Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2389 word(s) 2389 2022-02-22 07:44:32 |
2 format correct -27 word(s) 2362 2022-03-01 07:22:56 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Rescher, U. Antifungals for Host-Directed Antiviral Therapy. Encyclopedia. Available online: https://encyclopedia.pub/entry/19992 (accessed on 18 May 2024).
Rescher U. Antifungals for Host-Directed Antiviral Therapy. Encyclopedia. Available at: https://encyclopedia.pub/entry/19992. Accessed May 18, 2024.
Rescher, Ursula. "Antifungals for Host-Directed Antiviral Therapy" Encyclopedia, https://encyclopedia.pub/entry/19992 (accessed May 18, 2024).
Rescher, U. (2022, February 28). Antifungals for Host-Directed Antiviral Therapy. In Encyclopedia. https://encyclopedia.pub/entry/19992
Rescher, Ursula. "Antifungals for Host-Directed Antiviral Therapy." Encyclopedia. Web. 28 February, 2022.
Antifungals for Host-Directed Antiviral Therapy
Edit

Because of their epidemic and pandemic potential, emerging viruses are a major threat to global healthcare systems. While vaccination is in general a straightforward approach to prevent viral infections, immunization can also cause escape mutants that hide from immune cell and antibody detection. Thus, other approaches than immunization are critical for the management and control of viral infections. Viruses are prone to mutations leading to the rapid emergence of resistant strains upon treatment with direct antivirals. In contrast to the direct interference with pathogen components, host-directed therapies aim to target host factors that are essential for the pathogenic replication cycle or to improve the host defense mechanisms, thus circumventing resistance. These relatively new approaches are often based on the repurposing of drugs which are already licensed for the treatment of other unrelated diseases. 

antifungals host-directed drug therapy drug repurposing azoles polyenes echinocandins

1. Polyenes—Disruptors of Fungal Membrane

Polyenes exert their antifungal activity by binding ergosterol in the fungal cell membrane. The resulting membrane disintegration by pore formation increases the permeability, and the membrane leakage leads to the subsequent death of the fungal cell. A well-known, highly effective polyene is amphotericin B, which is widely used in the clinics against invasive fungal infections. Amphotericin B was shown to also induce oxidative stress to fungal cells and modulate the immune system [1].
A potential antiviral capacity of amphotericin B has been reported against a variety of viruses, including human immunodeficiency virus (HIV) [2], Japanese encephalitis virus (JEV) [3], herpes simplex virus (HSV) [4], rubella virus [3] and vesicular stomatitis virus (VSV) [5]. Amphotericin B inhibits the infectivity of JEV in a concentration-dependent manner up to 200-fold at a postinfection step likely at viral replication and/or synthesis of viral proteins without affecting virus adsorption to host cell surfaces [3].
In vitro results highlight the possibility of using amphotericin B against HIV infections, providing a dual effect, against the virus itself and the opportunistic fungal infections that often accompany HIV infections due to the patients’ compromised immune system [2]. Furthermore, amphotericin B potentiated the antiviral efficacy of acyclovir against pseudorabies virus (PRV) without a direct effect on PRV replication in the absence of acyclovir [6]. The mechanism of action remains unclear; however, a recent high-throughput virtual screening approach showed amphotericin B, among other antiparasitic drugs, to possess potential inhibitory features against 10 SARS-CoV-2 molecular targets including the RNA-dependent RNA polymerase [7].

2. Flucytosine—A Selective Inhibitor of Fungal Nuclear Acid Synthesis

Flucytosine as such has no antifungal capacity but is converted by the fungal enzyme cytosine deaminase (which is not present in mammalian cells) into 5-fluorouracil, which is further metabolized. The incorporation of 5-fluorouracil and its metabolites into fungal DNA and RNA then causes aberrant fungal RNA and DNA synthesis [8][9]

