Next Article in Journal
Can Optimizing the Mechanical Environment Deliver a Clinically Significant Reduction in Fracture Healing Time?
Next Article in Special Issue
Triterpenoid–PEG Ribbons Targeting Selectivity in Pharmacological Effects
Previous Article in Journal
Immunotherapeutic Potential of m6A-Modifiers and MicroRNAs in Controlling Acute Myeloid Leukaemia
Previous Article in Special Issue
20-Hydroxyecdysone, from Plant Extracts to Clinical Use: Therapeutic Potential for the Treatment of Neuromuscular, Cardio-Metabolic and Respiratory Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Products, Alone or in Combination with FDA-Approved Drugs, to Treat COVID-19 and Lung Cancer

1
School of Physics and Physical Engineering, Qufu Normal University, Qufu 273165, China
2
Key Laboratory of Green Natural Products and Pharmaceutical Intermediates in Colleges and Universities of Shandong Province, School of Chemistry and Chemical Engineering, Qufu Normal University, Qufu 273165, China
3
Key Laboratory of Life-Organic Analysis of Shandong Province, School of Chemistry and Chemical Engineering, Qufu Normal University, Qufu 273165, China
4
School of Pharmaceutical Sciences, Tsinghua University, Beijing 100084, China
*
Author to whom correspondence should be addressed.
Biomedicines 2021, 9(6), 689; https://doi.org/10.3390/biomedicines9060689
Submission received: 18 May 2021 / Revised: 10 June 2021 / Accepted: 15 June 2021 / Published: 18 June 2021

Abstract

:
As a public health emergency of international concern, the highly contagious coronavirus disease 2019 (COVID-19) pandemic has been identified as a severe threat to the lives of billions of individuals. Lung cancer, a malignant tumor with the highest mortality rate, has brought significant challenges to both human health and economic development. Natural products may play a pivotal role in treating lung diseases. We reviewed published studies relating to natural products, used alone or in combination with US Food and Drug Administration-approved drugs, active against severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) and lung cancer from 1 January 2020 to 31 May 2021. A wide range of natural products can be considered promising anti-COVID-19 or anti-lung cancer agents have gained widespread attention, including natural products as monotherapy for the treatment of SARS-CoV-2 (ginkgolic acid, shiraiachrome A, resveratrol, and baicalein) or lung cancer (daurisoline, graveospene A, deguelin, and erianin) or in combination with FDA-approved anti-SARS-CoV-2 agents (cepharanthine plus nelfinavir, linoleic acid plus remdesivir) and anti-lung cancer agents (curcumin and cisplatin, celastrol and gefitinib). Natural products have demonstrated potential value and with the assistance of nanotechnology, combination drug therapies, and the codrug strategy, this “natural remedy” could serve as a starting point for further drug development in treating these lung diseases.

1. Introduction

As a traditional source for modern pharmaceutical discovery and potential drug leads, natural products have played an integral role in treating patients due to their unique structural, chemical, and biological diversity [1,2,3]. The current race to identify efficacious drugs, natural products with promising therapeutic effects has attracted significant attention, especially for the prevention and treatment of lung diseases, such as pulmonary fibrosis [4], asthma [5], acute lung injury [6], chronic obstructive pulmonary disease [7], defective pulmonary innate immunity [8], coronavirus disease 2019 (COVID-19) [9], and lung cancer [10]. Among the myriad of known lung maladies, COVID-19 and lung cancer are currently the most important public health concerns and burdens worldwide [11,12].
The highly contagious COVID-19 pandemic, caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), has spread quickly across all continents [13,14]. Presently, this global pandemic has posed a significant threat to the lives of billions of individuals through human-to-human transmission [15,16]. In this scenario, the rapid discovery of efficacious agents against the fast-spreading COVID-19 pandemic is currently a top priority of research across the world [17]. Lung cancer, globally, is a malignant tumor with the highest mortality rate (accounting for 18% of all cancer deaths), and the five-year survival rate is very low (only 10% to 20%) [18]. Non-small cell lung cancer (NSCLC), a subtype of lung cancer with the highest incidence rate (accounting for about 85% of lung cancer [19]), has brought significant threats and challenges to human life and health as well as social and economic development. In this context, more aggressive drug trial protocols investigating anti-lung cancer agents are another top research priority [20].
Significant progress had been made in the understanding of natural products active against COVID-19 and lung cancer. However, there has been no hierarchical review (natural product, monotherapy, or in combination with a US Food and Drug Administration (FDA)-approved drug) covering the use of natural products (including natural product-based nanoparticles) as high-quality therapeutic agents for the treatment of COVID-19 or lung cancer in the literature. To underline systematically the potential importance of natural products, including their biological activity and underlying molecular mechanisms, this review will focus on the current knowledge of potential anti-COVID-19 or anti-lung cancer agents. To explore the therapeutic value of natural products better, we have focused on the current progress in representative chemical components against SARS-CoV-2 and lung cancer based on evidence from promising in vitro studies published from 1 January 2020 to 31 May 2021 by interrogating online databases (such as Google Scholar, ACS Publications, Wiley, MDPI, Web of Science, Science Direct, Springer, PubMed, and X-MOL), rather than taking an exhaustively literature-driven approach. Our purpose is to provide a promising “natural remedy” for the treatment of lung cancer and COVID-19.

2. Natural Products as Monotherapy for the Treatment of SARS-CoV-2

Natural products have demonstrated potential value, which supports this strategy as an indispensable research focus in the fight against the COVID-19 epidemic [21,22]. The chemical structures of the components described in this section are shown in Figure 1. The SARS-CoV-2 main protease (Mpro), also called the 3C-like protease (3CLpro), has a vital function in viral replication and is, therefore, a preferred drug target [23]. The papain-like protease (PLpro), another prime therapeutic target, plays an essential role in maturing viral RNA polyproteins and dysregulation of host inflammation [24]. Ginkgolic acid, a phenolic acid, is an essential component of the traditional herbal medicine Ginkgo biloba (EGb) [25]. A study has demonstrated that ginkgolic acid is characterized by half-maximal inhibitory concentration (IC50) values of 1.79 μM and 16.3 μM against SARS-CoV-2 Mpro and SARS-CoV-2 PLpro, respectively [26]. The study unambiguously showed that ginkgolic acid exerts good dual-inhibitory effects through its irreversible binding to SARS-CoV-2 cysteine proteases [26].
Angiotensin-converting enzyme 2 (ACE2), an essential ingredient of the renin–angiotensin–aldosterone system (RAAS), is a critical host cell surface receptor for viral infection [27]. The glycosylated spike protein (S protein) plays an essential role in mediating viral entry via interactions with the ACE2 cell surface receptor [28]. Hypocrellin A and shiraiachrome A, two-axial chiral perylenequinones, have been reported to exhibit potent effects on the infected monkey Vero E6 cell line by inhibiting the activity of the SARS-CoV-2 S protein at EC50 values of 0.22 μM and 0.21 μM, respectively, while at doses of up to 10 μM, these presented no observable cytotoxicity against these cells [29].
Transmembrane protease serine 2 (TMPRSS2), a critical factor enabling SARS-CoV-2 infection, can interact with ACE2 [30]. It has been reported that platycodin D, a triterpenoid saponin isolated from Platycodon grandiflorum, prevents TMPRSS2-driven infection in vitro by impairing membrane fusion [31]. Platycodin D has IC50 values of 0.69 μM and 0.72 μM for SARS-CoV-2 pseudovirus (pSARS-CoV-2) overexpression of ACE2 (ACE2+) and ACE2/TMPRSS2+, respectively, and IC50 values of 1.19 μM and 4.76 μM for SARS-CoV-2 in TMPRSS2-negative Vero cells and TMPRSS2-positive Calu-3 cells, respectively [31]. Resveratrol, a remarkable phytoalexin, may effectively inhibit the replication of SARS-CoV-2 S protein in Vero E6 cells at an EC50 of 4.48 μM [32], and has an excellent safety tracking record, with no cytotoxicity even up to a concentration of 150 µM [33].
The RNA-dependent RNA polymerase (RdRp) of SARS-CoV-2 is another promising target that regulates the replication of the viral genome [34]. Corilagin, a non-nucleoside inhibitor, is a gallotannin isolated from the medicinal plant Phmllanthi Fructus [35]. Corilagin has been reported to inhibit SARS-CoV-2 infection with an EC50 value of 0.13 μM in a concentration-dependent manner by preventing the conformational change of RdRp and inhibits SARS-CoV-2 replication [36]. Furthermore, corilagin, as identified via molecular dynamics simulation-guided studies, could also be used as an endogenous Mpro candidate, with an 88% anti-SARS-CoV-2 Mpro activity at concentrations of 20 μM in vitro [37].
Bafilomycin B2, which can be isolated from Streptomyces sp. HTL16, indicates enhanced inhibitory potency against SARS-CoV-2 at IC50 values of 5.11 nM (in the full-time approach) and 8.32 nM (in the pretreatment-of-virus approach) in Vero E6 cells, respectively [38]. While bafilomycin B2 has demonstrated potential effectiveness in inhibiting the viral entry process, evidence of its utility as anti-SARS-CoV-2 agents in vivo is currently insufficient.
The above evidence supports the potential value of the above natural products as therapeutic agents for the treatment of the novel SARS-CoV-2 infection, suggesting more validation studies (both in vitro and in animal models as well as on humans) could be encouraged to perform. Besides the above-mentioned molecules, several other natural products have also been shown to exhibit potent anti-SARS-CoV-2 activities in vitro. Table 1 summarizes a range of studies investigating the in vitro effects of anti-SARS-CoV-2 agents since 2020.
Traditional Chinese medicines have attracted considerable attention due to their ability to effectively inhibit SARS-CoV-2 [63,64,65]. For example, the Qingfei Paidu decoction (QFPD) has shown an ability to treat COVID-19 patients at all stages with excellent clinical efficacy (cure rate >90%) [66,67]. Shuanghuanglian oral liquid or injection (SHL), another well-known traditional Chinese medicine, dose-dependently inhibits SARS-CoV-2 Mpro replication [68]. In addition to the above-mentioned QFPD and SHL, several other traditional Chinese medicines (such as Kegan Liyan oral liquid and Toujie Quwen granule) listed in Table 2 contain Scutellaria baicalensis Georgi (Chinese name: Huangqin), whose major component is baicalein, exerts a marked anti-SARS-CoV-2 effect (IC50 of 0.94 μM, and SI > 212) [69]. Furthermore, it is crucial to investigate how herbal medicine affects SARS-CoV-2 infection by studying its active ingredients. To elucidate the underlying molecular mechanisms, a crystal structure of SARS-CoV-2 Mpro complexed with baicalein was constructed at a resolution of 2.2 Å (the Protein Data Bank (PDB) ID: 6M2N) [68]. Analysis of the core of the substrate-binding pocket revealed multiple interactions (such as hydrogen bonding with Leu141/Gly143 and Ser144/His163, π–π interactions with Cys145 and His4, and hydrophobic interactions with Met49 and His41), which effectively blocked SARS-CoV-2 replication via noncovalent incorporation [68]. The relevant studies [70,71,72] provided direct data for a better understanding of the molecular mechanisms of Chinese herbal medicine by studying its active ingredients.

3. Natural Products as Monotherapy for the Treatment of Lung Cancer

There is no doubt that natural products have always been recognized as promising anti-lung cancer agents. Daurisoline, an autophagy blocker, is a bisbenzylisoquinoline alkaloid extracted from the herbal medicine Nelumbo nucifera Gaertn [74]. The chemical structures of the molecules discussed in this section are shown in Figure 2. Daurisoline increases the degradation of β-catenin by targeting heat shock protein 90 (HSP90) directly and decreases the expression of MYC proto-oncogene (c-MYC) and cyclin D1, which resulted in cell cycle arrest at the G1 phase in human lung cancer A549 cells and Hop62 cells lines to exert its anti-lung cancer activity [75]. More importantly, in animals, daurisoline has been reported to be a promising anti-lung cancer agent (by inhibiting tumor growth in lung cancer xenografts) with no observable side effects, thus highlighting a potential role for daurisoline in the treatment of lung cancer [75]. Another recent study has shown that daurisoline can effectively inhibit SARS-CoV-2 replication at IC50 values of 3.664 μM and 0.875 μM in Vero E6 cells and in human pulmonary alveolar epithelial cells (HPAEpiC), respectively [49].
Graveospene A, isolated from the leaves of Casearia graveolens, is a new clerodane diterpenoid that has been reported to induce apoptosis in A549 cells with an IC50 value of 1.9 μM by inducing cell cycle arrest in phase G0/G1 [76]. Deguelin, a protein kinase B (AKT) kinase inhibitor, is isolated from the African plant Mundulea sericea (Leguminosae) and is commonly used to inhibit the growth of several types of human cancer cell lines [77]. Deguelin promoted the phosphorylation of myeloid cell leukemia sequence-1 (Mcl-1) protein and induced the inhibition of the wildtype and mutated epidermal growth factor receptor (EGFR)-Akt signaling pathway, which resulted in activation of downstream GSK3β/FBW7 and profound anti-NSCLC activity with no obvious side effects in vivo [78].
Licochalcone A is a natural flavonoid derived from Xinjiang licorice and Glycyrrhiza inflata. Licochalcone A is known to possess a broad spectrum of activities with important pharmacological effects in various cancer cell lines [79]. Licochalcone A can significantly increase autophagic cytotoxicity (in both A549 and H460 cell lines) and downregulated the expression of c-IAP1, c-IAP2, XIAP, survivin, c-FLIPL, and RIP1, apoptosis-related proteins via inhibiting the activity of phosphorylated extracellular signal-regulated kinase (ERK) and autophagy [80]. In addition, licochalcone A has been reported to abolish the expression of programmed death ligand-1 (PD-L1) by increasing reactive oxygen species (ROS) levels in a time-dependent manner and interfering with protein translation in cancer cells [81]. Further, licochalcone A can inhibit PD-L1 translation likely through the inhibition of the phosphorylation of 4EBP1 and activation of the PERK-eIF2α signaling pathway [81]. Licochalcone A plays a vital role in reversing the ectopic expression of key microRNA (miR-144-3p, miR-20a-5p, miR-29c-3p, let-7d-3p, and miR-328-3p) to elicit lung cancer chemopreventive activities both in vivo and in vitro [82]. In addition, licochalcone A has been reported to inhibit EGFR signaling and reduced the expression of Survivin protein in a cap-dependent translation manner to exhibit profound activity in mutated NSCLC cells [83].
Erianin, a novel dibenzyl compound, can be isolated from the traditional herbal medicine Dendrobium chrysotoxum Lindl and has been proposed as an apoptosis-inducing agent in human lung cancer cells [84]. The main mechanisms of its anti-lung cancer activity involve the induction of ferroptosis by activating Ca2+/calmodulin signaling, inhibition of cell proliferation and metastasis, and induction of cell cycle arrest in phase G2/M [85].
Tutuilamide A, isolated from marine cyanobacteria Schizothrix sp., is a novel cyclic peptide reported to exhibit moderate cytotoxicity activity in the H-460 human lung cancer cell line with an IC50 value of 0.53 μM [86]. Tutuilamide A, with the help of the vinyl chloride side chain, showed enhanced inhibitory potency with high selectivity (IC50 0.73 nM) for human neutrophil elastase, which is associated mainly with the migration and metastasis of lung cancer cells [87]. Besides the above-mentioned molecules, Table 3 also exhibits other natural products (including their underlying molecular mechanisms) with notable anti-lung cancer activities reported since 2020.

