2,182
Views
6
CrossRef citations to date
0
Altmetric
Review Article

Treading a HOSTile path: Mapping the dynamic landscape of host cell–rotavirus interactions to explore novel host-directed curative dimensions

, , , &
Pages 1022-1062 | Received 21 Sep 2020, Accepted 10 Mar 2021, Published online: 05 Apr 2021

References

  • Crawford SE, Ramani S, Tate JE, et al. Rotavirus infection. Nat Rev Dis Primers. 2017;3(1):17083.
  • Jonesteller CL, Burnett E, Yen C, et al. Effectiveness of rotavirus vaccination: a systematic review of the first decade of global postlicensure data, 2006-2016. Clin Infect Dis. 2017;65(5):840–850.
  • Leshem E, Tate JE, Steiner CA, et al. national estimates of reductions in acute Gastroenteritis-Related hospitalizations and associated costs in us children after implementation of rotavirus vaccines. J Pediatric Infect Dis Soc. 2018;7(3):257–260.
  • Tate JE, Burton AH, Boschi-Pinto C, et al., World Health Organization–Coordinated Global Rotavirus Surveillance Network. Global, regional, and national estimates of rotavirus mortality in children <5 years of age, 2000-2013. Clin Infect Dis. 2016;62(Suppl 2):S96–S105. .
  • Troeger C, Khalil IA, Rao PC, et al. rotavirus vaccination and the global burden of rotavirus diarrhea among children younger than 5 years. JAMA Pediatr. 2018;172(10):958–965.
  • Steele AD, Victor JC, Carey ME, et al. Experiences with rotavirus vaccines: can we improve rotavirus vaccine impact in developing countries? Hum Vaccin Immunother. 2019;15(6):1215–1227.
  • Chen JZ, Settembre EC, Aoki ST, et al. Molecular interactions in rotavirus assembly and uncoating seen by high-resolution cryo-EM. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(26):10644–10648
  • Li Z, Baker ML, Jiang W, et al. Rotavirus architecture at subnanometer resolution. J Virol. 2009;83(4):1754–1766.
  • Settembre EC, Chen JZ, Dormitzer PR, et al. Atomic model of an infectious rotavirus particle. Embo J. 2011;30(2):408–416.
  • Bányai K, Kemenesi G, Budinski I, et al. Candidate new rotavirus species in Schreiber’s bats, Serbia. Infect Genet Evol. 2017;48:19–26.
  • Matthijnssens J, Otto PH, Ciarlet M, et al. VP6-sequence-based cutoff values as a criterion for rotavirus species demarcation. Arch Virol. 2012;157(6):1177–1182.
  • Mihalov-Kovács E, Gellért Á, Marton S, et al. Candidate new rotavirus species in sheltered dogs, Hungary. Emerg Infect Dis. 2015;21(4):660–663.
  • RCWG. 2017. Newly assigned genotypes. https://rega.kuleuven.be/cev/viralmetagenomics/virus-classification/rcwg.
  • Lopez S, Arias CF. Early steps in rotavirus cell entry. Curr Top Microbiol Immunol. 2006;309:39–66.
  • Desselberger U. Rotaviruses. Virus Res. 2014;190:75–96.
  • Estes MK, Greenberg HB. Rotaviruses and their replication. Virology F, Knipe DN, Howley PM… Eds. Philadelphia, PA, USA: Wolters Kluwer Health, LippincottWilliams &Wilkins; 2013. 1347–1401.
  • Silvestri LS, Taraporewala ZF, Patton JT. Rotavirus replication: plus-sense templates for double-stranded RNA synthesis are made in viroplasms. J Virol. 2004;78(14):7763–7774.
  • López S, Sánchez-Tacuba L, Moreno J, et al. Rotavirus Strategies Against the Innate Antiviral System. Annu Rev Virol. 2016;3(1):591–609.
  • Eichwald C, Rodriguez JF, Burrone OR. Characterization of rotavirus NSP2/NSP5 interactions and the dynamics of viroplasm formation. J Gen Virol. 2004;85(Pt 3):625–634.
  • Fabbretti E, Afrikanova I, Vascotto F, et al. Two non-structural rotavirus proteins, NSP2 and NSP5, form viroplasm-like structures in vivo. J Gen Virol. 1999;80(Pt 2):333–339.
  • Patton JT, Silvestri LS, Tortorici MA, et al. Rotavirus genome replication and morphogenesis: role of the viroplasm. In Reoviruses: entry, assembly and morphogenesis (pp. 169–187). Berlin, Heidelberg: Springer; 2006.
  • Gridley CL, Patton JT. Regulation of rotavirus polymerase activity by inner capsid proteins. Curr Opin Virol. 2014;9:31–38.
  • Lu X, McDonald SM, Tortorici MA, et al. Mechanism for coordinated RNA packaging and genome replication by rotavirus polymerase VP1. Structure. 2008;16(11):1678–1688. . (London, England: 1993)
  • Patton JT, Chen D. RNA-binding and capping activities of proteins in rotavirus open cores. J Virol. 1999;73(2):1382–1391.
  • Chen D, Ramig RF. Rescue of infectivity by sequential in vitro transcapsidation of rotavirus core particles with inner capsid and outer capsid proteins. Virology. 1993;194(2):743–751.
  • Navarro A, Williamson L, Angel M, et al. Rotavirus replication and reverse genetics. In: Svensson L, Desselberger U, Greenberg HB, et al., editors. Viral gastroenteritis. Molecular epidemiology and pathogenesis. Cambridge: Academic Press; 2016. p. 121–143.
  • Crawford SE, Hyser JM, Utama B, et al. Autophagy hijacked through viroporin-activated calcium/calmodulin-dependent kinase kinase-β signaling is required for rotavirus replication. Proceedings of the National Academy of Sciences. 2012;109(50):E3405–E3413
  • Crawford SE, Criglar JM, Liu Z, et al. COPII vesicle transport is required for rotavirus NSP4 interaction with the autophagy protein LC3 II and trafficking to viroplasms. J Virol. 2019;94(1):e01341–19.
  • Trask SD, McDonald SM, Patton JT. Structural insights into the coupling of virion assembly and rotavirus replication. Nature reviews. Microbiology. 2012;10(3):165–177.
  • Au KS, Mattion NM, Estes MK. A subviral particle binding domain on the rotavirus nonstructural glycoprotein NS28. Virology. 1993;194(2):665–673.
  • Maass DR, Atkinson PH. Rotavirus proteins VP7, NS28, and VP4 form oligomeric structures. J Virol. 1990;64(6):2632–2641.
  • Poruchynsky MS, Atkinson PH. Rotavirus protein rearrangements in purified membrane-enveloped intermediate particles. J Virol. 1991;65(9):4720–4727.
  • Stirzaker SC, Both GW. The signal peptide of the rotavirus glycoprotein VP7 is essential for its retention in the ER as an integral membrane protein. Cell. 1989;56(5):741–747.
  • Sastri NP, Crawford SE, Estes MK. Pleiotropic properties of rotavirus nonstructural protein 4 (NSP4) and their effects on viral replication and pathogenesis. In: Svensson L, Desselberger U, Greenberg HB, et al., editors. Viral gastroenteritis. Molecular epidemiology and pathogenesis. Cambridge: Academic Press; 2016. p. 145–174.
  • López T, Camacho M, Zayas M, et al. Silencing the morphogenesis of rotavirus. J Virol. 2005;79(1):184–192.
  • Cuadras MA, Bordier BB, Zambrano JL, et al. Dissecting rotavirus particle-raft interaction with small interfering RNAs: insights into rotavirus transit through the secretory pathway. J Virol. 2006;80(8):3935–3946.
  • Delmas O, Durand-Schneider AM, Cohen J, et al. Spike protein VP4 assembly with maturing rotavirus requires a postendoplasmic reticulum event in polarized caco-2 cells. J Virol. 2004;78(20):10987–10994.
  • Jourdan N, Maurice M, Delautier D, et al. Rotavirus is released from the apical surface of cultured human intestinal cells through nonconventional vesicular transport that bypasses the Golgi apparatus. J Virol. 1997;71(11):8268–8278.