3. Echinocandins—Noncompetitive Inhibitors of (1,3)-β-D-Glucan Synthase

Echinocandins attack the fungal cell wall by inhibiting the (1,3)-β-D-glucan synthesis and thereby triggering osmotic stress and subsequent cell lysis. They are fungicidal against molds and yeast (most Candida species) and are generally considered well-tolerable due to little adverse effects and drug–drug interactions. Therefore, they are the preferred treatment option for invasive candidiasis [10][11]. Micafungin is an FDA-approved echinocandin that exhibits broad antifungal activity against a variety of Candida species. Of note, a potential antiviral capacity against chikungunya virus (CHIKV) [12], enteroviruses [13] and dengue virus [14], among others, has been reported in recent years. Mosquito-borne CHIKV belongs to the alphaviruses and is a global health problem. Micafungin was able to attenuate the cytopathic effects of CHIKV, reduce viral replication, release and spread, and impair viral stability. Micafungin also had antiviral effects against the alphaviruses Sindbis virus (SINV) and Semliki Forest virus (SFV) [12]. Enterovirus 71 (EV71), the major causative agent of hand-foot-and-mouth disease (HFMD), was shown sensitive to micafungin treatment as well. Micafungin effectively diminished EV71 proliferation and replication already at a micromolar dose [13]. Against other enteroviruses like Coxsackievirus group B type 3 (CVB3) and human rhinovirus (HRV), its antiviral capacity was only moderate. The authors proposed a virion-independent mechanism of action targeting intracellular processes such as translation, polyprotein processing or replication [13]. Recently, micafungin and its analogs, caspofungin and anidulafungin, were suggested for treatment of dengue virus (DENV) infection. In this case, the mechanism of action depends on the direct binding to the envelope protein DENV-2, thereby destabilizing and destroying the virion [14]. Of note, recent in silico studies argued for a binding of the echinocandins micafungin and pneumocandin B0 to the 3C-like protease (3CLpro) of porcine epidemic diarrhea virus (PEDV), as well as to the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) main protease (Mpro), with micafungin having a higher calculated binding affinity towards 3CLpro and pneumocandin B0 binding preferably to Mpro, warranting further investigation into the use of echinocandin-type antifungal drugs as antiviral agents that act via binding to defined viral molecular targets [15].

4. Azoles—Inhibitors of Ergosterol Biosynthesis

The antifungal azoles are classified by the numbers of nitrogen atoms in the azole ring and include imidazoles (e.g., ketoconazole, miconazole, and clotrimazole) and triazoles (e.g., itraconazole, posaconazole, and fluconazole) [16]. Their efficacy and relative safety have led to their widespread clinical use for the antifungal therapy against aspergillosis, candidiasis, and cryptococcosis [17]. While the clinical use of imidazoles (the only exception is ketoconazole) is limited to the treatment of superficial mycoses, triazoles are used for superficial and systemic fungal infections [18]. Azoles exert their antifungal activities through multiple modes of action. By inhibiting the two cytochrome P450 enzymes that function in the ergosterol biosynthesis (CYP51, lanosterol 14α-demethylase) and the conversion of ergostatrienol into ergostatetraenol, (CYP61, 22-desaturase), azoles lead to a depletion of ergosterol. The resulting accumulation of toxic sterol precursors impairs membrane fluidity, asymmetry, and membrane integrity in fungal cells [19].
Interestingly, some of the azoles (itraconazole, posaconazole, voriconazole, and ketoconazole) also impact the mammalian cholesterol metabolism at higher concentrations [20]. Itraconazole and posaconazole influence cellular cholesterol levels by impairing different pivotal steps in the cholesterol homeostasis. Both antifungals include (i) the inhibition of the lanosterol 14α-demethylase and thereby the homeostasis and de novo synthesis of cholesterol [21][22], (ii) the inhibition of the cholesterol-transferring membrane protein Niemann–Pick C1 (NPC1) that results in the accumulation of cholesterol in the endolysosomal system [23], and they further (iii) interact with the oxysterol-binding protein (OSBP), which blocks the shuttling of cholesterol and phosphatidylinositol-4-phosphate (PI4P) between membranes [24]