4. Natural Products in Combination with the FDA-Approved Drugs Inhibit SARS-CoV-2

The bisbenzylisoquinoline alkaloid cepharanthine can be isolated from the traditional herbal medicine Stephania cephalantha Hayata [132]. Cepharanthine exhibits a range of promising bioactivity. It has IC50 values of 0.026 μM, 9.5 μg/mL, and 0.83 μM against the human immunodeficiency virus type 1 (HIV-1) [133], SARS-CoV [134], and human coronavirus OC43 (HCoV-OC43) [135], respectively. This alkaloid inhibits SARS-CoV-2 entry in vitro at an IC50 of 0.35 μM without any evident toxicity profile (selectivity index, [SI] > 70) [136]. Furthermore, the cell death cascade induced by the cellular stress response is another key target for SARS-CoV-2 [137]. It is worth noting that this bisbenzylisoquinoline alkaloid, with a good safety profile, is an approved drug in Japan since the 1950s and is used to treat acute and chronic diseases [132], highlighting that cepharanthine can serve as a potential therapeutic candidate for the treatment of SARS-CoV-2 infection.
Nelfinavir (Viracept), the first HIV-1 protease inhibitor developed by Agouron Pharmaceuticals, was approved by the FDA in March 1997 for the treatment of HIV-AIDS [138]. Recently, nelfinavir was shown to be effective at inhibiting SARS-CoV-2 Mpro infection (IC50 = 3.3 μM) with a low level of toxicity (SI = 3.7) [139]. In addition, nelfinavir inhibited SARS-CoV-2 replication in vitro with an EC50 of 1.13 μM [140]. Nelfinavir was also effective at dose-dependently inhibiting SARS-CoV-2 S protein—complete inhibition at the concentration of 10 μM—with no evidence of cellular cytotoxicity [141]. Remarkably, nelfinavir can also improve lung pathology caused by SARS-CoV-2 infection [142]. Nonetheless, nelfinavir might not benefit SARS-CoV-2-infected patients by reducing viral loads in the lungs, just as it does not reduce viral load in hamsters [142].
Taken together, numerous studies have demonstrated the in vitro anti-SARS-CoV-2 activity of cepharanthine (via inhibition of SARS-CoV-2 S protein) and nelfinavir (via inhibition SARS-CoV-2 Mpro and partly S protein). To reveal the synergistic efficacy (Figure 3) of the above two molecules in SARS-CoV-2 infected patients, based on models of pharmacokinetics, pharmacodynamics, and viral-dynamics, Ohashi et al. constructed a mathematical prediction model of the therapeutic effects and revealed that the combination of cepharanthine (intravenous) and nelfinavir (oral) showed excellent synergistic effects in COVID-19 patients (with viral clearance occurring 1.23 days earlier than with nelfinavir alone; cepharanthine alone showed a minimal effect) [136]. Considering all these factors, including the critical value of cepharanthine and nelfinavir in anti-SARS-CoV-2 infection, both in vitro and in animal models and mathematical prediction modeling, further research is needed to explore whether these molecules exert synergistically augmented activity for the treatment of SARS-CoV-2 infection in patients. It is worth noting that further research is needed to explore whether they have anti-SARS-CoV-2 activity in vivo.
Remdesivir (GS-5734, Veklury®), an RdRp inhibitor developed by Gilead Science, was the first, and currently the only, anti-SARS-CoV-2 drug approved by the FDA (approval on 22 October 2020) for the treatment of COVID-19 [143,144,145]. Remdesivir exhibits broad-spectrum activity against multiple viral infections in vitro, including SARS-CoV, Middle East respiratory syndrome coronavirus (MERS-CoV), Ebola virus (EBOV), and SARS-CoV-2, with EC50 values of 0.069 μM, 0.090 μM, 0.012 μM, and 0.77 μM, respectively [146,147,148,149]. Furthermore, remdesivir has also been thoroughly explored in animal models. Remdesivir reduced lung viral loads in MERS-CoV-infected rhesus monkeys [150] and transgenic Ces1c/ hDPP4 mice [147], protected Nipah virus-infected African green monkeys [151] and rhesus macaques from SARS-CoV-2 infection [152]. Moreover, since 2016, the efficacy and safety of remdesivir have been clinically investigated for the treatment of EBOV infection [153]. Nonetheless, the FDA-approved remdesivir does not appear highly effective in the fight against the COVID-19 pandemic [154,155,156]. In this scenario, the combination of remdesivir with other small molecules, including natural products and natural-product-inspired potential anti-SARS-CoV-2 agents, may exhibit a synergistic effect, compared to remdesivir alone in COVID-19 patients.
Linoleic acid, an inflammatory response modulator [157] isolated from the traditional meal Vicia faba [158], significantly suppresses MERS-CoV replication [159]. Toelzer et al. hypothesized that the combination of remdesivir and linoleic acid, an essential diunsaturated fatty acid, may be superior for treating COVID-19 patients over remdesivir alone [160]. Indeed, the combination of linoleic acid (50 μM) and remdesivir (20 to 200 nM) exerted a synergistic effect on SARS-CoV-2 replication in human Caco-2 ACE2+ cells in vitro [160].
The synergistic mechanisms involved in the combination of linoleic acid and remdesivir shown in Figure 4. To clarify the underlying inhibitory mechanisms of action of linoleic acid, a cryo-electron microscopy (cryo-EM) model of SARS-CoV-2 S protein complexed with linoleic acid was determined at 2.85 Å resolution (Electron Microscopy Data (EMD) ID: 11145) [160]. Further analysis of the linoleic acid binding pocket within the S protein revealed that the hydrocarbon tail of linoleic acid binds to hydrophobic amino acids. At the same time, the acidic headgroup interacts with a positively charged anchor (Arg408 and Gln409) to lock the S protein irreversibly. The hydrophobic pocket with a tube-like shape of the S protein allows a good fit for linoleic acid, and results in reduced ACE2 interactions, and thus sets the stage for an intervention strategy that targets linoleic acid binding to SARS-CoV-2 S protein [160].
As for remdesivir, it is a phosphoramidate prodrug, which requires conversion from the parent drug into the active triphosphate form (GS-443902) [161]. In cells, the triphosphate form, GS-443902, can block SARS-CoV-2 replication by evading the “proofreading” activity of viral RNA sequences [162]. In addition, Yin et al. [34] revealed the cryo-EM structure of SARS-CoV-2 RdRp in complex with remdesivir (using its triphosphate metabolite GS-443902) at 2.5 Å resolution (PDB ID: 7BV2) [34]. The cryo-EM structure unambiguously demonstrated that GS-443902 could positioned itself at the center of the catalytic site of the primer RNA, covalently binding to the primer at the 1+ position of the template strand to terminate chain elongation. Three strong H-bonds with active site residues (ribose -OH groups: Asp623, Ser682, and Asn691; sugar 2′-OH: Asn691) were identified [34]. Further research is warranted to establish whether linoleic acid and remdesivir exert synergistic anti-SARS-CoV-2 effects in vivo. At present, a more well-designed combination drug therapy that exhibits better additive or synergistic effects against COVID-19 is a promising strategy. However, for COVID-19, the nanodrug strategy (containing natural products and FDA-approved drugs) remains an open question, and undoubtedly, it has a long way to go.

5. Natural Products in Combination with the FDA-Approved Anti-Lung Cancer Drugs

As regards lung cancer, significant progress has been made in the research of natural product-based nanomedicines [163,164] and combination drug therapies [165,166], which can provide some reference for the related drug discovery and development for COVID-19. In this section, we mainly focused on the nanodrug strategy (containing natural products and FDA-approved drugs) to reveal its unique advantage in the research and development of anti-lung cancer drugs.
Curcumin is one of the main products of the Curcuma longa L. (turmeric) rhizome extract and has been proposed for its antimicrobial, antimutagenic, antiproliferative, and neuroprotective activities [167]. Curcumin is considered an ideal scaffold for lung cancer drug discovery due to its potent antitumor effects against NSCLC [168]. In particular, several crucial molecular pathways involved in the efficacy of curcumin as an anti-lung cancer drug involve the vascular endothelial growth factor (VEGF), EGFR, nuclear factor-κB (NF-κB), and mammalian target of rapamycin (mTOR) pathways [169]. Nonetheless, the biomedical application of curcumin is currently hindered by its poor aqueous solubility and low bioavailability [170]. In contrast, cisplatin, already marketed as the first platinum-based complex approved by the US FDA, has been used therapeutically for a broad range of cancers such as lung, lymphomas, melanoma, head, and neck cancer [171]. Unfortunately, the routine clinical practice of cisplatin is often coupled with severe toxic side effects (such as nephrotoxicity [172], severe hearing loss [173], and cardiotoxicity [174]) and intrinsic or acquired drug resistance [175].
Indeed, an efficacy study in NSCLC cells evidenced improved effects of the drug combination of curcumin and cisplatin [176]. An in vitro study showed that curcumin enhanced cisplatin-induced therapeutic efficacy in lung cancer cell lines A549, H460, and H1299 by regulating the Cu-Sp1-CTR1 regulatory loop. Furthermore, the promotion of active targeting ability with β-cyclodextrin (β-CD)-modified hyaluronic acid (HA) was identified as an effective strategy to address cellular uptake, intracellular trafficking, and therapy performance of the drug delivery systems [177]. Taking all these factors into account, Bai et al. [178] designed and constructed a β-cyclodextrin-modified hyaluronic acid-based pH- and esterase-dual-responsive supramolecular codrug combining curcumin and cisplatin (Figure 5). In detail, the designed guest moiety Cur-Pt was prepared via esterification reactions between curcumin, oxoplatin, and a molecule of succinic acid. The scheduled host moiety β-CD-modified hyaluronic acid (HA-CD) was prepared via amidation of the carboxylate salt sodium hyaluronate with free amine mono-6-deoxy-6-ethylenediamino-β-CD (β-CD-EDA). Eventually, the desired curcumin and cisplatin nanoparticles (HCPNs) were developed through a host–guest inclusion strategy and subsequent self-assembled. Interestingly, in this targeting system, curcumin acted as both the guest molecule and the chemical anticancer drug.
In vitro evaluation revealed that the HCPNs could be internalized by cancer cells. Once inside the cell, curcumin is released under acidic endosomal conditions (pH-responsive), and cisplatin is released via reducing of oxoplatin under higher expressed glutathione (GSH) conditions (esterase-responsive). Moreover, cell-based experiments revealed the effective cellular toxicity (high efficiency, the IC50 value of 5.4 μM in A549 cells) and active targeting ability (low toxicity, with low expression levels in normal LO-2 cells) of this novel drug-delivery system. Given the observed positive synergistic effect in the study, the authors concluded that HCPNs exhibited improved effects, compared with either monotherapy with curcumin or cisplatin [178]. The drug delivery and sustained release behavior of Cur from HCPNs were investigated in vitro at pH 7.4 after 48 h (11% Cur was released) and pH 5.0 after 48 h (79% Cur was released), respectively, proving the better stability than Cur alone [178]. Meanwhile, although the Tian group did not proceed further with their in vivo studies; we suggest additional in vivo studies should be performed to identify the pharmacokinetic or pharmacodynamic profile of the HCPNs and the synergistic activity against lung cancer of this codrug.
The disulfide bond, a promising redox-reactive switch in vivo, plays an essential role in many biological processes [179]. To reduce adverse effects resulting from chemotherapy regimens, the disulfide-based drug design has attracted great enthusiasm in the synthesis of prodrug or codrug, and especially for the preparation of functional nanodrugs due to their high selectivity and biocompatibility [180,181]. The nontoxic nanodrugs are activated by the excess of GSH in the tumor microenvironment, which provides an essential strategy for lung cancer-targeting treatment [182].
Celastrol, a typical pentacyclic triterpenoid, can be extracted from traditional herbal medicines of the Celastraceae family [183]. Celastrol is considered another up-and-coming natural product for lung cancer treatment due to its potent anti-NSCLC activity via its suppression of Axl protein expression [184], initiating tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-mediated apoptotic cell death [185], and suppressing cell invasion [186]. However, the clinical translation and biomedical application of celastrol are hindered due to its low bioavailability and physiological instability [187].
Gefitinib, approved by US FDA, has been used therapeutically as the first-line agent in patients with advanced lung cancer [188]. Unfortunately, the routine clinical practice of gefitinib is often coupled with severe adverse effects, such as pulmonary toxicity [189], respiratory failure, and severe comorbidities [190]. Following a reasonable design, Wu et al. developed a GSH-responsive nanodrug (identified as CEL@G-SS-NIR in Figure 6), which possesses unique therapeutic efficacy for NSCLC in mice models by inhibiting upstream and downstream EGFR signaling pathways [191]. The nanodrug CEL@G-SS-NIR was prepared in two steps: preparation of the prodrug and acquisition of the nanocomplex. As shown in Figure 6, the main molecule G-SS-NIR of the nanodrug CEL@G-SS-NIR was synthesized through a two-step reaction. First, the key intermediate G-SS was synthesized successfully in the presence of gefitinib (G), 2-hydroxyethyl disulfide (-SS-), and tiphosgene via covalent linkage. Next, the near-infrared (NIR-OH) chromophore was bound to the side chain of the G-SS to form the prodrug G-SS-NIR. The amphiphilic G-SS-NIR readily self-assembled into spherical nanomicelles in an aqueous medium (driven by the disulfide bond and the π–π interaction) and was encapsulated concomitantly the hydrophobic serine-threonine protein kinase (Akt) inhibitor celastrol (marked as CEL) to form CEL@G-SS-NIR.
This novel nanodrug CEL@G-SS-NIR possesses a suitable size (average diameter 119 ± 6 nm), outstanding overall drug loading (64.0 ± 1.4 wt.%), and excellent stability in the blood circulation, and has a rapid release rate of the free molecules (gefitinib, celastrol, and NIR-OH) at tumor region due to the breaking of the disulfide bonds in the presence of high levels of GSH [191].
In vitro, the nanodrug CEL@G-SS-NIR formulation could effectively target the tumor region due to its enhanced permeability and retention effect and also allowed fluorescent imaging in vivo, at a predetermined timepoint after tail vein injection, in orthotopic lung tumors [191]. In the treatment protocol, the mice were randomly divided into five groups (five mice per treatment group), and after a single treatment cycle, the CEL@G-SS-NIR group (13.4 mg/kg, intravenously, for 20 days), compared to the control groups, exhibited stronger NSCLC tumor-suppressive effects [191]. As for the response mechanism involved, the entire process can be divided into four steps: (i) CEL@G-SS-NIR accumulates in the lung tumor region, (ii) CEL@G-SS-NIR releases the drug celastrol and the protonated intermediates (and) through the deprotonated glutathione (GS¯) nucleophilic attack of the disulfide bond on G-SS-NIR bonds, (iii) this further induces the synchronous releases of the parent drug gefitinib and the fluorescent dye NIR-OH via an intramolecular cyclization reaction (thiolate anion moiety reacts with the adjacent carbonyl group), and (iv) finally, the synergistic anticancer activity is activated by suppressing the phosphatidylinositol 3-kinase/serine threonine protein kinase (PI3K/Akt) signaling pathway by celastrol and downregulating EGFR signaling pathway by gefitinib. Simultaneously, a fluorescent and multispectral optoacoustic tomography imaging signal is generated by NIR-OH [192]. This study showed that disulfide-based and targeted fluorescent nano-prodrugs for treating NSCLC and tracking drug delivery systems are particularly advantageous.