  • Musalem C, Espejo RT. Release of progeny virus from cells infected with simian rotavirus SA11. J Gen Virol. 1985;66(Pt 12):2715–2724.
  • Delorme C, Brüssow H, Sidoti J, et al. Glycosphingolipid binding specificities of rotavirus: identification of a sialic acid-binding epitope. J Virol. 2001;75(5):2276–2287.
  • Haselhorst T, Blanchard H, Frank M, et al. STD NMR spectroscopy and molecular modeling investigation of the binding of N-acetylneuraminic acid derivatives to rhesus rotavirus VP8* core. Glycobiology. 2007;17(1):68–81.
  • Haselhorst T, Fleming FE, Dyason JC, et al. Sialic acid dependence in rotavirus host cell invasion. Nat Chem Biol. 2009;5(2):91–93.
  • Martínez MA, López S, Arias CF, et al. Gangliosides have a functional role during rotavirus cell entry. J Virol. 2013;87(2):1115–1122.
  • Dormitzer PR, Sun ZY, Wagner G, et al. The rhesus rotavirus VP4 sialic acid binding domain has a galectin fold with a novel carbohydrate binding site. Embo J. 2002;21(5):885–897.
  • Isa P, López S, Segovia L, et al. Functional and structural analysis of the sialic acid-binding domain of rotaviruses. J Virol. 1997;71(9):6749–6756.
  • Venkataram Prasad BV, Shanker S, Hu L, et al. Structural basis of glycan interaction in gastroenteric viral pathogens. Curr Opin Virol. 2014;7:119–127.
  • Böhm R, Fleming FE, Maggioni A, et al. Revisiting the role of histo-blood group antigens in rotavirus host-cell invasion. Nat Commun. 2015;6(1):5907.
  • Hu L, Crawford SE, Czako R, et al. Cell attachment protein VP8* of a human rotavirus specifically interacts with A-type histo-blood group antigen. Nature. 2012;485(7397):256–259.
  • Huang P, Xia M, Tan M, et al. Spike protein VP8* of human rotavirus recognizes histo-blood group antigens in a type-specific manner. J Virol. 2012;86(9):4833–4843.
  • Liu Y, Huang P, Tan M, et al. Rotavirus VP8*: phylogeny, host range, and interaction with histo-blood group antigens. J Virol. 2012;86(18):9899–9910.
  • Nordgren J, Sharma S, Bucardo F, et al. Both Lewis and secretor status mediate susceptibility to rotavirus infections in a rotavirus genotype-dependent manner. Clin Infect Dis. 2014;59(11):1567–1573.
  • Payne DC, Currier RL, Staat MA, et al. Epidemiologic association between FUT2 secretor status and severe rotavirus gastroenteritis in children in the United States. JAMA Pediatr. 2015;169(11):1040–1045.
  • Van Trang N, Vu HT, Le NT, et al. Association between norovirus and rotavirus infection and histo-blood group antigen types in Vietnamese children. J Clin Microbiol. 2014;52(5):1366–1374.
  • Hester SN, Chen X, Li M, et al. Human milk oligosaccharides inhibit rotavirus infectivity in vitro and in acutely infected piglets. Br J Nutr. 2013;110(7):1233–1242.
  • Laucirica DR, Triantis V, Schoemaker R, et al. Milk oligosaccharides inhibit human rotavirus infectivity in MA104 cells. J Nutr. 2017;147(9):1709–1714.
  • Yu Y, Lasanajak Y, Song X, et al. Human milk contains novel glycans that are potential decoy receptors for neonatal rotaviruses. Mol Cell Proteomics. 2014;13(11):2944–2960.
  • Ramani S, Stewart CJ, Laucirica DR, et al. Human milk oligosaccharides, milk microbiome and infant gut microbiome modulate neonatal rotavirus infection. Nat Commun. 2018;9(1):5010.
  • Isa P, Realpe M, Romero P, et al. Rotavirus RRV associates with lipid membrane microdomains during cell entry. Virology. 2004;322(2):370–381.
  • Cui J, Fu X, Xie J, et al. Critical role of cellular cholesterol in bovine rotavirus infection. Virol J. 2014;11(1):98.
  • Dou X, Li Y, Han J, et al. Cholesterol of lipid rafts is a key determinant for entry and post-entry control of porcine rotavirus infection. BMC Vet Res. 2018;14(1):45.
  • Guerrero CA, Zárate S, Corkidi G, et al. Biochemical characterization of rotavirus receptors in MA104 cells. J Virol. 2000;74(20):9362–9371.
  • Gutiérrez M, Isa P, Sánchez-San Martin C, et al. Different rotavirus strains enter MA104 cells through different endocytic pathways: the role of clathrin-mediated endocytosis. J Virol. 2010;84(18):9161–9169.
  • Graham KL, Halasz P, Tan Y, et al. Integrin-using rotaviruses bind alpha2beta1 integrin alpha2 I domain via VP4 DGE sequence and recognize alphaXbeta2 and alphaVbeta3 by using VP7 during cell entry. J Virol. 2003;77(18):9969–9978.
  • Zárate S, Romero P, Espinosa R, et al. VP7 mediates the interaction of rotaviruses with integrin alphavbeta3 through a novel integrin-binding site. J Virol. 2004;78(20):10839–10847.
  • Civra A, Giuffrida MG, Donalisio M, et al. Identification of equine lactadherin-derived peptides that inhibit rotavirus infection via integrin receptor competition. J Biol Chem. 2015;290(19):12403–12414.
  • Salas-Cárdenas SP, Olaya-Galán NN, Fernández K, et al. Decreased rotavirus infection of MA104 cells via probiotic extract binding to Hsc70 and β3 integrin receptors. Universitas scientiarum. 2018;23(2):219–239.
  • López T, López S, Arias CF. The tyrosine kinase inhibitor genistein induces the detachment of rotavirus particles from the cell surface. Virus Res. 2015;210:141–148.
  • Halasz P, Holloway G, Turner SJ, et al. Rotavirus replication in intestinal cells differentially regulates integrin expression by a phosphatidylinositol 3-kinase-dependent pathway, resulting in increased cell adhesion and virus yield. J Virol. 2008;82(1):148–160.
  • Nava P, López S, Arias CF, et al. The rotavirus surface protein VP8 modulates the gate and fence function of tight junctions in epithelial cells. J Cell Sci. 2004;117(Pt 23):5509–5519.
  • Torres-Flores JM, Silva-Ayala D, Espinoza MA, et al. The tight junction protein JAM-A functions as coreceptor for rotavirus entry into MA104 cells. Virology. 2015;475:172–178.
  • Guerrero CA, Bouyssounade D, Zárate S, et al. Heat shock cognate protein 70 is involved in rotavirus cell entry. J Virol. 2002;76(8):4096–4102.
  • Zárate S, Cuadras MA, Espinosa R, et al. Interaction of rotaviruses with Hsc70 during cell entry is mediated by VP5. J Virol. 2003;77(13):7254–7260.
  • Pfeiffer JK, Virgin HW. Transkingdom control of viral infection and immunity in the mammalian intestine. Science (New York, N.Y.). Viral immun. 2016;351(6270). 10.1126/science.aad5872
  • Pérez-Vargas J, Romero P, López S, et al. The peptide-binding and ATPase domains of recombinant hsc70 are required to interact with rotavirus and reduce its infectivity. J Virol. 2006;80(7):3322–3331.
  • Calderon MN, Guerrero CA, Acosta O, et al. Inhibiting rotavirus infection by membrane-impermeant thiol/disulfide exchange blockers and antibodies against protein disulfide isomerase. Intervirology. 2012;55(6):451–464.
  • Arias CF, Silva-Ayala D, Isa P, et al. Rotavirus attachment, internalization, and vesicular traffic. In: Svensson L, Desselberger U, Greenberg HB, et al., editors. Viral gastroenteritis. Molecular epidemiology and pathogenesis. Cambridge: Academic Press; 2016. p. 103–119.
  • Díaz-Salinas MA, Romero P, Espinosa R, et al. The spike protein VP4 defines the endocytic pathway used by rotavirus to enter MA104 cells. J Virol. 2013;87(3):1658–1663.