4.1. Itraconazole Directly Interacts with the Endolysosomal Cholesterol Transporter NPC1

Niemann–Pick disease, type C1 (NPC1) protein NPC1 was first identified and characterized as a membrane protein that when mutated causes Niemann–Pick disease, type C1, a rare autosomal neurovisceral lipid storage disorder [25]. NPC1 protein is an endolysosomal integral membrane protein and mediates endolysosomal cholesterol transport [26]. A dysfunctional NPC1 protein, which is found mutated in 95% of the NPC patients, disturbs the intracellular lipid transport, leading to the excessive accumulation of lipid products including cholesterol in the endolysosomal compartment [25]. Interestingly, a possible function of NPC1 as a drug target in antiviral strategies has been explored in several recent publications [27][28]. The Ebola virus entry has been shown to directly rely on NPC1 function via binding of the Ebola virus glycoprotein (GP) to NPC1 [29]. The Ebola virus GP is cleaved by endosomal proteases to unmask the NPC1 binding site, and GP–NPC1 engagement within lysosomes promotes the viral escape into the host cytoplasm [30]. Consistent with this vital dependence of the Ebola virus replication on NPC1 protein, cells lacking NPC1 are nonpermissive for the virus entry and NPC1 knockout mice are protected from lethal Ebola virus infection [29]. Blocking NPC1 has also been reported to cause accumulation of the HIV-1 viral gag protein in the endolysosomal compartment [31][32], resulting in a profound suppression of virion release [32]. The finding that the HIV-1 accessory protein Nef induces host cell genes involved in cholesterol biosynthesis and homeostasis [33] emphasizes the strong dependency of HIV-1 on the host cell cholesterol levels, suggesting that NPC1 is a candidate drug target in the treatment of HIV-1 infections. Interestingly, NPC1 also emerged as a candidate drug target for other enveloped viruses, namely IAV and SARS-CoV-2. Viral replication rates were decreased in cells in which NPC1 was functionally blocked, and the increased endolysosomal cholesterol levels were suggested to interfere with the proper insertion of the fusogenic IAV hemagglutinin domains and the SARS-CoV-2 spike protein, thus affecting virus uncoating [34][35][36].
In many of these and other in vitro studies exploring the importance of NPC1 in diverse cellular functions, the cell-permeable hydrophobic polyamine U18666A, a small-molecule NPC1 inhibitor, was used [37][38]. However, the substantial toxicity of this compound limits a clinical use [39][40]. Of note, itraconazole has been shown to also directly bind and inhibit NPC1 [23], and thus might serve as an attractive candidate for NPC1 targeting strategies via drug repurposing. In favor of this notion, itraconazole-treated cells generated lesser IAV and Ebola virus progeny [36][41], and a beneficial treatment outcome was indeed confirmed in a mouse IAV infection model in vivo [36]. Itraconazole-mediated induction of type I interferons (IFNs), which is considered a fundamental step in establishing antiviral immunity, might also contribute to the observed antiviral effects [36]. While itraconazole also proved its antiviral potential in a 3D cell culture model for SARS-CoV-2 infection [34], an antiviral effect was not seen in the hamster infection model [42].

