M.; Sun, R.; Gao, D.; Bao, J.; Ren, S.; Yang, B.; Lin, J.; et al. Myricetin Inhibits SARS-CoV-2 Viral Replication by Targeting Mpro and Ameliorates Pulmonary Inflammation. Front. Pharmacol. 2021, 12, 669642. [Google Scholar] [CrossRef]
  56. Nawrot-Hadzik, I.; Zmudzinski, M.; Matkowski, A.; Preissner, R.; Kęsik-Brodacka, M.; Hadzik, J.; Drag, M.; Abel, R. Reynoutria Rhizomes as a Natural Source of SARS-CoV-2 Mpro Inhibitors–Molecular Docking and In Vitro Study. Pharmaceuticals 2021, 14, 742. [Google Scholar] [CrossRef]
  57. Chaves, O.A.; Fintelman-rodrigues, N.; Wang, X.; Sacramento, C.Q.; Temerozo, J.R.; Ferreira, A.C.; Mattos, M.; Pereira-dutra, F.; Bozza, T.; Castro-faria-neto, H.C.; et al. Commercialy available flavonols are better SARS-CoV-2 inhibitors than isoflavone and flavones. Viruses 2022, 14, 1458. [Google Scholar] [CrossRef] [PubMed]
  58. Abd El Hafez, M.S.M.; AbdEl-Wahab, M.G.; Seadawy, M.G.; El-Hosseny, M.F.; Beskales, O.; Saber Ali Abdel-Hamid, A.; El Demellawy, M.A.; Ghareeb, D.A. Characterization, In-Silico, and In-Vitro Study of a New Steroid Derivative from Ophiocoma Dentata as a Potential Treatment for COVID-19. Sci. Rep. 2022, 12, 5846. [Google Scholar] [CrossRef] [PubMed]
  59. Hamdy, R.; Mostafa, A.; Abo Shama, N.M.; Soliman, S.S.M.; Fayed, B. Comparative Evaluation of Flavonoids Reveals the Superiority and Promising Inhibition Activity of Silibinin against SARS-CoV-2. Phyther. Res. 2022, 36, 2921–2939. [Google Scholar] [CrossRef] [PubMed]
  60. Elhusseiny, S.M.; El-mahdy, T.S.; Elleboudy, N.S.; Yahia, I.S. In Vitro Anti SARS-CoV-2 Activity and Docking In Vitro Anti SARS-CoV-2 Activity and Docking Analysis of Pleurotus ostreatus, Lentinula edodes and Agaricus bisporus Edible Mushrooms. Infect. Drug Resist. 2022, 15, 3459–3475. [Google Scholar]
  61. Jorge, A.; Lopes, O.; Calado, G.P.; Nascimento, Y.; Alves, S.; Ara, D.; França, L.M.; Marcus, A.; Paes, D.A. Plant Metabolites as SARS-CoV-2 Inhibitors Candidates: In Silico and In Vitro Studies. Pharmaceutics 2022, 15, 1045. [Google Scholar]
  62. Kaur, J.; Kumar, V.; Garg, S.; Subramani, C.; Agarwal, S.; Wang, J.; Zhang, H.; Kaul, A.; Singh, R.; Kaul, S.C.; et al. Molecular Mechanism of Anti-SARS-CoV2 Activity of Ashwagandha-Derived Withanolides. Int. J. Biol. Macromol. 2021, 184, 297–312. [Google Scholar]
  63. Kim, Y.S.; Kim, B.; Kwon, E.B.; Chung, H.S.; Choi, J.G. Mulberrofuran G, a Mulberry Component, Prevents SARS-CoV-2 Infection by Blocking the Interaction between SARS-CoV-2 Spike Protein S1 Receptor-Binding Domain and Human Angiotensin-Converting Enzyme 2 Receptor. Nutrients 2022, 14, 4170. [Google Scholar] [CrossRef] [PubMed]
  64. Bernardi, M.; Ghaani, M.R.; Bayazeid, O. Phenylethanoid Glycosides as a Possible COVID-19 Protease Inhibitor: A Virtual Screening Approach. J. Mol. Model. 2021, 27, 341. [Google Scholar] [CrossRef] [PubMed]
  65. Chtita, S.; Fouedjou, R.T.; Belaidi, S.; Djoumbissie, L.A.; Ouassaf, M.; Qais, F.A.; Bakhouch, M.; Efendi, M.; Tok, T.T.; Bouachrine, M.; et al. In Silico Investigation of Phytoconstituents from Cameroonian Medicinal Plants towards COVID-19 Treatment. Struct. Chem. 2022, 33, 1799–1813. [Google Scholar] [CrossRef]
  66. Hajji, H.; Alaqarbeh, M.; Lakhlifi, T.; Ajana, M.A.; Alsakhen, N.; Bouachrine, M. Computational Approach Investigation Bioactive Molecules from SaussureaCostus Plant as SARS-CoV-2 Main Protease Inhibitors Using Reverse Docking, Molecular Dynamics Simulation, and Pharmacokinetic ADMET Parameters. Comput. Biol. Med. 2022, 150, 106209. [Google Scholar] [CrossRef]
  67. Parida, P.K.; Paul, D.; Chakravorty, D. The Natural Way Forward: Molecular Dynamics Simulation Analysis of Phytochemicals from Indian Medicinal Plants as Potential Inhibitors of SARS-CoV-2 Targets. Phyther. Res. 2020, 34, 3420–3433. [Google Scholar] [CrossRef]
  68. Adem, S.; Eyupoglu, V.; Ibrahim, I.M.; Sarfraz, I.; Rasul, A.; Ali, M.; Elfiky, A.A. Multidimensional in silico strategy for identification of natural polyphenols-based SARS-CoV-2 main protease (Mpro) inhibitors to unveil a hope against COVID-19. Comput. Biol. Med. 2022, 145, 105452. [Google Scholar] [CrossRef]
  69. Elmi, A.; Sayem, S.A.J.; Ahmed, M.; Abdoul-Latif, F. Natural Compounds from Djiboutian Medicinal Plants as Inhibitors of COVID-19 by In Silico Investigations. Int. J. Curr. Pharm. Res. 2020, 12, 52–57. [Google Scholar] [CrossRef]
  70. Ebada, S.S.; Al-Jawabri, N.A.; Youssef, F.S.; El-Kashef, D.H.; Knedel, T.O.; Albohy, A.; Korinek, M.; Hwang, T.L.; Chen, B.H.; Lin, G.H.; et al. Anti-Inflammatory, Antiallergic and COVID-19 Protease Inhibitory Activities of Phytochemicals from the Jordanian Hawksbeard: Identification, Structure-Activity Relationships, Molecular Modeling and Impact on Its Folk Medicinal Uses. RSC Adv. 2020, 10, 38128–38141. [Google Scholar] [CrossRef]
  71. Rakib, A.; Paul, A.; Chy, N.U.; Sami, S.A.; Baral, S.K.; Majumder, M.; Tareq, A.M.; Amin, M.N. Biochemical and Computational Approach of Selected Phytocompounds from Tinospora Crispa in the Management of COVID-19. Molecules 2020, 25, 3936. [Google Scholar] [CrossRef]
  72. Verma, S.; Patel, C.N.; Chandra, M. Identification of Novel Inhibitors of SARS-CoV-2 Main Protease (Mpro) from Withania Sp. by Molecular Docking and Molecular Dynamics Simulation. J. Comput. Chem. 2021, 42, 1861–1872. [Google Scholar] [CrossRef]
  73. Mujwar, S.; Sun, L.; Fidan, O. In Silico Evaluation of Food-Derived Carotenoids against 2 Drug Targets: Crocin Is a Promising Dietary Supplement Candidate for COVID-19. J. Food Biochem. 2022, 46, e14219. [Google Scholar] [CrossRef]
  74. Johnson, T.O. A Computational Approach to Elucidate the Interactions of Chemicals from Artemisia Annua Targeted Toward SARS-CoV-2 Main Protease Inhibition for COVID-19 Treatment. Front. Med. 2022, 9, 907483. [Google Scholar] [CrossRef]
  75. Prajapati, J.; Patel, R.; Goswami, D.; Saraf, M.; Rawal, R.M. Sterenin M as a Potential Inhibitor of SARS-CoV-2 Main Protease Identified from MeFSAT Database Using Molecular Docking, Molecular Dynamics Simulation and Binding Free Energy Calculation. Comput. Biol. Med. 2021, 135, 104568. [Google Scholar] [CrossRef]
  76. Zarei, A.; Ramazani, A.; Pourmand, S.; Sattari, A.; Rezaei, A. In Silico Evaluation of COVID-19 Main Protease Interactions with Honeybee Natural Products for Discovery of High Potential Antiviral Compounds. Nat. Prod. Res. 2022, 36, 4254–4260. [Google Scholar] [CrossRef]
  77. Vijayakumar, B.G.; Ramesh, D.; Joji, A.; Jayachandraprakasan, J.; Kannan, T. In Silico Pharmacokinetic and Molecular Docking Studies of Natural Flavonoids and Synthetic Indole Chalcones Against Essential Proteins of SARS-CoV-2. Eur. J. Pharmacol. 2020, 886, 173448. [Google Scholar] [CrossRef] [PubMed]
  78. Tung, B.T.; Nhung, N.H.; Thi, T.; Hang, T.; Linh, V.K.; Ngoc, L.M. Evaluating Natural Product Compound Inhibitors of SARS-CoV-2 Main Protease and Spike Protein Target by Molecular Docking Approach. Vietnam J. Chem. 2021, 59, 846–861. [Google Scholar]
  79. De Leon, V.N.O.; Manzano, J.A.H.; Pilapil, D.Y.H.; Fernandez, R.A.T.; Ching, J.K.A.R.; Quimque, M.T.J.; Agbay, J.C.M.; Notarte, K.I.R.; Macabeo, A.P.G. Anti-HIV Reverse Transcriptase Plant Polyphenolic Natural Products with In Silico Inhibitory Properties on Seven Non-Structural Proteins Vital in SARS-CoV-2 Pathogenesis. J. Genet. Eng. Biotechnol. 2021, 19, 1–17. [Google Scholar] [CrossRef] [PubMed]
  80. Mpiana, P.T.; Ngbolua, K.t.N.; Tshibangu, D.S.T.; Kilembe, J.T.; Gbolo, B.Z.; Mwanangombo, D.T.; Inkoto, C.L.; Lengbiye, E.M.; Mbadiko, C.M.; Matondo, A.; et al. Identification of Potential Inhibitors of SARS-CoV-2 Main Protease from Aloe Vera Compounds: A Molecular Docking Study. Chem. Phys. Lett. 2020, 754, 137751. [Google Scholar] [CrossRef]
  81. Ullah, S.; Munir, B.; Al-Sehemi, A.G.; Muhammad, S.; Haq, I.U.; Aziz, A.; Ahmed, B.; Ghaffar, A. Identification of Phytochemical Inhibitors of SARS-CoV-2 Protease 3CLpro from Selected Medicinal Plants as per Molecular Docking, Bond Energies and Amino Acid Binding Energies. Saudi J. Biol. Sci. 2022, 29, 103274. [Google Scholar] [CrossRef] [PubMed]
