Next Article in Journal
Advantages and Prospects of Tag/Catcher Mediated Antigen Display on Capsid-Like Particle-Based Vaccines
Previous Article in Journal
Genomic Characterization and Phylogenetic Classification of Bovine Coronaviruses Through Whole Genome Sequence Analysis
 
 
Correction published on 30 June 2020, see Viruses 2020, 12(7), 712.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Flavonoids as Antiviral Agents for Enterovirus A71 (EV-A71)

Centre for Virus and Vaccine Research, Sunway University, Bandar Sunway, Subang Jaya, Selangor 47500, Malaysia
*
Author to whom correspondence should be addressed.
Viruses 2020, 12(2), 184; https://doi.org/10.3390/v12020184
Submission received: 26 December 2019 / Revised: 21 January 2020 / Accepted: 22 January 2020 / Published: 6 February 2020
(This article belongs to the Section Viral Immunology, Vaccines, and Antivirals)

Abstract

:
Flavonoids are natural biomolecules that are known to be effective antivirals. These biomolecules can act at different stages of viral infection, particularly at the molecular level to inhibit viral growth. Enterovirus A71 (EV-A71), a non-enveloped RNA virus, is one of the causative agents of hand, foot and mouth disease (HFMD), which is prevalent in Asia. Despite much effort, no clinically approved antiviral treatment is available for children suffering from HFMD. Flavonoids from plants serve as a vast reservoir of therapeutically active constituents that have been explored as potential antiviral candidates against RNA and DNA viruses. Here, we reviewed flavonoids as evidence-based natural sources of antivirals against non-picornaviruses and picornaviruses. The detailed molecular mechanisms involved in the inhibition of EV-A71 infections are discussed.

1. Introduction

Flavonoids are naturally occurring polyphenolic biomolecules widely found in plants and are responsible for a wide variety of biological functions [1]. Pharmacological properties of flavonoids include antioxidant, anti-inflammatory, anticancer, antimicrobial and immunomodulatory functions [2]. More than 6000 compounds have been identified as flavonoids based on their basic structures consisting of C6-C3-C6 and they are divided into several classes such as flavonols, flavanones, isoflavones, flavones, and anthocyanidins. Due to a wide range of biological activities being exhibited by flavonoids, they have become molecules of interest for natural drug discovery research. In this review, antiviral mechanisms of flavonoids, mainly against Enterovirus A-71 (EV-A71), will be discussed to provide evidence-based efficacy of antiviral properties.

2. Flavonoids against Viruses

Flavonoids have been studied against a wide range of DNA and RNA viruses. In general, flavonoids work by several mechanisms. They can block attachment and entry of viruses into cells, interfere with various stages of viral replication processes or translation and polyprotein processing to prevent the release of the viruses to infect other cells. Different flavonoids have been found to inhibit the virus through various mechanisms. Based on antiviral mechanisms of action, flavonoids can be prophylactic inhibitors, therapeutic inhibitors or indirect inhibitors by interaction with the immune system. Flavonoids that inhibit viral activity can be further divided into the following sub-categories.
  • Flavonoids that bind to specific extracellular regions of the virus such as viral proteins present on the capsids.
  • Flavonoids that prevent attachment or entry of the virus into host cells. In some cases, flavonoids can bind to virions and modify the virus structure. Though the virus can still internalize, the process of viral uncoating is stalled.
  • Early-stage replication inhibitors.
  • Transcription and translation blockers.
  • Inhibition of late stages of maturation such as inhibition of assembly/packaging and release.
  • Flavonoids that can inhibit viral infections by interfering with host factors that are required for successful infection or modulating the immune system to reduce the viral load.

2.1. Flavonoids against Non-Picornaviruses

Research has been carried out on plants that are being used in traditional medicines in different parts of the world. Many natural compounds of various chemical classes have been identified as potential therapeutic agents. Several flavonoids have been studied to evaluate their antiviral potential. For example, quercetin and quercitrin were first demonstrated by Cutting et al. (1949) to prophylactically inhibit rabies virus in mice [3]. Moreover, quercetin was found to exhibit its antiviral activity against vesicular stomatitis virus (VSV) by activation of macrophages [4]. Biologically active components of Elderberry extract dihydromyricetin and tetra-O-methyl quercetin were found to inhibit influenza A virus in vitro, most likely by binding to viral mannose-rich hemagglutinin domains which are important for the virus to enter the host cells [5]. Mechanistic studies later showed that it was able to inhibit neuraminidase activity of influenza A virus by interaction with the subunit 2 of the hemagglutinin that resulted in virus entry inhibition [6]. Quercitrin, a rhamnoside derivative of quercetin, was found to inhibit the initial stage of viral replication of the influenza A virus, evident by a decrease in mRNA synthesis of the virus. The study also concluded that quercitrin was unable to interact directly with virus particles [7]. Quercetin was shown to inactivate the NS3 helicase and NS5 protease of hepatitis C virus (HCV) [8,9]. Quercetin also blocked viral binding and entry of herpes simplex virus (HSV)-1, HSV-2, and drug-resistant HSV-1 to Madin–Darby canine kidney NBL-2 (MDCK) cells [6]. Another flavonoid derivative, 3-β-O-d-glucoside, was able to protect mice from Ebola virus infection when given prior to virus challenge. However, the mechanism by which this flavonoid inhibited viral entry is unclear [10].
In a similar manner, sulfated rutin was shown to block the entry of human immunodeficiency viruse (HIV)-1 without interactions with the host cell membrane. Cell fusion and entry assays revealed that sulfated rutin could drastically inhibit HIV-1 infection when cells were treated during the early adsorption phase. The probable mechanism proposed by investigators was the inhibition of HIV glycoprotein-mediated cell-cell fusion step [11]. In the same study, it was identified that sulfated rutin could also inhibit HSV, the mechanism of inhibition is still unknown.
Apigenin was reported as an antiviral against HCV by host factor modulation. It caused a reduction in mature miRNA122 production that regulated HCV infection in vitro [12]. Baicalin was found to interfere with the interaction between the HIV-1 envelope protein and host immune cells. This anti-HIV activity was achieved by inhibition of the HIV-1 envelope glycoprotein (gp120)-mediated fusion with T cells and monocytes expressing CD4/CXCR4 or CD4/CCR5 [13]. In addition, baicalin was also reported to inhibit dengue virus (DENV-2) by the virucidal mechanism. It blocked the attachment of DENV-2 to the Vero cells but did not show any activity during the virus entry stage [14]. Baicalin was able to induce the production of interferon-gamma (IFN-ɣ) in natural killer (NK), CD4+ and CD8+ T cells during in vitro influenza A virus infection. It was suggested that baicalin was able to directly bind NS1–p85β (RNA binding domain), which was the main mediator to downregulate IFN-ɣ [15]. The induction of IFN-ɣ triggered the immune system to activate the Janus kinase/signal transducer and activator of transcription protein 1 (JAK/STAT1) pathway, which led to the expression of IFN-ɣ-inducible genes. This expression of IFN-ɣ genes and signaling pathway facilitated the reduction of viral loads by immune modulation [16]. Baicalein, the parent compound of baicalin, has also been reported as antiviral against Japanese encephalitis virus (JEV) and DENV-2. It is postulated that baicalein inhibited JEV by probable interaction with structural proteins of virus and this prevented entry of the virus into the cells [17]. However, the antiviral mechanism against DENV-2 was identified as replication inhibition in silico by interaction with viral NS3/NS2B and NS5 proteins [18].
Flavonoids, genistein and ginkgetin, were reported to inhibit assembly and release of HIV and influenza A virus, respectively. Genistein inhibited Vpu protein that is involved in the formation of ion channels in infected cells and thus controlled the release of HIV [19]. Ginkgetin inhibited sialidase activity of influenza A virus and thus inhibited virus assembly and release [20]. Fisetin possessed post-infection anti-CHIKV activity by the inhibition of NS proteins 1 and 3 in a dose-dependent manner. Moreover, it was observed that CHIKV E2 protein and its precursor pE2 were also downregulated during treatment with fisetin [21]. Similarly, a decreased replication process of DENV-2 was observed when Vero cells were treated with fisetin after DENV-2 infection. It was postulated that this antiviral activity could be due to the direct binding of fisetin to the viral RNA, thus impeding polymerase activity [22]. Kaempferol and its structural flavonol derivatives were studied to identify the suitable pharmacophore to design antiviral against the influenza A virus. This study revealed that most of the flavonols with structural similarity with kaempferol were able to noncompetitively inhibit neuraminidase enzyme. However, the exact interacting partner is still unknown [23]. Moreover, kaempferol was also reported to inhibit 3a channels of coronavirus and blocked the release of virus progeny [24]. Luteolin was found to interfere with the entry of influenza A virus and severe acute respiratory syndrome coronavirus (SARS-CoV) by interacting with hemagglutinins of influenza A virus and S2 protein of SARS CoV viruses, respectively [25,26].
Another group of flavonoids, viz. methoxyflavone, isoscutellarein, and 8-methoxy-isoscutellarein, were reported to inhibit early replication of influenza A virus by reduction of sialidase activity, inhibition of lysosomal fusion and RNA polymerase activity [27]. Naringenin, a flavanone, also possessed antiviral activities against HCV [28], DENV-2, DENV-4 [29] and CHIKV [30] by replication inhibition at RNA and protein levels. Other flavonoid such as silymarin which has a broad-spectrum antiviral activity [31,32,33] was able to block entry and fusion of HCV viral pseudo-particles (pp) when Huh-7.5 cells were pre-treated with silymarin [34]. Moreover, it was shown to interfere with the replication process of DEN-V, CHIKV and influenza A virus [35,36,37].
Tea catechins were also reported to inhibit the influenza A virus by binding to the hemagglutinins and restricted virus adsorption, preventing the penetration of the virus into the cells [38]. Tea catechins were studied in clinical trials by Matsumoto and co-workers (2011). They conducted a randomized controlled trial to study the effects of tea catechins in the prevention of influenza virus infections in healthcare workers in elderly homes. Tea catechins were effective in preventing the clinical incidence of influenza infections in the treatment group as compared to the placebo group [39]. Epigallocatechin (EGC), a tea catechin flavonoid, was reported to inhibit the replication of influenza A and B viruses by in vitro acidification of the lysosomal and endosomal environment through clathrin-mediated endocytosis [40]. Apart from having antiviral activity against the influenza virus, another flavonoid from tea catechins, epigallocatechin gallate (EGCG), was shown to directly bind to CD4+ T-cells. This binding blocked the binding of the HIV-1 envelope protein (gp120) and caused viral entry inhibition [41]. EGCG was also demonstrated to bind to the glycoprotein B and glycoprotein D of HSV to render it non-infective [42]. Moreover, EGCG was shown to have potent antiviral activity against CHIKV by blocking its entry, probably by competing for cellular co-receptors of target cells such as heparan sulfate or sialic acid [43]. EGCG was also found to inhibit the replication of Epstein–Barr virus (EBV) by inhibition of Zta, Rta and EA-D genes. This inhibition interrupted the MEK/ERK1/2 and PI3-K/Akt signaling pathway of the lytic cycle EBV [44]. EGCG also inhibited hepatitis B virus (HBV) replication by impairment of the production of pre-core mRNA and replicative intermediates of DNA [45]. Zika virus was also inhibited by EGCG, as it was able to interact with the lipid envelope and blocked the entry of virus into cells [46]. It also caused acidification in lysosomes to make an unfavorable environment for virus replication [47]. The antiviral activities of several other flavonoids according to the mechanisms of inhibition against various viruses are summarized in Table 1. Structures of these flavonoids can be found in Supplementary material Table S1: Molecular structures of antiviral flavonoids.