4.2. Itraconazole Interferes with OSBP and OSBP-Related Proteins (ORP) Functionality

Azoles also impair cellular lipid metabolism via an inhibitory effect on oxysterol-binding protein 1 (OSBP) and on other proteins that belong to the OSBP-related proteins (ORP) family, and this property might add to their antiviral use. OSBP was first identified as an intracellular protein that binds cytosolic 25-hydroxycholesterol [43]. Beside its capacity to bind 25-hydroxycholesterol, OSBP orchestrates the formation of endoplasmic reticulum (ER)–Golgi complex membrane contact sites and thereby shuttles sterols into the Golgi and phosphatidylinositol-4-phosphate (PI4P) back to the ER [44]. OSBP and the family of OSBP-related proteins (ORP) share a lipid-binding domain that binds either a sterol or a nonsterol ligand as well as a PI4P-binding N-terminal pleckstrin homology (PH) domain [45][46]. Another binding motif found in many ORPs, including OSBP, is the FFAT-motif which interacts with the ER-resident VAMP-associated proteins (VAP) receptors [45][46]. Both motifs are involved in shaping the ER–Golgi or, in the case of some viruses, the ER–replication organelle (RO) contact sites and are considerably engaged in lipid transport through different organelles.
Enterovirus, dengue virus, and hepatitis C virus replication are reported to depend on ORP and OSBP [47][24]. Pharmacologic inhibition, siRNA knockdown, and rescue of replication by overexpression have demonstrated the importance of ORPs and OSBP for enterovirus replication and propagation [24]. The virus-induced accumulation of PI4P lipids drives the recruitment of OSBP to these contact sides, and the OSBP-mediated transport of cholesterol and PI4P is pivotal for the formation and functionality of the enterovirus RO [24][48][49][50]. In line, OSBP knockdown and treatment with 25-hydroxycholesterol, an inhibitor of the cholesterol-PI4P exchange, negatively affects virus replication [51]; however, the precise molecular mechanism remains unclear. Itraconazole directly binds OSBP [24], leading to increased PI4P levels at the Golgi (in uninfected cells) or the RO (in infected cells), while the accumulation of cholesterol at the RO is blocked [24].

4.3. Targeting mTOR Signaling via Itraconazole

Itraconazole not only impairs lipid homeostasis, but also affects different signaling pathways, including mammalian target of rapamycin (mTOR), hedgehog, and Wnt signaling pathways that are hijacked by a broad range of viruses to drive the production of infectious particles.
The mammalian target of rapamycin (mTOR) signaling cascade is a pivotal signaling pathway that regulates apoptosis and counteracts stress-induced autophagy (such as, e.g., that elicited by viruses). Although different cellular locations for mTOR complex 1 (mTORC1) and mTORC2 have been reported, mTORC1 lysosomal localization appears critical for its ability to sense and respond to cell starvation [52]. Cholesterol was recently identified to promote the recruitment of mTORC1 to the lysosomal membrane [53] and the mTOR signaling cascade is regulated in a cholesterol-dependent manner [54].
Several viruses have evolved strategies to subvert the mTORC1 signaling network to drive their replication and propagation [55][56][57][58][59][60][61][62][63]. The Semliki Forest virus (SFV), Sindbis virus (SINV), and Chikungunya virus (CHIKV), members of the alphavirus family, cause different diseases but have in common that they encode nonstructural proteins (nsP) [64]. The activation of the PI3K/Akt/mTOR pathway is mediated through the phosphorylated and membrane-attached protein nsP3, which forms the viral replication complex upon virus internalization [65][66][67]. The activation of mTOR signaling is also fundamental for infection with the Flaviviridae West Nile virus (WNV), Japanese encephalitis virus (JEV) and dengue virus (DENV) [68][69][70]Flaviviridae infection increases mTOR activity through a PI3K-dependent mechanism to maintain translation of its positive-sense RNA genome and also delays WNV-induced apoptosis [58][71][72]. The hepatitis C virus, another Flaviviridae member, increases phosphorylation of mTOR through the nonstructural protein 5A (NS5A) [73]. NS5A seems to activate PI3K/Akt signaling by directly binding PI3K [74][75]. The activation of the mTORC1 pathway by HCV has been linked to antiapoptotic signals that ensure cell survival and maintain persistence by promoting steady-state levels of virus replication [76][77]. Among the β-herpesvirus, the human cytomegalovirus (HCMV) maintains mTORC1 activation [78][79] through the expression of the two HCMV immediate early proteins, IEP72 and IEP86 [80]. Another persistent virus that tightly regulates mTOR signaling pathways is the human immunodeficiency virus type 1 (HIV-1). In dendritic cells, the HIV-1 envelope glycoprotein activates mTOR to prevent autophagy and to increase virus infection. Pharmacological treatment with rapamycin decreased viral spreading [81]. Another study implied that the HIV-1 protein Nef initiates mTOR activation which can be blocked by inhibitors of mTOR or PI3K [82][83], suggesting that drugs that modify the mTORC1 signaling pathway could act as anti-HIV-1 agents [84][85].
Well-balanced mTOR signaling is vitally important for IAV infection [86][87]. Thus, the pharmacological inhibition of the mTOR signaling axis might serve as potential antiviral target. In contrast to other azoles, itraconazole additionally inhibits mTOR signaling through affecting the upstream 5′-AMP-dependent protein kinase (AMPK) [88], which is activated upon an increased AMP/ATP ratio and serves as a regulator of cellular energy levels [89]. Once activated, AMPK inhibits mTOR signaling [90]. The activation of AMPK through itraconazole is a result of direct binding and inhibition of the mitochondrial Voltage-Dependent Anion Channel 1 (VDAC1), a critical regulator of mitochondrial metabolism, resulting in a drop in cellular energy levels [91]. Itraconazole treatment also impairs vascular endothelial growth factor receptor 2 (VEGFR2) functionality in endothelial cells, which is mostly due to altered VEGFR2 glycosylation, trafficking, and signaling [92]. As some viruses like human papillomaviruses or hepatitis viruses promote angiogenesis to facilitate optimal supply by nutrients [93][94][95], this might be an additional beneficial effect of itraconazole in antiviral strategy.