  82. Khanh, N.T.; Hoa, T.T. In Silico Studies of Natural Products from Medicinal Plants to Identify Potential Inhibitors for SARS-CoV-2 3C-like Protease. Vietnam J. Chem. 2021, 59, 557–562. [Google Scholar]
  83. Sisakht, M.; Mahmoodzadeh, A.; Darabian, M. Plant-Derived Chemicals as Potential Inhibitors of SARS-CoV-2 Main Protease (6LU7), A Virtual Screening Study. Phyther. Res. 2021, 35, 3262–3274. [Google Scholar] [CrossRef] [PubMed]
  84. Chowdhury, P.; Pathak, P. Neuroprotective Immunity by Essential Nutrient “Choline” for the Prevention of SARS-CoV-2 Infections: An In Silico Study by Molecular Dynamics Approach. Chem. Phys. Lett. 2020, 761, 138057. [Google Scholar] [CrossRef] [PubMed]
  85. Jamhour, R.M.A.Q.; Al-Nadaf, A.H.; Wedian, F.; Al-Mazaideh, G.M.; Mustafa, M.; Huneif, M.A.; Mahmoud, S.Y.; Farrag, E.S.; Al-Rimawi, F.; Salman, H.A.; et al. Phytochemicals As a Potential Inhibitor of COVID-19: An In-Silico Perspective. Russ. J. Phys. Chem. A 2022, 96, 1589–1597. [Google Scholar] [CrossRef]
  86. Tallei, T.E.; Tumilaar, S.G.; Niode, N.J.; Fatimawali; Kepel, B.J.; Idroes, R.; Effendi, Y.; Sakib, S.A.; Emran, T.B. Potential of Plant Bioactive Compounds as SARS-CoV-2 Main Protease (Mpro) and Spike (S) Glycoprotein Inhibitors: A Molecular Docking Study. Scientifica 2020, 2020, 6307457. [Google Scholar] [CrossRef]
  87. Shady, N.H.; Hayallah, A.M.; Mohamed, M.F.A.; Ghoneim, M.M.; Chilingaryan, G.; Al-Sanea, M.M.; Fouad, M.A.; Kamel, M.S.; Abdelmohsen, U.R. Targeting 3CLPRO and SARS-CoV-2 RDRP by Amphimedon sp. Metabolites: A Computational Study. Molecules 2021, 26, 3775. [Google Scholar] [CrossRef]
  88. Kumar, V.; Parate, S.; Yoon, S.; Lee, G.; Lee, K.W. Computational Simulations Identified Marine-Derived Natural Bioactive Compounds as Replication Inhibitors of SARS-CoV-2. Front. Microbiol. 2021, 12, 647295. [Google Scholar] [CrossRef]
  89. Gyebi, G.A.; Ogunyemi, O.M.; Ibrahim, I.M.; Afolabi, S.O.; Adebayo, J.O. Dual Targeting of Cytokine Storm and Viral Replication in COVID-19 by Plant-Derived Steroidal Pregnanes: An In Silico Perspective. Comput. Biol. Med. 2021, 134, 104406. [Google Scholar] [CrossRef]
  90. Mehmood, A.; Khan, S.; Khan, S.; Ahmed, S.; Ali, A.; Xue, M.; Ali, L.; Hamza, M.; Munir, A.; ur Rehman, S.; et al. In Silico Analysis of Quranic and Prophetic Medicinals Plants for the Treatment of Infectious Viral Diseases Including Corona Virus. Saudi J. Biol. Sci. 2021, 28, 3137–3151. [Google Scholar] [CrossRef]
  91. Elsbaey, M.; Ibrahim, M.A.A.; Bar, F.A.; Elgazar, A.A. Chemical Constituents from Coconut Waste and Their In-Silico Evaluation as Potential Antiviral Agents against SARS-CoV-2. S. Afr. J. Bot. 2021, 141, 278–289. [Google Scholar] [CrossRef] [PubMed]
  92. Irfan, A.; Imran, M.; Khalid, N.; Hussain, R.; Basra, M.A.R.; Khaliq, T.; Shahzad, M.; Hussien, M.; Shah, A.T.; Qayyum, M.A.; et al. Isolation of Phytochemicals from Malva neglectaWallr and their Quantum Chemical, Molecular Docking Exploration as Active Drugs against COVID-19. J. Saudi Chem. Soc. 2021, 25, 101358. [Google Scholar] [CrossRef]
  93. Panagiotopoulos, A.A.; Karakasiliotis, I.; Kotzampasi, D.M.; Dimitriou, M.; Sourvinos, G.; Kampa, M.; Pirintsos, S.; Castanas, E.; Daskalakis, V. Natural Polyphenols Inhibit the Dimerization of the SARS-CoV-2 Main Protease: The Case of Fortunellin and Its Structural Analogs. Molecules 2021, 26, 6068. [Google Scholar] [CrossRef] [PubMed]
  94. Abodunrin, O.P.; Onifade, O.F.; Adegboyega, A.E. Therapeutic Capability of Five Active Compounds in Typical African Medicinal Plants Against Main Proteases of SARS-CoV-2 by Computational Approach. Inform. Med. Unlocked 2022, 31, 100964. [Google Scholar] [CrossRef] [PubMed]
  95. Yalçın, S.; Yalçınkaya, S.; Ercan, F. In Silico Detection of Inhibitor Potential of Passiflora Compounds Against SARS-CoV-2(COVID-19) Main Protease by Using Molecular Docking and Dynamic Analyses. J. Mol. Struct. 2021, 1240, 130556. [Google Scholar] [CrossRef]
  96. Rauf, A.; Rashid, U.; Khalil, A.A.; Khan, S.A.; Anwar, S.; Alafnan, A.; Alamri, A.; Rengasamy, K.R. Docking-Based Virtual Screening and Identification of Potential COVID-19 Main Protease Inhibitors from Brown Algae. S. Afr. J. Bot. 2021, 143, 428–434. [Google Scholar] [CrossRef]
  97. Ibrahim, M.A.A.; Abdelrahman, A.H.M.; Atia, M.A.M.; Mohamed, T.A.; Moustafa, M.F.; Hakami, A.R.; Khalifa, S.A.M.; Alhumaydhi, F.A.; Alrumaihi, F.; Abidi, S.H.; et al. Blue Biotechnology: Computational Screening of Sarcophyton Cembranoid Diterpenes for SARS-CoV-2 Main. Mar. Drugs 2021, 19, 391. [Google Scholar] [CrossRef]
  98. Panikar, S.; Shoba, G.; Arun, M.; Sahayarayan, J.J.; Usha Raja Nanthini, A.; Chinnathambi, A.; Alharbi, S.A.; Nasif, O.; Kim, H.J. Essential Oils as an Effective Alternative for the Treatment of COVID-19: Molecular Interaction Analysis of Protease (Mpro) with Pharmacokinetics and Toxicological Properties. J. Infect. Public Health 2021, 14, 601–610. [Google Scholar] [CrossRef]
  99. Srivastava, A.K.; Kumar, A.; Srivastava, H.; Misra, N. The Role of Herbal Plants in the Inhibition of SARS-CoV-2 Main Protease: A Computational Approach. J. Indian Chem. Soc. 2022, 99, 100640. [Google Scholar] [CrossRef]
  100. Vijayakumar, M.; Janani, B.; Kannappan, P.; Renganathan, S.; Al-ghamdi, S.; Alsaidan, M.; Abdelaziz, M.A.; Peer, A.; Shahid, M.; Ramesh, T. In Silico Identification of Potential Inhibitors Against Main Protease of SARS-CoV-2 6LU7 from Andrographis Panniculata via Molecular Docking, Binding Energy Calculations and Molecular Dynamics Simulation Studies. Saudi J. Biol. Sci. 2022, 29, 18–29. [Google Scholar] [CrossRef]
  101. Bora, A.; Pacureanu, L.; Crisan, L. In Silico Study of Some Natural Flavonoids as Potential Agents against COVID-19: Preliminary Results. Chem. Proc. 2021, 3, 25. [Google Scholar]
  102. Imran, M.; Iqbal, S.; Hussain, A.; Uddin, J.; Kashif, M.; Mahmood, K. In Silico Screening, SAR and Kinetic Studies of Naturally Occurring Flavonoids Against SARS-CoV-2 Main Protease. Arab. J. Chem. 2022, 15, 103473. [Google Scholar] [CrossRef]
  103. Muhammad, I.; Rahman, N.; Niaz, S.; Basharat, Z.; Rastrelli, L.; Jayanthi, S.; Efferth, T.; Khan, H. Screening of Potent Phytochemical Inhibitors against SARS-CoV-2 Protease and Its Two Asian Mutants. Comput. Biol. Med. 2021, 133, 104362. [Google Scholar] [CrossRef]
  104. Baildya, N.; Ashik, A.; Nath, N.; Dutta, T.; Chattopadhyay, A.P. Screening of Potential Drug from Azadirachta Indica (Neem) Extracts for SARS-CoV-2: An Insight from Molecular Docking and MD-Simulation Studies. J. Mol. Struct. 2021, 1227, 129390. [Google Scholar] [CrossRef] [PubMed]
  105. Joshi, T.; Joshi, T.; Pundir, H.; Sharma, P.; Chandra, S. Predictive Modeling by Deep Learning, Virtual Screening and Molecular Dynamics Study of Natural Compounds against SARS-CoV-2 Main Protease. J. Biomol. Struct. Dyn. 2021, 39, 6728–6746. [Google Scholar] [CrossRef] [PubMed]
  106. Bharadwaj, S.; El-kafrawy, S.A.; Alandijany, T.A.; Bajrai, L.H.; Shah, A.A.; Dubey, A.; Sahoo, A.K.; Yadava, U.; Kamal, M.A.; Azhar, E.I.; et al. Structure-Based Identification of Natural Products as SARS-CoV-2 Mpro Antagonist from Echinacea angustifolia Using Computational Approaches. Viruses 2021, 13, 305. [Google Scholar] [CrossRef] [PubMed]
  107. Chowdhury, P. In Silico Investigation of Phytoconstituents from Indian Medicinal Herb ‘Tinospora Cordifolia (Giloy)’ against SARS-CoV-2 (COVID-19) by Molecular Dynamics Approach. J. Biomol. Struct. Dyn. 2021, 39, 6792–6809. [Google Scholar] [CrossRef]
  108. Shree, P.; Mishra, P.; Selvaraj, C.; Singh, S.K.; Chaube, R.; Garg, N.; Tripathi, Y.B. Targeting COVID-19 (SARS-CoV-2) Main Protease through Active Phytochemicals of Ayurvedic Medicinal Plants—Withania somnifera (Ashwagandha), Tinospora cordifolia (Giloy) and Ocimum sanctum (Tulsi)—A Molecular Docking Study. J. Biomol. Struct. Dyn. 2022, 40, 190–203. [Google Scholar] [CrossRef]