  • Sánchez-San Martín C, López T, Arias CF, et al. Characterization of rotavirus cell entry. J Virol. 2004;78(5):2310–2318.
  • Silva-Ayala D, López T, Gutiérrez M, et al. Genome-wide RNAi screen reveals a role for the ESCRT complex in rotavirus cell entry. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(25):10270–10275
  • Fan B, Zhu L, Chang X, et al. Mortalin restricts porcine epidemic diarrhea virus entry by downregulating clathrin-mediated endocytosis. Vet Microbiol. 2019;239:108455.
  • Díaz-Salinas MA, Silva-Ayala D, López S, et al. Rotaviruses reach late endosomes and require the cation-dependent mannose-6-phosphate receptor and the activity of cathepsin proteases to enter the cell. J Virol. 2014;88(8):4389–4402.
  • Abdelhakim AH, Salgado EN, Fu X, et al. Structural correlates of rotavirus cell entry. PLoS Pathog. 2014;10(9):e1004355.
  • Wolf M, Vo PT, Greenberg HB. Rhesus rotavirus entry into a polarized epithelium is endocytosis dependent and involves sequential VP4 conformational changes. J Virol. 2011;85(6):2492–2503.
  • Zambrano JL, Sorondo O, Alcala A, et al. Rotavirus infection of cells in culture induces activation of RhoA and changes in the actin and tubulin cytoskeleton. PloS One. 2012;7(10):e47612.
  • Soliman M, Cho EH, Park JG, et al. Rotavirus-Induced early activation of the RhoA/ROCK/MLC signaling pathway mediates the disruption of tight junctions in polarized MDCK cells. Sci Rep. 2018;8(1):13931.
  • Li B, Ding S, Feng N, et al. Drebrin restricts rotavirus entry by inhibiting dynamin-mediated endocytosis. Proceedings of the National Academy of Sciences of the United States of America. 2017;114(18):E3642–E3651
  • Trejo-Cerro O, Aguilar-Hernández N, Silva-Ayala D, et al. The actin cytoskeleton is important for rotavirus internalization and RNA genome replication. Virus Res. 2019;263:27–33.
  • Cossart P, Helenius A. Endocytosis of viruses and bacteria. Cold Spring Harb Perspect Biol. 2014;6(8):a016972.
  • Mercer J, Schelhaas M, Helenius A. Virus entry by endocytosis. Annu Rev Biochem. 2010;79(1):803–833.
  • Wolf M, Deal EM, Greenberg HB. Rhesus rotavirus trafficking during entry into MA104 cells is restricted to the early endosome compartment. J Virol. 2012;86(7):4009–4013.
  • Frankel EB, Audhya A. ESCRT-dependent cargo sorting at multivesicular endosomes. Seminars in cell & developmental biology. 2018;74:4–10
  • Stenmark H. Rab GTPases as coordinators of vesicle traffic. Nature reviews. Mol Cell Biol. 2009;10(8):513–525.
  • Díaz-Salinas MA, Casorla LA, López T, et al. Most rotavirus strains require the cation-independent mannose-6-phosphate receptor, sortilin-1, and cathepsins to enter cells. Virus Res. 2018;245:44–51.
  • Braulke T, Bonifacino JS. Sorting of lysosomal proteins. Biochim Biophys Acta. 2009;1793(4):605–614.
  • Turk V, Stoka V, Vasiljeva O, et al. Cysteine cathepsins: from structure, function and regulation to new frontiers. Biochim Biophys Acta. 2012;1824(1):68–88.
  • Tsai B. Penetration of nonenveloped viruses into the cytoplasm. Annu Rev Cell Dev Biol. 2007;23(1):23–43.
  • Yoder JD, Dormitzer PR. Alternative intermolecular contacts underlie the rotavirus VP5* two- to three-fold rearrangement. Embo J. 2006;25(7):1559–1568.
  • Soliman M, Seo JY, Kim DS, et al. Activation of PI3K, Akt, and ERK during early rotavirus infection leads to V-ATPase-dependent endosomal acidification required for uncoating. PLoS Pathog. 2018;14(1):e1006820.
  • Civra A, Francese R, Gamba P, et al. 25-Hydroxycholesterol and 27-hydroxycholesterol inhibit human rotavirus infection by sequestering viral particles into late endosomes. Redox Biol. 2018;19:318–330.
  • Sobo K, Le Blanc I, Luyet PP, et al. Late endosomal cholesterol accumulation leads to impaired intra-endosomal trafficking. PloS One. 2007;2(9):e851.
  • Schneider WM, Chevillotte MD, Rice CM. Interferon-stimulated genes: a complex web of host defenses. Annu Rev Immunol. 2014;32(1):513–545.
  • Fensterl V, Chattopadhyay S, Sen GC. No love lost between viruses and interferons. Annu Rev Virol. 2015;2(1):549–572.
  • Sen A, Pruijssers AJ, Dermody TS, et al. The early interferon response to rotavirus is regulated by PKR and depends on MAVS/IPS-1, RIG-I, MDA-5, and IRF3. J Virol. 2011;85(8):3717–3732.
  • Broquet AH, Hirata Y, McAllister CS, et al. RIG-I/MDA5/MAVS are required to signal a protective IFN response in rotavirus-infected intestinal epithelium. J Iimmunol. 2011;186(3):1618–1626. . (Baltimore, Md.: 1950)
  • Pott J, Stockinger S, Torow N, et al. Age-dependent TLR3 expression of the intestinal epithelium contributes to rotavirus susceptibility. PLoS Pathog. 2012;8(5):e1002670.
  • Deal EM, Jaimes MC, Crawford SE, et al. Rotavirus structural proteins and dsRNA are required for the human primary plasmacytoid dendritic cell IFNalpha response. PLoS Pathog. 2010;6(6):e1000931.
  • Deal EM, Lahl K, Narváez CF, et al. Plasmacytoid dendritic cells promote rotavirus-induced human and murine B cell responses. J Clin Invest. 2013;123(6):2464–2474.
  • Douagi I, McInerney GM, Hidmark AS, et al. Role of interferon regulatory factor 3 in type I interferon responses in rotavirus-infected dendritic cells and fibroblasts. J Virol. 2007;81(6):2758–2768.
  • González AM, Azevedo MS, Jung K, et al. Innate immune responses to human rotavirus in the neonatal gnotobiotic piglet disease model. Immunology. 2010;131(2):242–256.
  • Lopez-Guerrero DV, Meza-Perez S, Ramirez-Pliego O, et al. Rotavirus infection activates dendritic cells from Peyer’s patches in adult mice. J Virol. 2010;84(4):1856–1866.
  • Sen A, Feng N, Ettayebi K, et al. IRF3 inhibition by rotavirus NSP1 is host cell and virus strain dependent but independent of NSP1 proteasomal degradation. J Virol. 2009;83(20):10322–10335.
  • Uzri D, Greenberg HB. Characterization of rotavirus RNAs that activate innate immune signaling through the RIG-I-like receptors. PloS One. 2013;8(7):e69825.
  • Li Y, Yu P, Qu C, et al. MDA5 against enteric viruses through induction of interferon-like response partially via the JAK-STAT cascade. Antiviral Res. 2020;176:104743.
  • Arnold MM. The rotavirus interferon antagonist NSP1: many targets, many questions. J Virol. 2016;90(11):5212–5215.
  • Sen A, Greenberg HB. Innate immune responses to rotavirus infection. In: Svensson L, Desselberger U, Greenberg HB, et al., editors. Viral gastroenteritis. Molecular epidemiology and pathogenesis. Cambridge: Academic Press; 2016. p. 243–263.
  • Qin L, Ren L, Zhou Z, et al. Rotavirusnonstructural protein 1 antagonizes innate immune response by interacting with retinoic acid inducible gene I. Virol J. 2011;8(1):526.
  • Nandi S, Chanda S, Bagchi P, et al. MAVS protein is attenuated by rotavirus nonstructural protein 1. PloS One. 2014;9(3):e92126.
  • Nandi S, Chanda S, Bagchi P, et al. Correction: MAVS protein is attenuated by rotavirus nonstructural protein 1. PloS One. 2015;10(6):e0131956.