References

  1. Mesa-Arango, A.C.; Scorzoni, L.; Zaragoza, O. It only takes one to do many jobs: Amphotericin B as antifungal and immunomodulatory drug. Front. Microbiol. 2012, 3, 286.
  2. Konopka, K.; Guo, L.S.S.; Düzgüneş, N. Anti-HIV activity of amphotericin B-cholesteryl sulfate colloidal dispersion in vitro. Antiviral Res. 1999, 42, 197–209.
  3. Kim, H.; Kim, S.J.; Park, S.N.; Oh, J.W. Antiviral effect of amphotericin B on Japanese encephalitis virus replication. J. Microbiol. Biotechnol. 2004, 14, 121–127.
  4. Shiota, H.; Jones, B.R.; Schaffner, C.P. Anti-Herpes Simplex Virus (HSV) Effect of Amphotericin B Methyl Ester In Vivo. In Parasites, Fungi, and Viruses; Springer US: Boston, MA, USA, 1976; pp. 339–346.
  5. Jordan, G.W.; Humphreys, S.; Zee, Y.C. Effect of amphotericin B methyl ester on vesicular stomatitis virus morphology. Antimicrob. Agents Chemother. 1978, 13, 340–341.
  6. Malewicz, B.; Momsen, M.; Jenkin, H.M.; Borowski, E. Potentiation of antiviral activity of acyclovir by polyene macrolide antibiotics. Antimicrob. Agents Chemother. 1984, 25, 772–774.
  7. Marak, B.N.; Dowarah, J.; Khiangte, L.; Singh, V.P. Step toward repurposing drug discovery for COVID-19 therapeutics through in silico approach. Drug Dev. Res. 2021, 82, 374–392.
  8. Di Mambro, T.; Guerriero, I.; Aurisicchio, L.; Magnani, M.; Marra, E. The yin and yang of current antifungal therapeutic strategies: How can we harness our natural defenses? Front. Pharmacol. 2019, 10, 80.
  9. Carmona, E.M.; Limper, A.H. Overview of Treatment Approaches for Fungal Infections. Clin. Chest Med. 2017, 38, 393–402.
  10. Gintjee, T.J.; Donnelley, M.A.; Thompson, G.R. Aspiring Antifungals: Review of Current Antifungal Pipeline Developments. J. Fungi 2020, 6, 28.
  11. Van Daele, R.; Spriet, I.; Wauters, J.; Maertens, J.; Mercier, T.; Van Hecke, S.; Brüggemann, R. Antifungal drugs: What brings the future? Med. Mycol. 2019, 57, S328–S343.
  12. Ho, Y.-J.; Liu, F.-C.; Yeh, C.-T.; Yang, C.M.; Lin, C.-C.; Lin, T.-Y.; Hsieh, P.-S.; Hu, M.-K.; Gong, Z.; Lu, J.-W. Micafungin is a novel anti-viral agent of chikungunya virus through multiple mechanisms. Antiviral Res. 2018, 159, 134–142.
  13. Kim, C.; Kang, H.; Kim, D.E.; Song, J.H.; Choi, M.; Kang, M.; Lee, K.; Kim, H.S.; Shin, J.S.; Jeong, H.; et al. Antiviral activity of micafungin against enterovirus 71. Virol. J. 2016, 13, 1–9.
  14. Chen, Y.-C.; Lu, J.-W.; Yeh, C.-T.; Lin, T.-Y.; Liu, F.-C.; Ho, Y.-J. Micafungin Inhibits Dengue Virus Infection through the Disruption of Virus Binding, Entry, and Stability. Pharmaceuticals 2021, 14, 338.