  109. Mohammadi, M.; Rupa, F.H.; Khan, M.F.; Rashid, R.B.; Rashid, M.A. Identification of Natural Compounds with Anti-SARS-CoV-2 Activity using Machine Learning, Molecular Docking and Molecular Dynamics Simulation Studies. Dhaka Univ. J. Pharm. Sci. 2022, 21, 1–13. [Google Scholar]
  110. Bui, T.Q.; Thi, N.; Hai, T.; Van Chen, T.; Quy, P.T.; Nguyen, L.; Du, H. Theoretical Study on Inhibitability of Some Natural Alkaloids against Influenza Virus Hemagglutinin and SARS-CoV-2 Main Protease. Vietnam J. Chem. 2022, 60, 502–517. [Google Scholar]
  111. Durgam, L.; Guruprasad, L. Computational Studies on the Design of NCI Natural Products as Inhibitors to SARS-CoV-2 Main Protease. J. Biomol. Struct. Dyn. 2022, 41, 3741–3751. [Google Scholar] [CrossRef]
  112. El-hawary, S.S.; Issa, M.Y.; Ebrahim, H.S.; Mohammed, A.F.; Hayallah, A.M.; El-kadder, E.M.A.; Sayed, A.M. Potential of (Citrus Nobilis Lour × Citrus Deliciosa Tenora) Metabolites on COVID-19 Virus Main Protease Supported by In Silico Analysis Potential of (Citrus Nobilis Lour  Citrus Deliciosa Tenora) Metabolites on COVID-19 Virus Main Protease Supported. Nat. Prod. Res. 2022, 36, 2843–2847. [Google Scholar] [CrossRef] [PubMed]
  113. Rao, P.; Shukla, A.; Parmar, P.; Rawal, R.M.; Patel, B.; Saraf, M.; Goswami, D. Reckoning a Fungal Metabolite, Pyranonigrin A as a Potential Main protease (Mpro) Inhibitor of Novel SARS-CoV-2 Virus Identified Using Docking and Molecular Dynamics Simulation. Biophys. Chem. 2020, 264, 106425. [Google Scholar] [CrossRef] [PubMed]
  114. Roshni, J.; Vaishali, R.; Ganesh, K.S.; Dharani, N.; Alzahrani, K.J.; Banjer, H.J.; Alghamdi, A.H.; Theyab, A.; Ahmed, S.S.S.J.; Patil, S. Multi-Target Potential of Indian Phytochemicals Against SARS-CoV-2 a Docking, Molecular Dynamics and MM-GBSA Approach Extended to Omicron B. 1. 1. 529. J. Infect. Public Health 2022, 15, 662–669. [Google Scholar] [CrossRef]
  115. Moharana, M.; Das, A.; Narayan, S.; Kumar, S.; Khan, F. Evaluation of Binding Performance of Bioactive Compounds Against Main Protease and Mutant Model Spike Receptor Binding Domain of SARS-CoV-2: Docking, ADMET Properties and Molecular Dynamics Simulation Study. J. Indian Chem. Soc. 2022, 99, 100417. [Google Scholar] [CrossRef]
  116. Thuy, B.T.P.; Nhan, V.D.; Quang, N.M.; Duoc, N.T.; Tat, P.V. Evaluation of SARS-CoV-2 Inhibition of Some Compounds in Cymbopogon Citratus Oil Combining Docking and Molecular Dynamics Simulations. Vietnam J. Chem. 2021, 59, 790–799. [Google Scholar]
  117. Muhseen, Z.T.; Hameed, A.R.; Al-Hasani, H.M.; Ahmad, S.; Li, G. Computational Determination of Potential Multiprotein Targeting Natural Compounds for Rational Drug Design Against SARS-CoV-2. Molecules 2021, 26, 674. [Google Scholar] [CrossRef]
  118. Goyzueta-Mamani, L.D.; Barazorda-Ccahuana, H.L.; Mena-Ulecia, K.; Chávez-Fumagalli, M.A. Antiviral Activity of Metabolites from Peruvian Plants Against SARS-CoV-2: An In-Silico Approach. Molecules 2021, 26, 3882. [Google Scholar] [CrossRef]
  119. Africa, J.G.G.; Arturo, H.C.P.; Bernardo, L.J.M.; Kyle, J.; Ching, A.R.; Casandra, O.; De Cruz, E.; Hernandez, J.B.E.; Magsipoc, R.J.Y.; Sales, C.T.C.; et al. In Silico Triple Targeting of SARS-CoV-2 3CL Pro, PL pro, and RdRp by Philippine Antitubercular Natural Products Libraries. Philipp. J. Sci. 2022, 151, 35–58. [Google Scholar]
  120. Hang, T.T.T.; Khanh, D.T.H.; Tung, B.T. In Silico Screening of Natural Antivirals as Potential Inhibitors of SARS-CoV-2 Virus. Vietnam J. Chem. 2022, 60, 211–222. [Google Scholar]
  121. Shaldam, M.A.; Yahya, G.; Mohamed, N.H.; Abdel-daim, M.M.; Naggar, Y. In Silico Screening of Potent Bioactive Compounds from Honeybee Products Against COVID-19 Target Enzymes. Environ. Sci. Pollut. Res. 2021, 28, 40507–40514. [Google Scholar] [CrossRef]
  122. Melinda, Y.N.; Widada, J.; Dwi, T.; Febriansah, R.; Damayanti, E.; Mustofa, M. Heliyon Metabologenomics Approach to the Discovery of Novel Compounds from Streptomyces sp. GMR22 as Anti-SARS-CoV-2 Drugs. Heliyon 2021, 7, 08308. [Google Scholar] [CrossRef]
  123. Dutta, M.; Nezam, M.; Chowdhury, S.; Rakib, A.; Paul, A. Appraisals of the Bangladeshi Medicinal Plant Calotropis Gigantea Used by Folk Medicine Practitioners in the Management of COVID-19: A Biochemical and Computational Approach. Front. Mol. Biosci. 2021, 8, 625391. [Google Scholar] [CrossRef]
  124. Kumar, M.; Ghosh, S.K.; Das, A. Natural Product Compounds in Alpinia Officinarum and Ginger are Potent SARS-CoV-2 Papain-like Protease Inhibitors. ChemRxiv 2020, 1, 12071997. [Google Scholar]
  125. Suleimen, Y.M.; Jose, R.A.; Suleimen, R.N.; Arenz, C.; Ishmuratova, M.; Toppet, S.; Dehaen, W.; Alsfouk, A.A.; Elkaeed, E.B.; Eissa, I.H. Isolation and In Silico Anti-SARS-CoV-2 Papain-Like Protease Potentialities of Two Rare 2-Phenoxychromone Derivatives from Artemisia spp. Molecules 2022, 27, 1216. [Google Scholar] [CrossRef] [PubMed]
  126. Jose, S.; Gupta, M.; Sharma, U.; Quintero-Saumeth, J.; Dwivedi, M. Potential of Phytocompounds from Brassica Oleracea Targeting S2-Domain of SARS-CoV-2 Spike Glycoproteins: Structural and Molecular Insights. J. Mol. Struct. 2022, 1254, 132369. [Google Scholar] [CrossRef] [PubMed]
  127. Mondal, S.; Karmakar, A.; Mallick, T.; Begum, N.A. Exploring the Efficacy of Naturally Occurring Biflavone Based Antioxidants towards the Inhibition of the SARS-CoV-2 Spike Glycoprotein Mediated Membrane Fusion. Virology 2021, 556, 133–139. [Google Scholar] [CrossRef] [PubMed]
  128. Kim, Y.S.; Chung, H.S.; Noh, S.G.; Lee, B.; Chung, H.Y.; Choi, J.G. Geraniin Inhibits the Entry of SARS-CoV-2 by Blocking the Interaction between Spike Protein RBD and Human ACE2 Receptor. Int. J. Mol. Sci. 2021, 22, 8604. [Google Scholar] [CrossRef] [PubMed]
  129. Kar, P.; Saleh-E-In, M.M.; Jaishee, N.; Anandraj, A.; Kormuth, E.; Vellingiri, B.; Angione, C.; Rahman, P.K.S.M.; Pillay, S.; Sen, A.; et al. Computational Profiling of Natural Compounds as Promising Inhibitors against the Spike Proteins of SARS-CoV-2 Wild-Type and the Variants of Concern, Viral Cell-Entry Process, and Cytokine Storm in COVID-19. J. Cell. Biochem. 2022, 123, 964–986. [Google Scholar] [CrossRef] [PubMed]
  130. Bhowmik, A.; Biswas, S.; Hajra, S.; Saha, P. In Silico Validation of Potent Phytochemical Orientin as Inhibitor of SARS-CoV-2 Spike and Host Cell Receptor GRP78 Binding. Heliyon 2021, 7, 05923. [Google Scholar] [CrossRef]
  131. Kiran, G.; Karthik, L.; Devi, M.S.S.; Sathiyarajeswaran, P.; Kanakavalli, K.; Kumar, K.M.; Kumar, D.R. In Silico Computational Screening of KabasuraKudineer-of Fi Cial Siddha Formulation and JACOM against SARS-CoV-2 Spike Protein. J. Ayurveda Integr. Med. 2022, 13, 100324. [Google Scholar] [CrossRef]
  132. Mhatre, S.; Gurav, N.; Shah, M.; Patravale, V. Entry-Inhibitory Role of Catechins Against SARS-CoV-2 and Its UK Variant. Comput. Biol. Med. 2021, 135, 104560. [Google Scholar] [CrossRef]
  133. Shadrack, D.M.; Vuai, S.A.H.; Sahini, M.G.; Onoka, I. In Silico Study of the Inhibition of SARS-CoV-2 Viral Cell Entry by Neem Tree Extracts. RSC Adv. 2021, 11, 26524–26533. [Google Scholar] [CrossRef] [PubMed]
  134. Cheke, R.S.; Jain, G.; Narkhede, R.; Shinde, S.D.; Ambhore, J.P. Natural Product Emerging as Potential SARS Spike Glycoproteins-ACE2 Inhibitors to Combat COVID-19 Attributed by In-Silico Investigations. Biointerface Res. Appl. Chem. 2021, 11, 10628–10639. [Google Scholar]
  135. Prasad, S.K.; Pradeep, S.; Shimavallu, C.; Kollur, S.P.; Syed, A.; Marraiki, N.; Egbuna, C.; Gaman, M.; Kosakowska, O.; Cho, W.C.; et al. Evaluation of Annona Muricata Acetogenins as Potential Anti-SARS-CoV-2 Agents through Computational Approaches. Front. Chem. 2021, 8, 1281. [Google Scholar] [CrossRef] [PubMed]
  136. Huang, J.; Wang, J.; Li, Y.; Wang, Z.; Chu, M.; Wang, Y. Tuftsin: A Natural Molecule Against SARS-CoV-2 Infection. Front. Mol. Biosci. 2022, 9, 859162. [Google Scholar] [CrossRef] [PubMed]
  137. Tallei, T.E.; Yelnetty, A.; Idroes, R.; Kusumawaty, D. An Analysis Based on Molecular Docking and Molecular Dynamics Simulation Study of Bromelain as Anti-SARS-CoV-2 Variants. Front. Pharmacol. 2021, 12, 2192. [Google Scholar] [CrossRef] [PubMed]