2.2. Flavonoids against Non-EV-A71 Picornaviruses

Picornaviruses are non-enveloped RNA viruses belonging to the picornaviridae family. Their natural hosts are vertebrates such as reptiles, fish, amphibians, birds and mammals. Viral infections such as poliomyelitis, hepatitis A, common cold and hand, foot and mouth disease are caused by picornaviruses. Flavonoids, in particular, have shown promising results in the laboratory against picornaviruses. Some of these flavonoids are discussed.
Quercetin, a well-studied flavonoid, exhibited anti-rhinoviral (RV) activity by inhibition of the viral transcription and translation processes. Quercetin decreased endocytosis of RV and reduced phosphorylation of Akt (effector of phosphoinositol 3-kinase) when epithelial cells were pretreated with quercetin. In addition, it also reduced viral infection when added 6 h post-infection in vitro. Moreover, it also repressed interferon and interleukin-8 response, resulting in lower viral RNA and capsid protein production. Quercetin demonstrated similar effects in vivo with decreased viral loads and suppression of viral immune mediators [50]. Quercetin was shown to protect mice when challenged with the quercetin-treated Mengo virus. It was interesting to note that there was no antiviral activity detected when quercetin was tested in L-929 cells [3]. This observation led to the study of the role of interferon induction in antiviral activity. However, there was no significant detection of interferons in drug-treated mice. This resulted in speculating that the antiviral activity was achieved by macrophage activation instead of interferon induction. In the same study, quercetin was also found to inhibit encephalomyocarditis virus (EMCV) by macrophage activation [4]. A methyl derivative of quercetin was shown to inhibit poliovirus by blocking the synthesis of viral genomic RNA and proteins [51,52]. Similarly, trimethylquercetin was shown to inhibit the production of progeny coxsackievirus B4 (CV-B4) and protected the mice from death upon lethal challenge [53]. In another independent study, dihydroquercetin isolated from Larix sibirica was found to mediate the in vivo CV-B4 infection by inhibition of ROS induced oxidative stress [54].
Similarly, luteolin and its derivatives have been studied against picornaviruses. For example, luteolin was found to inhibit the RNA synthesis of Coxsackievirus A16 (CV-A16) [55]. Moreover, it was also reported to inhibit poliovirus at replication stage. However, the mechanism of inhibition of poliovirus by luteolin is still unknown [56]. In contrast to luteolin, its derivative eupafolin was shown to inhibit CV-A16 by inhibition of viral attachment. It was shown to reduce the IL-6 and RANTES which led to the inactivation of virus-induced upregulated the ERK1/2, c-Jun, and STAT3 signaling pathways [57].
In a similar manner, kaemferol and its derivatives were also identified to be potent inhibitors of picornaviruses. Methylkaempferol was reported to inhibit the late replication step of the poliovirus. The inhibition mechanism is attributed to the potential of this flavonoid to inhibit the positive-strand synthesis of viral RNA [58]. Furthermore, Kaempferol-3-O-[2″,6″-di-O-Z-p-coumaroyl]-β-d-glucopyranoside and its derivatives were identified as replication inhibitors of Rhinovirus 1B (HRV-1B) and Coxsackievirus B3 (CV-B3) in vitro [59]. The group of methoxy flavones were reported to inhibit poliovirus-1 by downregulation of signaling pathway and apoptosis [60]. Furthermore, entry of rhinovirus was shown to be inhibited in silico by inhibition of the protein grid of virus [61]. Another flavonoid sakuranetin was also identified as a potent inhibitor of rhinovirus-3 by its antioxidant properties [62]. Several other flavonoids were reported to inhibit the activity of various picornaviruses [60,63,64,65,66]. However, their exact molecular mechanisms of actions are still unknown. Examples of bioflavonoids from various chemical classes exhibiting antiviral activities against picornaviruses are summarized in Table 2. Structures of these flavonoids can be found in Supplementary material Table S1: Molecular structures of antiviral flavonoids.

3. Flavonoids that Target EV-A71

Hand, foot and mouth disease (HFMD) is a global childhood infectious disease. It is predominant in Asia, with annual outbreaks ranging from two to three million cases. EV-A71 is one of the main etiological agents of HFMD, prevalent in children under the age of 6 years. EV-A71 belongs to the Picornaviridae family and is classified as a non-enveloped RNA virus. The 7.4 kb genome of EV-A71 encodes four structural viral proteins (VP1 to VP4) and seven non-structural proteins (2A to 2C and 3A to 3D). These structural and non-structural proteins are generally being considered for designing effective antiviral agents against EV-A71.
The life cycle of EV-A71 initiates with binding to host receptors such as SCARB2, vimentin and annexin 2. Once attached, the virus enters into the cell through endocytosis and uncoating starts in an endosome. Thereafter, the translation of mRNA and polyprotein processing occur. Human hnRNP A1 is an interaction partner of viral internal ribosome entry site (IRES) whereas hnRNP A2 is highly similar to hnRNP A1. Upon interaction of IRES with hnRNP A1 and hnRNP A2, the viral translation process is activated. Structural and non-structural viral proteins are formed via cleavage by 2A and 3C proteases. This is followed by the replication and assembly of viral RNA into the capsid, maturation of viral particles and release of virus from the cells [69].
Many flavonoids have been studied to identify their potential as antiviral agents against EV-A71. The majority of the flavonoids studied demonstrated the antiviral activity by interference with the EV-A71 replication cycle (Figure 1). However, a few flavonoids have been reported to exert their antiviral activity by immune mediation (Figure 2). Examples of flavonoids with antiviral activities against EV-A71 are discussed.

3.1. Apigenin

Apigenin is a flavone that has been tested for its antiviral potential against Fuyang and BrCr strains of EV-A71. It was able to reduce infection at the inhibitory concentration (IC50) of 10.3 μM. Apigenin inhibited EV-A71 infection by suppressing viral internal ribosome entry site (IRES) activity. Apigenin was found to disrupt the viral RNA association with hnRNP A1 and A2 proteins [70].
Moreover, apigenin was found to inhibit cellular apoptosis by caspase-3 cleavage, which is considered as a prime process to release viral progeny. It is also a known scavenger of reactive oxygen species (ROS) and was found to reduce infection-induced ROS generation in EV-A71 damaged cells. Furthermore, it decreased the cytokine levels involved in infection. Apigenin was confirmed to inhibit the activation of the c-Jun N-terminal kinase (JNK) pathway required for the replication process and release of progeny EV-A71. Another downstream p38 MAP kinase signaling pathway was found to be partially inhibited by apigenin. All these inhibitory mechanisms were suggested to suppress viral infection [71].
In a recent study by Dai et al. (2019), apigenin exhibited potent antiviral activity against EV-A71 genotype C4 strain in vitro and in vivo. It was able to reduce the cytopathic effect to 50% at the concentration of 24.74 µM in 293S cells. Moreover, apigenin significantly reduced viral RNA and protein synthesis in vitro. When evaluated for its potential to protect newborn BALB/c mice from intracranial lethal challenge with EV-A71, apigenin was found to protect mice at the dose of 50 mg/Kg with 88.89% survival rate. Significant improvement in clinical scores and body mass were also observed in the treatment group as compared to the control group [72].

3.2. Baicalin

Baicalin is a flavone reported to inhibit the BrCr-Tr strain of EV-A71 in post-infection assays with an IC50 of 4.96 μg/mL. Baicalin was shown to inhibit 75% viral infection at 8 h post-infection. Inhibition by baicalin at early stages of infection was due to a decrease in mRNA and protein levels of 3D polymerase. Moreover, it was also demonstrated that baicalin reduced the expressions of FasL and caspase-3, hence inhibiting EV-A71 induced apoptosis in Rhabdomyosarcoma cells (RD). Additionally, it was identified to inhibit proinflammatory cytokines by the reduction in NF-κB p65 expression [73].

3.3. Chrysin and Its Derivative

Chrysin is a flavone widely found in leaves, fruits, and vegetables. Chrysin and its ester (diisopropyl chrysin-7-yl phosphate) were evaluated against the clinical isolate EV-A71 Shzh-98 and were found to inhibit the viral replication process by blocking the activity of the 3C protein. Chrysin ester was found to be the more potent inhibitor of 3C protein than the parent chrysin. The activity was predicted in silico by simulation and was confirmed by in vitro protease inhibition assays. These flavonoids were able to decrease the production of the viral capsids, infectious virions and RNA of EV-A71 [74].

3.4. Fisetin

Fisetin, a flavonol, was reported to possess antiviral activity against the clinical isolate EV-A71 CMUH01 with an IC50 of 84.5 μM. Fisetin was able to inhibit the enzymatic activity of recombinant 3C protease of the virus in a cell-based assay. To further confirm its activity in vitro, antiviral assays were performed in RD cells. Fisetin was found to block the cleavage activity of 3C and thereby inhibited the viral replication [75].

3.5. Formononetin

The isoflavone formononetin was tested against several isolates of EV-A71 including BrCr, H, JS-52 and Shzh 98. Inflammatory pathways such as ERK, p38, JNK, MAPK are usually upregulated in EV-A71 infections and help in viral reproduction. Formononetin was shown to act on the MAPK pathway and reduced the activation of the downstream regulator of infection-induced inflammation mediators, prostaglandin E2 (PEG2) and cyclooxygenase 2 (COX2). This, in turn, reduced the viral infection. Moreover, it also suppressed the signaling cascade of the ERK, p38 and JNK pathways to bring down the infection [76].
In another study, formononetin was found to inhibit EV-A71 strains G082, SH12-036 and SH12-276. Formononetin was able to inhibit attachment and entry of EV-A71 in Vero cells. To further evaluate the exact target, the virus was passaged with increasing concentrations of formononetin and sequenced after the 10th passage. A common mutation causing K58T in VP4 was observed and this mutation was able to confer viral resistance towards formononetin. This resistance was further confirmed by the absence of K58T mutation in the wild type virus. The mutant virus was evaluated for attachment in Vero cells and the K58T mutation was found to inhibit viral attachment. It was hypothesized that formononetin could possibly cause an overall conformational change in the viral capsid [77].
Formononetin was also evaluated for its antiviral potential against EV-A71 subgenotype C4 strain in 239S cells. It exhibited antiviral activity with an IC50 of 12.5 µM and significantly reduced viral replication and protein synthesis. It also protected newborn mice with a survival rate of 75% and significantly improved clinical score and animal body weight at a dose of 10 mg/Kg [72].