6. Conclusions and Future Perspectives

COVID-19 and lung cancer, the two most critical lung diseases presenting high mortality rates, have posed a great challenge and a serious threat to human health and economic development. Since 2020, as is well-known, the scientific community has made great efforts and remarkable inroads in developing promising anti-SARS-CoV-2 and anti-lung cancer agents through various approaches. In this scenario, numerous natural products have fueled significant attention and have shown good results as potential therapeutics for the above-mentioned lung diseases. This review highlighted state-of-the-art of important natural products (including their underlying molecular mechanisms), covering studies published between 1 January 2020 and 31 May 2021, in the treatment of the above-mentioned lung diseases. We found that natural products can be applied in vitro as monotherapy for the treatment of SARS-CoV-2 (ginkgolic acid, resveratrol, and baicalein) and lung cancer (graveospene A, deguelin, and erianin), as well as in combination with the FDA-approved drug inhibit SARS-CoV-2 (cepharanthine plus nelfinavir, linoleic acid plus remdesivir) and as codrug formulations with anti-lung cancer activity in vitro (codrug of curcumin and cisplatin). The evidence revealed herein that natural products could serve as a starting point for further drug development both in COVID-19 and lung cancer. It is worth noting, however, that some natural products could be pan-assay interference compounds, which can give false readouts, and close attention should be paid to decrease futile attempts [193,194]. There is currently very little direct data associated with the clinical effect of natural products against SARS-CoV-2 infection. To understand better and explore systematically the activity of natural products, more validation studies, with high-quality evidence (both in vitro and in animal models as well as on humans), are now needed.
To improve the use of natural products, many intensive research efforts (both in vitro and in vivo) are still needed to explore the limitations of these agents, such as poor water solubility, limited oral absorption, low bioavailability, and the poor first-pass effect, which represent the first step to develop promising anti-COVID-19 or anti-lung cancer agents. It is clear that a long way is still ahead for us to realize natural product-based drug discovery and development, as only phase 1–3 clinical trials can ensure that any small molecule inhibitor can be used as a drug.
More aggressive and well-designed combination drug therapies that exhibit better additive or synergistic effects against COVID-19 and lung cancer are a promising strategy. For example, shiraiachrome A exhibits potent effects in Vero E6 cells by inhibiting the activity of the SARS-CoV-2 S protein at EC50 values of 0.21 μM; bafilomycin B2 presents enhanced inhibitory potency against SARS-CoV-2 at IC50 values of 5.11 nM in Vero E6 cells by inhibiting the viral entry process; ginkgolic acid has IC50 values of 1.79 μM and 16.3 μM against SARS-CoV-2 Mpro and SARS-CoV-2 PLpro. Combining the properties of the above-mentioned natural products with FDA-approved drugs (for example, with nelfinavir or remdesivir) could achieve optimal COVID-19 treatment through multitargeted mechanisms of action. In addition, a codrug of a natural product with an FDA-approved drug could achieve a combination booster through multitargeted activity. However, the codrug strategy remains an open question in the treatment of patients with COVID-19. Thus, we suggest researchers pay considerable attention to the development of emerging codrug therapy strategies.
In contrast, precisely fabricated nanodrugs may be a more potent weapon to enhance biocompatibility, minimize toxicity as well as side effects, achieve long-term circulation in the body, as well as sustained release, overcome undesired adverse effects, and expand the modalities of administration (intravenous injection or inhalation). However, for COVID-19, the nanodrug strategy (containing natural products and FDA-approved drugs) remains another open question. Fortunately, significant progress has been made in the research of lung cancer nanomedicines, which can provide some reference for the related drug discovery and development for COVID-19. There is no doubt that there is a long way to go and many difficulties to overcome. Nonetheless, natural products have their advantages. We sincerely hope natural products will be proven a safe and effective “natural remedy” for the treatment of the above-mentioned lung diseases with the assistance of multiple techniques and strategies.

Author Contributions

Conceptualization, Z.W.; investigation, L.Y.; resources, L.Y.; writing—original draft preparation, L.Y.; writing—review and editing, Z.W.; project administration, Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the project of the PhD research start-up funds of Qufu Normal University, China (Grant No. 614901, and 615201). The authors, therefore, gratefully acknowledge the Qufu Normal University for the financial supports.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ACE2angiotensin-converting enzyme 2
ACE2+overexpression of ACE2
AKTprotein kinase B
ATPadenosine triphosphate
β-CDβ-cyclodextrin
3CLpro3C-Like protease
COVID-19coronavirus disease 2019
cryo-EMcryo-electron microscopy
EBOVEbola virus
EC50half-maximal effective concentration
EGFRepidermal growth factor receptor
EMTepithelial-mesenchymal transition
ERKextracellular signal-regulated kinase
FDAUS Food and Drug Administration
GSHglutathione
HAhyaluronic acid
HCoV-229Ehuman coronavirus 229E
HCPNscurcumin and cisplatin nanoparticles
HIV-1human immunodeficiency virus type 1
IC50half-maximal inhibitory concentration
MAPKmitogen-activated protein kinase
Mcl-1myeloid cell leukemia sequence-1
MERS-CoVMiddle East respiratory syndrome coronavirus
MMPsmatrix-degrading metalloproteinases
Mpromain protease
mTORmammalian target of rapamycin
NSCLCnon-small cell lung cancer
PDBProtein Data Bank
PD-L1programmed death ligand-1
PI3Kphosphoinositide 3-kinase
PLpropapain-like protease
pSARS-CoV-2SARS-CoV-2 pseudovirus
QFPDQingfei Paidu decoction
RdRpRNA-dependent RNA polymerase
ROSreactive oxygen species
S proteinspike protein
SARS-CoVsevere acute respiratory syndrome coronavirus
SARS-CoV-2severe acute respiratory syndrome coronavirus 2
SHLShuanghuanglian oral liquid or injection
SIselectivity index
TMPRSS2transmembrane protease serine 2