  138. Dharmashekara, C.; Pradeep, S.; Prasad, S.K.; Jain, A.S.; Syed, A.; Shiva, K.; Patil, S.S.; Beelagi, M.S.; Srinivasa, C.; Shivamallu, C. Virtual Screening of Potential Phyto-Candidates as Therapeutic Leads against SARS-CoV-2 Infection. Environ. Chall. 2021, 4, 100136. [Google Scholar] [CrossRef]
  139. Zhang, Y.; Yao, Y.; Yang, Y.; Wu, H. Investigation of Anti-SARS, MERS, and COVID-19 Effect of Jinhua Qinggan Granules Based on a Network Pharmacology and Molecular Docking Approach. Nat. Prod. Commun. 2021, 16, 1–9. [Google Scholar] [CrossRef]
  140. Kashyap, D.; Jakhmola, S.; Tiwari, D.; Kumar, R.; Moorthy, N.H.N.; Elangovan, M.; Brás, N.F.; Jha, H.C. Plant Derived Active Compounds as Potential Anti SARS-CoV-2 Agents: An In-Silico Study. J. Biomol. Struct. Dyn. 2021, 40, 10629–10650. [Google Scholar] [CrossRef]
  141. Siddiqui, S.; Ahmad, R.; Alaidarous, M.; Zia, Q.; Ahmad, S.; Bader, M.; Aditi, A.; Anchal, S. Phytoconstituents from Moringa Oleifera Fruits Target ACE2 and Open Spike Glycoprotein to Combat SARS-2: An Integrative Phytochemical and Computational Approach. J. Food Biochem. 2022, 46, 14062. [Google Scholar] [CrossRef] [PubMed]
  142. Mir, S.A.; Firoz, A.; Alaidarous, M.; Alshehri, B.; Bin Dukhyil, A.A.; Banawas, S.; Alsagaby, S.A.; Alturaiki, W.; Bhat, G.A.; Kashoo, F.; et al. Identification of SARS-CoV-2 RNA-Dependent RNA Polymerase Inhibitors from the Major Phytochemicals of Nigella Sativa: An In Silico Approach. Saudi J. Biol. Sci. 2022, 29, 394–401. [Google Scholar] [CrossRef]
  143. Husain, I.; Ahmad, R.; Siddiqui, S.; Chandra, A.; Misra, A.; Srivastava, A.; Ahamad, T.; Faheem, M.; Siddiqi, Z.; Trivedi, A.; et al. Structural interactions of phytoconstituent(s) from cinnamon, bay leaf, oregano, and parsley with SARS-CoV-2 nucleocapsid protein: A comparative assessment for development of potential antiviral nutraceuticals. J. Food Biochem. 2022, 46, 14262. [Google Scholar] [CrossRef]
  144. Suravajhala, R.; Parashar, A.; Choudhir, G.; Kumar, A.; Malik, B. Molecular Docking and Dynamics Studies of Curcumin with COVID-19 Proteins. Netw. Model. Anal. Health Inform. Bioinform. 2021, 10, 44. [Google Scholar] [CrossRef]
  145. Muhseen, Z.T.; Hameed, A.R.; Al-Hasani, H.M.H.; Tahir ul Qamar, M.; Li, G. Promising Terpenes as SARS-CoV-2 Spike Receptor-Binding Domain (RBD) Attachment Inhibitors to the Human ACE2 Receptor: Integrated Computational Approach. J. Mol. Liq. 2020, 320, 114493. [Google Scholar] [CrossRef]
  146. Hu, L.C.; Ding, C.H.; Li, H.Y.; Li, Z.Z.; Chen, Y.; Li, L.P.; Li, W.Z.; Liu, W.S. Identification of Potential Target Endoribonuclease NSP15 Inhibitors of SARS-CoV-2 from Natural Products through High-Throughput Virtual Screening and Molecular Dynamics Simulation. J. Food Biochem. 2022, 46, 14085. [Google Scholar] [CrossRef]
  147. Naik, B.; Gupta, N.; Ojha, R.; Singh, S.; Prajapati, V.K.; Prusty, D. High Throughput Virtual Screening Reveals SARS-CoV-2 Multi-Target Binding Natural Compounds to Lead Instant Therapy for COVID-19 Treatment. Int. J. Biol. Macromol. 2020, 160, 1–17. [Google Scholar] [CrossRef]
  148. Bandyopadhyay, S.; Abiodun, O.A.; Ogboo, C.; Kola-mustapha, A.T.; Ifeanyi, E.; Edemhanria, L.; Kumari, A.; Jaganathan, R.; Niyi, S. Polypharmacology of Some Medicinal Plant Metabolites Against SARS-CoV-2 and Host Targets: Molecular Dynamics Evaluation of NSP9 RNA Binding Protein. J. Biomol. Struct. Dyn. 2021, 40, 11467–11483. [Google Scholar] [CrossRef]
  149. Khan, M.T.; Ali, A.; Wang, Q.; Irfan, M.; Khan, A.; Zeb, M.T.; Zhang, Y.; Chinnasamy, S.; Wei, D. Marine natural compounds as potent inhibitors against the main protease of SARS-CoV-2. A molecular dynamic study. J. Biomol. Struct. Dyn. 2020, 39, 3627–3637. [Google Scholar] [CrossRef] [PubMed]
  150. Sigman, D.S.; Mooser, G. Chemical Studies of Enzyme Active Sites. Annu. Rev. Biochem. 1975, 40, 889–931. [Google Scholar] [CrossRef] [PubMed]
  151. Parida, P.K.; Paul, D.; Chakravorty, D. Nature’s Therapy for COVID-19: Targeting the Vital Non-Structural Proteins (NSP) from SARS-CoV-2 with Phytochemicals from Indian Medicinal Plants. Phytomedicine Plus 2021, 1, 100002. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structure of some prescribed repurposed drugs for SARS-CoV-2.
Figure 1. Structure of some prescribed repurposed drugs for SARS-CoV-2.
Molecules 28 04860 g001
Figure 2. Natural compounds reported as SARS-CoV-2 inhibitors through in vivo studies.
Figure 2. Natural compounds reported as SARS-CoV-2 inhibitors through in vivo studies.
Molecules 28 04860 g002
Figure 3. Natural compounds reported as SARS-CoV-2 inhibitors through in vitro studies.
Figure 3. Natural compounds reported as SARS-CoV-2 inhibitors through in vitro studies.
Molecules 28 04860 g003
Figure 4. Natural compounds reported as SARS-CoV-2 inhibitors through in vitro and in silico studies.
Figure 4. Natural compounds reported as SARS-CoV-2 inhibitors through in vitro and in silico studies.
Molecules 28 04860 g004
Figure 5. Natural compounds as anti-SARS-CoV-2 agents.
Figure 5. Natural compounds as anti-SARS-CoV-2 agents.
Molecules 28 04860 g005
Figure 6. Natural compounds as Mpro inhibitors.
Figure 6. Natural compounds as Mpro inhibitors.
Molecules 28 04860 g006
Figure 7. Natural compounds acting as anti-COVID-19 agents by prohibiting Mpro.
Figure 7. Natural compounds acting as anti-COVID-19 agents by prohibiting Mpro.
Molecules 28 04860 g007
Figure 8. Chemical structures of SARS-CoV-2 main protease inhibitors.
Figure 8. Chemical structures of SARS-CoV-2 main protease inhibitors.
Molecules 28 04860 g008
Figure 9. SARS-CoV-2 Mpro inhibition from natural sources.
Figure 9. SARS-CoV-2 Mpro inhibition from natural sources.
Molecules 28 04860 g009aMolecules 28 04860 g009b
Figure 10. Chemical structures of potent inhibitors of Mpro.
Figure 10. Chemical structures of potent inhibitors of Mpro.
Molecules 28 04860 g010
Figure 11. Natural products as spike protein inhibitors.
Figure 11. Natural products as spike protein inhibitors.
Molecules 28 04860 g011aMolecules 28 04860 g011b
Figure 12. Natural products as inhibitors of replication enzyme (RdRp) of SARS-CoV-2.
Figure 12. Natural products as inhibitors of replication enzyme (RdRp) of SARS-CoV-2.
Molecules 28 04860 g012aMolecules 28 04860 g012b
Figure 13. Natural compound as nucleocapsid inhibitor of SARS-CoV-2.
Figure 13. Natural compound as nucleocapsid inhibitor of SARS-CoV-2.
Molecules 28 04860 g013
Figure 14. Structures of the SARS-CoV-2 helicase inhibitors.
Figure 14. Structures of the SARS-CoV-2 helicase inhibitors.
Molecules 28 04860 g014
Figure 15. Natural compounds as SARS-CoV-2 anti-methyltransferase agents.
Figure 15. Natural compounds as SARS-CoV-2 anti-methyltransferase agents.
Molecules 28 04860 g015
Figure 16. Exoribonuclease inhibitor of SARS-CoV-2.
Figure 16. Exoribonuclease inhibitor of SARS-CoV-2.
Molecules 28 04860 g016
Figure 17. Natural products acting as SARS-CoV-2 NSP inhibitors.
Figure 17. Natural products acting as SARS-CoV-2 NSP inhibitors.
Molecules 28 04860 g017
Figure 18. Envelope protein inhibitor of SARS-CoV-2.
Figure 18. Envelope protein inhibitor of SARS-CoV-2.
Molecules 28 04860 g018
Table 2. In silico results for compounds for which no in vitro studies are available.
Table 2. In silico results for compounds for which no in vitro studies are available.
Serial No.Compound NameBinding PotentialInteractions at the Enzyme Active Site *References
Docking ScoreBinding Free
Energy
H-Bond InteractionsHydrophobic and Other Interactions
Major Protease (Mpro/3CLpro)
1Forsythiaside A (46)−8.08 kcal/mol−121 ± 19 kJ/molTHR190, PRO168HIS41, ASP187[64]
2Pycnanthuquinone C (47)-−7.8 kcal/molHIS41GLU166, HIS164, CYS145, LEU27[65]
3Pycnanthuquinone B (48)-−8.3 kcal/molGLY143, SER144HIS41, HIS164[65]
4Cyanopyrin (49)−4.78 kcal/mol−7.4 kcal/molTHR111, ASP153, SER158PHE294, ILE106[66]
5Withanolide R (50)−9.63 kcal/mol−141.96 kJ/molGLN189, HIS41MET165, PRO168, ASP187, ALA191[67]
6Hesperidin (51)−178.5910 kJ/mol-GLU166, GLN192, ARG 188, GLN189, MET49, ASP187, TYR54, LEU141, SER144, HIS163,
THR26
-[68]
7Rutin (29)−176.2740 kJ/mol
 