3.6. Hydroxyflavone and Its Derivatives

The hydroxyflavones are synthetic molecules, the core structure of which is also a backbone of flavonols. Wang et al. (2014) predicted that 7-hydroxyflavone and its phosphate ester diisopropyl-flavon7-yl phosphate were able to inhibit the activity of 3C protease of EV-A71 in silico. They further evaluated in vitro potential of these flavonoids as antivirals against Shzh-98 isolate of EV-A71. It was found that 7-hydroxyflavone and diisopropyl-flavon7-yl phosphate both inhibited virus-induced cytopathic effect in RD cells with an IC50 of 23.45 and 13.63 μM, respectively. In addition, both flavonoids significantly reduced viral protein synthesis [78].

3.7. Kaempferol

The flavonol kaempferol was reported to inhibit the EV-A71 clinical isolate Chuh-050530-5. Kaempferol treated cells were observed to reduce virus yield by 6 log units 24 hpi. Kaempferol was able to exert its action through induction of trans-factors such as the far upstream element-binding proteins (FUBP) and hnRPs. These trans-factors were identified to bind the highly conserved 5‘ UTR region of the virus genome and attenuated the viral infection. Moreover, it also disrupted the activity of IRES by an overall change in the composition of the trans-factors. This resulted in reduced viral infection by inhibition of translation and replication [79].
Kaempferol was found to inhibit EV-A71 subgenotype C4 strain with an IC50 of 52.75 μM. It significantly reduced RNA copy number and protein synthesis 16 hpi. Newborn BALB/c mice were also protected from lethal challenge with the virus with 50 mg/Kg of kaempferol. The clinical score and body weight were found to be similar to that of the solvent treated group (negative control) [72].

3.8. Luteolin and Its Derivatives

Luteolin is one of the examples of flavonoids that acted at an early step of the EV-A71 (subgenotype C4b) replication cycle by post-attachment inhibition of the virus. The IC50 was found to be 10 μM [55]. Moreover, when luteolin was tested against the prototype BrCr strain and the Fuyang0805 isolate of EV-A71, it was able to prevent virus-induced ROS generation, cellular apoptosis and inflammatory cytokine production such as IL-6 and CCL5 [71]. In another independent study, Dai et al. (2019) showed that luteolin exhibited similar inhibition patterns with an IC50 of 13.5 μM. It was identified to reduce viral replication and protein synthesis. Moreover, in the murine model, it showed 91.67% protection of newborn mice at a dose of 10 mg/Kg [72].
Eupafolin, a methoxy derivative of luteolin, was also found to inhibit EV-A71 with an IC50 of 1.39 µM in an independent study. It was shown to inhibit IL-6 and chemokine RANTES which were upregulated during the in vitro viral infection. This inhibition resulted in the down-regulation of ERK, activated protein-1 (AP-1) and STAT3 mediated inflammatory signaling pathways [57].

3.9. Penduletin

Penduletin is a methylated flavonol abundantly found in the leaves of Laggera pterodonta. The compound was tested for its antiviral activity against EV-A71 strain GZ-08-02. The flavonoid was a potent inhibitor of EV-A71 with an IC50 of 0.17 and 0.20 µM in Vero and RD cells, respectively. Mechanistic studies revealed that penduletin specifically inhibited the production of infectious RNA viral progeny. Moreover, it was also able to reduce VP1 expression in post-infection assays, suggesting its ability to halt protein synthesis during viral replication process [80].
Penduletin also exhibited potent in vitro and in vivo antiviral activity against EV-A71 subgenotype C4 strain. The IC50 was found to be 0.63 µM in 293 S cells and 66.67% newborn BALB/c mice survived the lethal challenge with EV-A71 when mice were treated with 5 mg/Kg dose of penduletin [72].

3.10. Peracetate Pulicarine

Peracetate pulicarine is one of the less-studied flavonoids against viruses. It is a methoxyphenyl acetate, a derivative of a flavonoid which was found to possess anti-EV-A71 activity in a recent screening of a flavonoid library. The antiviral activity of pulicarine was associated with inhibition of the 3D polymerase expression in EV-A71 replication process [81].

3.11. Prunin

Prunin is a flavanone glycoside recently found to be potent antiviral against enteroviruses A and B including clinical isolates and strains of H, B5 and C4 genotypes of EV-A71. Prunin was able to suppress the activity of IRES and disrupted protein and RNA synthesis with an IC50 of 115.3 nM. It was further evaluated in the murine model using newborn BALB/C mice. Prunin was able to protect from the lethal challenge of EV-A71 (strain 41) with a 100% survival rate at the doses of 3 mg/Kg and 10 mg/Kg per mice. Apart from suppression of IRES activity in vitro, prunin successfully protected mice from the infiltration of immune cells and inflammation of tissues. Moreover, prunin was also able to reduce viral antigen in hindlimb muscles of mice [68].

3.12. Quercetin and Its Derivatives

Quercetin has been found to possess broad-spectrum antiviral activity by several mechanisms. When tested against EV-A71, quercetin was able to target the post-attachment stage of EV-A71 infection in vitro by disrupting viral RNA [55]. In another study, quercetin was predicted to interact with catalytic residues of the 3C protease and could inhibit the activity of viral the 3C protease in silico. To further confirm the findings, quercetin was evaluated for its antiviral activity using the SK-EV006/Malaysia/97 isolate of EV-A71. The IC50 of 8.8 μM was observed in Vero cells and 12.1 μM in RD cells. Quercetin was also investigated for its anti-apoptotic potential against virus-induced cell death. Quercetin was able to prevent the viral spread induced by apoptosis, thus making it a dual action antiviral agent [82].
Rutin is another glycoside flavonol derivative of quercetin. It was identified as antiviral against EV-A71 using the recombinant 3C protease screening assay. To confirm its activity in vitro, rutin was tested against EV-A71 strain CMUH01 (B5). Rutin was able to inhibit the replication stage of infection with an IC50 of 110 μM [75]. In another study, rutin was found to reduce the infectivity of the C4 subgenotype EV-A71 with an IC50 of 200 μM. The anticipated mechanism was recognized as the suppression of the MEK1-ERK signaling pathway [83]. Direct inhibition of the 3C protease and interference with inflammatory pathways were assumed to contribute to the reduction of virus titres.
Quercetagetin, a synthetic derivative of quercetin, was identified to inhibit the RNA replication machinery and viral translation. The flavonoid was found to reduce the activity of the 3D polymerase in vitro [81]. Similarly, chrysosplenetin, a methylated derivative of quercetagetin, was identified to potently inhibit EV-A71 by reducing the production of viral RNA and proteins [80].
Quercetin and its methoxy derivative isorhamnetin and chrysosplenetin were evaluated for their in vitro and in vivo potential as antivirals against EV-A71 (subgenotype C4). Quercetin and chrysosplenetin displayed potent antiviral activities in vitro with an IC50 of 1.2 µM and 0.68 µM, respectively while isorhamnetin showed inhibition with an IC50 of 60.7 µM. All three flavonoids reduced RNA and protein synthesis in vitro. In the same study, these flavonoids were evaluated to identify their protective effect in mice from lethal challenge of EV-A71. Isorhamnetin was found to confer the best protection at the dose of 10 mg/Kg with 100% survival while quercetin could only protect 50% of the animals using the same dose. Despite being the most potent antiviral amongst the three flavonoids when tested in vitro, chrysosplenetin was the least effective in vivo. Only 30% of the animals survived after treatment with 5 mg/Kg of chrysosplenetin. All three flavonoids were able to improve the clinical score and bodyweight of the treated animals [72]. It should be taken into consideration that in this study, the route of virus inoculation was intracranial while flavonoids were administered by intraperitoneal (IP) route. This difference of injection routes could result in varying efficacy of flavonoids to cross the blood-brain barrier and potentially protect mice from virus. In vivo antiviral activity of different flavonoids against EV-A71 has been summarized in Table 3.

3.13. Flavonoids Isolated from Scutellaria Baicalensis Georgi

Flavone derivatives such as mosloflavones, norwogonin and oroxylin A are abundant in Scutellaria baicalensis Georgi. These flavones were isolated as pure compounds and were reported to inhibit early stages of EV-A71 replication with an IC50 of 37.72, 31.83 and 14.19 µg/mL, respectively. All three flavonoids were able to limit replication process, leading to inhibition of viral protein (VP2) synthesis. However, oroxylin was found to possess superior activity as compared to mosloflavones and norwogonin. Furthermore, these flavonoids were evaluated for their potential to inhibit viral attachment and entry. None of these flavonoids were found to be involved in the inhibition of attachment or entry of the virus in Vero cells [84].

3.14. Flavonoids with Unknown Mechanisms of Action against EV-A71

There are more examples of flavonoids that were reported to inhibit EV-A71. However, the mechanisms by which they modulated the infectivity of EV-A71 are still unknown. Examples include nobiletin, morin hydrate, myricetin, taxifolin, diosmetin, dihydromyricetin [85], hesperidin [79], thio flavones [86] and galangin [55].

4. Structure–Activity Relationship among Potential Antiviral Flavonoids against EV-A71

Flavonoid classes differ from each other by various substitutions particularly phenolic or hydroxyl (OH) groups on its flavan nucleus. Flavonoids that have successfully accounted for antiviral activity against EV-A71 mostly belong to the flavonoid sub-classes flavonols and flavones. Flavones are dual fused aromatic rings (A and C) with an extension of a benzene ring (B) at position 2 of ring C (Figure 3) [87]. The skeleton “ring A fused with ring C” has previously reported to possess antiviral activity against the dengue virus NS2B-NS3 complex [88]. Moreover, hydroxylation on ring A, particularly on positions 5 and 7, was found to be favorable for antiviral activity against the influenza A virus [89]. Based on previous findings, it can be expected that the same structure could attribute towards antiviral activity against EV-A71.
It can be observed that the anti-EV-A71 flavonoids with methylation have drastically improved the antiviral activity, such as in the case of chrysosplenetin, eupafolin, penduletin and prunin (Table 4). Moreover, presence of methoxy substitution at the position 7 of ring A, along with hydroxyl or methoxy group at position 6 rendered higher activity with lowest IC50 values (chrysosplenetin, eupafolin, penduletin) amongst all evaluated flavonoids. Prunin, a glucoside of naringenin, having large glucose moiety at position 7 instead of methoxy substitution, interestingly showed potent antiviral activity as well. From these examples, it can be speculated that positions 6 and 7 of ring A require bulky group substitutions in order to possess potent activity against EV-A71. Possible interactions between different substitutions and EV-A71 could be studied in silico to determine the nature of bonds involved in antiviral activity. It must be noted the antiviral activity of flavonoids was evaluated against specific EV-A71 genotype/subgenotype. It is unknown if there is a broad spectrum of antiviral activities against other EV-A71 genotypes/subgenotypes.

5. Limitations of Flavonoids as Antivirals against EV-A71

Flavonoids are abundant in fruits and vegetables and therefore are considered as safe and non-toxic for human consumption. However, the bioavailability of these flavonoids in their crude form is very poor. Use of isolated, purified or chemically synthesized small molecular weight flavonoids has increased prospects to reach the target. To further enhance the stability, solubility and improved systemic distribution, a nano-drug delivery system can be employed [90]. Despite possessing the potential to treat EV-A71 in vitro and in mice, flavonoids have failed to reach clinical trials. Some of the possible reasons could include infancy of in vivo studies on flavonoids against EV-A71 as there have been only two reports which were published in 2019, claiming absolute protection of mice by prunin and isorhamnetin [68,72]. Moreover, lack of long-term toxicity data on purified flavonoids contributes challenges to further progress these potential molecules to clinical trials. Another important factor could be the development of resistance by mutant viruses against these flavonoids. Like any other drug candidate, purified flavonoids require regulatory approval from the Food and Drug Administration (FDA) before it can be marketed as antivirals. Detailed pharmacokinetic, pharmacodynamics and long-term toxicology studies on these flavonoids are necessary before they can be offered as a therapeutic option.

6. Conclusions

Flavonoids such as penduletin, eupafolin, baicalin, luteolin, quercetin and chrysosplenetin have been shown to be antiviral agents against EV-A71 based on their IC50 values (<10 µM). These flavonoids were also found to be safe and non-cytotoxic to human cells as indicated by CC50 values. Flavonoids such as prunin could confer 100% protection of neonatal mice while luteolin showed 91.67% protection against lethal challenges with EV-A71. However, flavonoids with higher IC50 values can also serve as a potential antiviral against EV-A71. For instance, kaempferol and isorhamnetin exhibited higher IC50 in vitro but both conferred significant protection of mice from viral challenge. The flavonoids that have not yet been tested in vivo could be further evaluated to identify their effectiveness.
Flavonoids, with their ability to target various stages of viral infection, are becoming a more focused topic to explore their potential as antivirals in the current era. Apart from their classical antioxidant properties, some flavonoids have been shown to inhibit viruses at the molecular level both in vitro and in vivo. In light of the current finding of the in vivo efficacy of some flavonoids against EV-A71, it can be concluded that flavonoids have great potential to be developed as therapeutic candidates against EV-A71. However, to make these flavonoids available as antivirals, a more in depth understanding of their pharmacological properties and clinical outcomes is warranted.

Supplementary Materials

The following are available online at https://www.mdpi.com/1999-4915/12/2/184/s1, Table S1: Molecular structures of antiviral flavonoids. Structures of flavonoids discussed in Tables 1 and 2 are provided in supplementary Table S1.

Author Contributions

Conceptualization, C.L.P. and S.L.; software, S.L.; resources, C.L.P.; data curation, S.L.; writing—original draft preparation, S.L.; writing—review and editing, C.L.P.; visualization, S.L.; supervision, C.L.P.; funding acquisition, C.L.P. All authors have read and agreed to the published version of the manuscript.

Funding

The publication of this paper was funded by Sunway University to support research projects in the Centre for Virus and Vaccine Research (CVVR) for the year 2019.

Acknowledgments

Figures in this manuscript have been created with biorender.com (Toronto, ON, Canada).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Samanta, A.; Das, G.; Das, S.K. Roles of flavonoids in plants. Int. J. Pharm. Sci. Technol. 2011, 100, 12–35. [Google Scholar]
  2. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An overview. J. Nutr. Sci. 2016, 5, e47. [Google Scholar] [CrossRef] [Green Version]
  3. Cutting, W.; Dreisbach, R.; Neff, B. Antiviral chemotherapy; flavones and related compounds. Stanf. Med. Bull. 1949, 7, 137. [Google Scholar]
  4. Veckenstedt, A.; Pusztai, R. Mechanism of antiviral action of quercetin against cardiovirus infection in mice. Antivir. Res. 1981, 1, 249–261. [Google Scholar] [CrossRef]
  5. Roschek, B., Jr.; Fink, R.C.; McMichael, M.D.; Li, D.; Alberte, R.S. Elderberry flavonoids bind to and prevent H1N1 infection in vitro. Phytochemistry 2009, 70, 1255–1261. [Google Scholar] [CrossRef]
  6. Wu, W.; Li, R.; Li, X.; He, J.; Jiang, S.; Liu, S.; Yang, J. Quercetin as an antiviral agent inhibits influenza A virus (IAV) entry. Viruses 2016, 8, 6. [Google Scholar] [CrossRef]
  7. Choi, H.J.; Song, J.H.; Park, K.S.; Kwon, D.H. Inhibitory effects of quercetin 3-rhamnoside on influenza A virus replication. Eur. J. Pharm. Sci. 2009, 37, 329–333. [Google Scholar] [CrossRef]
  8. Bachmetov, L.; Gal-Tanamy, M.; Shapira, A.; Vorobeychik, M.; Giterman-Galam, T.; Sathiyamoorthy, P.; Golan-Goldhirsh, A.; Benhar, I.; Tur-Kaspa, R.; Zemel, R. Suppression of hepatitis C virus by the flavonoid quercetin is mediated by inhibition of NS3 protease activity. J. Viral. Hepat. 2012, 19, e81–e88. [Google Scholar] [CrossRef] [PubMed]
  9. Lu, N.; Khachatoorian, R.; French, S.W. Quercetin: Bioflavonoids as part of interferon-free hepatitis C therapy? Expert Rev. Anti Infect. 2012, 10, 619–621. [Google Scholar] [CrossRef] [PubMed]
  10. Qiu, X.; Kroeker, A.; He, S.; Kozak, R.; Audet, J.; Mbikay, M.; Chrétien, M. Prophylactic efficacy of quercetin 3-β-Od-glucoside against Ebola virus infection. Antimicrob. Agents Chemother. 2016, 60, 5182–5188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Tao, J.; Hu, Q.; Yang, J.; Li, R.; Li, X.; Lu, C.; Chen, C.; Wang, L.; Shattock, R.; Ben, K. In vitro anti-HIV and-HSV activity and safety of sodium rutin sulfate as a microbicide candidate. Antivir. Res. 2007, 75, 227–233. [Google Scholar] [CrossRef] [PubMed]
  12. Shibata, C.; Ohno, M.; Otsuka, M.; Kishikawa, T.; Goto, K.; Muroyama, R.; Kato, N.; Yoshikawa, T.; Takata, A.; Koike, K. The flavonoid apigenin inhibits hepatitis C virus replication by decreasing mature microRNA122 levels. Virology 2014, 462, 42–48. [Google Scholar] [CrossRef] [PubMed]
  13. Li, B.Q.; Fu, T.; Dongyan, Y.; Mikovits, J.A.; Ruscetti, F.W.; Wang, J.M. Flavonoid baicalin inhibits HIV-1 infection at the level of viral entry. Biochem. Biophys. Res. Commun. 2000, 276, 534–538. [Google Scholar] [CrossRef] [PubMed]
  14. Moghaddam, E.; Teoh, B.T.; Sam, S.S.; Lani, R.; Hassandarvish, P.; Chik, Z.; Yueh, A.; Abubakar, S.; Zandi, K. Baicalin, a metabolite of baicalein with antiviral activity against dengue virus. Sci. Rep. 2014, 4, 5452. [Google Scholar] [CrossRef] [Green Version]
  15. Nayak, M.K.; Agrawal, A.S.; Bose, S.; Naskar, S.; Bhowmick, R.; Chakrabarti, S.; Sarkar, S.; Chawla-Sarkar, M. Antiviral activity of baicalin against influenza virus H1N1-pdm09 is due to modulation of NS1-mediated cellular innate immune responses. J. Antimicrob. Chemother. 2014, 69, 1298–1310. [Google Scholar] [CrossRef] [Green Version]
  16. Chu, M.; Xu, L.; Zhang, M.-B.; Chu, Z.-Y.; Wang, Y.-D. Role of Baicalin in anti-influenza virus A as a potent inducer of IFN-gamma. Biomed. Res. Int. 2015, 2015, 11. [Google Scholar] [CrossRef] [Green Version]
  17. Johari, J.; Kianmehr, A.; Mustafa, M.R.; Abubakar, S.; Zandi, K. Antiviral activity of baicalein and quercetin against the Japanese encephalitis virus. Int. J. Mol. Sci. 2012, 13, 16785–16795. [Google Scholar] [CrossRef]
  18. Hassandarvish, P.; Rothan, H.A.; Rezaei, S.; Yusof, R.; Abubakar, S.; Zandi, K. In silico study on baicalein and baicalin as inhibitors of dengue virus replication. RSC Adv. 2016, 6, 31235–31247. [Google Scholar] [CrossRef]
  19. Sauter, D.; Schwarz, S.; Wang, K.; Zhang, R.; Sun, B.; Schwarz, W. Genistein as antiviral drug against HIV ion channel. Planta Med. 2014, 80, 682–687. [Google Scholar] [CrossRef]
  20. Miki, K.; Nagai, T.; Suzuki, K.; Tsujimura, R.; Koyama, K.; Kinoshita, K.; Furuhata, K.; Yamada, H.; Takahashi, K. Anti-influenza virus activity of biflavonoids. Bioorg. Med. Chem. Lett. 2007, 17, 772–775. [Google Scholar] [CrossRef]
  21. Lani, R.; Hassandarvish, P.; Shu, M.-H.; Phoon, W.H.; Chu, J.J.H.; Higgs, S.; Vanlandingham, D.; Bakar, S.A.; Zandi, K. Antiviral activity of selected flavonoids against Chikungunya virus. Antivir. Res. 2016, 133, 50–61. [Google Scholar] [CrossRef] [PubMed]
  22. Zandi, K.; Teoh, B.-T.; Sam, S.-S.; Wong, P.-F.; Mustafa, M.R.; AbuBakar, S. Antiviral activity of four types of bioflavonoid against dengue virus type-2. Virol. J. 2011, 8, 560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Jeong, H.J.; Ryu, Y.B.; Park, S.-J.; Kim, J.H.; Kwon, H.-J.; Kim, J.H.; Park, K.H.; Rho, M.-C.; Lee, W.S. Neuraminidase inhibitory activities of flavonols isolated from Rhodiola rosea roots and their in vitro anti-influenza viral activities. Bioorg. Med. Chem. 2009, 17, 6816–6823. [Google Scholar] [CrossRef] [PubMed]
  24. Schwarz, S.; Sauter, D.; Wang, K.; Zhang, R.; Sun, B.; Karioti, A.; Bilia, A.R.; Efferth, T.; Schwarz, W. Kaempferol derivatives as antiviral drugs against the 3a channel protein of coronavirus. Planta Med. 2014, 80, 177–182. [Google Scholar] [CrossRef] [Green Version]
  25. Yi, L.; Li, Z.; Yuan, K.; Qu, X.; Chen, J.; Wang, G.; Zhang, H.; Luo, H.; Zhu, L.; Jiang, P. Small molecules blocking the entry of severe acute respiratory syndrome coronavirus into host cells. J. Virol. 2004, 78, 11334–11339. [Google Scholar] [CrossRef] [Green Version]
  26. Yan, H.; Ma, L.; Wang, H.; Wu, S.; Huang, H.; Gu, Z.; Jiang, J.; Li, Y. Luteolin decreases the yield of influenza A virus in vitro by interfering with the coat protein I complex expression. J. Nat. Med. 2019, 73, 487–496. [Google Scholar] [CrossRef]
  27. Nagai, T.; Miyaichi, Y.; Tomimori, T.; Suzuki, Y.; Yamada, H. Inhibition of influenza virus sialidase and anti-influenza virus activity by plant flavonoids. Antivir. Res. 1990, 38, 1329–1332. [Google Scholar] [CrossRef] [Green Version]
  28. Nahmias, Y.; Goldwasser, J.; Casali, M.; van Poll, D.; Wakita, T.; Chung, R.T.; Yarmush, M.L. Apolipoprotein B–dependent hepatitis C virus secretion is inhibited by the grapefruit flavonoid naringenin. Hepatology 2008, 47, 1437–1445. [Google Scholar] [CrossRef] [Green Version]
  29. Frabasile, S.; Koishi, A.C.; Kuczera, D.; Silveira, G.F.; Verri, W.A., Jr.; Dos Santos, C.N.D.; Bordignon, J. The citrus flavanone naringenin impairs dengue virus replication in human cells. Sci. Rep. 2017, 7, 41864. [Google Scholar] [CrossRef]
  30. Keivan, Z.; Boon-Teong, T.; Sing-Sin, S.; Pooi-Fong, W.; Mustafa, M.R.; Sazaly, A. In vitro antiviral activity of fisetin, rutin and naringenin against dengue virus type-2. J. Med. Plants Res. 2011, 5, 5534–5539. [Google Scholar]
  31. Ferenci, P.; Scherzer, T.M.; Kerschner, H.; Rutter, K.; Beinhardt, S.; Hofer, H.; Schöniger–Hekele, M.; Holzmann, H.; Steindl–Munda, P. Silibinin is a potent antiviral agent in patients with chronic hepatitis C not responding to pegylated interferon/ribavirin therapy. Gastroenterology 2008, 135, 1561–1567. [Google Scholar] [CrossRef]
  32. Lin, A.-S.; Shibano, M.; Nakagawa-Goto, K.; Tokuda, H.; Itokawa, H.; Morris-Natschke, S.L.; Lee, K.-H. Cancer preventive agents. 7. Antitumor-promoting effects of seven active flavonolignans from milk thistle (Silybum marianum.) on epstein-barr virus activation. Pharm. Biol. 2007, 45, 735–738. [Google Scholar] [CrossRef]
  33. Camini, F.C.; da Silva, T.F.; da Silva Caetano, C.C.; Almeida, L.T.; Ferraz, A.C.; Vitoreti, V.M.A.; de Mello Silva, B.; de Queiroz Silva, S.; de Magalhães, J.C.; de Brito Magalhães, C.L. Antiviral activity of silymarin against Mayaro virus and protective effect in virus-induced oxidative stress. Antivir. Res. 2018, 158, 8–12. [Google Scholar] [CrossRef]
  34. Wagoner, J.; Negash, A.; Kane, O.J.; Martinez, L.E.; Nahmias, Y.; Bourne, N.; Owen, D.M.; Grove, J.; Brimacombe, C.; McKeating, J.A. Multiple effects of silymarin on the hepatitis C virus lifecycle. J. Hepatol. 2010, 51, 1912–1921. [Google Scholar] [CrossRef] [Green Version]
  35. Lani, R.; Hassandarvish, P.; Chiam, C.W.; Moghaddam, E.; Chu, J.J.H.; Rausalu, K.; Merits, A.; Higgs, S.; Vanlandingham, D.; Bakar, S.A. Antiviral activity of silymarin against chikungunya virus. Sci. Rep. 2015, 5, 11421. [Google Scholar] [CrossRef] [Green Version]
  36. Qaddir, I.; Rasool, N.; Hussain, W.; Mahmood, S. Computer-aided analysis of phytochemicals as potential dengue virus inhibitors based on molecular docking, ADMET and DFT studies. J. Vector Borne Dis. 2017, 54, 255. [Google Scholar]
  37. Song, J.; Choi, H. Silymarin efficacy against influenza A virus replication. Phytomedicine 2011, 18, 832–835. [Google Scholar] [CrossRef]
  38. Song, J.M.; Lee, K.H.; Seong, B.L. Antiviral effect of catechins in green tea on influenza virus. Antivir. Res. 2005, 68, 66–74. [Google Scholar] [CrossRef]
  39. Matsumoto, K.; Yamada, H.; Takuma, N.; Niino, H.; Sagesaka, Y.M. Effects of Green Tea Catechins and Theanine on Preventing Influenza Infection among Healthcare Workers: A Randomized Controlled Trial. BMC Complem. Altern. Med. 2011, 11, 15. [Google Scholar] [CrossRef]
  40. Imanishi, N.; Tuji, Y.; Katada, Y.; Maruhashi, M.; Konosu, S.; Mantani, N.; Terasawa, K.; Ochiai, H. Additional inhibitory effect of tea extract on the growth of influenza A and B viruses in MDCK cells. J. Microbiol. Immunol. 2002, 46, 491–494. [Google Scholar] [CrossRef]
  41. Kawai, K.; Tsuno, N.H.; Kitayama, J.; Okaji, Y.; Yazawa, K.; Asakage, M.; Hori, N.; Watanabe, T.; Takahashi, K.; Nagawa, H. Epigallocatechin gallate, the main component of tea polyphenol, binds to CD4 and interferes with gp120 binding. J. Allergy Clin. Immunol. 2003, 112, 951–957. [Google Scholar] [CrossRef]
  42. Isaacs, C.E.; Wen, G.Y.; Xu, W.; Jia, J.H.; Rohan, L.; Corbo, C.; Di Maggio, V.; Jenkins, E.C.; Hillier, S. Epigallocatechin gallate inactivates clinical isolates of herpes simplex virus. Antimicrob. Agents Chemother. 2008, 52, 962–970. [Google Scholar] [CrossRef] [Green Version]
  43. Weber, C.; Sliva, K.; von Rhein, C.; Kümmerer, B.M.; Schnierle, B.S. The green tea catechin, epigallocatechin gallate inhibits chikungunya virus infection. Antivir. Res. 2015, 113, 1–3. [Google Scholar] [CrossRef] [PubMed]
  44. Chang, L.-K.; Wei, T.-T.; Chiu, Y.-F.; Tung, C.-P.; Chuang, J.-Y.; Hung, S.-K.; Li, C.; Liu, S.-T. Inhibition of Epstein–Barr virus lytic cycle by (−)-epigallocatechin gallate. Biochem. Biophys. Res. Commun. 2003, 301, 1062–1068. [Google Scholar] [CrossRef]
  45. He, W.; Li, L.-X.; Liao, Q.-J.; Liu, C.-L.; Chen, X.-L. Epigallocatechin gallate inhibits HBV DNA synthesis in a viral replication-inducible cell line. World J. Gastroenterol. 2011, 17, 1507. [Google Scholar] [CrossRef] [PubMed]
  46. Carneiro, B.M.; Batista, M.N.; Braga, A.C.S.; Nogueira, M.L.; Rahal, P. The green tea molecule EGCG inhibits Zika virus entry. Virology 2016, 496, 215–218. [Google Scholar] [CrossRef]
  47. Zhong, L.; Hu, J.; Shu, W.; Gao, B.; Xiong, S. Epigallocatechin-3-gallate opposes HBV-induced incomplete autophagy by enhancing lysosomal acidification, which is unfavorable for HBV replication. Cell Death Dis. 2015, 6, e1770. [Google Scholar] [CrossRef] [Green Version]
  48. Nagai, T.; Moriguchi, R.; Suzuki, Y.; Tomimori, T.; Yamada, H. Mode of action of the anti-influenza virus activity of plant flavonoid, 5, 7, 4′-trihydroxy-8-methoxyflavone, from the roots of Scutellaria baicalensis. Antivir. Res. 1995, 26, 11–25. [Google Scholar] [CrossRef]
  49. Ahmadi, A.; Hassandarvish, P.; Lani, R.; Yadollahi, P.; Jokar, A.; Bakar, S.A.; Zandi, K. Inhibition of chikungunya virus replication by hesperetin and naringenin. RSC Adv. 2016, 6, 69421–69430. [Google Scholar] [CrossRef]
  50. Ganesan, S.; Faris, A.N.; Comstock, A.T.; Wang, Q.; Nanua, S.; Hershenson, M.B.; Sajjan, U.S. Quercetin inhibits rhinovirus replication in vitro and in vivo. Antivir. Res. 2012, 94, 258–271. [Google Scholar] [CrossRef]
  51. Gonzalez, M.; Martínez-Abarca, F.; Carrasco, L. Flavonoids: Potent inhibitors of poliovirus RNA synthesis. Antivir. Chem. Chemother. 1990, 1, 203–209. [Google Scholar] [CrossRef]
  52. Vrijsen, R.; Everaert, L.; Van Hoof, L.; Vlietinck, A.; Berghe, D.V.; Boeye, A. The poliovirus-induced shut-off of cellular protein synthesis persists in the presence of 3-methylquercetin, a flavonoid which blocks viral protein and RNA synthesis. Antivir. Res. 1987, 7, 35–42. [Google Scholar] [CrossRef]
  53. Ishitsuka, H.; Ohsawa, C.; Ohiwa, T.; Umeda, I.; Suhara, Y. Antipicornavirus flavone Ro 09-0179. Antimicrob. Agents Chemother. 1982, 22, 611–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Galochkina, A.V.; Anikin, V.B.; Babkin, V.A.; Ostrouhova, L.A.; Zarubaev, V.V. Virus-inhibiting activity of dihydroquercetin, a flavonoid from Larix sibirica, against coxsackievirus B4 in a model of viral pancreatitis. Arch. Virol. 2016, 161, 929–938. [Google Scholar] [CrossRef]
  55. Xu, L.; Su, W.; Jin, J.; Chen, J.; Li, X.; Zhang, X.; Sun, M.; Sun, S.; Fan, P.; An, D. Identification of luteolin as enterovirus 71 and coxsackievirus A16 inhibitors through reporter viruses and cell viability-based screening. Viruses 2014, 6, 2778–2795. [Google Scholar] [CrossRef] [Green Version]
  56. Vrijsen, R.; Everaert, L.; Boeyé, A. Antiviral activity of flavones and potentiation by ascorbate. J. Gen. Virol. 1988, 69, 1749–1751. [Google Scholar] [CrossRef]
  57. Wang, C.-Y.; Huang, S.-C.; Lai, Z.-R.; Ho, Y.-L.; Jou, Y.-J.; Kung, S.-H.; Zhang, Y.; Chang, Y.-S.; Lin, C.-W. Eupafolin and ethyl acetate fraction of Kalanchoe gracilis stem extract show potent antiviral activities against enterovirus 71 and coxsackievirus A16. Evid. Based Complement. Altern. Med. 2013, 2013. [Google Scholar] [CrossRef] [Green Version]
  58. Robin, V.; Irurzun, A.; Amoros, M.; Boustie, J.; Carrasco, L. Antipoliovirus flavonoids from Psiadia dentata. Antivir. Chem. Chemother. 2001, 12, 283–291. [Google Scholar] [CrossRef]
  59. Kim, N.; Park, S.; Nhiem, N.X.; Song, J.-H.; Ko, H.-J.; Kim, S.H. Cycloartane-type triterpenoid derivatives and a flavonoid glycoside from the burs of Castanea crenata. Phytochemistry 2019, 158, 135–141. [Google Scholar] [CrossRef]
  60. Zhang, Q.; Yang, Q.; Yang, W.; Wu, K.; Wu, J. Virucidal Capacity of Novel ProtecTeaV Sanitizer Formulations Containing Lipophilic Epigallocatechin-3-Gallate (EGCG). Antivir. Antiretrovir. Res. 2016, 1, 1–5. [Google Scholar]
  61. Kant, K.; Lal, U.R.; Ghosh, M. In-silico discovery of natural lead hits from the genus of arisaema against Human Rhino Virus. In Proceedings of the 20th International Electronic Conference on Synthetic Organic Chemistry, 1–31 November 2016. [Google Scholar]
  62. Choi, H.-J. In Vitro Antiviral Activity of Sakuranetin against Human Rhinovirus 3. Osong Public Health Res. Perspect. 2017, 8, 415. [Google Scholar] [CrossRef] [PubMed]
  63. Desideri, N.; Conti, C.; Sestili, I.; Tomao, P.; Stein, M.L.; Orsi, N. In vitro Evaluation of the Anti-Picornavirus Activities of New Synthetic Flavonoids. Antivir. Chem. Chemother. 1995, 6, 298–306. [Google Scholar] [CrossRef]
  64. Li, Y.; Leung, K.-T.; Yao, F.; Ooi, L.S.; Ooi, V.E. Antiviral Flavans from the Leaves of Pithecellobium C lypearia. J. Nat. Prod. 2006, 69, 833–835. [Google Scholar] [CrossRef] [PubMed]
  65. Semple, S.; Nobbs, S.; Pyke, S.; Reynolds, G.; Flower, R. Antiviral flavonoid from Pterocaulon sphacelatum, an Australian Aboriginal medicine. J. Ethnopharmacol. 1999, 68, 283–288. [Google Scholar] [CrossRef]
  66. Van, N.T.H.; Vien, T.A.; Nhiem, N.X.; Van Kiem, P.; Van Minh, C.; Long, P.Q.; Anh, L.T.; Cuong, N.M.; Song, J.H.; Ko, H.J. Chemical components of Ardisia splendens leaves and their activity against Coxsackie A16 viruses. Nat. Prod. Commun. 2014, 9. [Google Scholar] [CrossRef] [Green Version]
  67. Ortega, J.T.; Serrano, M.L.; Suárez, A.I.; Baptista, J.; Pujol, F.H.; Cavallaro, L.V.; Campos, H.R.; Rangel, H.R. Antiviral activity of flavonoids present in aerial parts of Marcetia taxifolia against Hepatitis B virus, Poliovirus, and Herpes Simplex Virus in vitro. Excli J. 2019, 18, 1037. [Google Scholar]
  68. Gunaseelan, S.; Wong, K.Z.; Min, N.; Sun, J.; Ismail, N.K.B.M.; Tan, Y.J.; Lee, R.C.H.; Chu, J.J.H. Prunin suppresses viral IRES activity and is a potential candidate for treating enterovirus A71 infection. Sci. Transl. Med. 2019, 11. [Google Scholar] [CrossRef]
  69. Solomon, T.; Lewthwaite, P.; Perera, D.; Cardosa, M.J.; McMinn, P.; Ooi, M.H. Virology, epidemiology, pathogenesis, and control of enterovirus 71. Lancet Infect. Dis. 2010, 10, 778–790. [Google Scholar] [CrossRef]
  70. Zhang, W.; Qiao, H.; Lv, Y.; Wang, J.; Chen, X.; Hou, Y.; Tan, R.; Li, E. Apigenin inhibits enterovirus-71 infection by disrupting viral RNA association with trans-acting factors. PLoS ONE 2014, 9, e110429. [Google Scholar] [CrossRef]
  71. Lv, X.; Qiu, M.; Chen, D.; Zheng, N.; Jin, Y.; Wu, Z. Apigenin inhibits enterovirus 71 replication through suppressing viral IRES activity and modulating cellular JNK pathway. Antivir. Res. 2014, 109, 30–41. [Google Scholar] [CrossRef]
  72. Dai, W.; Bi, J.; Li, F.; Wang, S.; Huang, X.; Meng, X.; Sun, B.; Wang, D.; Kong, W.; Jiang, C. Antiviral Efficacy of Flavonoids against Enterovirus 71 Infection in Vitro and in Newborn Mice. Viruses 2019, 11, 625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Li, X.; Liu, Y.; Wu, T.; Jin, Y.; Cheng, J.; Wan, C.; Qian, W.; Xing, F.; Shi, W. The antiviral effect of baicalin on enterovirus 71 in vitro. Viruses 2015, 7, 4756–4771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Wang, J.; Zhang, T.; Du, J.; Cui, S.; Yang, F.; Jin, Q. Anti-enterovirus 71 effects of chrysin and its phosphate ester. PLoS ONE 2014, 9, e89668. [Google Scholar] [CrossRef] [Green Version]
  75. Lin, Y.-J.; Chang, Y.-C.; Hsiao, N.-W.; Hsieh, J.-L.; Wang, C.-Y.; Kung, S.-H.; Tsai, F.-J.; Lan, Y.-C.; Lin, C.-W. Fisetin and rutin as 3C protease inhibitors of enterovirus A71. J. Virol. Methods 2012, 182, 93–98. [Google Scholar] [CrossRef]
  76. Wang, H.; Zhang, D.; Ge, M.; Li, Z.; Jiang, J.; Li, Y. Formononetin inhibits enterovirus 71 replication by regulating COX-2/PGE 2 expression. Virol. J. 2015, 12, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Li, G.; Gao, Q.; Yuan, S.; Wang, L.; Altmeyer, R.; Lan, K.; Yin, F.; Zou, G. Characterization of three small molecule inhibitors of enterovirus 71 identified from screening of a library of natural products. Antivir. Res. 2017, 143, 85–96. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, J.; Su, H.; Zhang, T.; Du, J.; Cui, S.; Yang, F.; Jin, Q. Inhibition of Enterovirus 71 replication by 7-hydroxyflavone and diisopropyl-flavon7-yl Phosphate. PLoS ONE 2014, 9, e92565. [Google Scholar] [CrossRef]
  79. Tsai, F.-J.; Lin, C.-W.; Lai, C.-C.; Lan, Y.-C.; Lai, C.-H.; Hung, C.-H.; Hsueh, K.-C.; Lin, T.-H.; Chang, H.C.; Wan, L. Kaempferol inhibits enterovirus 71 replication and internal ribosome entry site (IRES) activity through FUBP and HNRP proteins. Food Chem. 2011, 128, 312–322. [Google Scholar] [CrossRef]
  80. Zhu, Q.-C.; Wang, Y.; Liu, Y.-P.; Zhang, R.-Q.; Li, X.; Su, W.-H.; Long, F.; Luo, X.-D.; Peng, T. Inhibition of enterovirus 71 replication by chrysosplenetin and penduletin. Eur. J. Pharm. Sci. 2011, 44, 392–398. [Google Scholar] [CrossRef]
  81. Min, N.; Leong, P.T.; Lee, R.C.H.; Khuan, J.S.E.; Chu, J.J.H. A flavonoid compound library screen revealed potent antiviral activity of plant-derived flavonoids on human enterovirus A71 replication. Antivir. Res. 2018, 150, 60–68. [Google Scholar] [CrossRef]
  82. Yao, C.; Xi, C.; Hu, K.; Gao, W.; Cai, X.; Qin, J.; Lv, S.; Du, C.; Wei, Y. Inhibition of enterovirus 71 replication and viral 3C protease by quercetin. Virol. J. 2018, 15, 116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Wang, C.; Wang, P.; Chen, X.; Wang, W.; Jin, Y. Saururus chinensis (Lour.) Baill blocks enterovirus 71 infection by hijacking MEK1–ERK signaling pathway. Antivir. Res. 2015, 119, 47–56. [Google Scholar] [CrossRef] [PubMed]
  84. Choi, H.J.; Song, H.-H.; Lee, J.-S.; Ko, H.-J.; Song, J.-H. Inhibitory effects of norwogonin, oroxylin A, and mosloflavone on enterovirus 71. Biomol. Ther. 2016, 24, 552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Yin, Y.; Xu, Y.; Ou, Z.; Yang, X.; Liu, H. An antiviral drug screening system for enterovirus 71 based on an improved plaque assay: A potential high-throughput method. J. Med. Virol. 2019, 91, 1440–1447. [Google Scholar] [CrossRef]
  86. Zhang, D.; Ji, X.; Gao, R.; Wang, H.; Meng, S.; Zhong, Z.; Li, Y.; Jiang, J.; Li, Z. Synthesis and antiviral activities of a novel class of thioflavone and flavonoid analogues. Acta Pharm. Sin. B 2012, 2, 575–580. [Google Scholar] [CrossRef] [Green Version]
  87. Amic, D.; Davidovic-Amic, D.; Beslo, D.; Rastija, V.; Lucic, B.; Trinajstic, N. SAR and QSAR of the antioxidant activity of flavonoids. Curr. Med. Chem. 2007, 14, 827–845. [Google Scholar] [CrossRef]
  88. Sarwar, M.W.; Riaz, A.; Dilshad, S.M.R.; Al-Qahtani, A.; Nawaz-Ul-Rehman, M.S.; Mubin, M. Structure activity relationship (SAR) and quantitative structure activity relationship (QSAR) studies showed plant flavonoids as potential inhibitors of dengue NS2B-NS3 protease. BMC Struct. Biol. 2018, 18, 6. [Google Scholar] [CrossRef] [Green Version]
  89. Wang, T.-Y.; Li, Q.; Bi, K.-S. Bioactive flavonoids in medicinal plants: Structure, activity and biological fate. Asian J. Pharm. Sci. 2018, 13, 12–23. [Google Scholar] [CrossRef]
  90. Saraf, S. Applications of novel drug delivery system for herbal formulations. Fitoterapia 2010, 81, 680–689. [Google Scholar]
Figure 1. The antiviral activity of flavonoids in the life cycle of EV-A71. Flavonoids are reported to exhibit antiviral activity against EV-A71 and they are categorized according to the mechanism of inhibition at different stages of the virus life cycle. NR: not reported (Flavonoids that affect uncoating, assembly and release of EV-A71 are not reported). HnRNP = heterogeneous nuclear ribonucleoproteins, IRES = internal ribosome entry site, IFN = interferon, JAK = janus kinase, TYK = tyrosine kinase, and STAT = signal transducer and activator of transcription proteins.
Figure 1. The antiviral activity of flavonoids in the life cycle of EV-A71. Flavonoids are reported to exhibit antiviral activity against EV-A71 and they are categorized according to the mechanism of inhibition at different stages of the virus life cycle. NR: not reported (Flavonoids that affect uncoating, assembly and release of EV-A71 are not reported). HnRNP = heterogeneous nuclear ribonucleoproteins, IRES = internal ribosome entry site, IFN = interferon, JAK = janus kinase, TYK = tyrosine kinase, and STAT = signal transducer and activator of transcription proteins.
Viruses 12 00184 g001
Figure 2. Inhibition of molecular signaling pathways involved in EV-A71 infection by flavonoids. Flavonoids are classified according to their ability at the molecular level to inhibit signaling pathways involved in EV-A71-induced apoptosis, inflammation and infection associated cytokine production. GDP = guanosine diphosphate, GTP = guanosine triphosphate, RAF = rapidly accelerated fibrosarcoma, MEK/MAPK = mitogen-activated protein kinase, ERK = extracellular signal-regulated kinase, SRSF/MnK2a = serine/threonine-protein kinase. JNK = c-Jun N-terminal kinase, ASK = apoptosis signal-regulating kinase, PEG = prostaglandins, COX = cyclooxygenase, and FasL = Fas ligand.
Figure 2. Inhibition of molecular signaling pathways involved in EV-A71 infection by flavonoids. Flavonoids are classified according to their ability at the molecular level to inhibit signaling pathways involved in EV-A71-induced apoptosis, inflammation and infection associated cytokine production. GDP = guanosine diphosphate, GTP = guanosine triphosphate, RAF = rapidly accelerated fibrosarcoma, MEK/MAPK = mitogen-activated protein kinase, ERK = extracellular signal-regulated kinase, SRSF/MnK2a = serine/threonine-protein kinase. JNK = c-Jun N-terminal kinase, ASK = apoptosis signal-regulating kinase, PEG = prostaglandins, COX = cyclooxygenase, and FasL = Fas ligand.
Viruses 12 00184 g002
Figure 3. General chemical structure of flavone nucleus.
Figure 3. General chemical structure of flavone nucleus.
Viruses 12 00184 g003
Table 1. Examples of antiviral flavonoids against non-picornaviruses.
Table 1. Examples of antiviral flavonoids against non-picornaviruses.
FlavonoidVirusVirus FamilyModelStage of Virus InhibitionSuggested MechanismReference
ApigeninHepatitis C virus (HCV)FlaviviridaeIn vitroHost factor modulationReduction in mature miRNA122[12]
BaicalinDengue virus-2 (DENV-2)FlaviviridaeIn vitroAttachmentBlockade of attachment of the virus to Vero cells[14]
Human immunodeficiency virus-1 (HIV-1)RetroviridaeIn vitroFusionInhibition of the fusion of virus envelope protein with T cells and monocytes expressing CD4/CXCR4 or CD4/CCR5[13]
Influenza A virusOrthomyxoviridaeIn vitroIndirect: Immune-mediated infection controlDirectly binds to NS1–p85β (RNA binding domain) to down-regulate IFN-ɣ and activates the JAK/STAT1 pathway that reduced the viral load[15,16]
BaicaleinChikungunya virus (CHIKV)TogaviridaeIn vitroProphylaxisInhibition of attachment by inhibiting extracellular particles such as nsP1, nsP3, and E2 proteins[21]
DENV-2FlaviviridaeIn silicoReplicationBinds to the NS3/NS2B and NS5 proteins[18]
Japanese encephalitis virus (JEV)FlaviviridaeIn vitroEntryUnknown
Postulated to be an accumulation of the compound in cells to prevent entry or interaction with structural and/or non-structural protein(s)
[17]
Epigallocatechin (ECG)Influenza A and B virusesOrthomyxoviridaeIn vitroReplicationAcidification of the lysosomal and endosomal environment through clathrin-mediated endocytosis[40]
Epigallocatechin gallate (EGGC)CHIKVTogaviridaeIn vitroEntryCompetitor for cellular co-receptors of target cells such as heparan sulfate or sialic acid[43]
Epstein–Barr virus (EBV)Herpesviridae ReplicationInhibition of Zta, Rta and EA-D genes by interrupting the MEK/ERK1/2 and PI3-K/Akt signaling pathway of the lytic cycle of a virus[44]
Hepatitis B virus (HBV)HepadnaviridaeIn vitroReplicationImpair the production of pre-core mRNA and replicative intermediates of DNA[45]
Acidification in lysosomes to make an unfavorable environment for virus replication[47]
Herpes simplex virus (HSV)HerpesviridaeIn vitroEntryBinds to glycoprotein B and D of virus[42]
HIV-1RetroviridaeIn vitroEntryDirectly binds to CD4+ T-cells and blocks binding of envelope protein gp120 to cells[41]
Zika virusFlaviviridaeIn vitroEntryInteraction with the lipid envelope of virus[46]
FisetinCHIKVTogaviridaeIn vitroReplicationInhibition of NS protein 1 and 3 and downregulation of E2 protein and its precursor pE2[21]
DENV-2FlaviviridaeIn vitroReplicationDirectly binds to the viral RNA to impede polymerases activity[22]
GenisteinHIVRetroviridaeIn vitroAssembly and releaseInhibition of Vpu protein involved in the formation of ion channels in infected cells[19]
GinkgetinInfluenza A virusOrthomyxoviridaeIn vitroAssembly and releaseInhibition of sialidase[20]
KaempferolCoronavirusCoronaviridaeIn vitroAssembly and releaseInhibition of the release of progeny virus by blocking 3a channels of virus[24]
Influenza A virusOrthomyxoviridaeIn silicoEntryInhibition of neuraminidase enzyme[23]
LuteolinInfluenza A virusOrthomyxoviridaeIn vitroEntryInteraction with hemagglutinins of virus[25]
Severe acute respiratory syndrome coronavirus (SARS-CoV)CoronaviridaeIn vitroEntryBinds to the S2 protein of virus[25]
Methoxyflavone, isoscutellarein, and 8-methoxy-isoscutellareinInfluenza A virusOrthomyxoviridaeIn vitroEarly replicationReduction in sialidase activity, lysosomal fusion and RNA polymerase activity[27,48]
NaringeninCHIKVTogaviridaeIn vitroReplicationReduction in RNA and proteins[49]
DENV-2 and 4FlaviviridaeIn vitroReplicationReduction in RNA levels[29,30]
HCVFlaviviridaeIn vitroReplicationInhibition of RNA and core protein[28]
QuercetinHCVFlaviviridaeIn vitroTranscriptionInactivation of the NS3 helicase and NS5 protease[8,9]
HSV-1, HSV-2, drug-resistant HSV-1HerpesviridaeIn vitroBinding and entryN/R[6]
Influenza A virusOrthomyxoviridaeIn vitroBinding and entryInhibition of neuraminidase activity by interaction with the viral subunit 2 of the hemagglutinin[6]
Rabies virusRhabdoviridaeIn vivoCell protectionN/R[3]
Vesicular stomatitis virus (VSV)RhabdoviridaeIn vivoIndirect: Immune-mediated infection controlActivation of macrophages[4]
Quercetin 3-β-O-d-glucosideEbola virusFiloviridaeIn vivoProphylaxisUnknown[10]
QuercitrinInfluenza A virusOrthomyxoviridaeIn vitroEarly replicationReduction in mRNA synthesis[7]
Rabies virusRhabdoviridaeIn vivoProphylaxisUnknown[3]
Rutin (sulfated)HIV-1RetroviridaeIn vitroFusionInhibition of glycoprotein-mediated cell-cell fusion[11]
HSVHerpesviridaeIn vitroAdsorptionUnknown[11]
SilibininHCVFlaviviridaePhase II Clinical trialReplication or immune-mediated infection controlUnknown.
Postulated to be IFN-JAK/STAT independent immune-mediated antiviral mechanisms such as Regulation by interferon regulatory factor 3, Toll-like receptor 7 and p38 protein kinase pathways
[31]
SilymarinCHIKVTogaviridaeIn vitroReplicationInhibition of viral proteins[35]
DENV-2FlaviviridaeIn silicoReplicationInhibition of NS4B protein[36]
EBVHerpesviridaeIn vitroEarly antigen inactivationN/R[32]
HCVFlaviviridaeIn vitroEntry and fusionInhibition of viral pseudoparticles (pp) fusion with liposomes[34]
Influenza A virusOrthomyxoviridaeIn vitroLate replicationInhibition of viral mRNA synthesis[37]
Mayaro virusTogaviridaeIn vitroReplicationInhibition of reactive oxygen species (ROS) and reduction in levels of malondialdehyde (MDA)[33]
Tea catechinsInfluenza A virusOrthomyxoviridaeIn vitroAdsorption and entryInteraction with hemagglutinins of virus[38]
Tetra-O-methyl quercetinInfluenza A virusOrthomyxoviridaeIn vitroEntryInteraction with mannose-rich hemagglutinin domains of virus[5]
N/R = not reported, NS = nonstructural. Structures of flavonoids are provided in Supplementary Materials Table S1.
Table 2. Examples of antiviral flavonoids against picornaviruses.
Table 2. Examples of antiviral flavonoids against picornaviruses.
FlavonoidPicornavirusModelStage of Virus InhibitionSuggested MechanismReference
3-MethylquercetinPoliovirusIn vitroLate replicationBlocked of genomic RNA synthesis[51]
Reduction in viral protein and RNA synthesis[52]
3-MethylkaempferolPoliovirus-1In vitroReplicationInhibition of positive-strand of viral RNA[58]
5,3′-Dihydroxy-3,6,7,8,4′-pentamethoxyflavone and 5-hydroxy-3,6,7,3′,4′-pentamethoxyflavonePoliovirus-1In vitroReplicationPostulated to be inhibition of cellular processes (apoptosis and downstream signaling pathways)[67]
5,7,4′-Trihydroxy-3′-methoxyflavoneRhinovirus
(HRV)
In silicoEntryInhibition by binding to human rhinovirus protein grid[61]
6-Chloro-4′-oxazolinylflavanonePoliovirus-2In vitroReplicationN/R[63]
HRV-1BIn vitroReplicationN/R[63]
7-O-galloyltricetifavan and 7,4′-di-OgalloyltricetifavanCoxsackievirus B3
(CV-B3)
In vitroN/RN/R[64]
Chrysosplenol CPoliovirusIn vitroReplicationN/R[65]
Desmanthin-1Coxsackieviruses A16
(CV-A16)
In vitroReplicationN/R[66]
DihydroquercetinCoxsackievirus B4
(CV-B4)
In vivoIndirect: Immune-mediated infection controlReduction in viral immune mediators (ROS-mediated signaling and oxidative stress[54]
Epigallocatechin-3-GallatePoliovirus-1In vitroVirucidal effect (irreversible)N/R[60]
EupafolinCV-A16In vitroAttachmentReduction in IL-6 and RANTES and inactivation of downstream signaling pathways (ERK1/2, c-Jun, and STAT3)[57]
Kaempferol-3-O-[2″,6″-di-O-Z-p-coumaroyl]-β-d-glucopyranoside and derivativesCV-B3 In vitroReplicationN/R[59]
HRV-1B In vitroReplicationN/R[59]
LuteolinCV-A16In vitroReplicationInhibition of viral RNA synthesis[55]
PoliovirusIn vitroReplicationN/R[56]
MyricitrinCV-A16In vitroReplicationN/R[66]
Pachypodol (RO 09-0179)CVIn vitroEarly replicationInterference with viral replications between the uncoating and RNA synthesis stage[53]
PoliovirusIn vitroLate replicationBlocked the synthesis of positive-strand RNA[51]
HRVIn vitroEarly replicationInterference with viral replications between the uncoating and RNA synthesis stage[53]
PruninEnteroviruses A and BIn vitro and in vivoTranslation and replicationInhibition of IRES activity and protein synthesis[68]
QuercetinEncephalomyocarditis virus (EMCV)In vivoIndirect: Immune-mediated infection controlActivation of macrophages[4]
Mengo virusIn vivoIndirect: Immune-mediated infection controlActivation of macrophages[4]
HRVIn vitroTranscription and translationReduction in endocytosis of virus and phosphorylation of Akt (effector of phosphoinositol 3-kinase). Repression of interferon and interleukin-8 response resulted in lower viral RNA and capsid protein production.[50]
HRVIn vivoIndirect: Immune mediated infection controlSuppression of viral immune mediators[50]
RO 09-0298CV-B1In vivoN/RN/R[53]
SakuranetinHRV-3In vitroReplicationAntioxidant activity through inhibition of viral adsorption[62]
Structures of flavonoids are provided in Supplementary Materials Table S1.
Table 3. Antiviral activity of flavonoids against Enterovirus A71 in newborn mice.
Table 3. Antiviral activity of flavonoids against Enterovirus A71 in newborn mice.
FlavonoidIn Vitro EC50 (µM)Lethal Dose of Challenge VirusIn Vivo Dose of FlavonoidSurvival RateDuration of TreatmentReference
Apigenin24.74600,000 TCID5050 mg/Kg88.89%Once a day for 7 days, starting from 2 h post-infection[72]
Chrysosplenetin0.68600,000 TCID505 and 1 mg/Kg30%Once a day for 7 days, starting from 2 h post-infection[72]
Formononetin12.5600,000 TCID5010 mg/Kg75%Once a day for 7 days, starting from 2 h post-infection[72]
Isorhamnetin60.7600,000 TCID5010 mg/Kg100%Once a day for 7 days, starting from 2 h post-infection[72]
Kaempferol52.75600,000 TCID5050 mg/Kg88.89%Once a day for 7 days, starting from 2 h post-infection[72]
Luteolin13.5600,000 TCID5010 mg/Kg91.67%Once a day for 7 days, starting from 2 h post-infection[72]
Penduletin0.63600,000 TCID505 mg/Kg66.67%Once a day for 7 days, starting from 2 h post-infection[72]
Prunin0.1152 × 107 PFU3 and 10 mg/Kg100%Once a day for 7 days, starting from 1 or 6 h post infection[68]
Quercetin1.2600,000 TCID5010 mg/Kg50%Once a day for 7 days, starting from 2 h post infection[72]
Note: Flavonoids were administered into the newborn mice by the intraperitoneal route.
Table 4. In vitro antiviral activities of flavonoids against Enterovirus A71.
Table 4. In vitro antiviral activities of flavonoids against Enterovirus A71.
FlavonoidStructureEV-A71 Strain * (Genotype/Subgenotype)Antiviral Activity/IC50Cytoxicity/CC50Reference
Apigenin Viruses 12 00184 i001Fuyang 0805
(C4a)
10.3 μM 79.0 μM
(RD cells)
[70]
Fuyang0805
(C4a)
BrCr (A)
Not reported>200 μM
(RD and Vero cells)
[71]
Baicalin Viruses 12 00184 i002BrCr (A)4.96 μg/mL823.53 µg/mL
(RD cells)
[73]
Chrysin
Ester of chrysin (CR)
Viruses 12 00184 i003SHZH-98
(C4)
C = 13.86 μM>200 μM
(RD cells)
[74]
CR = 24.12 μM>200 μM
(RD cells)
Chrysosplenetin Viruses 12 00184 i004GZ-08-02
(Accession # FJ360545)
0.17 μM
(Vero cells)
18.27 μM
(Vero cells)
[80]
0.20 μM
(RD cells)
13.90 μM
(RD cells)
Eupafolin Viruses 12 00184 i005Not reported0.44 µg/mL
(RD cells)
355.87 µg/mL
(RD cells)
[57]
Fisetin Viruses 12 00184 i006CMUH01
(B5)
85 μM>1000 μM
(RD cells)
[75]
Formononetin Viruses 12 00184 i007SHZH-98 (C4)
JS-52 (C4)
H
BrCr (A)
3.45–3.95 μM
17.87  ±  8.51 μM
11.11  ±  9.23 μM
6.47  ±  4.40 μM
149.38 μM
(Vero cells)
198.80 μM
(SK-N-SH cells)
[76]
Galangin Viruses 12 00184 i008C4bNot reportedNot reported[55]
Hesperetin Viruses 12 00184 i009Cmuh-050530-5
(Accession # HM807310)
Not reported>50 μM
(RD cells)
[79]
Hesperidin Viruses 12 00184 i010Cmuh-050530-5
(Accession # HM807310)
Not reported>50 μM
(RD cells)
[79]
Hydroxyflavone (HF) and its phosphate ester Viruses 12 00184 i011SHZH-98
(C4)
23.45 μM>200 μM
(RD cells)
[78]
Viruses 12 00184 i012SHZH-98
(C4)
13.63 μM>200 μM
(RD cells)
[78]
Kaempferol Viruses 12 00184 i013Cmuh-050530-5
(Accession # HM807310)
Not reported
6 log reduction at 24 hrs
>50 μM
(RD cells)
[79]
Luteolin Viruses 12 00184 i014C4b10 μM148.02 μM
(RDS cells)
[55]
292.00 μM
(RD cells)
Fuyang0805 (C4a)
BrCr (A)
Not reported178.65 μM
(Vero cells)
[71]
157 μM
(Vero cells)
200 μM
(RD cells)
Penduletin Viruses 12 00184 i015GZ-08-02
(Accession # FJ360545)
0.17 µM
(Vero cells)
111.46 µM
(Vero cells)
[80]
0.37 µM
(RD cells)
74.18 µM
(RD cells)
Peracetate pulicarine Viruses 12 00184 i0165865/sin/000009
(B4)
Not reported.>20 µg/mL
(RD cells)
[81]
5511-SIN-00
(B5)
2.5 Log reduction
Prunin Viruses 12 00184 i017EV-A71
clinical isolates, EV-A71 strains H, B5 and C4 genotypes
115.3 nM2715 nM
(RD cells)
[68]
Quercetagetin Viruses 12 00184 i0185865/sin/000009
(B4)
Not reported.>20 µg/mL
(RD cells)
[81]
5511-SIN-00
(B5)
3.5 Log reduction
Quercetin Viruses 12 00184 i019SK-EV006/Malaysia/97
(Accession # AB469182)
12.1 μM
(RD cells)
>200 μM
(RD and Vero cells)
[82]
8.8 μM
(Vero cells)
Rutin Viruses 12 00184 i020CMUH01
(B5)
110 μM>1000 μM
(RD cells)
[75]
Thio flavones
(Multiple)
4b
7d
7i
8b
9b
Viruses 12 00184 i021
4b = R1 = OCH3, R2 = H
Viruses 12 00184 i022
7d = R1 = 4-methoxyphenyl, R2 = H
7i = R1 = CH3, R2 = Cl
Viruses 12 00184 i023
8b = R = n-propyl
Viruses 12 00184 i024
9b = R1 = OH, R2 = H, R3 = OH, R4 = OH
SHZH-98
(C4)
4b = 16.9 μM4d = 29.23 μM[86]
7d = 8.27 μM7d = 107.34 μM
7i = 39.63 μM7i = 133.15 μM
8b = 100.86 μM8b = 174.41 μM
9b = 5.48 μM9b = 23.75 μM (Vero cells)
* Accession number is mentioned where genotype/subgenotype is not reported; IC50 is the inhibitory concentration of flavonoid required to cause 50% inhibition of virus; CC50 is the cytotoxic concentration of flavonoid required to cause 50% cell death.

Share and Cite

MDPI and ACS Style

Lalani, S.; Poh, C.L. Flavonoids as Antiviral Agents for Enterovirus A71 (EV-A71). Viruses 2020, 12, 184. https://doi.org/10.3390/v12020184

AMA Style

Lalani S, Poh CL. Flavonoids as Antiviral Agents for Enterovirus A71 (EV-A71). Viruses. 2020; 12(2):184. https://doi.org/10.3390/v12020184

Chicago/Turabian Style

Lalani, Salima, and Chit Laa Poh. 2020. "Flavonoids as Antiviral Agents for Enterovirus A71 (EV-A71)" Viruses 12, no. 2: 184. https://doi.org/10.3390/v12020184

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