References

  1. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs over the nearly four decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef] [PubMed]
  2. Atanasov, A.G.; Zotchev, S.B.; Dirsch, V.M.; Supuran, C.T. Natural products in drug discovery: Advances and opportunities. Nat. Rev. Drug Discov. 2021, 20, 200–216. [Google Scholar] [CrossRef]
  3. Porras, G.; Chassagne, F.; Lyles, J.T.; Marquez, L.; Dettweiler, M.; Salam, A.M.; Samarakoon, T.; Shabih, S.; Farrokhi, D.R.; Quave, C.L. Ethnobotany and the role of plant natural products in antibiotic drug discovery. Chem. Rev. 2021, 121, 3495–3560. [Google Scholar] [CrossRef]
  4. Wang, L.; Li, S.; Yao, Y.; Yin, W.; Ye, T. The role of natural products in the prevention and treatment of pulmonary fibrosis: A review. Food Funct. 2021, 12, 990–1007. [Google Scholar] [CrossRef]
  5. Wang, W.; Yao, Q.; Teng, F.; Cui, J.; Dong, J.; Wei, Y. Active ingredients from Chinese medicine plants as therapeutic strategies for asthma: Overview and challenges. Biomed. Pharmacother. 2021, 137, 111383. [Google Scholar] [CrossRef]
  6. He, Y.Q.; Zhou, C.C.; Yu, L.Y.; Wang, L.; Deng, J.L.; Tao, Y.L.; Zhang, F.; Chen, W.S. Natural product derived phytochemicals in managing acute lung injury by multiple mechanisms. Pharmacol. Res. 2021, 163, 105224. [Google Scholar] [CrossRef]
  7. Santoro, A.; Tomino, C.; Prinzi, G.; Cardaci, V.; Fini, M.; Macera, L.; Russo, P.; Maggi, F. Microbiome in chronic obstructive pulmonary disease: Role of natural products against microbial pathogens. Curr. Med. Chem. 2020, 27, 2931–2948. [Google Scholar] [CrossRef] [PubMed]
  8. Belchamber, K.B.R.; Donnelly, L.E. Targeting defective pulmonary innate immunity-A new therapeutic option? Pharmacol. Therapeut. 2020, 209, 107500. [Google Scholar] [CrossRef]
  9. Christy, M.P.; Uekusa, Y.; Gerwick, L.; Gerwick, W.H. Natural products with potential to treat RNA virus pathogens including SARS-CoV-2. J. Nat. Prod. 2021, 84, 161–182. [Google Scholar] [CrossRef] [PubMed]
  10. Wen, T.; Song, L.; Hua, S. Perspectives and controversies regarding the use of natural products for the treatment of lung cancer. Cancer Med. 2021, 10, 2396–2422. [Google Scholar] [CrossRef]
  11. Nalbandian, A.; Sehgal, K.; Gupta, A.; Madhavan, M.V.; McGroder, C.; Stevens, J.S.; Cook, J.R.; Nordvig, A.S.; Shalev, D.; Sehrawat, T.S.; et al. Post-acute COVID-19 syndrome. Nature Med. 2021, 27, 601–615. [Google Scholar] [CrossRef] [PubMed]
  12. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics. CA Cancer J. Clin. 2021, 71, 7–33. [Google Scholar]
  13. Eurosurveillance Editorial Team. Note from the editors: World Health Organization declares novel coronavirus (2019-nCoV) sixth public health emergency of international concern. Euro. Surveil. 2020, 25, 200131e. [Google Scholar]
  14. Wang, C.; Horby, P.W.; Hayden, F.G.; Gao, G.F. A novel coronavirus outbreak of global health concern. Lancet 2020, 395, 470–473. [Google Scholar] [CrossRef] [Green Version]
  15. Chan, J.F.; Yuan, S.; Kok, K.H.; To, K.K.; Chu, H.; Yang, J.; Xing, F.F.; Liu, J.L.; Yip, C.C.; Poon, R.W.; et al. A familial cluster of pneumonia associated with the 2019 novel coronavirus indicating person-to-person transmission: A study of a family cluster. Lancet 2020, 395, 514–523. [Google Scholar] [CrossRef] [Green Version]
  16. Phan, L.T.; Nguyen, T.V.; Luong, Q.C.; Nguyen, T.V.; Nguyen, H.T.; Le, H.Q.; Nguyen, T.T.; Cao, T.M.; Pham, Q.D. Importation and human-to-human transmission of a novel coronavirus in Vietnam. N. Engl. J. Med. 2020, 382, 872–874. [Google Scholar] [CrossRef] [Green Version]
  17. 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] [PubMed]
  18. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  19. Singh, S.S.; Mattheolabakis, G.; Gu, X.; Withers, S.; Dahal, A.; Jois, S. A grafted peptidomimetic for EGFR heterodimerization inhibition: Implications in NSCLC models. Eur. J. Med. Chem. 2021, 216, 113312. [Google Scholar] [CrossRef]
  20. Dawkins, J.B.N.; Webster, R.M. The small-cell lung cancer drug market. Nat. Rev. Drug Discov. 2020, 19, 507–508. [Google Scholar] [CrossRef] [PubMed]
  21. Verma, S.; Twilley, D.; Esmear, T.; Oosthuizen, C.B.; Reid, A.M.; Nel, M.; Lall, N. Anti-SARS-CoV natural products with the potential to inhibit SARS-CoV-2 (COVID-19). Front. Pharmacol. 2020, 11, 561334. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Z.; Yang, L. Turning the tide: Natural products and natural-product-inspired chemicals as potential counters to SARS-CoV-2 infection. Front. Pharmacol. 2020, 11, 1013. [Google Scholar] [CrossRef] [PubMed]
  23. Qiao, J.; Li, Y.S.; Zeng, R.; Liu, F.L.; Luo, R.H.; Huang, C.; Wang, Y.F.; Zhang, J.; Quan, B.; Shen, C.; et al. SARS-CoV-2 Mpro inhibitors with antiviral activity in a transgenic mouse model. Science 2021, 371, 1374–1378. [Google Scholar] [CrossRef]
  24. Shin, D.; Mukherjee, R.; Grewe, D.; Bojkova, D.; Baek, K.; Bhattacharya, A.; Schulz, L.; Widera, M.; Mehdipour, A.R.; Tascher, G.; et al. Papain-like protease regulates SARS-CoV-2 viral spread and innate immunity. Nature 2020, 587, 657–662. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, J.; Feng, Z.; Liu, W.; Wang, Y.; Wang, G.; Yu, W.; Yang, G.; Yang, T.; Wang, Y.; Li, M. Exogenous hormone on episperm development and ginkgolic acid accumulation in Ginkgo biloba L. Ind. Crop. Prod. 2021, 160, 113140. [Google Scholar] [CrossRef]
  26. Chen, Z.; Cui, Q.; Cooper, L.; Zhang, P.; Lee, H.; Chen, Z.; Wang, Y.; Liu, X.; Rong, L.; Du, R. Ginkgolic acid and anacardic acid are specific covalent inhibitors of SARS-CoV-2 cysteine proteases. Cell Biosci. 2021, 11, 45. [Google Scholar] [CrossRef] [PubMed]
  27. Medina-Enríquez, M.M.; Lopez-León, S.; Carlos-Escalante, J.A.; Aponte-Torres, Z.; Cuapio, A.; Wegman-Ostrosky, T. ACE2: The molecular doorway to SARS-CoV-2. Cell Biosci. 2020, 10, 148. [Google Scholar] [CrossRef]
  28. Benton, D.J.; Wrobel, A.G.; Xu, P.; Roustan, C.; Martin, S.R.; Rosenthal, P.B.; Skehel, J.J.; Gamblin, S.J. Receptor binding and priming of the spike protein of SARS-CoV-2 for membrane fusion. Nature 2020, 588, 327–330. [Google Scholar] [CrossRef] [PubMed]
  29. Li, Y.T.; Yang, C.; Wu, Y.; Lv, J.J.; Feng, X.; Tian, X.; Zhou, Z.; Pan, X.; Liu, S.; Tian, L.W. Axial chiral binaphthoquinone and perylenequinones from the stromata of hypocrella bambusae are SARS-CoV-2 entry inhibitors. J. Nat. Prod. 2021, 84, 436–443. [Google Scholar] [CrossRef]
  30. Stopsack, K.H.; Mucci, L.A.; Antonarakis, E.S.; Nelson, P.S.; Kantoff, P.W. TMPRSS2 and COVID-19: Serendipity or opportunity for intervention? Cancer Discov. 2020, 10, 779–782. [Google Scholar] [CrossRef] [Green Version]
  31. Kim, T.Y.; Jeon, S.; Jang, Y.; Gotina, L.; Won, J.; Ju, Y.H.; Kim, S.; Jang, M.W.; Won, W.; Park, M.G.; et al. Platycodin D prevents both lysosome- and TMPRSS2-driven SARS-CoV-2 infection in vitro by hindering membrane fusion. Exp. Mol. Med. 2021. [Google Scholar] [CrossRef]
  32. Yang, M.; Wei, J.; Huang, T.; Lei, L.; Shen, C.; Lai, J.; Yang, M.; Liu, L.; Yang, Y.; Liu, G.; et al. Resveratrol inhibits the replication of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) in cultured Vero cells. Phytother. Res. 2021, 35, 1127–1129. [Google Scholar] [CrossRef]
  33. Ter Ellen, B.M.; Dinesh Kumar, N.; Bouma, E.M.; Troost, B.; van de Pol, D.P.I.; van der Ende-Metselaar, H.H.; Apperloo, L.; van Gosliga, D.; van den Berge, M.; Nawijn, M.C.; et al. Resveratrol and pterostilbene potently inhibit SARS-CoV-2 replication in vitro. bioRxiv 2020. [Google Scholar] [CrossRef]
  34. Yin, W.; Mao, C.; Luan, X.; Shen, D.D.; Shen, Q.; Su, H.; Wang, X.; Zhou, F.; Zhao, W.; Gao, M.; et al. Structural basis for inhibition of the RNA-dependent RNA polymerase from SARS-CoV-2 by remdesivir. Science 2020, 368, 1499–1504. [Google Scholar] [CrossRef] [PubMed]
  35. Wu, C.; Huang, H.; Choi, H.Y.; Ma, Y.; Zhou, T.; Peng, Y.; Pang, K.; Shu, G.; Yang, X. Anti-esophageal cancer effect of corilagin extracted from Phmllanthi fructus via the mitochondrial and endoplasmic reticulum stress pathways. J. Ethnopharmacol. 2021, 269, 113700. [Google Scholar] [CrossRef] [PubMed]
  36. Li, Q.; Yi, D.; Lei, X.; Zhao, J.; Zhang, Y.; Cui, X.; Xiao, X.; Jiao, T.; Dong, X.; Zhao, X.; et al. Corilagin inhibits SARS-CoV-2 replication by targeting viral RNA-dependent RNA polymerase. Acta Pharm. Sin. B 2021. [Google Scholar] [CrossRef]
  37. Loschwitz, J.; Jäckering, A.; Keutmann, M.; Olagunju, M.; Eberle, R.J.; Coronado, M.A.; Olubiyi, O.O.; Strodel, B. Novel inhibitors of the main protease enzyme of SARS-CoV-2 identified via molecular dynamics simulation-guided in vitro assay. Bioorg. Chem. 2021, 111, 104862. [Google Scholar] [CrossRef]
  38. Xie, X.; Lu, S.; Pan, X.; Zou, M.; Li, F.; Lin, H.; Hu, J.; Fan, S.; He, J. Antiviral bafilomycins from a feces-inhabiting Streptomyces sp. J. Nat. Prod. 2021, 84, 537–543. [Google Scholar] [CrossRef]
  39. Abdallah, H.; El-Halawany, A.; Sirwi, A.; El-Araby, A.; Mohamed, G.; Ibrahim, S.; Koshak, A.; Asfour, H.; Awan, Z.; Elfaky, M.A. Repurposing of Some Natural Product Isolates as SARS-COV-2 Main Protease Inhibitors via In Vitro Cell Free and Cell-Based Antiviral Assessments and Molecular Modeling Approaches. Pharmaceuticals 2021, 14, 213. [Google Scholar] [CrossRef] [PubMed]
  40. 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 Aactivity 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]
  41. Hu, Y.; Liu, M.; Qin, H.; Lin, H.; An, X.; Shi, Z.; Song, L.; Yang, X.; Fan, H.; Tong, Y. Artemether, artesunate, arteannuin B, echinatin, licochalcone B and andrographolide effectively inhibit SARS-CoV-2 and related viruses in vitro. Front. Cell. Infect. Microbiol. 2021. [Google Scholar] [CrossRef]
  42. Cao, R.; Hu, H.; Li, Y.; Wang, X.; Xu, M.; Liu, J.; Zhang, H.; Yan, Y.; Zhao, L.; Li, W.; et al. Anti-SARS-CoV-2 potential of artemisinins in vitro. ACS Infect. Dis. 2020, 6, 2524–2531. [Google Scholar] [CrossRef] [PubMed]
  43. Touret, F.; Gilles, M.; Barral, K.; Nougairède, A.; Decroly, E.; de Lamballerie, X.; Coutard, B. In vitro screening of a FDA approved chemical library reveals potential inhibitors of SARS-CoV-2 replication. Sci. Rep. 2020, 10, 13093. [Google Scholar] [CrossRef] [PubMed]
  44. Hong, S.; Seo, S.H.; Woo, S.J.; Kwon, Y.; Song, M.; Ha, N.C. Epigallocatechin gallate inhibits the uridylate-specific endoribonuclease Nsp15 and efficiently neutralizes the SARS-CoV-2 strain. J. Agric. Food Chem. 2021. [Google Scholar] [CrossRef] [PubMed]
  45. Raj, V.; Park, J.G.; Cho, K.H.; Choi, P.; Kim, T.; Ham, J.; Lee, J. Assessment of antiviral potencies of cannabinoids against SARSCoV-2 using computational and in vitro approaches. Int. J. Biol. Macromol. 2021, 168, 474–485. [Google Scholar] [CrossRef]
  46. Nguyen, L.C.; Yang, D.; Nicolaescu, V.; Best, T.; Chen, S.; Friesen, J.B.; Drayman, N.; Mohamed, A.; Dann, C.; Silva, D.; et al. Cannabidiol inhibits SARS-CoV-2 replication and promotes the host innate immune response. bioRxiv 2021. [Google Scholar] [CrossRef]
  47. Zhu, Y.; Xie, D.Y. Docking characterization and in vitro inhibitory activity of flavan-3-ols and dimeric proanthocyanidins against the main protease activity of SARS-CoV-2. Front. Plant Sci. 2020, 11, 601316. [Google Scholar] [CrossRef] [PubMed]
  48. Du, R.; Cooper, L.; Chen, Z.; Lee, H.; Rong, L.; Cui, Q. Discovery of chebulagic acid and punicalagin as novel allosteric inhibitors of SARS-CoV-2 3CLpro. Antivir. Res. 2021, 190, 105075. [Google Scholar] [CrossRef] [PubMed]
  49. Wang, P.; Luo, R.; Zhang, M.; Wang, Y.; Song, T.; Tao, T.; Li, Z.; Jin, L.; Zheng, H.; Chen, W.; et al. A cross-talk between epithelium and endothelium mediates human alveolar-capillary injury during SARS-CoV-2 infection. Cell Death Dis. 2020, 11, 1042. [Google Scholar] [CrossRef]
  50. Du, A.; Zheng, R.; Disoma, C.; Li, S.; Chen, Z.; Li, S.; Liu, P.; Zhou, Y.; Shen, Y.; Liu, S.; et al. Epigallocatechin-3-gallate, an active ingredient of Traditional Chinese Medicines, inhibits the 3CLpro activity of SARS-CoV-2. Int. J. Biol. Macromol. 2021, 176, 1–12. [Google Scholar] [CrossRef]