−9.09 kcal/mol
-
 
-
CYS145, ASN142, PHE140, GLU166, GLN192, THR190, ASP187, TYR54, HIS164, GLU166, LEU8-
 
-
[68]
 
[69]
8Diosmin (52)−174.126 kJ/mol-GLN192, THR190, ARG188, HIS164, GLN189, GLU166, GLY143, SER144, CYS145-[68]
9Apiin (53)−171.008 kJ/mol-THR190, SER144, LEU141, HIS163, CYS145, THR26-[68]
10Jaceidin (54)−7.3 kcal/mol-LEU141, GLY143, SER144, CYS145, ARG188, -[70]
11Pachypodol (55)−7.1 kcal/mol-GLY143, SER144, CYS145-[70]
12Chrysosplenetin (56)−7.1 kcal/mol-LEU141, GLY143, SER144,
CYS145
-[70]
13Imidazoline-4-one-2-imino-1-(4-methoxy-6-dimethylamino-1,3,5-triazin-2-yl) (57)−7.013 kcal/mol-GLU166, GLN192-[71]
14Spiro[4,5]dec-6-en-1-ol, 2,6,10,10-tetramethyl (58)−6.369 kcal/mol--MET165, HIS41, MET49, CYS145[71]
153-hydroxy-5-cholen-24-oic acid (59)−6.251 kcal/mol-THR25, THR190, GLN192MET49, HIS41, MET165[71]
16Withacoagulin H (60)−11.1 kcal/mol−63.463 kJ/molGLY143, ARG188, THR190, GLN192HIS41, CYS145, THR25, THR26, LEU27, MET49, ASN142, SER144, HIS164, MET165, GLN166, GLN189[72]
17Ajugin E (61)−11.5 kcal/mol−56.14 kJ/molGLN192, THR190HIS41, CYS145, THR25, MET49, ASN142, GLY143, HIS164, MET165, GLU166, PRO168, ARG188, GLN189[72]
18Withacoagulin (62)−10.8 kcal/mol−44.496 kJ/molTHR190, GLN192, GLY143HIS41, CYS145, THR25, THR26, LEU27, MET49, ASN142, SER144, HIS164, MET165, ARG188, GLN189[72]
19Crocin (63)−8.0 kcal/mol-ALA116, SER123, SER139, PHE140, LEU141, ASN142, GLU166, HIS172, GLN189 [73]
20Rhamnocitrin (64)−7.83 kcal/mol−49.53 kcal/molGLU166, THR190HIS41, CYS145, CYS44, TYR54, PRO52, MET49, ASP187, ARG188, GLN189, ALA191, GLN192, PRO168, LEU167, MET165, HIS164[74]
21Sterenin M (65)−8.431 kcal/mol−49.57 kcal/molGLU166, PHE140, HIS163, GLY143, ASN142, THR26, HIS41LEU141, HIS172, SER144, LEU27, THR25, THR24, CYS44, TYR54, ASP187, ARG188, MET49, GLN189, HIS164, MET165[75]
223,4,5-tricaffeoylquinic acid (66)−9.0 kcal/mol-THR190, ARG188, THR26, ASP187, GLY143, ASN142MET165, MET49, PRO168, HIS172[76]
23Quercetin (40)−9.2 kcal/mol-GLU290, GLU288, THR199LYS137, ARG131, LEU272, TYR239, LEU287[77]
24Peonidin (67)−8.4 kcal/mol-GLU290, ASP289, LYS5GLU288
LEU287, LEU271, GLY27, THR199, LEU272, GLU290
[77]
25Catechin (41)−7.67 kcal/mol-GLU166, HIE164-[69]
26Kaempferol (23)−7.215 kcal/mol-GLU166, GLN189, HIE164-[69]
27Amentoflavone (42)−9.8 kcal/mol
 