  • Ding S, Zhu S, Ren L, et al. Rotavirus VP3 targets MAVS for degradation to inhibit type III interferon expression in intestinal epithelial cells. eLife. 2018;7:e39494.
  • Bagchi P, Bhowmick R, Nandi S, et al. Rotavirus NSP1 inhibits interferon induced non-canonical NFκB activation by interacting with TNF receptor associated factor 2. Virology. 2013;444(1–2):41–44.
  • Davis KA, Morelli M, Patton JT. Rotavirus NSP1 requires casein kinase II-mediated phosphorylation for hijacking of cullin-RING ligases. MBio. 2017;8(4):e01213–17.
  • Di Fiore IJ, Pane JA, Holloway G, et al. NSP1 of human rotaviruses commonly inhibits NF-κB signalling by inducing β-TrCP degradation. J Gen Virol. 2015;96(Pt 7):1768–1776.
  • Ding S, Mooney N, Li B, et al. Comparative proteomics reveals strain-specific β-TrCP degradation via rotavirus NSP1 hijacking a host Cullin-3-Rbx1 complex. PLoS Pathog. 2016;12(10):e1005929.
  • Graff JW, Ettayebi K, Hardy ME. Rotavirus NSP1 inhibits NFkappaB activation by inducing proteasome-dependent degradation of beta-TrCP: a novel mechanism of IFN antagonism. PLoS Pathog. 2009;5(1):e1000280.
  • Morelli M, Ogden KM, Patton JT. Silencing the alarms: innate immune antagonism by rotavirus NSP1 and VP3. Virology. 2015;479:75–84.
  • Holloway G, Truong TT, Coulson BS. Rotavirus antagonizes cellular antiviral responses by inhibiting the nuclear accumulation of STAT1, STAT2, and NF-kappaB. J Virol. 2009;83(10):4942–4951.
  • Holloway G, Coulson BS. Innate cellular responses to rotavirus infection. J Gen Virol. 2013;94(Pt 6):1151–1160.
  • Arnold MM, Barro M, Patton JT. Rotavirus NSP1 mediates degradation of interferon regulatory factors through targeting of the dimerization domain. J Virol. 2013;87(17):9813–9821.
  • Graff JW, Mitzel DN, Weisend CM, et al. Interferon regulatory factor 3 is a cellular partner of rotavirus NSP1. J Virol. 2002;76(18):9545–9550.
  • Barro M, Patton JT. Rotavirus nonstructural protein 1 subverts innate immune response by inducing degradation of IFN regulatory factor 3. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(11):4114–4119
  • Barro M, Patton JT. Rotavirus NSP1 inhibits expression of type I interferon by antagonizing the function of interferon regulatory factors IRF3, IRF5, and IRF7. J Virol. 2007;81(9):4473–4481.
  • Stark GR, Darnell JE,J. The JAK-STAT pathway at twenty. Immunity. 2012;36(4):503–514.
  • Sen A, Rott L, Phan N, et al. Rotavirus NSP1 protein inhibits interferon-mediated STAT1 activation. J Virol. 2014;88(1):41–53.
  • Sen A, Sharma A, Greenberg HB. Rotavirus degrades multiple interferon (IFN) type receptors to inhibit ifn signaling and protects against mortality from endotoxin in suckling mice. J Virol. 2017;92(1):e01394–17.
  • Feng N, Kim B, Fenaux M, et al. Role of interferon in homologous and heterologous rotavirus infection in the intestines and extraintestinal organs of suckling mice. J Virol. 2008;82(15):7578–7590.
  • Feng N, Yasukawa LL, Sen A, et al. Permissive replication of homologous murine rotavirus in the mouse intestine is primarily regulated by VP4 and NSP1. J Virol. 2013;87(15):8307–8316.
  • Sen A, Rothenberg ME, Mukherjee G, et al. Innate immune response to homologous rotavirus infection in the small intestinal villous epithelium at single-cell resolution. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(50):20667–20672
  • Bass DM. Interferon gamma and interleukin 1, but not interferon alfa, inhibit rotavirus entry into human intestinal cell lines. Gastroenterology. 1997;113(1):81–89.
  • Chanda SD, Banerjee A, Nandi S, et al. Cordycepin an adenosine analogue executes anti rotaviral effect by stimulating induction of type i interferon. J Virol Antivir Res. 2015;4(2):2.
  • Ding S, Diep J, Feng N, et al. STAG2 deficiency induces interferon responses via cGAS-STING pathway and restricts virus infection. Nat Commun. 2018;9(1):1485.
  • Hakim MS, Chen S, Ding S, et al. Basal interferon signaling and therapeutic use of interferons in controlling rotavirus infection in human intestinal cells and organoids. Sci Rep. 2018;8(1):8341.
  • Ogden KM, Hu L, Jha BK, et al. Structural basis for 2ʹ-5ʹ-oligoadenylate binding and enzyme activity of a viral RNase L antagonist. J Virol. 2015;89(13):6633–6645.
  • Sánchez-Tacuba L, Rojas M, Arias CF, et al. Rotavirus controls activation of the 2ʹ-5ʹ-Oligoadenylate Synthetase/RNase L pathway using at least two distinct mechanisms. J Virol. 2015;89(23):12145–12153.
  • Silverman RH. Viral encounters with 2ʹ,5ʹ-oligoadenylate synthetase and RNase L during the interferon antiviral response. J Virol. 2007;81(23):12720–12729.
  • Silverman RH, Weiss SR. Viral phosphodiesterases that antagonize double-stranded RNA signaling to RNase L by degrading 2-5A. J Interferon Cytokine Res. 2014;34(6):455–463.
  • Zhang R, Jha BK, Ogden KM, et al. (2013). Homologous 2ʹ,5ʹ-phosphodiesterases from disparate RNA viruses antagonize antiviral innate immunity. Proceedings of the National Academy of Sciences of the United States of America, 110( 32),13114–13119.
  • Bartel DP. Metazoan MicroRNAs. Cell. 2018;173(1):20–51.
  • Bivalkar-Mehla S, Vakharia J, Mehla R, et al. Viral RNA silencing suppressors (RSS): novel strategy of viruses to ablate the host RNA interference (RNAi) defense system. Virus Res. 2011;155(1):1–9.
  • Ding SW, Han Q, Wang J, et al. Antiviral RNA interference in mammals. Curr Opin Immunol. 2018;54:109–114.
  • Berkhout B. RNAi-mediated antiviral immunity in mammals. Curr Opin Virol. 2018;32:9–14.
  • Jeang KT. RNAi in the regulation of mammalian viral infections. BMC Biol. 2012;10(1):58.
  • Rojas M, Arias CF, López S. Protein kinase R is responsible for the phosphorylation of eIF2α in rotavirus infection. J Virol. 2010;84(20):10457–10466.
  • Zhu S, Ding S, Wang P, et al. Nlrp9b inflammasome restricts rotavirus infection in intestinal epithelial cells. Nature. 2017;546(7660):667–670.
  • Mukhopadhyay U, Chanda S, Patra U, et al. Biphasic regulation of RNA interference during rotavirus infection by modulation of Argonaute2. Cell Microbiol. 2019;21(12):e13101.
  • Dhillon P, Rao CD. Rotavirus induces formation of remodeled stress granules and P bodies and their sequestration in viroplasms to promote progeny virus production. J Virol. 2018;92(24):e01363–18.
  • Oceguera A, Peralta AV, Martínez-Delgado G, et al. Rotavirus RNAs sponge host cell RNA binding proteins and interfere with their subcellular localization. Virology. 2018;525:96–105.
  • Déctor MA, Romero P, López S, et al. Rotavirus gene silencing by small interfering RNAs. EMBO Rep. 2002;3(12):1175–1180.
  • Arias CF, Dector MA, Segovia L, et al. RNA silencing of rotavirus gene expression. Virus Res. 2004;102(1):43–51.
  • López T, Rojas M, Ayala-Bretón C, et al. Reduced expression of the rotavirus NSP5 gene has a pleiotropic effect on virus replication. J Gen Virol. 2005;86(Pt 6):1609–1617.