  51. Choy, K.T.; Wong, A.Y.L.; Kaewpreedee, P.; Sia, S.F.; Chen, D.D.; Hui, K.P.Y.; Chu, D.K.W.; Chan, M.C.W.; Cheung, P.P.H.; Huang, X.; et al. Remdesivir, lopinavir, emetine, and homoharringtonine inhibit SARS-CoV-2 replication in vitro. Antivir. Res. 2020, 178, 104786. [Google Scholar] [CrossRef] [PubMed]
  52. Kumar, R.; Afsar, M.; Khandelwal, N.; Chander, Y.; Riyesh, T.; Dedar, R.K.; Gulati, B.R.; Pal, Y.; Barua, S.; Tripathi, B.N.; et al. Emetine suppresses SARS-CoV-2 replication by inhibiting interaction of viral mRNA with eIF4E. Antivir. Res. 2021, 189, 105056. [Google Scholar] [CrossRef]
  53. Ashhurst, A.; Tang, A.; Fajtova, P.; Yoon, M.; Aggarwal, A.; Stoye, A.; Larance, M.; Beretta, L.; Drelich, A.; Skinner, D.; et al. Potent in vitro anti-SARS-CoV-2 activity by gallinamide A and analogues via inhibition of cathepsin L. bioRxiv 2020. [Google Scholar] [CrossRef]
  54. Tietjen, I.; Cassel, J.; Register, E.T.; Zhou, X.Y.; Messick, T.E.; Keeney, F.; Lu, L.D.; Beattie, K.D.; Rali, T.; Ertl, H.C.J.; et al. The natural stilbenoid (-)-hopeaphenol inhibits cellular entry of SARS-CoV-2 USA-WA1/2020, B.1.1.7 and B.1.351 variants. bioRxiv 2021. [Google Scholar] [CrossRef]
  55. O’Keefe, S.; Roboti, P.; Duah, K.B.; Zong, G.; Schneider, H.; Shi, W.Q.; High, S. Ipomoeassin-F inhibits the in vitro biogenesis of the SARS-CoV-2 spike protein and its host cell membrane receptor. J. Cell Sci. 2021, 134, jcs257758. [Google Scholar] [CrossRef] [PubMed]
  56. Gangadevi, S.; Badavath, V.N.; Thakur, A.; Yin, N.; Jonghe, S.D.; Acevedo, O.; Jochmans, D.; Leyssen, P.; Wang, K.; Neyts, J.; et al. Kobophenol A inhibits binding of host ACE2 receptor with Spike RBD domain of SARS-CoV-2, a lead compound for blocking COVID-19. J. Phys. Chem. Lett. 2021, 12, 1793–1802. [Google Scholar] [CrossRef]
  57. Kuzikov, M.; Costanzi, E.; Reinshagen, J.; Esposito, F.; Vangeel, L.; Wolf, M.; Ellinger, B.; Claussen, C.; Geisslinger, G.; Corona, A.; et al. Identification of inhibitors of SARS-CoV-2 3CL-Pro enzymatic activity using a small molecule in-vitro repurposing screen. ACS Pharmacol. Transl. Sci. 2021. [Google Scholar] [CrossRef]
  58. Xiao, T.; Cui, M.; Zheng, C.; Wang, M.; Sun, R.; Gao, D.; Bao, J.; Ren, S.; Yang, B.; Lin, J.; et al. Myricetin inhibit SARS-CoV-2 viral replication by targeting Mpro and ameliorates pulmonary inflammation. Front. Pharmacol. 2021. [Google Scholar] [CrossRef]
  59. Clementi, N.; Scagnolari, C.; D’Amore, A.; Palombi, F.; Criscuolo, E.; Frasca, F.; Pierangeli, A.; Mancini, N.; Antonelli, G.; Clementi, M.; et al. Naringenin is a powerful inhibitor of SARS-CoV-2 infection in vitro. Pharmacol. Res. 2021, 163, 105255. [Google Scholar] [CrossRef] [PubMed]
  60. Jeon, S.; Ko, M.; Lee, J.; Choi, I.; Byun, S.Y.; Park, S.; Shum, D.; Kim, S. Identification of antiviral drug candidates against SARS-CoV-2 from FDA-approved drugs. Antimicrob. Agents Chemother. 2020, 64, e00819-20. [Google Scholar] [CrossRef]
  61. Jin, Z.; Du, X.; Xu, Y.; Deng, Y.; Liu, M.; Zhao, Y.; Zhang, B.; Li, X.; Zhang, L.; Peng, C.; et al. Structure of Mpro from SARS-CoV-2 and discovery of its inhibitors. Nature 2020, 582, 289–293. [Google Scholar] [CrossRef] [Green Version]
  62. Jang, M.; Park, Y.I.; Cha, Y.E.; Park, R.; Namkoong, S.; Lee, J.I.; Park, J. Tea polyphenols EGCG and theaflavin inhibit the activity of SARS-CoV-2 3CL-protease in vitro. Evid. Based Compl. Alt. Med. 2020, 2020, 5630838. [Google Scholar] [CrossRef] [PubMed]
  63. Choudhry, N.; Zhao, X.; Xu, D.; Zanin, M.; Chen, W.; Yang, Z.; Chen, J. Chinese Therapeutic Strategy for Fighting COVID-19 and Potential Small-Molecule Inhibitors against Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2). J. Med. Chem. 2020, 63, 13205–13227. [Google Scholar] [CrossRef]
  64. Wang, Z.; Yang, L. Chinese herbal medicine: Fighting SARS-CoV-2 infection on all fronts. J. Ethnopharmacol. 2021, 270, 113869. [Google Scholar] [CrossRef]
  65. Huang, K.; Zhang, P.; Zhang, Z.; Youn, J.Y.; Zhang, H.; Cai, H.L. Traditional Chinese Medicine (TCM) in the treatment of viral infections: Efficacies and mechanisms. Pharm. Ther. 2021, 225, 107843. [Google Scholar] [CrossRef] [PubMed]
  66. National Health Commission of the People’s Republic of China. Notice on the Issunance of Guidelines of Diagnosis and Treatment for 2019-nCoV Infected Pneumonia (Version 7). 2020. Available online: http://www.nhc.gov.cn/yzygj/s7653p/202003/46c9294a7dfe4cef80dc7f5912eb1989/files/ce3e6945832a438eaae415350a8ce964.pdf (accessed on 3 March 2020).
  67. Xinhua Net. Academician Xiaolin Tong: The Total Effective Rate of Qingfeipaidu Formula was 97%, none Transfer from Mild to Severe Cases. 2020. Available online: http://www.kunlunce.com/ssjj/fl1/2020-03-18/141570.html (accessed on 18 March 2020).
  68. Su, H.X.; Yao, S.; Zhao, W.F.; Li, M.J.; Liu, J.; Shang, W.J.; Xie, H.; Ke, C.Q.; Hu, H.C.; Gao, M.N.; et al. Anti-SARS-CoV-2 activities in vitro of Shuanghuanglian preparations and bioactive ingredients. Acta Pharmacol. Sin. 2020, 41, 1167–1177. [Google Scholar] [CrossRef] [PubMed]
  69. 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. Chemother. 2021, 36, 497–503. [Google Scholar] [CrossRef]
  70. Pooja, M.; Reddy, G.J.; Hema, K.; Dodoala, S.; Koganti, B. Unravelling high-affinity binding compounds towards transmembrane protease serine 2 enzyme in treating SARS-CoV-2 infection using molecular modelling and docking studies. Eur. J. Pharmacol. 2021, 890, 173688. [Google Scholar]
  71. Song, J.; Zhang, L.; Xu, Y.; Yang, D.; Zhang, L.; Yang, S.; Zhang, W.; Wang, J.; Tian, S.; Yang, S.; et al. The comprehensive study on the therapeutic effects of baicalein for the treatment of COVID-19 in vivo and in vitro. Biochem. Pharmacol. 2021, 183, 114302. [Google Scholar] [CrossRef]
  72. Ibrahim, M.A.A.; Mohamed, E.A.R.; Abdelrahman, A.H.M.; Allemailem, K.S.; Moustafa, M.F.; Shawky, A.M.; Mahzari, A.; Hakami, A.R.; Abdeljawaad, K.A.A.; Atia, M.A.M. Rutin and flavone analogs as prospective SARS-CoV-2 main protease inhibitors: In silico drug discovery study. J. Mol. Graph. Model. 2021, 105, 107904. [Google Scholar] [CrossRef] [PubMed]
  73. 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]
  74. Qian, J.Q. Cardiovascular pharmacological effects of bisbenzylisoquinoline alkaloid derivatives. Acta Pharm. Sin. 2002, 23, 1086–1092. [Google Scholar]
  75. Huang, X.H.; Yan, X.; Zhang, Q.H.; Hong, P.; Zhang, W.X.; Liu, Y.P.; Xu, W.W.; Li, B.; He, Q.Y. Direct targeting of HSP90 with daurisoline destabilizes β-catenin to suppress lung cancer tumorigenesis. Cancer Lett. 2020, 489, 66–78. [Google Scholar] [CrossRef]
  76. Liu, F.; Ma, J.; Shi, Z.; Zhang, Q.; Wang, H.; Li, D.; Song, Z.; Wang, C.; Jin, J.; Xu, J.; et al. Clerodane diterpenoids isolated from the leaves of Casearia graveolens. J. Nat. Prod. 2020, 83, 36–44. [Google Scholar] [CrossRef] [PubMed]
  77. Tuli, H.S.; Mittal, S.; Loka, M.; Aggarwal, V.; Aggarwal, D.; Masurkar, A.; Kaur, G.; Varol, M.; Sak, K.; Kumar, M.; et al. Deguelin targets multiple oncogenic signaling pathways to combat human malignancies. Pharmacol. Res. 2021, 166, 105487. [Google Scholar] [CrossRef]
  78. Gao, F.; Yu, X.; Li, M.; Zhou, L.; Liu, W.; Li, W.; Liu, H. Deguelin suppresses non-small cell lung cancer by inhibiting EGFR signaling and promoting GSK3β/FBW7-mediated Mcl-1 destabilization. Cell Death Dis. 2020, 11, 143. [Google Scholar] [CrossRef] [PubMed]
  79. Wu, C.P.; Lusvarghi, S.; Hsiao, S.H.; Liu, T.C.; Li, Y.Q.; Huang, Y.H.; Hung, T.H.; Ambudka, S.V. Licochalcone A selectively resensitizes ABCG2-overexpressing multidrug-resistant cancer cells to chemotherapeutic drugs. J. Nat. Prod. 2020, 83, 1461–1472. [Google Scholar] [CrossRef]
  80. Luo, W.; Sun, R.; Chen, X.; Li, J.; Jiang, J.; He, Y.; Shi, S.; Wen, H. ERK activation-mediated autophagy induction resists licochalcone A-induced anticancer activities in lung cancer cells in vitro. OncoTargets Ther. 2020, 13, 13437–13450. [Google Scholar] [CrossRef]
  81. Yuan, L.W.; Jiang, X.M.; Xu, Y.L.; Huang, M.Y.; Chen, Y.C.; Yu, W.B.; Su, M.X.; Ye, Z.H.; Chen, X.; Wang, Y.; et al. Licochalcone A inhibits interferon-gamma-induced programmed death-ligand 1 in lung cancer cells. Phytomedicine 2021, 80, 153394. [Google Scholar] [CrossRef]
  82. Li, B.; Zhou, D.; Li, S.; Feng, Y.; Li, X.; Chang, W.; Zhang, J.; Sun, Y.; Qing, D.; Chen, G.; et al. Licochalcone A reverses NNK-induced ectopic miRNA expression to elicit in vitro and in vivo chemopreventive effects. Phytomedicine 2020, 76, 153245. [Google Scholar] [CrossRef]
  83. Gao, F.; Li, M.; Yu, X.; Liu, W.; Zhou, L.; Li, W. Licochalcone A inhibits EGFR signalling and translationally suppresses survivin expression in human cancer cells. J. Cell Mol. Med. 2021, 25, 813–826. [Google Scholar] [CrossRef]
  84. Yang, A.; Li, M.Y.; Zhang, Z.H.; Wang, J.Y.; Xing, Y.; Ri, M.; Jin, C.H.; Xu, G.H.; Piao, L.X.; Jin, H.L.; et al. Erianin regulates programmed cell death ligand 1 expression and enhances cytotoxic T lymphocyte activity. J. Ethnopharmacol. 2021, 273, 113598. [Google Scholar] [CrossRef]
  85. Chen, P.; Wu, Q.; Feng, J.; Yan, L.; Sun, Y.; Liu, S.; Xiang, Y.; Zhang, M.; Pan, T.; Chen, X.; et al. Erianin, a novel dibenzyl compound in Dendrobium extract, inhibits lung cancer cell growth and migration via calcium/calmodulin-dependent ferroptosis. Signal Transduct. Tar. 2020, 5, 51. [Google Scholar] [CrossRef] [PubMed]
  86. Keller, L.; Canuto, K.M.; Liu, C.; Suzuki, B.M.; Almaliti, J.; Sikandar, A.; Naman, C.B.; Glukhov, E.; Luo, D.; Duggan, B.M.; et al. Tutuilamides A-C: Vinyl-chloride-containing cyclodepsipeptides from marine cyanobacteria with potent elastase inhibitory properties. ACS Chem. Biol. 2020, 15, 751–757. [Google Scholar] [CrossRef] [PubMed]
  87. Chen, Q.Y.; Luo, D.; Seabra, G.M.; Luesch, H. Ahp-Cyclodepsipeptides as tunable inhibitors of human neutrophil elastase and kallikrein 7: Total synthesis of tutuilamide A, serine protease selectivity profile and comparison with lyngbyastatin 7. Bioorg. Med. Chem. 2020, 28, 115756. [Google Scholar] [CrossRef] [PubMed]
  88. Hafner, S.; Lang, S.J.; Gaafary, M.E.; Schmiech, M.; Simmet, T.; Syrovets, T. The cardenolide glycoside acovenoside A interferes with epidermal growth factor receptor (EGFR) trafficking in non-small cell lung cancer cells. Front. Pharmacol. 2021, 11, 611657. [Google Scholar] [CrossRef] [PubMed]
  89. Han, A.R.; Lee, S.; Han, S.; Lee, Y.J.; Kim, J.B.; Seo, E.K.; Jung, C.H. Triterpenoids from the leaves of Centella asiatica inhibit ionizing radiation-induced migration and invasion of human lung cancer cells. Evid. Based Compl. Alt. 2020, 2020, 3683460. [Google Scholar] [CrossRef]
  90. Zhang, X.; Ruan, Q.; Zhai, Y.; Lu, D.; Li, C.; Fu, Y.; Zheng, Z.; Song, Y.; Guo, J. Baicalein inhibits non-small-cell lung cancer invasion and metastasis by reducing ezrin tension in inflammation microenvironment. Cancer Sci. 2020, 111, 3802–3812. [Google Scholar] [CrossRef] [PubMed]
  91. Zhang, Z.; Nong, L.; Chen, M.; Gu, X.; Zhao, W.; Liu, M.; Cheng, W. Baicalein suppresses vasculogenic mimicry through inhibiting RhoA/ROCK expression in lung cancer A549 cell line. Acta Bioch. Bioph. Sin. 2020, 52, 1007–1015. [Google Scholar] [CrossRef]
  92. Li, J.; Yan, L.; Luo, J.; Tong, L.; Gao, Y.; Feng, W.; Wang, F.; Cui, W.; Li, S.; Sun, Z. Baicalein suppresses growth of non-small cell lung carcinoma by targeting MAP4K3. Biomed. Pharmacother. 2021, 133, 110965. [Google Scholar] [CrossRef]
  93. Sui, X.; Han, X.; Chen, P.; Wu, Q.; Feng, J.; Duan, T.; Chen, X.; Pan, T.; Yan, L.; Jin, T.; et al. Baicalin induces apoptosis and suppresses the cell cycle progression of lung cancer cells through downregulating Akt/mTOR signaling pathway. Front. Mol. Biosci. 2021, 7, 602282. [Google Scholar] [CrossRef] [PubMed]
  94. Yan, Y.; Yao, L.; Sun, H.; Pang, S.; Kong, X.; Zhao, S.; Xu, S. Effects of wogonoside on invasion and migration of lung cancer A549 cells and angiogenesis in xenograft tumors of nude mice. J. Thorac. Dis. 2020, 12, 1552–1560. [Google Scholar] [CrossRef] [PubMed]