−8.6 kcal/mol
-
 
-
THR25, GLY143, GLN189, GLU166
 
CYS44, VAL186, ARG188, GLU166
MET49, HIS41, CYS145, THR24, THR45, MET165, TYR54, ARG188, ASP187, LEU167, PRO168, ASN142
 
THR25, HIS41, ASN142, CYS145
[78]
 
[79]
28Ferolide (68)−7.9 kcal/mol-HIS163, CYS145, LEU141, ASN142CYS145, HIS41, GLU166, SER144, PHE140, HIS164, MET165, GLN189, CYS144, MET49, THR25, GLY143[80]
29Ginkgolide A (69)−65.412 kcal/mol−63 kcal/molVAL72, LYS73, TYR135, GLY151, CYS144, HIS41[81]
30Gracillin (70)−9.2 kcal/mol-GLU166, LEU141, SER144, CYS145, GLY143, ASN142-[82]
31Proanthocyanidin (71)−9.2 kcal/mol-GLU166, HIS164, HIS163, TYR54HIS41, ASP187, MET165[82]
32Ginkgolide M (72)−11.2 kcal/mol-ASN142, CYS145, GLU166, GLY143, HIS163, PHE140[83]
33Mezerein (73)−11 kcal/mol-ASN142, CYS145, GLU166, GLY143, HIS163, HIS172, PHE140[83]
34Tubocuraine (74)−10.9 kcal/mol-ASN142, CYS145, GLU166, GLY14[83]
35Choline (75)−3.7 kcal/mol-GLY143, LEU141ASN142, PHE140, HIS163, SER144, GLU166[84]
36Volkensiflavone (76)−8.6 kcal/mol-CYS145, GLU166THR25, HIS41[79]
37Stigmasterol (77)−6.30 kcal/mol-PRO39CYS145, LEU27, VAL42, HIS41[85]
38Hesperidin (51)−8.3 kcal/mol-PHE140, GLU166, SER144, CYS145MET49, LEU141, HIS163, THR26[86]
39Amphimedoside C (78)-−127.3 kJ/molASN142, PHE140, LEU141, HIS163, SER144, CYS145ASN142, HIS172, GLN189[87]
40Fasciospongide A (79)-−104.37 kJ/molHIS41, THR190, GLN192, VAL186MET165, MET49, PRO168, ALA193, LEU167, PHE185, ARG188, VAL42, THR25, CYS145[88]
41Glaucogenin D (81)−7.9 kcal/mol-GLY143, HIS41MET165, THR24[89]
42Calcium elenolate (83)−7.0 kcal/mol-GLU291LYS152, PHE111[90]
43Maackin A (164)−8.4 kcal/mol−43.1 kcal/molTHR26, HIS41, MET49, GLN192THR25, GLY143, ASN142,
MET165, PRO168
[91]
44Anthracene dione (165)−5.73 kcal/mol-GLY143-[92]
45Fortunellin (166)−13.9 kcal/mol-LEU32, ASP33, ASP34, VAL35, TYR37, GLN83, LYS88, TYR101, LYS102, PHE103, VAL104, ARG105, ASP108, PHE159, CYS160, ASP176, LEU177, GLU178[93]
46Curcumin (86)−8.62 kcal/mol-THR26, GLY143, GLN189, THR190-[94]
47Kolaviron (87)−7.027 kcal/mol-GLU166, GLY143-
48Bisdemethoxycurcumin (88)−5.641 kcal/mol-THR26, THR190, GLN189, GLU166-[94]
496-gingerol (89)−4.975 kcal/mol-THR190, GLN189-[94]
50Artemisinin (26)−4.252 kcal/mol-GLN 189, GLU 166-[94]
51Luteolin (90)−8.3 kcal/mol-GLN110, THR111 ASN151-[95]
52Lucenin (91)−10.7 kcal/mol-PHE219, LEU220, ARG222, ASN274-[95]
53Olealonic acid (92)−9.5 kcal/mol-ASP289-[95]
54Isoorientin (93)−9.2 kcal/mol-PRO52, TYR54, PHE181-[95]
55Isochaphoside (94)−10.5 kcal/mol-ARG222, ASN274, PHE219-[95]
56Saponarin (95)−10.6 kcal/mol-ARG40, PHE181-[95]
57Schaftoside (96)−10.2 kcal/mol-ASP197, ASN238, ARG131, LYS137-[95]
587,2″-bieckol (98)−10.7855 kcal/mol-THR24, THR26, GLY143, GLU189-[96]
59Sarelengan B (105)−9.8 kcal/mol-HIS41, CYS145, GLU166-[97]
60Bislatumlide A (106)−9.6 kcal/mol−34.8 kcal/molGLY143, GLU166-[97]
61Eucalyptol (107)−5.86 kcal/mol--MET49, MET165, HIS164, ARG188, HIS41, PRO52, ASP187, GLN189, TYR54, PHE18[98]
62Nictoflorin (114)−9.18 kcal/mol THR190, GLY143-[99]
63Aloenin (115)−9.13 kcal/mol PHE140-[99]
64Andrographolide (116)−7.06 kcal/mol-PHE140, SER144-[100]
65Galangin (117)−8.066 kcal/mol-CYS145, PHE140, HIS164, HIS41HIS41, MET165, MET49, GLU166[101]
65Amentoflavone (42)−9.9 kcal/mol-HIS163, THR26, GLU166GLN189, ARG188, ASP187, TYR54, MET49, HIS164, HIS172, PHE140, SER144, LEU141, GLY143, LEU27, THR25, THR26, ASN142[102]
67Kazinol T (123)−14.355008 kcal/mol-HIS41GLY143, THR190, GLY143, HIS42, CYS145, GLY143, HIS164[103]
68Desacetyl gedunin (125)−7.3 kcal/mol-TYR 207, SER245, MET206, ARG166, VAL202, SER170, GLU203, LEU199, THR197, LEUA85, LYS232, MET208[104]
69Palmatine (126)−8.7 kcal/mol−71.47 kJ/molGLN189ASN142, GLY143, THR25, LEU27, ARG188,GLN192, PRO168, CYS145, MET165, GLN189, HIS41[105]
70Sauchinone (127)−8.9 kcal/mol−71.68 kJ/molMET49, HIS41ARG188, ASP187, PRO52, CYS44, HIS41, THR25, LEU141, GLU166, GLN189, CYS145, MET165[105]
71Quercetagetin 7-glucoside (130)−15.20 kcal/mol-CYS44, LEU141, CYS145, GLU166, GLN189, HIS41, CYS44LEU27, CYS44, MET49, PHE140, LEU141, CYS145, MET165, LEU167, PRO168[106]
72Berberine (134)−7.3 kcal/mol-THR25, SER46HIS163[107]
73Withanoside V (43)−10.32 kcal/mol-ASN84, ARG40, MET 82CYS85, ARG105, PHE134[108]
74Lignoid (140)−9.0 kcal/mol-THR25, HIS41, LEU141, SER144, CYS145, HIS163, GLN189LEU27, MET49, HIS163, MET165, ASP187, GLN189[109]
75Lycorine (141)−11.9 kcal/mol-GLU166, HIS41ASP188, GLN192, MET 165, GLN 189, HIS 164, HIS163, PHE 140, LEU 141, CYS 146, SER 144, ARG 188[110]
76Hemanthamine (142)−11.4 kcal/mol-MET165, HIS163, HIS41CYS145, PHE140, ASN142, SER144, LEU141, ARG188, HIS164, GLU166, ASP187, GLN189[110]
77NSC36398 (143)−8.1 kcal/mol-PHE140, LEU141, SER144, MET165, GLU166, GLN189LEU27, HIS41, LEU50, PHE140, LEU141, GLY143, SER144, CYS145, HIS163, HIS164 MET165, GLU166, LEU167, PRO168, GLN189[111]
78Quercetin 7-O-glucoside-3-O-rutinoside (147)−9.47 kcal/mol-PRO168, GLU166, CYS145, MET165, MET49, GLN189MET165, GLU166, GLN189, ASN142, MET49, PRO168, HIS41[112]
79Pyranonigrin A (151)−7.3 kcal/mol-LEU141, GLY143, SER144, CYS145, HIS163, GLU166 GLN189, ASN142MET165[113]
80Arboreol (153)−8.2 kcal/mol---[114]
81Acacetin (154)−7.77 kcal/mol-MET49, TYR54, CYS145MET49[115]
82Epoxy-linalool oxide (155)−12.80 kcal/mol-LEU141, CYS145, GLY143MET165, ASN142, HIS163, HIS41, GLU166, HIS164, SER144, LEU27, THR25, MET49, THR26[116]
83Glycyrrhizin (156)−8.7 kcal/mol-ASP289, ASN238TYR239, LYS137, ASP197, THR198, GLU290, SER284, TYR237, LEU287, LEU286, LEU272, GLY275, LEU271, GLN273, ASN274, MET276, TYR239[117]
84Rutin (29)-−40.293 kcal/molASN142, MET49, HIS41, ASP187[118]
85Vobtusine lactone (157)−8.3 kcal/mol-ASN119, SER46, THR25CYS44, HIS41, MET49[119]
86Sotetsuflavone (159)−9.6 kcal/mol-GLU166, GLN189, THR25PRO168, LEU167, GLY170, MET165, ASP187, ARG188, TYR54, THR45, THR24, ASN142[120]
87Kaempferol (23)−7.8 kcal/mol-SER144, LEU141, ASP187, TYR54, GLN189HIS163, HIS164, CYS145, GLU166, MET165, MET49, HIS41, ARG188[121]
88Echoside A (168)−8.4 kcal/mol-THR25, HIS41, CYS145, GLU 166MET165, MET49[122]
89Echoside B (169)−9.4 kcal/mol-ARG4, SER 284, and LYS5LYS5, GLU288, LEU282[122]
90Juniper camphor (172)−6.06 kcal/mol-MET165MET49, MET165, PRO168, ASP147[123]
Papain-Like Protease
91Amentoflavone (42)−10.8 kcal/mol-HIS342, LYS711, ARG712LYS711, ASP339, ARG558, ILE310, ILE580, ALA579, LEU742[79]
92Jezonofol
(167)
−9.0 kcal/mol−60.7 kcal/molARG166, GLU16, GLN174, TYR264SER170. GLY163, MET208, GLN203, LYS157, VAL202,
MET206,LEU199, TYR207, LEU185, GLU161, LEU162,
GLN269, TYR268
[91]
93Constanolactone B (80)-−92.57 kJ/molTYR268, ASP164ASN109, TYR11, GLY271, LEY162, CYS270, GLY163, GLN269, THR301, ARG166, PRO248, MET208, PRO247, ALA246, SER245[88]
94Glaucogenin D (81)−6.4 kcal/mol-HIS272TRP106, HIS272[89]
95Glaucogenin A (82)−6.4 kcal/mol-HIS272, ASP286TRP106, HIS272, TRP270[89]
96(E)-7-(4-hydroxy-3-methoxyphenyl)-1-phenylhept-4-en-3-one (84)-−47 kJ/mol (closed conformation)
 