  • Chen F, Wang H, He H, et al. Short hairpin RNA-mediated silencing of bovine rotavirus NSP4 gene prevents diarrhoea in suckling mice. J Gen Virol. 2011;92(Pt 4):945–951.
  • Lee C. Therapeutic modulation of virus-induced oxidative stress via the Nrf2-Dependent antioxidative pathway. Oxid Med Cell Longev. 2018;(2018:6208067.
  • Ramezani A, Nahad MP, Faghihloo E. The role of Nrf2 transcription factor in viral infection. J Cell Biochem. 2018;119(8):6366–6382.
  • Buccigrossi V, Laudiero G, Russo C, et al. Chloride secretion induced by rotavirus is oxidative stress-dependent and inhibited by Saccharomyces boulardii in human enterocytes. PLoS One. 2014;9(6):e99830.
  • Gac M, Bigda J, Vahlenkamp TW. Increased mitochondrial superoxide dismutase expression and lowered production of reactive oxygen species during rotavirus infection. Virology. 2010;404(2):293–303.
  • De UK, Mukherjee R, Nandi S, et al. Alterations in oxidant/antioxidant balance, high-mobility group box 1 protein and acute phase response in cross-bred suckling piglets suffering from rotaviral enteritis. Trop Anim Health Prod. 2014;46(7):1127–1133.
  • Sodhi CP, Katyal R, Rana SV, et al. Study of oxidative-stress in rotavirus infected infant mice. Indian J Med Res. 1996;104:245–249.
  • Guererero CA, Murillo A, Acosta O. Inhibition of rotavirus infection in cultured cells by N-acetyl-cysteine, PPARγ agonists and NSAIDs. Antiviral Res. 2012;96(1):1–12.
  • Guerrero CA, Pardo P, Rodriguez V, et al. Inhibition of rotavirus ECwt infection in ICR suckling mice by N-acetylcysteine, peroxisome proliferator-activated receptor gamma agonists and cyclooxygenase-2 inhibitors. Memórias Inst Oswaldo Cruz. 2013;108(6):741–754.
  • Guerrero CA, Torres DP, García LL, et al. N‐Acetylcysteine treatment of rotavirus‐associated diarrhea in children. Pharmacother J Human Pharmacol Drug Ther. 2014;34(11):e333–e340.
  • Patra U, Mukhopadhyay U, Sarkar R, et al. RA-839, a selective agonist of Nrf2/ARE pathway, exerts potent anti-rotaviral efficacy in vitro. Antiviral Res. 2019;161:53–62.
  • Patra U, Mukhopadhyay U, Mukherjee A, et al. Progressive rotavirus infection downregulates Redox-Sensitive transcription factor Nrf2 and Nrf2-Driven transcription units. Oxid Med Cell Longev. 2020;(2020:7289120.
  • Danthi P. Viruses and the Diversity of Cell Death. Annu Rev Virol. 2016;3(1):533–553.
  • Chaïbi C, Cotte-Laffitte J, Sandré C, et al. Rotavirus induces apoptosis in fully differentiated human intestinal Caco-2 cells. Virology. 2005;332(2):480–490.
  • Martin-Latil S, Mousson L, Autret A, et al. Bax is activated during rotavirus-induced apoptosis through the mitochondrial pathway. J Virol. 2007;81(9):4457–4464.
  • Chattopadhyay S, Mukherjee A, Patra U, et al. Tyrosine phosphorylation modulates mitochondrial chaperonin Hsp60 and delays rotavirus NSP4‐mediated apoptotic signaling in host cells. Cell Microbiol. 2017;19(3):e12670.
  • Mukherjee A, Patra U, Bhowmick R, et al. Rotaviral nonstructural protein 4 triggers dynamin‐related protein 1‐dependent mitochondrial fragmentation during infection. Cell Microbiol. 2018;20(6):e12831.
  • Bagchi P, Dutta D, Chattopadhyay S, et al. Rotavirus nonstructural protein 1 suppresses virus-induced cellular apoptosis to facilitate viral growth by activating the cell survival pathways during early stages of infection. J Virol. 2010;84(13):6834–6845.
  • Bagchi P, Nandi S, Nayak MK, et al. Molecular mechanism behind rotavirus NSP1-mediated PI3 kinase activation: interaction between NSP1 and the p85 subunit of PI3 kinase. J Virol. 2013;87(4):2358–2362.
  • Dutta D, Bagchi P, Chatterjee A, et al. The molecular chaperone heat shock protein-90 positively regulates rotavirus infection. Virology. 2009;391(2):325–333.
  • Bhowmick R, Halder UC, Chattopadhyay S, et al. Rotavirus-encoded nonstructural protein 1 modulates cellular apoptotic machinery by targeting tumor suppressor protein p53. J Virol. 2013;87(12):6840–6850.
  • Chanda S, Nandi S, Chawla‐Sarkar M. Rotavirus‐induced miR‐142‐5p elicits proviral milieu by targeting non‐canonical transforming growth factor beta signalling and apoptosis in cells. Cell Microbiol. 2016;18(5):733–747.
  • Ngo C, Man SM. NLRP9b: a novel RNA-sensing inflammasome complex. Cell Res. 2017;27(11):1302–1303.
  • Wang R, Moniruzzaman M, Shuffle E, et al. Immune regulation of the unfolded protein response at the mucosal barrier in viral infection. Clin Transl Immunology. 2018;7(4):e1014.
  • Mehrbod P, Ande SR, Alizadeh J, et al. The roles of apoptosis, autophagy and unfolded protein response in arbovirus, influenza virus, and HIV infections. Virulence. 2019;10(1):376–413.
  • Trujillo-Alonso V, Maruri-Avidal L, Arias CF, et al. Rotavirus infection induces the unfolded protein response of the cell and controls it through the nonstructural protein NSP3. J Virol. 2011;85(23):12594–12604.
  • Zambrano JL, Ettayebi K, Maaty WS, et al. Rotavirus infection activates the UPR but modulates its activity. Virol J. 2011;8(1):359.
  • Duarte M, Vende P, Charpilienne A, et al. Rotavirus infection alters splicing of the stress-related transcription factor XBP1. J Virol. 2019;93(5):e01739–18.
  • Reineke LC, Lloyd RE. Diversion of stress granules and P-bodies during viral infection. Virology. 2013;436(2):255–267.
  • Bhowmick R, Mukherjee A, Patra U, et al. Rotavirus disrupts cytoplasmic P bodies during infection. Virus Res. 2015;210:344–354.
  • Green VA, Pelkmans L. A systems survey of progressive host-cell reorganization during rotavirus infection. Cell Host Microbe. 2016;20(1):107–120.
  • Montero H, Rojas M, Arias CF, et al. Rotavirus infection induces the phosphorylation of eIF2α but prevents the formation of stress granules. J Virol. 2008;82(3):1496–1504.
  • Dhillon P, Tandra VN, Chorghade SG, et al. Cytoplasmic relocalization and colocalization with viroplasms of host cell proteins, and their role in rotavirus infection. J Virol. 2018;92(15):e00612–18.
  • Rubio RM, Mora SI, Romero P, et al. Rotavirus prevents the expression of host responses by blocking the nucleocytoplasmic transport of polyadenylated mRNAs. J Virol. 2013;87(11):6336–6345.
  • Jackson RJ, Hellen CU, Pestova TV. The mechanism of eukaryotic translation initiation and principles of its regulation. Nature reviews. Mol Cell Biol. 2010;11(2):113–127.
  • Groft CM, Burley SK. Recognition of eIF4G by rotavirus NSP3 reveals a basis for mRNA circularization. Mol Cell. 2002;9(6):1273–1283.
  • Piron M, Vende P, Cohen J, et al. Rotavirus RNA-binding protein NSP3 interacts with eIF4GI and evicts the poly(A) binding protein from eIF4F. Embo J. 1998;17(19):5811–5821.
  • Vende P, Piron M, Castagné N, et al. Efficient translation of rotavirus mRNA requires simultaneous interaction of NSP3 with the eukaryotic translation initiation factor eIF4G and the mRNA 3ʹ end. J Virol. 2000;74(15):7064–7071.