  95. Cheng, Z.Y.; Hsiao, Y.T.; Huang, Y.P.; Peng, S.F.; Huang, W.W.; Liu, K.C.; Hsia, T.C.; Way, T.D.; Chung, J.G. Casticin induces DNA damage and affects DNA repair associated protein expression in human lung cancer A549 cells. Molecules 2020, 25, 341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Cui, L.; Yang, G.; Ye, J.; Yao, Y.; Lu, G.; Chen, J.; Fang, L.; Lu, S.; Zhou, J. Dioscin elicits anti-tumour immunity by inhibiting macrophage M2 polarization via JNK and STAT3 pathways in lung cancer. J. Cell. Mol. Med. 2020, 24, 9217–9230. [Google Scholar] [CrossRef]
  97. Chen, A.; Jiang, P.; Zeb, F.; Wu, X.; Xu, C.; Chen, L.; Feng, Q. EGCG regulates CTR1 expression through its pro-oxidative property in non-small-cell lung cancer cells. J. Cell. Physiol. 2020, 235, 7970–7981. [Google Scholar] [CrossRef]
  98. Wei, R.; Wirkus, J.; Yang, Z.; Machuca, J.; Esparza, Y.; Mackenzie, G.G. EGCG sensitizes chemotherapeutic-induced cytotoxicity by targeting the ERK pathway in multiple cancer cell lines. Arch. Biochem. Biophys. 2020, 692, 108546. [Google Scholar] [CrossRef]
  99. Duan, J.; Li, Y.; Gao, H.; Yang, D.; He, X.; Fang, Y.; Zhou, G. Phenolic compound ellagic acid inhibits mitochondrial respiration and tumor growth in lung cancer. Food Funct. 2020, 11, 6332–6339. [Google Scholar] [CrossRef]
  100. Boonjing, S.; Pothongsrisit, S.; Wattanathamsan, O.; Sritularak, B.; Pongrakhananon, V. Erianthridin induces non-small cell lung cancer cell apoptosis through the suppression of extracellular signal-regulated kinase activity. Planta Med. 2021, 87, 283–293. [Google Scholar]
  101. Pothongsrisit, S.; Arunrungvichian, K.; Hayakawa, Y.; Sritularak, B.; Mangmool, S.; Pongrakhananon, V. Erianthridin suppresses non-small-cell lung cancer cell metastasis through inhibition of Akt/mTOR/p70 S6K signaling pathway. Sci. Rep. 2021, 11, 6618. [Google Scholar] [CrossRef]
  102. Choudhury, P.; Barua, A.; Roy, A.; Pattanayak, R.; Bhattacharyya, M.; Saha, P. Eugenol emerges as an elixir by targeting β-catenin, the central cancer stem cell regulator in lung carcinogenesis: An in vivo and in vitro rationale. Food Funct. 2021, 12, 1063–1078. [Google Scholar] [CrossRef] [PubMed]
  103. Yu, X.; Gao, F.; Li, W.; Zhou, L.; Liu, W.; Li, M. Formononetin inhibits tumor growth by suppression of EGFR-Akt-Mcl-1 axis in non-small cell lung cancer. J. Exp. Clin. Canc. Res. 2020, 39, 62. [Google Scholar] [CrossRef] [Green Version]
  104. Yang, B.; Yang, N.; Chen, Y.; Maomao, Z.; Lian, Y.; Xiong, Z.; Wang, B.; Feng, L.; Jia, X. An integrated strategy for effective-component discovery of Astragali Radix in the treatment of lung cancer. Front. Pharmacol. 2021, 12, 580978. [Google Scholar] [CrossRef]
  105. Kang, D.Y.; Sp, N.; Jo, E.S.; Rugamba, A.; Hong, D.Y.; Lee, H.G.; Yoo, J.S.; Liu, Q.; Jang, K.J.; Yang, Y.M. The inhibitory mechanisms of tumor PD-L1 expression by natural bioactive gallic acid in non-small-cell lung cancer (NSCLC) cells. Cancers 2020, 12, 727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Wang, D.; Bao, B. Gallic acid impedes non-small cell lung cancer progression via suppression of EGFR-dependent CARM1-PELP1 complex. Drug Des. Dev. Ther. 2020, 14, 1583–1592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Chen, H.; Miao, L.; Huang, F.; Yu, Y.; Peng, Q.; Liu, Y.; Li, X.; Liu, H. Glochidiol, a natural triterpenoid, exerts its anti-cancer effects by targeting the colchicine binding site of tubulin. Invest. N. Drug. 2021, 39, 578–586. [Google Scholar] [CrossRef]
  108. Min, H.Y.; Pei, H.; Hyun, S.Y.; Boo, H.J.; Jang, H.J.; Cho, J.; Kim, J.H.; Son, J.; Lee, H.Y. Potent anticancer effect of the natural steroidal saponin gracillin is produced by inhibiting glycolysis and oxidative phosphorylation-mediated bioenergetics. Cancers 2020, 12, 913. [Google Scholar] [CrossRef] [Green Version]
  109. Yang, J.; Cao, L.; Li, Y.; Liu, H.; Zhang, M.; Ma, H.; Wang, B.; Yuan, X.; Liu, Q. Gracillin isolated from Reineckia carnea induces apoptosis of A549 Cells via the mitochondrial pathway. Drug Des. Dev. Ther. 2021, 2021, 233–243. [Google Scholar] [CrossRef]
  110. Lv, L.; Zhang, W.; Li, T.; Jiang, L.; Lu, X.; Lin, J. Hispidulin exhibits potent anticancer activity in vitro and in vivo through activating ER stress in non-small-cell lung cancer cells. Oncol. Rep. 2020, 43, 1995–2003. [Google Scholar] [CrossRef] [Green Version]
  111. Zhao, X.; Lin, Y.; Jiang, B.; Yin, J.; Lu, C.; Wang, J.; Zeng, J. Icaritin inhibits lung cancer-induced osteoclastogenesis by suppressing the expression of IL-6 and TNF-a and through AMPK/mTOR signaling pathway. Anti-Cancer Drug. 2020, 31, 1004–1011. [Google Scholar] [CrossRef] [PubMed]
  112. Lee, J.H.; Park, S.Y.; Hwang, W.; Sung, J.Y.; Shim, M.L.C.J.; Kim, Y.N.; Yoon, K. Isoharringtonine induces apoptosis of non-small cell lung cancer cells in tumorspheroids via the intrinsic pathway. Biomolecules 2020, 10, 1521. [Google Scholar] [CrossRef] [PubMed]
  113. Fouzder, C.; Mukhuty, A.; Kundu, R. Kaempferol inhibits Nrf2 signalling pathway via downregulation of Nrf2 mRNA and induces apoptosis in NSCLC cells. Arch. Biochem. Biophys. 2021, 697, 108700. [Google Scholar] [CrossRef]
  114. Sheng, H.; Lv, W.; Zhu, L.; Wang, L.; Wang, Z.; Han, J.; Hu, J. Liriopesides B induces apoptosis and cell cycle arrest in human non-small cell lung cancer cells. Int. J. Mol. Med. 2020, 46, 1039–1050. [Google Scholar] [CrossRef] [PubMed]
  115. Chen, Y.C.; Huang, M.Y.; Zhang, L.L.; Feng, Z.L.; Jiang, X.M.; Yuan, L.W.; Huang, R.Y.; Liu, B.; Yu, H.; Wang, Y.T.; et al. Nagilactone E increases PD-L1 expression through activation of c-Jun in lung cancer cells. Chin. J. Nat. Med. 2020, 18, 517–525. [Google Scholar] [CrossRef]
  116. Zhang, L.L.; Guo, J.; Jiang, X.M.; Chen, X.P.; Wang, Y.T.; Li, A.; Lin, L.G.; Li, H.; Lu, J.J. Identification of nagilactone E as a protein synthesis inhibitor with anticancer activity. Acta Pharmacol. Sin. 2020, 41, 698–705. [Google Scholar] [CrossRef] [PubMed]
  117. Liu, X.; Zhang, Y.; Zhou, G.J.; Hou, Y.; Kong, Q.; Lu, J.J.; Zhang, Q.; Chen, X. Natural alkaloid 8-oxo-epiberberine inhibited TGF-β1-triggred epithelial-mesenchymal transition by interfering Smad3. Toxicol. Appl. Pharm. 2020, 404, 115179. [Google Scholar] [CrossRef] [PubMed]
  118. Li, X.; Huang, R.; Li, M.; Zhu, Z.; Chen, Z.; Cui, L.; Luo, H.; Luo, L. Parthenolide inhibits the growth of non-small cell lung cancer by targeting epidermal growth factor receptor. Cancer Cell Int. 2020, 20, 561. [Google Scholar] [CrossRef] [PubMed]
  119. Sun, L.; Yuan, W.; Wen, G.; Yu, B.; Xu, F.; Gan, X.; Tang, J.; Zeng, Q.; Zhu, L.; Chen, C.; et al. Parthenolide inhibits human lung cancer cell growth by modulating the IGF-1R/PI3K/Akt signaling pathway. Oncol. Rep. 2020, 44, 1184–1193. [Google Scholar] [CrossRef] [PubMed]
  120. Wu, L.M.; Liao, X.Z.; Zhang, Y.; He, Z.R.; Nie, S.Q.; Ke, B.; Shi, L.; Zhao, J.F.; Chen, W.H. Parthenolide augments the chemosensitivity of non-small-cell lung cancer to cisplatin via the PI3K/AKT signaling pathway. Front. Cell Dev. Biol. 2021, 8, 610097. [Google Scholar] [CrossRef] [PubMed]
  121. Meng, N.; Zhang, R.; Liu, C.; Wang, Q.; Wang, X.; Guo, X.; Wang, P.; Sun, J. PDB-1 from potentilla discolor bunge suppresses lung cancer cell migration and invasion via FAK/Src and MAPK signaling pathways. Med. Chem. Res. 2020, 29, 887–896. [Google Scholar] [CrossRef]
  122. Wu, Y.; Si, Y.; Xiang, Y.; Zhou, T.; Liu, X.; Wu, M.; Li, W.; Zhang, T.; Xiang, K.; Zhang, L.; et al. Polyphyllin I activates AMPK to suppress the growth of non-small-cell lung cancer via induction of autophagy. Arch. Biochem. Biophys. 2020, 687, 108285. [Google Scholar] [CrossRef]
  123. Lai, L.; Shen, Q.; Wang, Y.; Chen, L.; Lai, J.; Wu, Z.; Jiang, H. Polyphyllin I reverses the resistance of osimertinib in non-small cell lung cancer cell through regulation of PI3K/Akt signaling. Toxicol. Appl. Pharm. 2021, 419, 115518. [Google Scholar] [CrossRef] [PubMed]
  124. Guo, H.; Ding, H.; Tang, X.; Liang, M.; Li, S.; Zhang, J.; Cao, J. Quercetin induces pro-apoptotic autophagy via SIRT1/AMPK signaling pathway in human lung cancer cell lines A549 and H1299 in vitro. Thoracic Cancer 2021, 12, 1415–1422. [Google Scholar] [CrossRef] [PubMed]
  125. Xu, S.; Zhang, H.; Wang, A.; Ma, Y.; Gan, Y.; Li, G. Silibinin suppresses epithelial-mesenchymal transition in human non-small cell lung cancer cells by restraining RHBDD1. Cell. Mol. Biol. Lett. 2020, 25, 36. [Google Scholar] [CrossRef]
  126. Shen, K.H.; Hung, J.H.; Liao, Y.C.; Tsai, S.T.; Wu, M.J.; Chen, P.S. Sinomenine inhibits migration and invasion of human lung cancer cell through downregulating expression of miR-21 and MMPs. Int. J. Mol. Sci. 2020, 21, 3080. [Google Scholar] [CrossRef] [PubMed]
  127. Bai, S.; Wen, W.; Hou, X.; Wu, J.; Yi, L.; Zhi, Y.; Lv, Y.; Tan, X.; Liu, L.; Wang, P.; et al. Inhibitory effect of sinomenine on lung cancer cells via negative regulation of α7 nicotinic acetylcholine receptor. J. Leukocyte Biol. 2021, 109, 843–852. [Google Scholar] [CrossRef]
  128. Liu, W.; Yu, X.; Zhou, L.; Li, J.; Li, M.; Li, W.; Gao, F. Sinomenine inhibits non-small cell lung cancer via downregulation of hexokinases II-mediated aerobic glycolysis. OncoTargets Ther. 2020, 13, 3209–3221. [Google Scholar] [CrossRef] [Green Version]
  129. Zheng, W.; Huang, F.Y.; Dai, S.Z.; Wang, J.Y.; Lin, Y.Y.; Sun, Y.; Tan, G.H.; Huang, Y.H. Toxicarioside O inhibits cell proliferation and epithelial-mesenchymal transition by downregulation of Trop2 in lung cancer cells. Front. Oncol. 2020, 10, 609275. [Google Scholar] [CrossRef] [PubMed]
  130. Al-Rashed, S.; Baker, A.; Ahmad, S.S.; Syed, A.; Bahkali, A.H.; Elgorban, A.M.; Khan, M.S. Vincamine, a safe natural alkaloid, represents a novel anticancer agent. Bioorg. Chem. 2021, 107, 104626. [Google Scholar] [CrossRef] [PubMed]
  131. Gao, F.; Li, M.; Zhou, L.; Liu, W.; Zuo, H.; Li, W. Xanthohumol targets the ERK1/2-Fra1 signaling axis to reduce cyclin D1 expression and inhibit non-small cell lung cancer. Oncol. Rep. 2020, 44, 1365–1374. [Google Scholar] [CrossRef] [PubMed]
  132. Bailly, C. Cepharanthine: An update of its mode of action, pharmacological properties and medical applications. Phytomedicine 2019, 62, 152956. [Google Scholar] [CrossRef]
  133. Okamoto, M.; Ono, M.; Baba, M. Potent inhibition of HIV type 1 replication by an antiinflammatory alkaloid, cepharanthine, in chronically infected monocytic cells. AIDS Res. Hum. Retrovir. 1998, 14, 1239–1245. [Google Scholar] [CrossRef]
  134. Zhang, C.H.; Wang, Y.F.; Liu, X.J.; Lu, J.H.; Qian, C.W.; Wan, Z.Y. Antiviral activity of cepharanthine against severe acute respiratory syndrome coronavirus in vitro. Chin. Med. J. 2005, 118, 493–496. [Google Scholar]
  135. Kim, D.E.; Min, J.S.; Jang, M.S.; Lee, J.Y.; Shin, Y.S.; Song, J.H.; Kim, H.R.; Kim Se Jin, Y.H.; Kwon, S. Natural bis-benzylisoquinoline alkaloids-tetrandrine, fangchinoline, and cepharanthine, inhibit human coronavirus OC43 infection of MRC-5 human lung cells. Biomolecules 2019, 9, 696. [Google Scholar] [CrossRef] [Green Version]
  136. Ohashi, H.; Watashi, K.; Saso, W.; Shionoya, K.; Iwanami, S.; Hirokawa, T.; Shirai, T.; Kanaya, S.; Ito, Y.; Kim, K.S.; et al. Potential anti-COVID-19 agents, cepharanthine and nelfinavir, and their usage for combination treatment. Science 2021, 24, 102367. [Google Scholar] [CrossRef]
  137. Li, S.; Liu, W.; Chen, Y.; Wang, L.; An, W.; An, X.; Song, L.; Tong, Y.; Fan, H.; Lu, C. Transcriptome analysis of cepharanthine against a SARS-CoV-2-related coronavirus. Brief. Bioinform. 2021, 22, 1378–1386. [Google Scholar] [CrossRef] [PubMed]
  138. Bihani, S.C.; Hosur, M.V. Molecular basis for reduced cleavage activity and drug resistance in D30N HIV-1 protease. bioRxiv 2021. [Google Scholar] [CrossRef]
  139. Jan, J.T.; Cheng, T.J.R.; Juang, Y.P.; Ma, H.H.; Wu, Y.T.; Yang, W.B.; Cheng, C.W.; Chen, X.; Chou, T.H.; Shie, J.J.; et al. Identification of existing pharmaceuticals and herbal medicines as inhibitors of SARS-CoV-2 infection. Proc. Natl. Acad. Sci. USA 2021, 118, e2021579118. [Google Scholar] [CrossRef] [PubMed]