−28 kJ/mol (open conformation)
ARG166, ASP164, GLN269, TYR264, LEU163, GLY163, MET208
-
[124]
978-gingerol (85) −43 kJ/mol (closed conformation)
 
−15 kJ/mol (open conformation)
ARG166, ASP164, TYR268, GLN269, TYR264, PRO247, GLY163, PRO248
-
[124]
984,8-dihydroxysesamin (152)−10.3 kcal/mol---[114]
99Glycyrrhizin (156)−7.9 kcal/mol-ASN128, ASP179, GLN174ASP76, ASN156, ARG82, THR74, PHE79, TYR154, HIS175, HIS73, PHE69, VAL202, PHE173, LEU178, ALA176, ASN177[117]
100Rutin (29)-−21.713 kcal/molTYR268, GLN269[118]
101Deoxyvobtusine lactone (158)−10.8 kcal/mol-ARG558, ARG712ILE310, PHE735, LEU742, ASP339 THR583[119]
1026-demethoxy-4′-O-capillarsine (170)−18.86 kcal/mol-GLN270, PRO249TYR269, ASP165, PRO248, PRO249[125]
103Tenuflorin C (171)−18.37 kcal/mol-GLN270, TYR274, ALA247, LEU163TYR269, ASP165, PRO248, ASP165, MET209[125]
Spike Proteins
104Terpene NPACT01552 (173)−11 kcal/mol-GLY496, TYR453, GLN493, SER494, GLU484LYS417, LEU452, TYR489, PHE456, LEU455[126]
105Hinokiflavone (175)−9.60 kcal/mol-GLN1201, GLN926, GLU1195, LEU1197, ASN928, GLU918, ASN919GLU1195, LEU1197, GLU918[127]
106Robustaflavone (176)−9.40 kcal/mol-ASP936, GLN926, LYS1191, GLU1195, ASN928GLU1195, ASN928[127]
107Abietatriene (178)−9.8 ± 0.02 kcal/mol--TYR453, TYR495, ASN501, TYR505[128]
108Quercetin (40)−7.8 kcal/mol-ASN501, TYR505, GLY496GLN493, LYS417, GLU406, LEU455, TYR495, PHE497, GLN506[77]
109Hesperidin (51)-−10.4 kcal/molASN1023, SER1030, LHR1027ARG1039, ALA1026, LEU1024, PHE1042, ARG1039, THR1027, ARG1039, LEU1024, THR1027, ALA1020, ASN1023, GLN784, GLU780[86]
110Nabiximols (179)-−10.2 kcal/molGLN762, LYS776, GLU1017ARG765, ARG1014, ALA958, ALA766, LEU1012, GLY769[86]
111Amentoflavone (42)−9.9 kcal/mol-LEU861, LYS733, SER730, THR732ALA956, PRO862, HIS1058, ASP867, VAL860, PRO1057, MET731, VAL952, ASN955, PHE823[78]
112Crocin (63)−7.6 kcal/mol-ARG346, PHE347, SER349, TYR351, LEU441, LYS444, VAL445, GLY447, ASN448, TYR449, ASN450, PHE490[73]
113Pelargonidin-3-galactoside (180)−8.6 kcal/mol-GLN166, PHE83SER80, LYS103, PRO40, ASP165[90]
114Orientin (181)−6.2 kcal/mol-LYS333, ASP429, THR431, ASN435, ASN437, TYR438, ARG495ASN437[129]
115Chrysoeriol (182)−11.478 kcal/mol-CYS336, GLY339, ASP364PHE338, PHE342, PHE374, LEU335, VAL367, SER373[130]
116Luteolin (183)−11.392 kcal/mol-ASP364, VAL367, SER371, SER373, CYS336, VAL362PHE338, GLY339, PHE374, PHE342[130]
117Acacetin (154)−7.75 kcal/mol-THR26, THR24, HIS164, GLU166, LEU141, SER144, CYS145, GLT143HIS41[115]
118Epigallocatechin gallate (21)−6.3 kcal/mol-TYR1110, PHE1109, GLN1071, GLU918, THR716-[131]
119Geraniin (177)−8.1 kcal/mol-ARG403, TYR449, TYR453, GLN493, SER494, GLN498TYR495, GLY496[132]
120cis-Annonacin (192)−7.7 kcal/mol-PHE390, ARG393, GLN409, GLY496, TYR505, ARG403, TYR453ASN33, GLU37, PRO389, ARG403, GLU406, LYS417, TYR495, PHE497[133]
121Trigoneoside IB (195)−8.5 kcal/mol-THR260, ALA618, SER289, LYS946, SER279, THR261, THR616, GLY288, THR302, THE285, CYS288-[134]
122Arctiin (196)137.043 (LiDock score)-SER142THR193, ALA146, GLY144, SER196, GLY143, PRO195, PHE118, THR141, VAL194[135]
123Linarin (197)162.676 (LiDock score)-GLY42, VAL105, GLN168ARG44, GLY103, GLN102, TYR88, GLN43, GLN39, ALA85, GLU107, PRO41, SER170, GLU167, THR104, GLY103[135]
124Withanolide D (198)−9.8 kcal/mol-GLN954ASP950, VAL951, LYS947, PRO728, GLU1017, GLU773, ILE770, LEU1012, ILE1013, GLN1010, AGR1014, ALA766, GLY769[136]
125Morellic acid (199)−10.3 kcal/mol-THR549, THR547MET740, GLY744, ASN856, PHE855, PHE541, ILE587, THR573, LEU546, ASP571, LEU977, ARG1000, SER975, ASN978, ASN540, ASP745, GLY548[120]
126Echoside A (168)−7.9 kcal/mol-GLU1195, LEU1200, GLN1201, ASP1199LEU1197[122]
127Echoside B (169)−7.8 kcal/mol-SER943, ASP936, ARG1185LYS1191, ALA1190[122]
128Pyrrolidinone (200)−5.97 kcal/mol-SER730THR778, PHE782, VAL729, ILE870, ALA1056, AND GLY1059[137]
129Scutellarein (118)−8.9 kcal/mol---[114]
(a) RBD-ACE2
1303-p-coumaroylquinic acid (174)−8.9 kcal/mol−6.71 kcal/molTHR1006, GLN1005ALA766, LEU763, VAL1008, GLN1010, GLN1002, THR1009, THR1006[138]
131Rutin (29)−7.601 kcal/mol-ASN388, ASP389, ALA363, CYS361, SER359, ILE332, ASN331-[69]
132Catechin (41)−6.470 kcal/mol-SER359, ASN331, CYS 361, ILE332-[69]
133Kaempferol (23)−6.743 kcal/mol-ILE358, ASN388-[69]
134Azadirachtin H (184)−8.18 kcal/mol-ARG393, ARG408, LYS417, GLN409, ASP30, ASN33, HIS34, LYS31, ASP30, GLU406, LYS455, ARG403[139]
135Indigo blue (186)−11.2 kcal/mol-GLN 947, GLN 744PHE 741, THR 943[140]
136Echoside A (168)−7.5 kcal/mol-ARG393, PHE390, ASN394ALA99, LEU100, LEU73, LEU391, PHE40[122]
137Echoside B (169)−8.2 kcal/mol-SER47TRP349[122]
(b) ACE2
138Geraniin (177)−7.0 kcal/mol-ARG403, TYR449, TYR453, GLN493, SER494, GLN498TYR495, GLY496[132]
139Tuftsin (193)−6.9 kcal/mol-SER47ASN51, HIS345, ASP67[141]
140Epoxy-linalool oxide (155)−13.13 kcal/mol-GLN101, ASN103GLN81, GLN98, LEU85, HIS195, ASN194, ALA193, AND GLN102[116]
141Withanolide D (198)−8.1 kcal/mol--GLY352, ARG393, TRP349, ALA348, THR347, GLU402, HIS378, HIS401, ASP382, ASP350, PHE40[136]
142Morellic acid (199)−12.1 kcal/mol-THR371, ASP368, LYS363, ASP367CYS361, MET360, ASN149, TRP271, ASP269, THR276, THR445, HIS374, GLU406, GLU375, HIS505, TYR515[120]
143Pyrrolidinone (200)−5.24 kcal/mol-THR434ILE291, PRO415, PHE438[137]
(c) GRP78
144Orientin (181)−7.2 kcal/mol-MET433, LYS435, ARG439, GLU469LYS435, PRO438, PRO471, LYS556[129]
(d) NRP1
145Tuftsin (193)−8.1 kcal/mol-ASN544GLU550, LYS397, PRO398, LEU551[141]
(e) TMPRSS2
146Withanolide D (198)−9.7 kcal/mol (TMPRSS2)-ARG257, ARG241, TYR238GLU253, THR396, TYR250, ASN249, THR246, GLU243, MET239, TYR250[136]
RNA-dependent RNA polymerase
147α-hederin (201)−8.6 kcal/mol-ASP760, CYS622, ARG553THR556, ALA688, LYS500, ASP623, SER682[142]
148Kaempferol (23)−7.0 kcal/mol-MET87ASN414, ASN416, GLN18, GLN19, SER15, ASP846, LYS411, PRQ412, TYR546[90]
149Ellagitannin punicalin (202)−9.5 kcal/mol-ASN497, GLY590, ASP684, TYR689ILE494, LYS577, ASP684, ALA685[79]
150Cyanidin (203)−7.7 kcal/mol-ASP761, TRP617, TRP800, GLU811CYS622, TYR619, ASP760, ASP618, ALA762, GLY616, PHE812[77]
151Amphimedoside C (78)-−47.9 kJ/molASP623, LYS621, ALA554, ASP452, ARG553ARG553[87]
15214-debromoaraplysilin I (204)-−111.52 kJ/molALA762, ASP623, ARG555, ARG553TRP617, CYS622, PHE694, VAL764, GLY616, VAL763, ASN695, ASP618, TYR619, ASP760[88]
153Glaucogenin D (81)−7.3 kcal/mol-TRP619, ASP760-[89]
154Glaucogenin C (205)−7.3 kcal/mol-GLU811, ASP760LYS798[89]
155Crocin (63)−10.5 kcal/mol-ASN497, ASP499, LYS500, LYS545, ILE548, ARG836, ASP845, ARG858[73]
156Jezonofol (167)−10.5 kcal/mol−30.7 kcal/molARG836, A19, A11,
U9, LEU544, ASP845, U18, U17
ALA574, ILE548,
ILE847, TYE546,
LYS545, VAL557,
U20
[91]
157Ellagic acid (206)−6.4 kcal/mol-GLY808, THR817, PRO809, HIS816, TYR831HIS810, LYS807, GLU802,[121]
158Vobtusine lactone (157)−8.7 kcal/mol-SER682, ALA688THR687, ASP684, ASN497, ALA685, TYR689, ARG569, GLN573, ILE494, LEU576, LYS577, ALA580, SER759, GLY590[119]
159Echoside A (168)−7.3 kcal/mol-ASN781, SER709, LYS47LYS780, GLY774, ALA706[122]
160Echoside B (169)−8.0 kcal/mol-LYS47, SER709, LYS714, THR710LYS 780[122]
161Arboreol (153)−8.9 kcal/mol---[114]
Nucleocapsid
1624,8-dihydroxysesamin (152)−10.7 kcal/mol-PRO163, THR166, LEU162, GLY70GLU137, GLY165, GLN164, THR167, THR77, ASN76, GLN84, SER79, PRO81, PRO163, THR136[114]
163Arboreol (153)−10.6 kcal/mol-GLN84, PRO163, GLY70ILE75, GLN164, GLY165, THR166, GLU137, LEU162, PRO81, PRO163, THR136[114]
164Glycyrrhizin (156)−7.9 kcal/mol−30.05 kcal/molTHR92, ARG94, ARG89, TYR110, ARG150ARG90, ALA91, ASN49, THR50, ALS51, PHE54, THR55, TYR112, PRO118, PRO152, ALA157[117]
165Rutin (29) −34.342 kcal/mol--[118]
166Trigoneoside IB (195)−7.6 kcal/mol-ARG71, ARG70, THR60, SER29, ALA28, THR27, GLY125, ILE124, TYR90-[134]
167Apigenin (120)−8.11 kcal/mol-PHE67, PRO68, ARG69, GLY70, GLN71, TYR124, TRP133, VAL134, THR136, GLY138, ALA139[143]
168Curcumin (86)−8.75 kcal/mol-LEU161, GLN163, ALA173[144]
Endoribonuclease
169Hinokiflavone (175)−8.6 kcal/mol-MET243MET243, TYR262, GLU258, HIS362, ALA256[79]
170Rutin (29)−8.68 kcal/mol-LEU246, GLY248, GLN245, CYS291, THR341, HIS250LYS345, LEU346, TYR343[85]
171Glucopyranoside derivative (211)-−263.640 kJ/molGLU340, GLN245, HIS235, LYS290, GLY248, VAL292, SER294TRP333 MET331 LYS345, PRO344, VAL318, TYR343, GLN347, GLY247, HIS243, ASP240, LEU246, THR341, HIS250, CYS293[145]
1724-((2S,3R)-3-(hydroxymethyl)-5-((E)-3-hydroxyprop-1-en-1-yl)-7-methoxy-2,3-dihydrobenzofuran-2-yl)-2-methoxyphenol (212)−6.26 kcal/mol−87.52 kcal/mol--[146]
Helicase
173Crocin (63)−9.5 kcal/mol-ALA18, ILE20, CYS112, ASP113, TRP114, THR141, PHE145, GLY415, HIS482, ASP483, VAL484, SER485, TYR515, THR552, HIS554[73]
174Rhusflavanone (213)−9.2 kcal/mol-GLU341, ASP534ALA312, ALA313, VAL340[79]
175Morelloflavone (214)−9.2 kcal/mol-LYS288, ALA316, ARG443THR286, ALA316, LYS320, GLY538, SER539[79]
176Chromone (215)−6.24 kcal/mol−90.99 kcal/mol--[146]
177Chromone (216)−6.24 kcal/mol−90.99 kcal/mol--[146]
178Scirpusin A (217)−8.9 kcal/mol−41.9 kcal/molPRO514, ASN516ARG560, SP534,
ASN177, GLU201,
LYS202
[91]
Methyl transferase (NSP16–NSP10 complex)
179Robustaflavone (176)−10.6 kcal/mol (NSP16)
 