  • Harb M, Becker MM, Vitour D, et al. Nuclear localization of cytoplasmic poly(A)-binding protein upon rotavirus infection involves the interaction of NSP3 with eIF4G and RoXaN. J Virol. 2008;82(22):11283–11293.
  • Gratia M, Sarot E, Vende P, et al. Rotavirus NSP3 is a translational surrogate of the poly (A) binding protein-poly (A) complex. J Virol. 2015;89(17):8773–8782.
  • Montero H, Arias CF, Lopez S. Rotavirus nonstructural protein NSP3 is not required for viral protein synthesis. J Virol. 2006;80(18):9031–9038.
  • Arnold MM, Brownback CS, Taraporewala ZF, et al. Rotavirus variant replicates efficiently although encoding an aberrant NSP3 that fails to induce nuclear localization of poly(A)-binding protein. J Gen Virol. 2012;93(Pt 7):1483–1494.
  • López S, Oceguera A, Sandoval-Jaime C. Stress response and translation control in rotavirus infection. Viruses. 2016;8(6):162.
  • Chen X, Cao R, Zhong W. Host calcium channels and pumps in viral infections. Cells. 2019;9(1):94.
  • Zhou Y, Frey TK, Yang JJ. Viral calciomics: interplays between Ca2+ and virus. Cell Calcium. 2009;46(1):1–17.
  • Ball JM, Tian P, Zeng CQ, et al. Age-dependent diarrhea induced by a rotaviral nonstructural glycoprotein. Science. 1996;272(5258):101–104. . (New York, N.Y.)
  • Tian P, Hu Y, Schilling WP, et al. The nonstructural glycoprotein of rotavirus affects intracellular calcium levels. J Virol. 1994;68(1):251–257.
  • Tian P, Estes MK, Hu Y, et al. The rotavirus nonstructural glycoprotein NSP4 mobilizes Ca2+ from the endoplasmic reticulum. J Virol. 1995;69(9):5763–5772.
  • Zhang M, Zeng CQ, Morris AP, et al. A functional NSP4 enterotoxin peptide secreted from rotavirus-infected cells. J Virol. 2000;74(24):11663–11670.
  • Dong Y, Zeng CQ, Ball JM, et al. The rotavirus enterotoxin NSP4 mobilizes intracellular calcium in human intestinal cells by stimulating phospholipase C-mediated inositol 1,4,5-trisphosphate production. Proceedings of the National Academy of Sciences of the United States of America. 1997;94(8):3960–3965
  • Hyser JM, Collinson-Pautz MR, Utama B, et al. Rotavirus disrupts calcium homeostasis by NSP4 viroporin activity. mBio. 2010;1(5):e00265–10.
  • Hyser JM, Utama B, Crawford SE, et al. Activation of the endoplasmic reticulum calcium sensor STIM1 and store-operated calcium entry by rotavirus requires NSP4 viroporin activity. J Virol. 2013;87(24):13579–13588.
  • Pham T, Perry JL, Dosey TL, et al. The rotavirus NSP4 viroporin domain is a calcium-conducting ion channel. Sci Rep. 2017;7(1):43487.
  • Chang-Graham AL, Perry JL, Strtak AC, et al. Rotavirus calcium dysregulation manifests as dynamic calcium signaling in the cytoplasm and endoplasmic reticulum. Sci Rep. 2019;9(1):10822.
  • Díaz Y, Peña F, Aristimuño OC, et al. Dissecting the Ca2⁺ entry pathways induced by rotavirus infection and NSP4-EGFP expression in Cos-7 cells. Virus Res. 2012;167(2):285–296.
  • Pérez JF, Ruiz MC, Chemello ME, et al. Characterization of a membrane calcium pathway induced by rotavirus infection in cultured cells. J Virol. 1999;73(3):2481–2490.
  • Morris AP, Scott JK, Ball JM, et al. NSP4 elicits age-dependent diarrhea and Ca(2+)mediated I(-) influx into intestinal crypts of CF mice. A J Physiol. 1999;277(2):G431–G444.
  • Didsbury A, Wang C, Verdon D, et al. Rotavirus NSP4 is secreted from infected cells as an oligomeric lipoprotein and binds to glycosaminoglycans on the surface of non-infected cells. Virol J. 2011;8(1):551.
  • Gibbons TF, Storey SM, Williams CV, et al. Rotavirus NSP4: cell type-dependent transport kinetics to the exofacial plasma membrane and release from intact infected cells. Virol J. 2011;8(1):278.
  • Ousingsawat J, Mirza M, Tian Y, et al. Rotavirus toxin NSP4 induces diarrhea by activation of TMEM16A and inhibition of Na+ absorption. Pflugers Arch. 2011;461(5):579–589.
  • Jiang Y, Yu B, Yang H, et al. Shikonin Inhibits intestinal Calcium-Activated chloride channels and prevents rotaviral diarrhea. Front Pharmacol. 2016;7:270.
  • Ko EA, Jin BJ, Namkung W, et al. Chloride channel inhibition by a red wine extract and a synthetic small molecule prevents rotaviral secretory diarrhoea in neonatal mice. Gut. 2014;63(7):1120–1129.
  • Tradtrantip L, Ko EA, Verkman AS. Antidiarrheal efficacy and cellular mechanisms of a Thai herbal remedy. PLoS Negl Trop Dis. 2014;8(2):e2674.
  • Yu B, Jiang Y, Zhang B, et al. Resveratrol dimer trans-ε-viniferin prevents rotaviral diarrhea in mice by inhibition of the intestinal calcium-activated chloride channel. Pharmacol Res. 2018;129:453–461.
  • Yu B, Zhu X, Yang X, et al. Plumbagin prevents secretory diarrhea by inhibiting CaCC and CFTR channel activities. Front Pharmacol. 2019;10:1181.
  • Bialowas S, Hagbom M, Nordgren J, et al. Rotavirus and serotonin Cross-Talk in Diarrhoea. PloS One. 2016;11(7):e0159660.
  • Hagbom M, Istrate C, Engblom D, et al. Rotavirus stimulates release of serotonin (5-HT) from human enterochromaffin cells and activates brain structures involved in nausea and vomiting. PLoS Pathog. 2011;7(7):e1002115.
  • Hagbom M, Sharma S, Lundgren O, et al. Towards a human rotavirus disease model. Curr Opin Virol. 2012;2(4):408–418.
  • Hagbom M, Svensson L. Rotavirus disease mechanisms. In: Svensson L, Desselberger U, Greenberg HB, et al., editors. Viral gastroenteritis. Molecular epidemiology and pathogenesis. Cambridge: Academic Press; 2016. p. 189–218.
  • Kordasti S, Sjövall H, Lundgren O, et al. Serotonin and vasoactive intestinal peptide antagonists attenuate rotavirus diarrhoea. Gut. 2004;53(7):952–957.
  • Das JK, Kumar R, Salam RA, et al. The effect of antiemetics in childhood gastroenteritis. BMC Public Health. 2013;13(Suppl3):S9.
  • Freedman SB, Pasichnyk D, Black KJ, et al., Pediatric Emergency Research Canada Gastroenteritis Study Group. Gastroenteritis therapies in developed countries: systematic review and Meta-Analysis. PloS One. 2015;10(6):e0128754. .
  • Sen A, Sen N, Mackow ER. The formation of viroplasm-like structures by the rotavirus NSP5 protein is calcium regulated and directed by a C-terminal helical domain. J Virol. 2007;81(21):11758–11767.
  • Aoki ST, Settembre EC, Trask SD, et al. Structure of rotavirus outer-layer protein VP7 bound with a neutralizing Fab. Science. 2009;324(5933):1444–1447. . (New York, N.Y.)
  • Dormitzer PR, Greenberg HB, Harrison SC. Purified recombinant rotavirus VP7 forms soluble, calcium-dependent trimers. Virology. 2000;277(2):420–428.
  • Bhowmick R, Banik G, Chanda S, et al. Rotavirus infection induces G1 to S phase transition in MA104 cells via Ca+ 2/Calmodulin pathway. Virology. 2014;454:270–279.