  140. Yamamoto, N.; Matsuyama, S.; Hoshino, T.; Yamamoto, N. Nelfinavir inhibits replication of severe acute respiratory syndrome coronavirus 2 in vitro. bioRxiv 2020. [Google Scholar] [CrossRef]
  141. Musarrat, F.; Chouljenko, V.; Dahal, A.; Nabi, R.; Chouljenko, T.; Jois, S.D.; Kousoulas, K.G. The anti-HIV drug nelfinavir mesylate (Viracept) is a potent inhibitor of cell fusion caused by the SARSCoV-2 spike (S) glycoprotein warranting further evaluation as an antiviral against COVID-19 infections. J. Med. Virol. 2020, 92, 2087–2095. [Google Scholar] [CrossRef]
  142. Foo, C.S.; Abdelnabi, R.; Kaptein, S.J.F.; Zhang, X.; ter Horst, S.; Mols, R.; Delang, L.; Rocha-Pereira, J.; Coelmont, L.; Leyssen, P.; et al. Nelfinavir markedly improves lung pathology in SARS-CoV-2-infected Syrian hamsters despite lack of an antiviral effect. bioRxiv 2021. [Google Scholar] [CrossRef]
  143. Warren, T.K.; Jordan, R.; Lo, M.K.; Ray, A.S.; Mackman, R.L.; Soloveva, V.; Siegel, D.; Perron, M.; Bannister, R.; Hui, H.C.; et al. Therapeutic efficacy of the small molecule GS-5734 against Ebola virus in rhesus monkeys. Nature 2016, 531, 381–385. [Google Scholar] [CrossRef] [PubMed]
  144. Li, Y.; Cao, L.; Li, G.; Cong, F.; Li, Y.; Sun, J.; Luo, Y.; Chen, G.; Li, G.; Wang, P.; et al. Remdesivir metabolite GS-441524 effectively inhibits SARS-CoV-2 infection in mouse models. J. Med. Chem. 2021. [Google Scholar] [CrossRef]
  145. Rubin, D.; Chan-Tack, K.; Farley, J.; Sherwat, A. FDA approval of remdesivir—A step in the right direction. N. Engl. J. Med. 2020, 383, 2598–2600. [Google Scholar] [CrossRef] [PubMed]
  146. Sheahan, T.P.; Sims, A.C.; Graham, R.L.; Menachery, V.D.; Gralinski, L.E.; Case, J.a.B.; Leist, S.R.; Pyrc, K.; Feng, J.Y.; Trantcheva, I.; et al. Broad-spectrum antiviral GS-5734 inhibits both epidemic and zoonotic coronaviruses. Sci. Transl. Med. 2017, 9, eaal3653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Sheahan, T.P.; Sims, A.C.; Leist, S.R.; Schäfer, A.; Won, J.; Brown, A.J.; Montgomery, S.A.; Hogg, A.; Babusis, D.; Clarke, M.O.; et al. Comparative therapeutic efficacy of remdesivir and combination lopinavir, ritonavir, and interferon beta against MERS-CoV. Nat. Commun. 2020, 11, 222. [Google Scholar] [CrossRef] [Green Version]
  148. McMullan, L.K.; Flint, M.; Chakrabarti, A.; Guerrero, L.; Lo, M.K.; Porter, D.; Nichol, S.T.; Spiropoulou, C.F.; Albariño, C. Characterisation of infectious Ebola virus from the ongoing outbreak to guide response activities in the Democratic Republic of the Congo: A phylogenetic and in vitro analysis. Lancet Infect. Dis. 2019, 19, 1023–1032. [Google Scholar] [CrossRef] [Green Version]
  149. Wang, M.; Cao, R.; Zhang, L.; Yang, X.; Liu, J.; Xu, M.; Shi, Z.; Hu, Z.; Zhong, W.; Xiao, G. Remdesivir and chloroquine effectively inhibit the recently emerged novel coronavirus (2019-nCoV) in vitro. Cell Res. 2020, 30, 269–271. [Google Scholar] [CrossRef] [PubMed]
  150. De Wit, E.; Feldmann, F.; Cronin, J.; Jordan, R.; Okumura, A.; Thomas, T.; Scott, D.; Cihlar, T.; Feldmann, H. Prophylactic and therapeutic remdesivir (GS-5734) treatment in the rhesus macaque model of MERS-CoV infection. Proc. Natl. Acad. Sci. USA 2020, 117, 6771–6776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Lo, M.K.; Feldmann, F.; Gary, J.M.; Jordan, R.; Bannister, R.; Cronin, J.; Patel, N.R.; Klena, J.D.; Nichol, S.T.; Cihlar, T.; et al. Remdesivir (GS-5734) protects African green monkeys from Nipah virus challenge. Sci. Transl. Med. 2019, 11, eaau9242. [Google Scholar] [CrossRef]
  152. Williamson, B.N.; Feldmann, F.; Schwarz, B.; Meade-White, K.; Porter, D.P.; Schulz, J.; van Doremalen, N.; Leighton, I.; Yinda, C.K.; Pérez-Pérez, L.; et al. Clinical benefit of remdesivir in rhesus macaques infected with SARS-CoV-2. Nature 2020, 585, 273–276. [Google Scholar] [CrossRef]
  153. Jacobs, M.; Rodger, A.; Bell, D.J.; Bhagani, S.; Cropley, I.; Filipe, A.; Gifford, R.J.; Hopkins, S.; Hughes, J.; Jabeen, F.; et al. Late Ebola virus relapse causing meningoencephalitis: A case report. Lancet 2016, 388, 498–503. [Google Scholar] [CrossRef] [Green Version]
  154. Schwartz, I.S.; Heil, E.L.; McCreary, E.K. Remdesivir: A pendulum in a pandemic. BMJ 2020, 371, m4560. [Google Scholar] [CrossRef]
  155. Wang, Y.; Zhang, D.; Du, G.; Du, R.; Zhao, J.; Jin, Y.; Fu, S.; Gao, L.; Cheng, Z.; Lu, Q.; et al. Remdesivir in adults with severe COVID-19: A randomised, double-blind, placebo-controlled, multicentre trial. Lancet 2020, 395, 1569–1578. [Google Scholar] [CrossRef]
  156. McCreary, E.K.; Angus, D.C. Efficacy of remdesivir in COVID-19. JAMA 2020, 324, 1041–1042. [Google Scholar] [CrossRef] [PubMed]
  157. Yan, B.; Chu, H.; Yang, D.; Sze, K.H.; Lai, P.M.; Yuan, S.; Shuai, H.; Wang, Y.; Kao, R.Y.T.; Chan, J.F.W.; et al. Characterization of lipidomic profile of human coronavirus -infected cells: Implications for lipid metabolism remodeling upon coronavirus replication. Viruses 2019, 11, 73. [Google Scholar] [CrossRef] [Green Version]
  158. Khalil, M.I.; Salih, M.A.; Mustafa, A.A. Broad beans (Vicia faba) and the potential to protect from COVID-19 coronavirus infection. Sudan J. Paediatr. 2020, 20, 10–12. [Google Scholar] [CrossRef] [PubMed]
  159. Simopoulos, A.P.; Serhan, C.N.; Bazinet, R.P. The need for precision nutrition, genetic variation and resolution in Covid-19 patients. Mol. Aspects Med. 2021, 77, 100943. [Google Scholar] [CrossRef]
  160. Toelzer, C.; Gupta, K.; Yadav, S.K.N.; Borucu, U.; Davidson, A.D.; Williamson, M.K.; Shoemark, D.K.; Garzoni, F.; Staufer, O.; Milligan, R.; et al. Free fatty acid binding pocket in the locked structure of SARS-CoV-2 spike protein. Science 2020, 370, 725–730. [Google Scholar] [CrossRef] [PubMed]
  161. Wang, Z.; Yang, L. GS-5734: A potentially approved drug by FDA against SARS-CoV-2. N. J. Chem. 2020, 44, 12417–12429. [Google Scholar] [CrossRef]
  162. Yan, V.C.; Muller, F.L. Advantages of the parent nucleoside GS-441524 over remdesivir for Covid-19 treatment. ACS Med. Chem. Lett. 2020, 11, 1361–1366. [Google Scholar] [CrossRef] [PubMed]
  163. Yuan, R.; Huang, Y.; Chan, L.; He, D.; Chen, T. Engineering EHD1-targeted natural borneol nanoemulsion potentiates therapeutic efficacy of gefitinib against non-small lung cancer. ACS Appl. Mater. Interfaces 2020, 12, 45714–45727. [Google Scholar] [CrossRef] [PubMed]
  164. Wang, Y.; Yu, H.; Wang, S.; Gai, C.; Cui, X.; Xu, Z.; Li, W.; Zhang, W. Targeted delivery of quercetin by nanoparticles based on chitosan sensitizing paclitaxel-resistant lung cancer cells to paclitaxel. Mater. Sci. Eng. C 2021, 119, 111442. [Google Scholar] [CrossRef]
  165. Zang, H.; Qian, G.; Arbiser, J.; Owonikoko, T.K.; Ramalingam, S.S.; Fan, S.; Sun, S.Y. Overcoming acquired resistance of EGFR-mutant NSCLC cells to the third generation EGFR inhibitor, osimertinib, with the natural product honokiol. Mol. Oncol. 2020, 14, 882–895. [Google Scholar] [CrossRef] [Green Version]
  166. Dai, C.H.; Zhu, L.R.; Wang, Y.; Tang, X.P.; Du, Y.J.; Chen, Y.C.; Li, J. Celastrol acts synergistically with afatinib to suppress non-small cell lung cancer cell proliferation by inducing paraptosis. J. Cell. Physiol. 2021, 236, 4538–4554. [Google Scholar] [CrossRef]
  167. Silva, A.C.D.; Santos, P.D.D.F.; Silva, J.T.D.P.; Leimann, F.V.; Bracht, L.; Gonçalves, O.H. Impact of curcumin nanoformulation on its antimicrobial activity. Trends Food Sci. Tech. 2018, 72, 74–82. [Google Scholar] [CrossRef]
  168. Zhang, L.; Qiang, P.; Yu, J.; Miao, Y.M.; Chen, Z.Q.; Qu, J.; Zhao, Q.B.; Chen, Z.; Liu, Y.; Yao, X.; et al. Identification of compound CA-5f as a novel late-stage autophagy inhibitor with potent anti-tumor effect against non-small cell lung cancer. Autophagy 2019, 15, 391–406. [Google Scholar] [CrossRef] [Green Version]
  169. Ashrafizadeh, M.; Najafi, M.; Makvandi, P.; Zarrabi, A.; Farkhondeh, T.; Samarghandian, S. Versatile role of curcumin and its derivatives in lung cancer therapy. J. Cell Physiol. 2020, 235, 9241–9268. [Google Scholar] [CrossRef] [PubMed]
  170. Ganguly, R.; Kumar, S.; Kunwar, A.; Nath, S.; Sarma, H.D.; Tripathi, A.; Verma, G.; Chaudhari, D.P.; Aswal, V.K.; Melo, J.S. Structural and therapeutic properties of curcumin solubilized pluronic F127 micellar solutions and hydrogels. J. Mol. Liq. 2020, 314, 113591. [Google Scholar] [CrossRef]
  171. Ghosh, S. Cisplatin: The first metal based anticancer drug. Bioorg. Chem. 2019, 88, 102925. [Google Scholar] [CrossRef]
  172. Zha, M.; Tian, T.; Xu, W.; Liu, S.; Jia, J.; Wang, L.; Yan, Q.; Li, N.; Yu, J.; Huang, L. The circadian clock gene Bmal1 facilitates cisplatin-induced renal injury and hepatization. Cell Death Dis. 2020, 11, 446. [Google Scholar] [CrossRef]
  173. Hazlitt, R.A.; Min, J.; Zuo, J. Progress in the development of preventative drugs for cisplatin-induced hearing loss. J. Med. Chem. 2018, 61, 5512–5524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Ma, W.; Wei, S.; Zhang, B.; Li, W. Molecular mechanisms of cardiomyocyte death in drug-induced cardiotoxicity. Front. Cell Dev. Biol. 2020, 8, 434. [Google Scholar] [CrossRef] [PubMed]
  175. Wu, C.; Chen, L.; Tao, H.; Kong, L.; Hu, Y. Ring finger protein 38 induces the drug resistance of cisplatin in non-small cell lung cancer. Cell Biol. Int. 2020. [Google Scholar] [CrossRef] [PubMed]
  176. Zhang, W.; Shi, H.; Chen, C.; Ren, K.; Xu, Y.; Liu, X.; He, L. Curcumin enhances cisplatin sensitivity of human NSCLC cell lines through influencing Cu-Sp1-CTR1 regulatory loop. Phytomedicine 2018, 48, 51–61. [Google Scholar] [CrossRef] [PubMed]
  177. Yang, Y.; Zhang, Y.M.; Chen, Y.; Chen, J.T.; Liu, Y. Targeted polysaccharide nanoparticle for adamplatin prodrug delivery. J. Med. Chem. 2013, 56, 9725–9736. [Google Scholar] [CrossRef]
  178. Bai, Y.; Liu, C.P.; Chen, D.; Liu, C.F.; Zhuo, L.H.; Li, H.; Wang, C.; Bu, H.T.; Tian, W. β-Cyclodextrin-modified hyaluronic acid-based supramolecular self-assemblies for pH- and esterase- dual-responsive drug delivery. Carbohydr. Polym. 2020, 246, 116654. [Google Scholar] [CrossRef] [PubMed]
  179. Gao, J.; Ma, X.; Zhang, L.; Yan, J.; Cui, H.; Zhang, Y.; Wang, D.; Zhang, H. Self-Assembled disulfide bond bearing paclitaxel-camptothecin prodrug nanoparticle for lung cancer therapy. Pharmaceutics 2020, 12, 1169. [Google Scholar] [CrossRef] [PubMed]
  180. Lou, X.; Zhang, D.; Ling, H.; He, Z.; Sun, J.; Sun, M.; Liu, D. Pure redox-sensitive paclitaxel-maleimide prodrug nanoparticles: Endogenous albumin-induced size switching and improved antitumor efficiency. Acta Pharm. Sin. B 2021. [Google Scholar] [CrossRef]
  181. Wang, J.; Pei, Q.; Xia, R.; Liu, S.; Hu, X.; Xie, Z.; Jing, X. Comparison of redox responsiveness and antitumor capability of paclitaxel dimeric nanoparticles with different linkers. Chem. Mater. 2020, 32, 10719–10727. [Google Scholar] [CrossRef]
  182. Lan, J.S.; Liu, L.; Zeng, R.F.; Qin, Y.H.; Hou, J.W.; Xie, S.S.; Yue, S.; Yang, J.; Ho, R.J.Y.; Ding, Y.; et al. Tumor-specific carrier-free nanodrugs with GSH depletion and enhanced ROS generation for endogenous synergistic anti-tumor by a chemotherapy-photodynamic therapy. Chem. Eng. J. 2021, 407, 127212. [Google Scholar] [CrossRef]
  183. Venkatesha, S.H.; Dudics, S.; Astry, B.; Moudgil, K.D. Control of autoimmune inflammation by celastrol, a natural triterpenoid. Pathog. Dis. 2016, 74, ftw059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Lee, Y.J.; Kim, S.Y.; Lee, C. Axl is a novel target of celastrol that inhibits cell proliferation and migration, and increases the cytotoxicity of gefitinib in EGFR mutant non-small cell lung cancer cells. Mol. Med. Rep. 2019, 19, 3230–3236. [Google Scholar] [CrossRef]
  185. Nazim, U.M.; Yin, H.; Park, S.Y. Autophagy flux inhibition mediated by celastrol sensitized lung cancer cells to TRAIL-induced apoptosis via regulation of mitochondrial transmembrane potential and reactive oxygen species. Mol. Med. Rep. 2019, 19, 984–993. [Google Scholar] [CrossRef] [PubMed]
  186. Wang, Y.; Liu, Q.; Chen, H.; You, J.; Peng, B.; Cao, F.; Zhang, X.; Chen, Q.; Uzan, G.; Xu, L.; et al. Celastrol improves the therapeutic efficacy of EGFR-TKIs for non-small-cell lung cancer by overcoming EGFR T790M drug resistance. Anti-Cancer Drug. 2018, 29, 748–755. [Google Scholar] [CrossRef] [PubMed]