−7.7 kcal/mol (NSP10)
-
 
-
ASP6897, ASP6928
 
ASP4335
CYS6913, CYS6914, MET6929, ASP6931, PHE6947, GLY6869,
LEU6898
 
ARG4331, ILE4334, LYS4346
[79]
 
[79]
180Michellamine B (218)−10.6 kcal/mol (NSP16)-LYS6844, CYS6913, ASP6928, ASP6928, ASN6996ASN6841, ASP6897, GLY6869, MET6929, LEU6898, GLU7001[79]
181Chromone (219)−6.20 kcal/mol−75.24 kcal/mol--[146]
ADP phosphatase/ADPRP
182Crocin (63)−8.2 kcal/mol-ASP22, LYS44, GLY48, ALA154, PHE156[73]
Exoribonuclease
183Chromone (220)−7.09 kcal/mol−81.16 kcal/mol--[146]
Other Nonstructural proteins
184Hispaglabridin B (221)−8.5 kcal/mol−42.88 kcal/molAASP61, GLY62, THR63, LEU45, ARG11, LEU10ALA55, ARG100, VAL103, ILE66, VAL42[147]
185Licoflavone B (222)−8.1 kcal/mol−42.76 kcal/molMET13, GLY62GLY64, ILE66, LYS93, LEU95, ALA31, ARG11[147]
186Ochnaflavone (223)−9.1 kcal/mol−41.43 kcal/molGLY94, ARG40, ASP301THR68, ILE92, GLY39, GLY38, PHE57, LYS59, MET13, LYS93, LEU95, SER60, PHE41, VAL42, ILE66, ARG40, GLU331[147]
1872,3-dehydrosomnifericin (224)−12.3 kcal/mol-LEU126PHE132, ILE131, VAL49[148]
Envelope proteins
188Trigoneoside IB (195)−7.5 kcal/mol-ILE124, TYR90, ARG70, THR60, SER29, ALA28, THR27, GLY125-[134]
189Dicumarol (225)−7.4 kcal/mol---[114]
* The catalytic residues are shown in blue, and the binding site residues are shown in black.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rafiq, A.; Jabeen, T.; Aslam, S.; Ahmad, M.; Ashfaq, U.A.; Mohsin, N.u.A.; Zaki, M.E.A.; Al-Hussain, S.A. A Comprehensive Update of Various Attempts by Medicinal Chemists to Combat COVID-19 through Natural Products. Molecules 2023, 28, 4860. https://doi.org/10.3390/molecules28124860

AMA Style

Rafiq A, Jabeen T, Aslam S, Ahmad M, Ashfaq UA, Mohsin NuA, Zaki MEA, Al-Hussain SA. A Comprehensive Update of Various Attempts by Medicinal Chemists to Combat COVID-19 through Natural Products. Molecules. 2023; 28(12):4860. https://doi.org/10.3390/molecules28124860

Chicago/Turabian Style

Rafiq, Ayesha, Tooba Jabeen, Sana Aslam, Matloob Ahmad, Usman Ali Ashfaq, Noor ul Amin Mohsin, Magdi E. A. Zaki, and Sami A. Al-Hussain. 2023. "A Comprehensive Update of Various Attempts by Medicinal Chemists to Combat COVID-19 through Natural Products" Molecules 28, no. 12: 4860. https://doi.org/10.3390/molecules28124860

Article Metrics

Back to TopTop