  • Chattopadhyay S, Basak T, Nayak MK, et al. Identification of cellular calcium binding protein calmodulin as a regulator of rotavirus A infection during comparative proteomic study. PloS One. 2013;8(2):e56655.
  • Fan Y, Sanyal S, Bruzzone R. breaking bad: how viruses subvert the cell cycle. Front Cell Infect Microbiol. 2018;8:396.
  • Glück S, Buttafuoco A, Meier AF, et al. Rotavirus replication is correlated with S/G2 interphase arrest of the host cell cycle. PloS One. 2017;12(6):e0179607.
  • Eichwald C, Arnoldi F, Laimbacher AS, et al. Rotavirus viroplasm fusion and perinuclear localization are dynamic processes requiring stabilized microtubules. PloS One. 2012;7(10):e47947.
  • Campagna M, Eichwald C, Vascotto F, et al. RNA interference of rotavirus segment 11 mRNA reveals the essential role of NSP5 in the virus replicative cycle. J Gen Virol. 2005;86(Pt 5):1481–1487.
  • Campagna M, Budini M, Arnoldi F, et al. Impaired hyperphosphorylation of rotavirus NSP5 in cells depleted of casein kinase 1alpha is associated with the formation of viroplasms with altered morphology and a moderate decrease in virus replication. J Gen Virol. 2007;88(Pt 10):2800–2810.
  • Eichwald C, Jacob G, Muszynski B, et al. Uncoupling substrate and activation functions of rotavirus NSP5: phosphorylation of Ser-67 by casein kinase 1 is essential for hyperphosphorylation. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(46):16304–16309
  • Criglar JM, Hu L, Crawford SE, et al. A novel form of rotavirus NSP2 and phosphorylation-dependent NSP2-NSP5 interactions are associated with viroplasm assembly. J Virol. 2014;88(2):786–798.
  • Criglar JM, Anish R, Hu L, et al. Phosphorylation cascade regulates the formation and maturation of rotaviral replication factories. Proceedings of the National Academy of Sciences of the United States of America. 2018;115(51):E12015–E12023
  • Papa G, Venditti L, Arnoldi F, et al. Recombinant rotaviruses rescued by reverse genetics reveal the role of NSP5 hyperphosphorylation in the assembly of viral factories. J Virol. 2019;94(1):e01110–19.
  • Herker E, Ott M. Emerging role of lipid droplets in host/pathogen interactions. J Biol Chem. 2012;287(4):2280–2287.
  • Zhang J, Lan Y, Sanyal S. Modulation of lipid droplet Metabolism-A potential target for therapeutic intervention in flaviviridae infections. Front Microbiol. 2017;8:2286.
  • Cheung W, Gill M, Esposito A, et al. Rotaviruses associate with cellular lipid droplet components to replicate in viroplasms, and compounds disrupting or blocking lipid droplets inhibit viroplasm formation and viral replication. J Virol. 2010;84(13):6782–6798.
  • Criglar JM, Crawford SE, Zhao B, et al. A genetically engineered rotavirus NSP2 phosphorylation mutant impaired in viroplasm formation and replication shows an early interaction between vNSP2 and cellular lipid droplets. J Virol. 2020;94(15):e00972–20.
  • Gaunt ER, Cheung W, Richards JE, et al. Inhibition of rotavirus replication by downregulation of fatty acid synthesis. J Gen Virol. 2013;94(Pt 6):1310–1317.
  • Kim Y, George D, Prior AM, et al. Novel triacsin C analogs as potential antivirals against rotavirus infections. Eur J Med Chem. 2012;50:311–318.
  • Crawford SE, Desselberger U. Lipid droplets form complexes with viroplasms and are crucial for rotavirus replication. Curr Opin Virol. 2016;19:11–15.
  • Gaunt ER, Zhang Q, Cheung W, et al. Lipidome analysis of rotavirus-infected cells confirms the close interaction of lipid droplets with viroplasms. J Gen Virol. 2013;94(Pt 7):1576–1586.
  • Kim Y, Chang KO. Inhibitory effects of bile acids and synthetic farnesoid X receptor agonists on rotavirus replication. J Virol. 2011;85(23):12570–12577.
  • Hoffmann HH, Kunz A, Simon VA, et al. Broad-spectrum antiviral that interferes with de novo pyrimidine biosynthesis. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(14):5777–5782
  • Luthra P, Naidoo J, Pietzsch CA, et al. Inhibiting pyrimidine biosynthesis impairs Ebola virus replication through depletion of nucleoside pools and activation of innate immune responses. Antiviral Res. 2018;158:288–302.
  • Chen S, Ding S, Yin Y, et al. Suppression of pyrimidine biosynthesis by targeting DHODH enzyme robustly inhibits rotavirus replication. Antiviral Res. 2019;167:35–44.
  • Yin Y, Wang Y, Dang W, et al. Mycophenolic acid potently inhibits rotavirus infection with a high barrier to resistance development. Antiviral Res. 2016;133:41–49.
  • Chen S, Wang Y, Li P, et al. Drug screening identifies gemcitabine inhibiting rotavirus through alteration of pyrimidine nucleotide synthesis pathway. Antiviral Res. 2020;180:104823.
  • Bosco EE, Mulloy JC, Zheng Y. Rac1 GTPase: a “Rac” of all trades. Cell Mol Life Sci. 2009;66(3):370–374.
  • Yin Y, Chen S, Hakim MS, et al. 6-Thioguanine inhibits rotavirus replication through suppression of Rac1 GDP/GTP cycling. Antiviral Res. 2018;156:92–101.
  • Holloway G, Coulson BS. Rotavirus activates JNK and p38 signaling pathways in intestinal cells, leading to AP-1-driven transcriptional responses and enhanced virus replication. J Virol. 2006;80(21):10624–10633.
  • Yin Y, Dang W, Zhou X, et al. PI3K-Akt-mTOR axis sustains rotavirus infection via the 4E-BP1 mediated autophagy pathway and represents an antiviral target. Virulence. 2018;9(1):83–98.
  • Contin R, Arnoldi F, Mano M, et al. Rotavirus replication requires a functional proteasome for effective assembly of viroplasms. J Virol. 2011;85(6):2781–2792.
  • López T, Silva-Ayala D, López S, et al. Replication of the rotavirus genome requires an active ubiquitin-proteasome system. J Virol. 2011;85(22):11964–11971.
  • Graff JW, Ewen J, Ettayebi K, et al. Zinc-binding domain of rotavirus NSP1 is required for proteasome-dependent degradation of IRF3 and autoregulatory NSP1 stability. J Gen Virol. 2007;88(Pt 2):613–620.
  • Arnoldi F, De Lorenzo G, Mano M, et al. Rotavirus increases levels of lipidated LC3 supporting accumulation of infectious progeny virus without inducing autophagosome formation. PLoS One. 2014;9(4):4.
  • Lutz LM, Pace CR, Arnold MM. Rotavirus NSP1 associates with components of the Cullin RING ligase family of E3 ubiquitin ligases. J Virol. 2016;90(13):6036–6048.
  • Flotho A, Melchior F. Sumoylation: a regulatory protein modification in health and disease. Annu Rev Biochem. 2013;82(1):357–385.
  • Raman N, Nayak A, Muller S. The SUMO system: a master organizer of nuclear protein assemblies. Chromosoma. 2013;122(6):475–485.
  • Campagna M, Marcos-Villar L, Arnoldi F, et al. Rotavirus viroplasm proteins interact with the cellular SUMOylation system: implications for viroplasm-like structure formation. J Virol. 2013;87(2):807–817.
  • Bartek J, Lukas J. Mammalian G1- and S-phase checkpoints in response to DNA damage. Curr Opin Cell Biol. 2001;13(6):738–747.
  • Luftig MA. Viruses and the DNA damage response: activation and antagonism. Annu Rev Virol. 2014;1(1):605–625.
  • Zhou BB, Elledge SJ. The DNA damage response: putting checkpoints in perspective. Nature. 2000;408(6811):433–439.
  • Sarkar R, Patra U, Lo M, et al. Rotavirus activates a noncanonical ATM-Chk2 branch of DNA damage response during infection to positively regulate viroplasm dynamics. Cell Microbiol. 2020;22(3):e13149.