  187. Shukla, S.K.; Chan, A.; Parvathaneni, V.; Kanabar, D.D.; Patel, K.; Ayehunie, S.; Muth, A.; Gupta, V. Enhanced solubility, stability, permeation and anti-cancer efficacy of Celastrol-β-cyclodextrin inclusion complex. J. Mol. Liq. 2020, 318, 113936. [Google Scholar] [CrossRef]
  188. Cao, L.; Hong, W.; Cai, P.; Xu, C.; Bai, X.; Zhao, Z.; Huang, M.; Jin, J. Cryptotanshinone strengthens the effect of gefitinib against non-small cell lung cancer through inhibiting transketolase. Eur. J. Pharmacol. 2021, 890, 173647. [Google Scholar] [CrossRef] [PubMed]
  189. Du, J.; Li, G.; Jiang, L.; Zhang, X.; Xu, Z.; Yan, H.; Zhou, Z.; He, Q.; Yang, X.; Luo, P. Crosstalk between alveolar macrophages and alveolar epithelial cells/fibroblasts contributes to the pulmonary toxicity of gefitinib. Toxicol. Lett. 2021, 338, 1–9. [Google Scholar] [CrossRef]
  190. Xie, X.; Wang, X.; Wu, S.; Yang, H.; Liu, J.; Chen, H.; Ding, Y.; Ling, L.; Lin, H. Fatal toxic effects related to EGFR tyrosine kinase inhibitors based on 53 cohorts with 9569 participants. J. Thorac. Dis. 2020, 12, 4057–4069. [Google Scholar] [CrossRef] [PubMed]
  191. Xie, X.; Zhan, C.; Wang, J.; Zeng, F.; Wu, S. An activatable nano-prodrug for treating tyrosine-kinase-inhibitor-resistant non-small cell lung cancer and for optoacoustic and fluorescent imaging. Small 2020, 16, 2003451. [Google Scholar] [CrossRef]
  192. Xu, S.; Zhou, J.; Dong, X.; Zhao, W.; Zhu, Q. Fluorescent probe for sensitive discrimination of Hcy and Cys/GSH in living cells via dual-emission. Anal. Chim. Acta 2019, 1074, 123–130. [Google Scholar] [CrossRef]
  193. Baell, J.; Walters, M.A. Chemistry: Chemical con artists foil drug discovery. Nature 2014, 513, 481–483. [Google Scholar] [CrossRef] [PubMed]
  194. Heinrich, M.; Appendino, G.; Efferth, T.; Fürst, R.; Izzo, A.A.; Kayser, O.; Pezzuto, J.M.; Viljoen, A. Best practice in research—Overcoming common challenges in phytopharmacological research. J. Ethnopharmacol. 2020, 246, 112230. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Promising natural products for treating SARS-CoV-2.
Figure 1. Promising natural products for treating SARS-CoV-2.
Biomedicines 09 00689 g001
Figure 2. Promising natural products for treating lung cancer.
Figure 2. Promising natural products for treating lung cancer.
Biomedicines 09 00689 g002
Figure 3. Cepharanthine in combination with FDA-approved nelfinavir inhibits SARS-CoV-2.
Figure 3. Cepharanthine in combination with FDA-approved nelfinavir inhibits SARS-CoV-2.
Biomedicines 09 00689 g003
Figure 4. Linoleic acid in combination with FDA-approved remdesivir inhibit SARS-CoV-2.
Figure 4. Linoleic acid in combination with FDA-approved remdesivir inhibit SARS-CoV-2.
Biomedicines 09 00689 g004
Figure 5. Synthesis, bioconversion, and synergistic effects of the nanoparticles HCPNs (image reproduced with permission from [178].
Figure 5. Synthesis, bioconversion, and synergistic effects of the nanoparticles HCPNs (image reproduced with permission from [178].
Biomedicines 09 00689 g005
Figure 6. Synthesis, bioconversion, and synergistic effect of the nanodrug CEL@G-SS-NIR.
Figure 6. Synthesis, bioconversion, and synergistic effect of the nanodrug CEL@G-SS-NIR.
Biomedicines 09 00689 g006
Table 1. Other natural products with anti-SARS-CoV-2 activities in vitro.
Table 1. Other natural products with anti-SARS-CoV-2 activities in vitro.
No.NameStructureEC50 or IC50 (μM)StrainRefs
1Acetoside Biomedicines 09 00689 i0010.043Vero E6 cells[39]
2Anacardic acid Biomedicines 09 00689 i0022.07USA-WA1/2020[26]
3Andrographolide Biomedicines 09 00689 i0030.034Calu-3 cells[40,41]
4Apigenin-7-O-glucoside Biomedicines 09 00689 i0040.074Vero E6 cells[39]
5Artemisinin Biomedicines 09 00689 i00564.45Vero E6 cells[41,42]
6Azithromycin Biomedicines 09 00689 i0062.12Caco-2 cells[43]
7Baicalin Biomedicines 09 00689 i0077.98Vero E6 cells[44]
8Cannabidiol Biomedicines 09 00689 i0087.91Vero E6 cells[45,46]
9Catechin-3-O-gallate Biomedicines 09 00689 i0092.98Vero E6 cells[47]
10Chebulagic acid Biomedicines 09 00689 i0109.76Vero E6 cells[48]
11Daurisoline Biomedicines 09 00689 i0113.66Vero E6 cells[49]
12EGCG Biomedicines 09 00689 i0120.874Vero E6 cells[44,50]
13Emetine Biomedicines 09 00689 i0130.46Vero E6 cells[51,52]
14Epicatechin-3-O-gallate Biomedicines 09 00689 i0145.21Vero E6 cells[47]
15Gallinamide A Biomedicines 09 00689 i0150.028Vero E6 cells[53]
16Gallocatechin-3-O-gallate Biomedicines 09 00689 i0166.38Vero E6 cells[47]
17Hopeaphenol Biomedicines 09 00689 i0172.3B.1.351[54]
18Ipomoeassin F Biomedicines 09 00689 i018 semi-permeabilized mammalian cells[55]
19Kobophenol A Biomedicines 09 00689 i0191.81Vero E6 cells[56]
20Myricetin Biomedicines 09 00689 i0200.22Vero E6 cells[57,58]
21Naringenin Biomedicines 09 00689 i0210.092Vero E6 cells[39,59]
22Osajin Biomedicines 09 00689 i0223.87Vero E6 cells[60]
232,3′,4,5′,6-Pentahydroxybenzophenone Biomedicines 09 00689 i0230.102Vero E6 cells[39]
24Procyanidin B2 Biomedicines 09 00689 i02475.3Vero E6 cells[47]
25Punicalagin Biomedicines 09 00689 i0257.20Vero E6 cells[48]
26Sennoside B Biomedicines 09 00689 i0260.104Vero E6 cells[39]
27Shikonin Biomedicines 09 00689 i02715.75Vero E6 cells[61]
28Δ9-Tetrahydrocannabinol Biomedicines 09 00689 i02810.25Vero E6 cells[45]
29Tetrandrine Biomedicines 09 00689 i0293.00Vero E6 cells[60]
30Theaflavin Biomedicines 09 00689 i0308.44HEK293T human
embryonic kidney cells
[62]
Table 2. Registered clinical trials relating to traditional Chinese medicine prescriptions containing baicalein (active ingredient of Huangqin) for treatment of COVID-19 patients (Chinese Clinical Trial Registry, www.chictr.org/cn/ (accessed on 31 January 2021).
Table 2. Registered clinical trials relating to traditional Chinese medicine prescriptions containing baicalein (active ingredient of Huangqin) for treatment of COVID-19 patients (Chinese Clinical Trial Registry, www.chictr.org/cn/ (accessed on 31 January 2021).
Baicalein (The Active Ingredient of Huangqin)Molecular Mechanisms of BaicaleinHerbal Formula Containing HuangqinRegistration NumberSample Size of the Control Group
Biomedicines 09 00689 i031RdRp inhibitor via noncovalent incorporation [73], potent antagonists against TMPRSS2 [70], improving respiratory function, decreasing IL-1β and TNF-α levels, and inhibiting cell infiltration [71,72].Qingfei Paidu decoctionChiCTR2000029433120
ChiCTR2000030883100
ChiCTR2000032767782
Xinguan I decoctionChiCTR200002963750
Tanreqing capsulesChiCTR200002981336
Tanreqing injectionChiCTR200002943272
Kegan Liyan oral liquidChiCTR2000033720240
ChiCTR2000033745240
ChiCTR2000031982240
Shuanghuanglian oral liquidChiCTR200003313330
ChiCTR2000029605100
Toujie Quwen granuleChiCTR2000031888150
Table 3. The mechanism involved in anticancer activities of other natural products (reported since 2020).
Table 3. The mechanism involved in anticancer activities of other natural products (reported since 2020).
No.NameStructureMechanism of Anti-Lung CancerRefs
1Acovenoside A Biomedicines 09 00689 i032Inhibit the adenosine triphosphate (ATP)-dependent Na+/K+ exchange through the Na+/K+-ATPase[88]
2Asiatic acid Biomedicines 09 00689 i033Inhibited the ionizing radiation-induced migration and invasion[89]
3Baicalein Biomedicines 09 00689 i034Restrained ezrin tension by decreasing inducible nitric oxide synthase expression levels, suppress invasion, reduced vasculogenic mimicry formation [90,91,92]
4Baicalin Biomedicines 09 00689 i035Inhibited the invasion, migration, angiogenesis, and Akt/mTOR pathway [93,94]
5Casticin Biomedicines 09 00689 i036Induced the expressions and nuclear translocation of phosphorylation of H2AX[95]
6Dioscin Biomedicines 09 00689 i037Down-regulated signal transducer and activator of transcription 3 and c-Jun N-terminal kinase signaling pathways[96]
7EGCG Biomedicines 09 00689 i038Regulated CTR1 expression through the ERK1/2/NEAT1 signaling pathway[97,98]
8Ellagic acid Biomedicines 09 00689 i039Inhibited tumor growth, increased p-AMPK, and suppressed hypoxia-inducible factor 1α levels [99]
9Erianthridin Biomedicines 09 00689 i040Attenuated extracellular signal-regulated kinase activity and mediated apoptosis, matrix-degrading metalloproteinases (MMPs) expression[100,101]
10Eugenol Biomedicines 09 00689 i041Restriction of β-catenin nuclear transportation[102]
11Formononetin Biomedicines 09 00689 i042Inhibited EGFR-Akt signaling, which in turn activates GSK3β and promotes Mcl-1 phosphorylation in NSCLC cells[103,104]
12Gallic Acid Biomedicines 09 00689 i043Inhibited of EGFR activation and impairment, inhibition of phosphoinositide 3-kinase (PI3K) and AKT phosphorylation[105,106]
13Glochidiol Biomedicines 09 00689 i044Inhibited tubulin polymerization[107]
14Gracillin Biomedicines 09 00689 i045Inhibited both glycolysis and mitochondria-mediated bioenergetics, induced apoptosis through the mitochondrial pathway[108,109]
15Hispidulin Biomedicines 09 00689 i046Promoted apoptosis by hispidulin via increased generation of ROS[110]
16Icaritin Biomedicines 09 00689 i047Downregulated the immunosuppressive cytokine (TNF-α, IL10, IL6) and upregulated chemotaxis (CXCL9 and CXCL10)[111]
17Isoharringtonine Biomedicines 09 00689 i048Induced death tumor spheroids by activating the intrinsic apoptosis pathway[112]
18Kaempferol Biomedicines 09 00689 i049Inhibitor of nuclear factor erythroid 2-related factor 2[113]
19Liriopesides B Biomedicines 09 00689 i050Reduced proliferation, and induced apoptosis and cell cycle arrest, inhibited the progression of the cell cycle from the G1 to the S phase[114]
20Nagilactone E Biomedicines 09 00689 i051Activated the c-Jun N-terminal kinases, increased the phosphorylation, and promoted the localization of c-Jun in the nucleus [115,116]
218-Oxo-epiberberine Biomedicines 09 00689 i052Inhibited TGF-β1-induced epithelial-mesenchymal transition (EMT) possibly by interfering with Smad3[117]
22Parthenolide Biomedicines 09 00689 i053Reduced the phosphorylation of EGFR and downstream signaling pathways mitogen-activated protein kinase (MAPK)/ERK, inhibited PI3K/Akt/FoxO3α signaling[118,119,120]
23PDB-1 Biomedicines 09 00689 i054Suppressed lung cancer cell migration and invasion via FAK/Src and MAPK signaling pathways[121]
24Polyphyllin I Biomedicines 09 00689 i055Induced autophagy by activating AMPK and then inhibited mTOR signaling, promoted apoptosis, modulated the PI3K/Akt signaling[122,123]
25Quercetin Biomedicines 09 00689 i056Inhibited proliferation and induced apoptosis[124]
26Silibinin Biomedicines 09 00689 i057Inhibited cell proliferation, migration, invasion, and EMT expression[125]
27Sinomenine Biomedicines 09 00689 i058Downregulated expression of MMPs and miR-21, suppressed α7 nicotinic acetylcholine receptors expression[126,127,128]
28Toxicarioside O Biomedicines 09 00689 i059Decreased the expression of trophoblast cell surface antigen 2, resulting in inhibition of the PI3K/Akt pathway and EMT program[129]
29Vincamine Biomedicines 09 00689 i060Interaction with the apoptotic protein caspase-3[130]
30Xanthohumol Biomedicines 09 00689 i061Suppressed ERK1/2 signaling and reduced the protein levels of FOS-related antigen 1, decreased the mRNA level of cyclin D1[131]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, L.; Wang, Z. Natural Products, Alone or in Combination with FDA-Approved Drugs, to Treat COVID-19 and Lung Cancer. Biomedicines 2021, 9, 689. https://doi.org/10.3390/biomedicines9060689

AMA Style

Yang L, Wang Z. Natural Products, Alone or in Combination with FDA-Approved Drugs, to Treat COVID-19 and Lung Cancer. Biomedicines. 2021; 9(6):689. https://doi.org/10.3390/biomedicines9060689

Chicago/Turabian Style

Yang, Liyan, and Zhonglei Wang. 2021. "Natural Products, Alone or in Combination with FDA-Approved Drugs, to Treat COVID-19 and Lung Cancer" Biomedicines 9, no. 6: 689. https://doi.org/10.3390/biomedicines9060689

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop