Next Article in Journal
Durability of Immune Response after Application of a Third Dose of SARS-CoV-2 Vaccination in Liver Transplant Recipients
Previous Article in Journal
The Fourth Dose of mRNA COVID-19 Vaccine Following 12 Different Three-Dose Regimens: Safety and Immunogenicity to Omicron BA.4/BA.5
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

Hand, Foot, and Mouth Disease Challenges and Its Antiviral Therapeutics

1
Department of Epidemiology, College of Public Health, Zhengzhou University, Zhengzhou 450001, China
2
Henan Key Laboratory of Molecular Medicine, Zhengzhou University, Zhengzhou 450001, China
*
Author to whom correspondence should be addressed.
Vaccines 2023, 11(3), 571; https://doi.org/10.3390/vaccines11030571
Submission received: 20 January 2023 / Revised: 10 February 2023 / Accepted: 17 February 2023 / Published: 1 March 2023

1. Introduction

Hand, Foot, and Mouth Disease (HFMD) is an infectious disease caused by enteroviruses (EVs) and is extremely contagious and prevalent among infants and children under 5 years old [1]. Historically, outbreaks of HFMD have been caused primarily by enterovirus A71 (EVA71) and coxsackievirus A16 (CVA16). Recently, however, other EV serotypes (e.g., CVA10, CVA6, CVA2, CVA8) have been associated with an increasing number of HFMD outbreaks and sporadic cases globally. Similar to other viruses, the life cycle of EVs includes attachment, endocytosis, uncoating, translation, replication, assembly, and release (Figure 1) [2].
HFMD has caused several large outbreaks worldwide and has created a huge disease burden in the world, especially in the Asia-Pacific region [1,3,4]. The main manifestations of HFMD are fever, vesicular rashes on hand, feet and buttocks, and ulcers in the oral mucosa. Although usually mild and self-limiting, HFMD can occasionally cause severe complications associated with the central nervous system or fatal respiratory disease [5]. However, there are no effective antiviral drugs available for HFMD, and supportive care as well as symptomatic treatment are currently the main treatment strategies for critical cases [6]. Although three kinds of monovalent inactivated EVA71 vaccines have been licensed by China FDA, these vaccines cannot provide broad-spectrum protection for other EV serotypes [2,7]. Polyvalent vaccines and other types of vaccines (e.g., virus-like particles vaccines, viral protein subunit vaccines, recombinant VP1 and P1 vaccines) are currently under development [7,8]. Facing the severe situation of HFMD epidemic, it is urgent to develop antiviral drugs for HFMD. This Editorial summarizes the current developments in antiviral agents for HFMD and highlights the viable research directions of antiviral agents.

2. Recent Development in the Design of Antiviral Agents

To date, proteins and structures that play an important role in the life cycle of EVs have been widely studied as drug targets. An increasing number of compounds acting on these drug targets are being identified as potential antivirals for HFMD. The following is a summary of recent researches in this area.

2.1. Antiviral Agents Targeting the Five-Fold Axis

Virus-host receptor interaction is the first essential event during virus infection [9]. At least six host cellular receptors can be involved in this process, such as scavenger receptor class B member 2 (SCARB2), P-selectin glycoprotein ligand-1 (PSGL-1), nucleolin and cell surface heparan sulfate (HS). Conserved and positively charged amino acid clusters (such as Arg166 and Lys242) are symmetrically arranged near the five-fold axis and serve as binding sites for cellular receptors [10]. It has been discovered that these amino acids interact with drugs to effectively prevent viral attachment to the relevant receptors [11].
A group of negatively charged glycosaminoglycans, including heparin, polysulfated dextran sulfate and suramin, can prevent EVA71 infection. They interact with the positively charged amino acid cluster of EVA71 and affect the attachment of EVA71 to HS [12]. A peptide derived from EVA71 VP1, L-SP40 peptide, was found to exhibit significant antiviral activity against EVA71 by blocking viral attachment to nucleolin [13]. Moreover, it demonstrated 80% protection of neonatal mice against lethal challenge by EVA71 [13]. Sulforaphane derivative NF449, rosmarinic acid, and a series of tryptophan dendritic macromolecules specifically block EV interaction with PSGL-1 as well as HS and inhibit EV attachment to target cells [14,15,16]. Transmembrane protein 106A (TMEM106A) is a protein produced by interferon (IFN) stimulation. TMEM106A is required for IFN-mediated viral suppression and interferes with the binding of EVA71 to SCARB2 [17]. A number of sulphonated azo dyes, widely used as food additives, have been identified as having potent antiviral activity against EVs, Brilliant black BN (E151) being one of them. E151 interacts with the apex of the five-fold axis to prevent viral entry and is regulated by amino acids at VP1-98, 145, and 246 [18]. An extract of Ilex kaushue, 3,4-dicaffeinated quinic acid (3,4-DCQA), targets the five-fold axis to form a stable structure with the VP-98 and 246 residues through noncovalent and van der Waals interactions, specifically inhibiting the attachment of EVA71 to the host receptor HS [19].

2.2. Antiviral Agents Targeting the Hydrophobic Pocket

The circular depression around the five-fold axis is called canyon, which has been shown to bind to cellular receptors. When the cell receptor binds to the canyon, the receptor presses against the bottom of the canyon, which corresponds to the top of the pocket, leading the hydrophobic pocket to release pocket factor (PF). PF plays an important role in regulating particle stability, and the release of PF leads to viral uncoating and genome release, which is the switch initiating EV infection [20,21]. Similarly, when antiviral bind to the pocket, it also expands the bottom of the canyon and affects receptor binding [22]. Antivirals competing for the same binding site usually have a higher affinity than PF [23]. Therefore, antivirals can bind to the hydrophobic pocket and inhibit viral uncoating, whether the pocket contains PF or not [21].
The WIN series from Sterling-Winthrop is a prominent series of capsid-binding compounds. WIN compounds and their derivatives, including WIN54954, WIN51711, BPR0Z-194, pleconaril, vapendavir, and pirodavir, have anti-EVs capabilities [21,24,25,26,27]. Among these, the broad spectrum EVs inhibitor pleconaril is the most potent antiviral, with a 50% inhibition Concentration (IC50) of 0.014 μg/mL against coxsackievirus B3 (CVB3) infection [26]. However, pleconaril is ineffective against EVA71 [28]. Two new compounds, NLD and ALD, designed based on the crystallographic analysis of the structure of 3-(-4-pyridyl)-2-imidazolidinone derivatives GPP3, and show good activity against both EVA71 and CVA16. NLD and ALD inhibited EVA71 replication with IC50 of 0.025 and 8.54 nmol/L, making them the most potent EVA71 inhibitors reported so far [29]. An imidazolidinone derivative, PR66, protected EVA71-induced neurological symptoms in a mouse model [30].

2.3. Antiviral Agents Targeting Non-Structural Proteins of EVs

The 2A protein (2Apro) catalyzes the peptide bond cleavage between VP1 and P2. Through 2Apro, EVA71 cleaves eukaryotic translation initiation factor 4G (eIF4G) and hijacks the cellular translation machinery to produce viral proteins [31]. Meanwhile, it was found that blocking the RAS/RAF/MEK/ERK (MAPK) pathway would inhibit the hydrolytic activity of 2Apro [32]. Therefore, direct targeting of 2Apro and inhibiting the MAPK pathway are how these inhibitors exert antiviral effects.
A 6-amino-acid hit peptide LVLQTM, and a novel furoquinoline alkaloid CW-33, inhibit EVA71 activity by directly targeting 2Apro [33,34]. Sorafenib, a clinically approved anti-cancer multi-targeted kinase inhibitor, blocks the MAPK pathway and significantly inhibits the hydrolytic activity of EVA71 2Apro [35]. Researchers designed and synthesized a series of substituted 3-benzylcoumarins, two of which compounds 13 and 14, were more effective in inhibiting MAPK pathway activity and had stronger antiviral effects compared to sorafenib [31].
The 2Cpro is one of the highly conserved viral proteins in all picornaviruses and functions as ATPase, 3’-5’ ATPase-dependent RNA helicase and ATPase-dependent RNA chaperone. In addition, 2Cpro is believed to involve in the formation of replicative organelles (ROs) of the virus and innate immune evasion [36].
Mutations in the coding sequence of the CVB3 2Cpro conferred resistance to dibucaine, pirlindole, fluoxetine and zuclopenthixol, suggesting that 2Cpro is the target for this set of compounds [37]. A peptide named 2CL was designed based on the structure of EVA71 2Cpro, which effectively impaired the oligomerization of EVA71 2Cpro and inhibited the RNA helicase activities of 2Cpro encoded by EVA71 and CVA16 [38]. An aryl-substituted amide, R523062, inhibits multiple strains of EVA71. Thermal displacement binding assays showed that R523062 binds to 2Cpro [39]. Moreover, some 1H-pyrazolo[3,4-b]pyridine-4-carboxamide molecules and dibucaine derivatives are also considered to be effective 2Cpro inhibitors [40,41].
3Apro is a membrane-bound protein that plays an important role in membrane rearrangement by recruiting essential host factors to ROs [36]. Enviroxime is a broad-spectrum anti-EVs drug targeting 3Apro. However, the clinical development was halted due to the lack of therapeutic benefits and gastrointestinal side effects [42]. Subsequently, a series of enviroxime analogs were synthesized and used for anti-EVs. The best-known of these compounds are GW5074, TTP-8307, and T-00127-HEV2 (THEV) [43]. Some FDA-approved drugs (such as itraconazole and posaconazole) have also been found to be effective 3Apro inhibitors [44,45].
3Cpro is responsible for the proteolytic processing of the Gln-Gly peptide bond be- tween P2–P3, resulting in the production of nonstructural and structural proteins [46]. 3Cpro was found to possess RNA-binding activity. Mutations in 3Cpro influenced RNA-binding activity and proteolytic activity [36]. In addition, 3Cpro can regulate cell pyroptosis and apoptosis through different pathways [36].
Rupintrivir (AG7088) represents a typical peptidyl mimic, it has an α,β-unsaturated ester group as the functional group [47], which functions as a Michael acceptor for nucleophilic Cys residues in the catalytic center of 3Cpro and affects 3Cpro activity [48]. Molecular docking experiments confirmed that Quercetin and two cyanohydrin derivatives FOPMC, and FIOMC, bind non-covalently to 3Cpro and coordinate with surrounding amino acid residues through hydrophobic interactions [49,50]. NK-1.8k and NK-1.9k are two peptide aldehydes identified as potent and selective inhibitors targeting EV 3Cpro [47,51]. SLQ-4 and SLQ-5 are EVA71 3Cpro inhibitors with two strong electron-withdrawing groups (EWGs), the two EWGs confer excellent potency to the Michael receptor, allowing the drug to inhibit viral activity more effectively [46].
3Dpro, as an RNA-dependent RNA polymerase (RdRp), is responsible for RNA replication. EVA71 mediates cell cycle arrest in S-phase through 3Dpro. In addition, 3Dpro is involved in VPg uridylation and has been shown to play an important role in the activation of inflammatory responses as well as evasion of innate immunity [36]. Both NITD008 and FNC are nucleoside analogs that can effectively inhibit EVA71 by targeting and inhibiting 3Dpro activity [52,53]. However, as nucleoside analogs, these drugs may have many side effects. DTriP-22 was identified as a non-nucleoside analog that targeted 3Dpro. It inhibits the poly (U) elongation activity of EVA71 polymerase [54], but not the VPg uridylation activity. Differently, another 3Dpro inhibitor, BPR-3P0128, inhibited both of these activities [55].

2.4. Antiviral Agents Targeting Internal Ribosome Entry Site (IRES)

The 5′ untranslated region (5′UTR) is the most phylogenetically conserved portion of the EVs genome, it folds into a structured RNA element composed of six stem-loops. The first stem-loop (SLI) directs RNA strand synthesis, and stem-loop II-VI constitutes IRES, which promotes cap-independent translation by internally recruiting the 40S ribosomal subunits [56]. In contrast to host cap-dependent cellular translation, the genome of EVs lacks a 50 cap, so EVs initiate viral RNA translation through IRES. Highly structured IRESs have also been used to identify virus inhibitors [57].
A flavonoid, kaempferol, was found to affect IRES function and EVA71 replication by altering the composition of IRES trans-acting factors (ITAFs) [58]. It suggested that the mechanism of IRES inhibitors was generated by affecting the interaction of viral RNA with ITAFs. Apigenin interacts with ITAF heterogeneous nuclear ribonucleoproteins (hnRNPs) and interferes with viral RNA editing activity [59]. Multiple binding of Idarubicin to EVA71 5′UTR RNA impairs binding between EVA71 IRES RNA and the ITAF hnRNPA1, resulting in a significant reduction in IRES activity [60]. The cellular proteins, AU-rich element factor 1 (AUF1), bind to the bulge of SLII to repress IRES-dependent translation [61]. Pull-down experiments in cell culture support that the EVs inhibitor DMA-135 induces a conformational change in virus RNA structure that stabilizes a ternary complex with the AUF1 protein, thus repressing translation [62].

3. Prospects and Summary

The functions of proteins and structures that act as drug targets have been extensively studied. With the advent of novel drug screening methods, as well as drug synthesis methods, more and more drugs have been recognized as broad-spectrum inhibitors of EVs. Unfortunately, most drugs will fail at some stages of clinical development due to off-target effects and/or lack of antiviral efficacy [63]. Childhood infections with milder symptoms are the main group of HFMD [5]. Thus, more attention should be paid to the side effects and toxicity of the drugs. Screening drugs that have passed clinical trials is a good approach, but further safety assessment trials are still needed. Many plants have antiviral and anti-inflammatory properties, especially those with medicinal value, such as herbs. Natural compounds extracted from these plants are often very effective against HFMD [64]. Many previous researches had been conducted on antiviral drugs for EVA71 and CVA16. However, as the pathogenic spectrum of HFMD has changed in recent years, the development of broad-spectrum antiviral drugs is urgently required, especially for important HFMD pathogens such as CVA6, CVA10, CVB3, etc. At the same time, for existing drugs, researchers can improve the pharmacokinetic characteristics and efficacy of the drug by changing parts of the drug’s structure, based on studies about the drug’s structure. Resistant mutants have been selected for almost all direct-targeting antivirals in cell culture, suggesting that most drugs exhibit a low resistance barrier. Selecting drugs that do not make the virus resistant is also an issue that needs attention in future research. More and more researchers are no longer targeting viral proteins as drug targets but instead targeting host factors, which solves the problem of drug resistance effectively.
Collectively, HFMD is a major threat to global public health, especially in the Asia-Pacific region. Although much progress has been made in the research of effective antiviral agents for HFMD, there is still a long way to go before clinical trials can be adopted and truly benefit patients.

Author Contributions

Conceptualization, G.D., S.C. and Y.J.; writing—original draft preparation, Z.L.; writing—review and editing, W.J. and Y.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (NO. 82273695, NO. 82002147 and NO. 82073618); China Postdoctoral Science Foundation (NO. 2019M662543); Key Scientific Research Project of Henan Institution of Higher Education (NO. 21A310026). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xie, Y.; Hu, Q.; Jiang, W.; Ji, W.; Chen, S.; Jin, Y.; Duan, G. Laboratory Indicators for Identifying Hand, Foot, and Mouth Disease Severity: A Systematic Review and Meta-Analysis. Vaccines 2022, 10, 1829. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, J.; Hu, Y.; Zheng, M. Enterovirus A71 antivirals: Past, present, and future. Acta Pharm. Sin. B 2022, 12, 1542–1566. [Google Scholar] [CrossRef]
  3. Zhu, Z.; Zhu, S.; Guo, X.; Wang, J.; Wang, D.; Yan, D.; Tan, X.; Tang, L.; Zhu, H.; Yang, Z.; et al. Retrospective seroepidemiology indicated that human enterovirus 71 and coxsackievirus A16 circulated wildly in central and southern China before large-scale outbreaks from 2008. Virol. J. 2010, 7, 300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Shah, J.; Sijun, L.; Hui, Z.; Zeb, F.; Haq, I.U.; Ullah, A. Neurological Complications of Hand, Foot and Mouth Disease In Children: A Review. J. Ayub Med. Coll. Abbottabad 2020, 32, 562–569. [Google Scholar] [PubMed]
  5. Aswathyraj, S.; Arunkumar, G.; Alidjinou, E.K.; Hober, D. Hand, foot and mouth disease (HFMD): Emerging epidemiology and the need for a vaccine strategy. Med. Microbiol. Immunol. 2016, 205, 397–407. [Google Scholar] [CrossRef] [PubMed]
  6. Li, P.; Yu, J.; Hao, F.; He, H.; Shi, X.; Hu, J.; Wang, L.; Du, C.; Zhang, X.; Sun, Y.; et al. Discovery of Potent EV71 Capsid Inhibitors for Treatment of HFMD. ACS Med. Chem. Lett. 2017, 8, 841–846. [Google Scholar] [CrossRef]
  7. Li, M.L.; Shih, S.R.; Tolbert, B.S.; Brewer, G. Enterovirus A71 Vaccines. Vaccines 2021, 9, 199. [Google Scholar] [CrossRef]
  8. Mustafa, S.; Abd-Aziz, N.; Saw, W.T.; Liew, S.Y.; Yusoff, K.; Shafee, N. Recombinant Enterovirus 71 Viral Protein 1 Fused to a Truncated Newcastle Disease Virus NP (NPt) Carrier Protein. Vaccines 2020, 8, 742. [Google Scholar] [CrossRef]
  9. Kobayashi, K.; Koike, S. Cellular receptors for enterovirus A71. J. Biomed. Sci. 2020, 27, 23. [Google Scholar] [CrossRef]
  10. Tan, C.W.; Chan, Y.F.; Sim, K.M.; Tan, E.L.; Poh, C.L. Inhibition of enterovirus 71 (EV-71) infections by a novel antiviral peptide derived from EV-71 capsid protein VP1. PLoS ONE 2012, 7, e34589. [Google Scholar] [CrossRef] [Green Version]
  11. Tan, C.W.; Sam, I.C.; Lee, V.S.; Wong, H.V.; Chan, Y.F. VP1 residues around the five-fold axis of enterovirus A71 mediate heparan sulfate interaction. Virology 2017, 501, 79–87. [Google Scholar] [CrossRef]
  12. Tan, C.W.; Poh, C.L.; Sam, I.C.; Chan, Y.F. Enterovirus 71 uses cell surface heparan sulfate glycosaminoglycan as an attachment receptor. J. Virol. 2013, 87, 611–620. [Google Scholar] [CrossRef] [Green Version]
  13. Lalani, S.; Tan, S.H.; Tan, K.O.; Lim, H.X.; Ong, K.C.; Wong, K.T.; Poh, C.L. Molecular mechanism of L-SP40 peptide and in vivo efficacy against EV-A71 in neonatal mice. Life Sci. 2021, 287, 120097. [Google Scholar] [CrossRef]
  14. Nishimura, Y.; McLaughlin, N.P.; Pan, J.; Goldstein, S.; Hafenstein, S.; Shimizu, H.; Winkler, J.D.; Bergelson, J.M. The Suramin Derivative NF449 Interacts with the 5-fold Vertex of the Enterovirus A71 Capsid to Prevent Virus Attachment to PSGL-1 and Heparan Sulfate. PLoS Pathog. 2015, 11, e1005184. [Google Scholar] [CrossRef]
  15. Sun, L.; Lee, H.; Thibaut, H.J.; Lanko, K.; Rivero-Buceta, E.; Bator, C.; Martinez-Gualda, B.; Dallmeier, K.; Delang, L.; Leyssen, P.; et al. Viral engagement with host receptors blocked by a novel class of tryptophan dendrimers that targets the 5-fold-axis of the enterovirus-A71 capsid. PLoS Pathog. 2019, 15, e1007760. [Google Scholar] [CrossRef] [Green Version]
  16. Hsieh, C.F.; Jheng, J.R.; Lin, G.H.; Chen, Y.L.; Ho, J.Y.; Liu, C.J.; Hsu, K.Y.; Chen, Y.S.; Chan, Y.F.; Yu, H.M.; et al. Rosmarinic acid exhibits broad anti-enterovirus A71 activity by inhibiting the interaction between the five-fold axis of capsid VP1 and cognate sulfated receptors. Emerg. Microbes Infect. 2020, 9, 1194–1205. [Google Scholar] [CrossRef]
  17. Guo, X.; Zeng, S.; Ji, X.; Meng, X.; Lei, N.; Yang, H.; Mu, X. Type I Interferon-Induced TMEM106A Blocks Attachment of EV-A71 Virus by Interacting With the Membrane Protein SCARB2. Front. Immunol. 2022, 13, 817835. [Google Scholar] [CrossRef]
  18. Meng, T.; Jia, Q.; Wong, S.M.; Chua, K.B. In Vitro and In Vivo Inhibition of the Infectivity of Human Enterovirus 71 by a Sulfonated Food Azo Dye, Brilliant Black BN. J. Virol. 2019, 93, e00061-19. [Google Scholar] [CrossRef] [Green Version]
  19. Hsieh, C.F.; Chen, Y.L.; Lin, G.H.; Chan, Y.F.; Hsieh, P.W.; Horng, J.T. 3,4-Dicaffeoylquinic Acid from the Medicinal Plant Ilex kaushue Disrupts the Interaction Between the Five-Fold Axis of Enterovirus A-71 and the Heparan Sulfate Receptor. J. Virol. 2022, 96, e0054221. [Google Scholar] [CrossRef]
  20. Rossmann, M.G.; He, Y.; Kuhn, R.J. Picornavirus-receptor interactions. Trends. Microbiol. 2002, 10, 324–331. [Google Scholar] [CrossRef]
  21. Plevka, P.; Perera, R.; Yap, M.L.; Cardosa, J.; Kuhn, R.J.; Rossmann, M.G. Structure of human enterovirus 71 in complex with a capsid-binding inhibitor. Proc. Natl. Acad. Sci. USA 2013, 110, 5463–5467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Muckelbauer, J.K.; Kremer, M.; Minor, I.; Diana, G.; Dutko, F.J.; Groarke, J.; Pevear, D.C.; Rossmann, M.G. The structure of coxsackievirus B3 at 3.5 A resolution. Structure 1995, 3, 653–667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Rossmann, M.G. Viral cell recognition and entry. Protein Sci. 1994, 3, 1712–1725. [Google Scholar] [CrossRef] [PubMed]
  24. Heim, A.; Pfetzing, U.; Muller, G.; Grumbach, I.M. Antiviral activity of WIN 54954 in coxsackievirus B2 carrier state infected human myocardial fibroblasts. Antiviral Res. 1998, 37, 47–56. [Google Scholar] [CrossRef]
  25. Shih, S.R.; Tsai, M.C.; Tseng, S.N.; Won, K.F.; Shia, K.S.; Li, W.T.; Chern, J.H.; Chen, G.W.; Lee, C.C.; Lee, Y.C.; et al. Mutation in enterovirus 71 capsid protein VP1 confers resistance to the inhibitory effects of pyridyl imidazolidinone. Antimicrob. Agents Chemother. 2004, 48, 3523–3529. [Google Scholar] [CrossRef] [Green Version]
  26. Pevear, D.C.; Tull, T.M.; Seipel, M.E.; Groarke, J.M. Activity of pleconaril against enteroviruses. Antimicrob. Agents Chemother. 1999, 43, 2109–2115. [Google Scholar] [CrossRef] [Green Version]
  27. Tijsma, A.; Franco, D.; Tucker, S.; Hilgenfeld, R.; Froeyen, M.; Leyssen, P.; Neyts, J. The capsid binder Vapendavir and the novel protease inhibitor SG85 inhibit enterovirus 71 replication. Antimicrob. Agents Chemother. 2014, 58, 6990–6992. [Google Scholar] [CrossRef] [Green Version]
  28. Ma, C.; Hu, Y.; Zhang, J.; Musharrafieh, R.; Wang, J. A Novel Capsid Binding Inhibitor Displays Potent Antiviral Activity against Enterovirus D68. ACS Infect. Dis. 2019, 5, 1952–1962. [Google Scholar] [CrossRef]
  29. De Colibus, L.; Wang, X.; Spyrou, J.A.B.; Kelly, J.; Ren, J.; Grimes, J.; Puerstinger, G.; Stonehouse, N.; Walter, T.S.; Hu, Z.; et al. More-powerful virus inhibitors from structure-based analysis of HEV71 capsid-binding molecules. Nat. Struct. Mol. Biol. 2014, 21, 282–288. [Google Scholar] [CrossRef] [Green Version]
  30. Ho, J.Y.; Chern, J.H.; Hsieh, C.F.; Liu, S.T.; Liu, C.J.; Wang, Y.S.; Kuo, T.W.; Hsu, S.J.; Yeh, T.K.; Shih, S.R.; et al. In vitro and in vivo studies of a potent capsid-binding inhibitor of enterovirus 71. J. Antimicrob. Chemother. 2016, 71, 1922–1932. [Google Scholar] [CrossRef] [Green Version]
  31. Zhang, H.; Wang, X.; Wang, Y.; Pei, X.; Wang, C.; Niu, Y.; Xu, P.; Peng, Y. Substituted 3-benzylcoumarins 13 and 14 suppress enterovirus A71 replication by impairing viral 2A(pro) dependent IRES-driven translation. Antiviral. Res. 2018, 160, 10–16. [Google Scholar] [CrossRef]
  32. Duan, H.; Zhu, M.; Xiong, Q.; Wang, Y.; Xu, C.; Sun, J.; Wang, C.; Zhang, H.; Xu, P.; Peng, Y. Regulation of enterovirus 2A protease-associated viral IRES activities by the cell’s ERK signaling cascade: Implicating ERK as an efficiently antiviral target. Antiviral Res. 2017, 143, 13–21. [Google Scholar] [CrossRef]
  33. Falah, N.; Montserret, R.; Lelogeais, V.; Schuffenecker, I.; Lina, B.; Cortay, J.C.; Violot, S. Blocking human enterovirus 71 replication by targeting viral 2A protease. J. Antimicrob. Chemother. 2012, 67, 2865–2869. [Google Scholar] [CrossRef] [Green Version]
  34. Wang, C.Y.; Huang, A.C.; Hour, M.J.; Huang, S.H.; Kung, S.H.; Chen, C.H.; Chen, I.C.; Chang, Y.S.; Lien, J.C.; Lin, C.W. Antiviral Potential of a Novel Compound CW-33 against Enterovirus A71 via Inhibition of Viral 2A Protease. Viruses 2015, 7, 3155–3171. [Google Scholar] [CrossRef]
  35. Gao, M.; Duan, H.; Liu, J.; Zhang, H.; Wang, X.; Zhu, M.; Guo, J.; Zhao, Z.; Meng, L.; Peng, Y. The multi-targeted kinase inhibitor sorafenib inhibits enterovirus 71 replication by regulating IRES-dependent translation of viral proteins. Antiviral Res. 2014, 106, 80–85. [Google Scholar] [CrossRef]
  36. Yuan, J.; Shen, L.; Wu, J.; Zou, X.; Gu, J.; Chen, J.; Mao, L. Enterovirus A71 Proteins: Structure and Function. Front. Microbiol. 2018, 9, 286. [Google Scholar] [CrossRef] [Green Version]
  37. Ulferts, R.; de Boer, S.M.; van der Linden, L.; Bauer, L.; Lyoo, H.R.; Mate, M.J.; Lichiere, J.; Canard, B.; Lelieveld, D.; Omta, W.; et al. Screening of a Library of FDA-Approved Drugs Identifies Several Enterovirus Replication Inhibitors That Target Viral Protein 2C. Antimicrob. Agents Chemother. 2016, 60, 2627–2638. [Google Scholar] [CrossRef] [Green Version]
  38. Fang, Y.; Wang, C.; Wang, C.; Yang, R.; Bai, P.; Zhang, X.Y.; Kong, J.; Yin, L.; Qiu, Y.; Zhou, X. Antiviral Peptides Targeting the Helicase Activity of Enterovirus Nonstructural Protein 2C. J. Virol. 2021, 95, e02324-20. [Google Scholar] [CrossRef]
  39. Ma, C.; Hu, Y.; Zhang, J.; Wang, J. Pharmacological Characterization of the Mechanism of Action of R523062, a Promising Antiviral for Enterovirus D68. ACS Infect. Dis. 2020, 6, 2260–2270. [Google Scholar] [CrossRef]
  40. Zuo, J.; Kye, S.; Quinn, K.K.; Cooper, P.; Damoiseaux, R.; Krogstad, P. Discovery of Structurally Diverse Small-Molecule Compounds with Broad Antiviral Activity against Enteroviruses. Antimicrob. Agents Chemother. 2015, 60, 1615–1626. [Google Scholar] [CrossRef] [Green Version]
  41. Tang, Q.; Xu, Z.; Jin, M.; Shu, T.; Chen, Y.; Feng, L.; Zhang, Q.; Lan, K.; Wu, S.; Zhou, H.B. Identification of dibucaine derivatives as novel potent enterovirus 2C helicase inhibitors: In vitro, in vivo, and combination therapy study. Eur. J. Med. Chem. 2020, 202, 112310. [Google Scholar] [CrossRef] [PubMed]
  42. Miller, F.D.; Monto, A.S.; DeLong, D.C.; Exelby, A.; Bryan, E.R.; Srivastava, S. Controlled trial of enviroxime against natural rhinovirus infections in a community. Antimicrob. Agents Chemother. 1985, 27, 102–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Arita, M.; Kojima, H.; Nagano, T.; Okabe, T.; Wakita, T.; Shimizu, H. Phosphatidylinositol 4-kinase III beta is a target of enviroxime-like compounds for antipoliovirus activity. J. Virol. 2011, 85, 2364–2372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. De Palma, A.M.; Thibaut, H.J.; van der Linden, L.; Lanke, K.; Heggermont, W.; Ireland, S.; Andrews, R.; Arimilli, M.; Al-Tel, T.H.; De Clercq, E.; et al. Mutations in the nonstructural protein 3A confer resistance to the novel enterovirus replication inhibitor TTP-8307. Antimicrob. Agents Chemother. 2009, 53, 1850–1857. [Google Scholar] [CrossRef] [Green Version]
  45. Gao, Q.; Yuan, S.; Zhang, C.; Wang, Y.; Wang, Y.; He, G.; Zhang, S.; Altmeyer, R.; Zou, G. Discovery of itraconazole with broad-spectrum in vitro antienterovirus activity that targets nonstructural protein 3A. Antimicrob. Agents Chemother. 2015, 59, 2654–2665. [Google Scholar] [CrossRef] [Green Version]
  46. Liu, M.; Xu, B.; Ma, Y.; Shang, L.; Ye, S.; Wang, Y. Reversible covalent inhibitors suppress enterovirus 71 infection by targeting the 3C protease. Antiviral Res. 2021, 192, 105102. [Google Scholar] [CrossRef]
  47. Wang, Y.; Cao, L.; Zhai, Y.; Yin, Z.; Sun, Y.; Shang, L. Structure of the Enterovirus 71 3C Protease in Complex with NK-1.8k and Indications for the Development of Antienterovirus Protease Inhibitor. Antimicrob. Agents Chemother. 2017, 61, e00298-17. [Google Scholar] [CrossRef] [Green Version]
  48. Matthews, D.A.; Dragovich, P.S.; Webber, S.E.; Fuhrman, S.A.; Patick, A.K.; Zalman, L.S.; Hendrickson, T.F.; Love, R.A.; Prins, T.J.; Marakovits, J.T.; et al. Structure-assisted design of mechanism-based irreversible inhibitors of human rhinovirus 3C protease with potent antiviral activity against multiple rhinovirus serotypes. Proc. Natl. Acad. Sci. USA 1999, 96, 11000–11007. [Google Scholar] [CrossRef] [Green Version]
  49. 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] [Green Version]
  50. Xu, B.; Liu, M.; Ma, S.; Ma, Y.; Liu, S.; Shang, L.; Zhu, C.; Ye, S.; Wang, Y. 4-Iminooxazolidin-2-One as a Bioisostere of Cyanohydrin Suppresses EV71 Proliferation by Targeting 3C(pro). Microbiol. Spectr. 2021, 9, e0102521. [Google Scholar] [CrossRef]
  51. Wang, Y.; Cao, L.; Zhai, Y.; Ma, J.; Nie, Q.; Li, T.; Yin, Z.; Sun, Y.; Shang, L. Inhibition of enterovirus 71 replication by an alpha-hydroxy-nitrile derivative NK-1.9k. Antiviral Res. 2017, 141, 91–100. [Google Scholar] [CrossRef]
  52. Shang, L.; Wang, Y.; Qing, J.; Shu, B.; Cao, L.; Lou, Z.; Gong, P.; Sun, Y.; Yin, Z. An adenosine nucleoside analogue NITD008 inhibits EV71 proliferation. Antiviral Res. 2014, 112, 47–58. [Google Scholar] [CrossRef]
  53. Xu, N.; Yang, J.; Zheng, B.; Zhang, Y.; Cao, Y.; Huan, C.; Wang, S.; Chang, J.; Zhang, W. The Pyrimidine Analog FNC Potently Inhibits the Replication of Multiple Enteroviruses. J. Virol. 2020, 94, e00204-20. [Google Scholar] [CrossRef] [Green Version]
  54. Chen, T.C.; Chang, H.Y.; Lin, P.F.; Chern, J.H.; Hsu, J.T.; Chang, C.Y.; Shih, S.R. Novel antiviral agent DTriP-22 targets RNA-dependent RNA polymerase of enterovirus 71. Antimicrob. Agents Chemother. 2009, 53, 2740–2747. [Google Scholar] [CrossRef] [Green Version]
  55. Velu, A.B.; Chen, G.W.; Hsieh, P.T.; Horng, J.T.; Hsu, J.T.; Hsieh, H.P.; Chen, T.C.; Weng, K.F.; Shih, S.R. BPR-3P0128 inhibits RNA-dependent RNA polymerase elongation and VPg uridylylation activities of Enterovirus 71. Antiviral Res. 2014, 112, 18–25. [Google Scholar] [CrossRef]
  56. Thompson, S.R.; Sarnow, P. Enterovirus 71 contains a type I IRES element that functions when eukaryotic initiation factor eIF4G is cleaved. Virology 2003, 315, 259–266. [Google Scholar] [CrossRef] [Green Version]
  57. Yuan, J.; Stein, D.A.; Lim, T.; Qiu, D.; Coughlin, S.; Liu, Z.; Wang, Y.; Blouch, R.; Moulton, H.M.; Iversen, P.L.; et al. Inhibition of coxsackievirus B3 in cell cultures and in mice by peptide-conjugated morpholino oligomers targeting the internal ribosome entry site. J. Virol. 2006, 80, 11510–11519. [Google Scholar] [CrossRef] [Green Version]
  58. 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.; et al. 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]
  59. 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. Antiviral Res. 2014, 109, 30–41. [Google Scholar] [CrossRef]
  60. Hou, H.Y.; Lu, W.W.; Wu, K.Y.; Lin, C.W.; Kung, S.H. Idarubicin is a broad-spectrum enterovirus replication inhibitor that selectively targets the virus internal ribosomal entry site. J. Gen. Virol. 2016, 97, 1122–1133. [Google Scholar] [CrossRef] [Green Version]
  61. Lin, J.Y.; Li, M.L.; Brewer, G. mRNA decay factor AUF1 binds the internal ribosomal entry site of enterovirus 71 and inhibits virus replication. PLoS ONE 2014, 9, e103827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Davila-Calderon, J.; Patwardhan, N.N.; Chiu, L.Y.; Sugarman, A.; Cai, Z.; Penutmutchu, S.R.; Li, M.L.; Brewer, G.; Hargrove, A.E.; Tolbert, B.S. IRES-targeting small molecule inhibits enterovirus 71 replication via allosteric stabilization of a ternary complex. Nat. Commun. 2020, 11, 4775. [Google Scholar] [CrossRef] [PubMed]
  63. Lim, Z.Q.; Ng, Q.Y.; Ng, J.W.Q.; Mahendran, V.; Alonso, S. Recent progress and challenges in drug development to fight hand, foot and mouth disease. Expert Opin. Drug Discov. 2020, 15, 359–371. [Google Scholar] [CrossRef] [PubMed]
  64. Lee, Y.R.; Chang, C.M.; Yeh, Y.C.; Huang, C.F.; Lin, F.M.; Huang, J.T.; Hsieh, C.C.; Wang, J.R.; Liu, H.S. Honeysuckle Aqueous Extracts Induced let-7a Suppress EV71 Replication and Pathogenesis In Vitro and In Vivo and Is Predicted to Inhibit SARS-CoV-2. Viruses 2021, 13, 308. [Google Scholar] [CrossRef]
Figure 1. Schematic overview of EV life cycle.
Figure 1. Schematic overview of EV life cycle.
Vaccines 11 00571 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, Z.; Ji, W.; Chen, S.; Duan, G.; Jin, Y. Hand, Foot, and Mouth Disease Challenges and Its Antiviral Therapeutics. Vaccines 2023, 11, 571. https://doi.org/10.3390/vaccines11030571

AMA Style

Li Z, Ji W, Chen S, Duan G, Jin Y. Hand, Foot, and Mouth Disease Challenges and Its Antiviral Therapeutics. Vaccines. 2023; 11(3):571. https://doi.org/10.3390/vaccines11030571

Chicago/Turabian Style

Li, Zijie, Wangquan Ji, Shuaiyin Chen, Guangcai Duan, and Yuefei Jin. 2023. "Hand, Foot, and Mouth Disease Challenges and Its Antiviral Therapeutics" Vaccines 11, no. 3: 571. https://doi.org/10.3390/vaccines11030571

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