  • Ren L, Ding S, Song Y, et al. Profiling of rotavirus 3ʹUTR-binding proteins reveals the ATP synthase subunit ATP5B as a host factor that supports late-stage virus replication. J Biol Chem. 2019;294(15):5993–6006.
  • Ahmad L, Mostowy S, Sancho-Shimizu V. Autophagy-Virus Interplay: from cell biology to human disease. Front Cell Dev Biol. 2018;6:155.
  • Choi Y, Bowman JW, Jung JU. Autophagy during viral infection - a double-edged sword. Nature reviews. Microbiology. 2018;16(6):341–354.
  • Mukhopadhyay U, Chanda S, Patra U, et al. Synchronized orchestration of miR-99b and let-7g positively regulates rotavirus infection by modulating autophagy. Sci Rep. 2019;9(1):1318.
  • Tian G, Liang X, Chen D, et al. Vitamin D3 supplementation alleviates rotavirus infection in pigs and IPEC-J2 cells via regulating the autophagy signaling pathway. J Steroid Biochem Mol Biol. 2016;163:157–163.
  • Kindrachuk J. Selective inhibition of host cell signaling for rotavirus antivirals: PI3K/Akt/mTOR-mediated rotavirus pathogenesis. Virulence. 2018;9(1):5–8.
  • Maruri-Avidal L, López S, Arias CF. Endoplasmic reticulum chaperones are involved in the morphogenesis of rotavirus infectious particles. J Virol. 2008;82(11):5368–5380.
  • Mirazimi A, Nilsson M, Svensson L. The molecular chaperone calnexin interacts with the NSP4 enterotoxin of rotavirus in vivo and in vitro. J Virol. 1998;72(11):8705–8709.
  • Mirazimi A, Svensson L. Carbohydrates facilitate correct disulfide bond formation and folding of rotavirus VP7. J Virol. 1998;72(5):3887–3892.
  • Xu A, Bellamy AR, Taylor JA. BiP (GRP78) and endoplasmin (GRP94) are induced following rotavirus infection and bind transiently to an endoplasmic reticulum-localized virion component. J Virol. 1998;72(12):9865–9872.
  • Martínez JL, Arnoldi F, Schraner EM, et al. The Guanine Nucleotide Exchange Factor GBF1 participates in rotavirus replication. J Virol. 2019;93(19):e01062–19.
  • Martínez JL, Arias CF. Role of the Guanine Nucleotide Exchange Factor GBF1 in the replication of RNA viruses. Viruses. 2020;12(6):682.
  • Cuadras MA, Greenberg HB. Rotavirus infectious particles use lipid rafts during replication for transport to the cell surface in vitro and in vivo. Virology. 2003;313(1):308–321.
  • Mohan KV, Muller J, Atreya CD. Defective rotavirus particle assembly in lovastatin-treated MA104 cells. Arch Virol. 2008;153(12):2283–2290.
  • Trejo-Cerro Ó, Eichwald C, Schraner EM, et al. Actin-Dependent nonlytic rotavirus exit and infectious virus morphogenetic pathway in nonpolarized cells. J Virol. 2018;92(6):e02076–17.
  • Santiana M, Ghosh S, Ho BA, et al. Vesicle-Cloaked virus clusters are optimal units for Inter-organismal viral transmission. Cell Host Microbe. 2018;24(2):208–220.e8.
  • Iša P, Pérez-Delgado A, Quevedo IR, et al. Rotaviruses associate with distinct types of extracellular vesicles. Viruses. 2020;12(7):E763.
  • Ashburn TT, Thor KB. Drug repositioning: identifying and developing new uses for existing drugs. Nature reviews. Drug Discovery. 2004;3(8):673–683.
  • Law GL, Tisoncik-Go J, Korth MJ, et al. Drug repurposing: a better approach for infectious disease drug discovery?.Curr Opin Immunol. 2013;25(5):588–592.
  • Emparanza Knörr JI, Ozcoidi Erro I, Martínez Andueza MC, et al. Revisión sistemática sobre la eficacia de racecadotrilo en el tratamiento de la diarrea aguda [Systematic review of the efficacy of racecadotril in the treatment of acute diarrhoea]. Anales de pediatria (Barcelona, Spain: 2003). 2008;69(5):432–438.
  • Gharial J, Laving A, Were F. Racecadotril for the treatment of severe acute watery diarrhoea in children admitted to a tertiary hospital in Kenya. BMJ Open Gastroenterol. 2017;4(1):e000124.
  • Gordon M, Akobeng A. Racecadotril for acute diarrhoea in children: systematic review and meta-analyses. Arch Dis Child. 2016;101(3):234–240.
  • Kang G, Thuppal SV, Srinivasan R, et al. Racecadotril in the management of rotavirus and Non-rotavirus diarrhea in under-five children: two randomized, Double-blind, Placebo-controlled trials. Indian Pediatr. 2016;53(7):595–600.
  • La Frazia S, Ciucci A, Arnoldi F, et al. Thiazolides, a new class of antiviral agents effective against rotavirus infection, target viral morphogenesis, inhibiting viroplasm formation. J Virol. 2013;87(20):11096–11106.
  • Mahapatro S, Mahilary N, Satapathy AK, et al. Nitazoxanide in acute rotavirus diarrhea: a randomized control trial from a developing country. J Trop Med. 2017;(2017:7942515.
  • Rossignol JF. Nitazoxanide: a first-in-class broad-spectrum antiviral agent. Antiviral Res. 2014;110:94–103.
  • Rossignol JF, Abu-Zekry M, Hussein A, et al. Effect of nitazoxanide for treatment of severe rotavirus diarrhoea: randomised double-blind placebo-controlled trial. Lancet. 2006;368(9530):124–129. . (London, England)
  • Buccigrossi V, Russo C, Guarino A, et al. Mechanisms of antidiarrhoeal effects by diosmectite in human intestinal cells. Gut Pathog. 2017;9(1):23.
  • Das RR, Sankar J, Naik SS. Efficacy and safety of diosmectite in acute childhood diarrhoea: a meta-analysis. Arch Dis Child. 2015;100(7):704–712.
  • Dupont C, Foo JL, Garnier P, et al., Peru and Malaysia Diosmectite Study Groups. Oral diosmectite reduces stool output and diarrhea duration in children with acute watery diarrhea. Clin Gastroenterol Hepatol. 2009;7(4):456–462. .
  • Eichwald C, De Lorenzo G, Schraner EM, et al. Identification of a small molecule that compromises the structural integrity of viroplasms and rotavirus Double-Layered particles. J Virol. 2018;92(3):e01943–17.
  • Van Dycke J, Arnoldi F, Papa G, et al. A single nucleoside viral polymerase inhibitor against norovirus, rotavirus, and sapovirus-induced diarrhea. J Infect Dis. 2018;218(11):1753–1758.
  • Saxena K, Blutt SE, Ettayebi K, et al. Human intestinal enteroids: a new model to study human rotavirus infection, host restriction, and pathophysiology. J Virol. 2015;90(1):43–56.
  • Saxena K, Simon LM, Zeng XL, et al. A paradox of transcriptional and functional innate interferon responses of human intestinal enteroids to enteric virus infection. Proceedings of the National Academy of Sciences of the United States of America. 2017;114(4):E570–E579
  • Kanai Y, Komoto S, Kawagishi T, et al. Entirely plasmid-based reverse genetics system for rotaviruses. Proceedings of the National Academy of Sciences of the United States of America. 2017;114(9):2349–2354
  • Kanai Y, Kawagishi T, Nouda R, et al. Development of stable rotavirus reporter expression systems. J Virol. 2019;93(4):e01774–18.
  • Kawagishi T, Nurdin JA, Onishi M, et al. Reverse genetics system for a human group a rotavirus. J Virol. 2020;94(2):e00963–19.
  • Sánchez-Tacuba L, Feng N, Meade NJ, et al. An optimized reverse genetics system suitable for efficient recovery of simian, human and murine-like rotaviruses. J Virol. 2020;94(18). 10.1128/JVI.01294-20. JVI.01294-20. Advance online publication