  15. Vergoten, G.; Bailly, C. In silico analysis of echinocandins binding to the main proteases of coronaviruses PEDV (3CLpro) and SARS-CoV-2 (Mpro). Silico Pharmacol. 2021, 9, 1–10.
  16. Sheehan, D.J.; Hitchcock, C.A.; Sibley, C.M. Current and emerging azole antifungal agents. Clin. Microbiol. Rev. 1999, 12, 40–79.
  17. Francois, I.; Cammue, B.; Borgers, M.; Ausma, J.; Dispersyn, G.; Thevissen, K. Azoles: Mode of Antifungal Action and Resistance Development. Effect of Miconazole on Endogenous Reactive Oxygen Species Production in Candida albicans. Antiinfect. Agents Med. Chem. 2006, 5, 3–13.
  18. Lass-Flörl, C. Triazole antifungal agents in invasive fungal infections: A comparative review. Drugs 2011, 71, 2405–2419.
  19. Kelly, S.L.; Lamb, D.C.; Baldwin, B.C.; Corran, A.J.; Kelly, D.E. Characterization of Saccharomyces cerevisiae CYP61, Sterol Δ22-Desaturase, and Inhibition by Azole Antifungal Agents. J. Biol. Chem. 1997, 272, 9986–9988.
  20. Minnebruggen, G.V.; Francois, I.E.J.A.; Cammue, B.P.A.; Thevissen, K.; Vroome, V.; Borgers, M.; Shroot, B. A General Overview on Past, Present and Future Antimycotics. Open Mycol. J. 2010, 4, 22–32.
  21. Warrilow, A.G.; Parker, J.E.; Kelly, D.E.; Kelly, S.L. Azole Affinity of Sterol 14-Demethylase (CYP51) Enzymes from Candida albicans and Homo sapiens. Antimicrob. Agents Chemother. 2013, 57, 1352–1360.
  22. Munayyer, H.K.; Mann, P.A.; Chau, A.S.; Yarosh-Tomaine, T.; Greene, J.R.; Hare, R.S.; Heimark, L.; Palermo, R.E.; Loebenberg, D.; McNicholas, P.M. Posaconazole is a potent inhibitor of sterol 14α-demethylation in yeasts and molds. Antimicrob. Agents Chemother. 2004, 48, 3690–3696.
  23. Trinh, M.N.; Lu, F.; Li, X.; Das, A.; Liang, Q.; De Brabander, J.K.; Brown, M.S.; Goldstein, J.L.; Chang, T.-Y.; Maxfield, F.R. Triazoles inhibit cholesterol export from lysosomes by binding to NPC1. Proc. Natl. Acad. Sci. USA 2017, 114, 89–94.
  24. Strating, J.R.P.M.; van der Linden, L.; Albulescu, L.; Bigay, J.; Arita, M.; Delang, L.; Leyssen, P.; van der Schaar, H.M.; Lanke, K.H.W.; Thibaut, H.J.; et al. Itraconazole Inhibits Enterovirus Replication by Targeting the Oxysterol-Binding Protein. Cell Rep. 2015, 10, 600–615.
  25. Vanier, M.; Millat, G. Niemann-Pick disease type C. Clin. Genet. 2003, 64, 269–281.
  26. Pfeffer, S.R. NPC intracellular cholesterol transporter 1 (NPC1)-mediated cholesterol export from lysosomes. J. Biol. Chem. 2019, 294, 1706–1709.
  27. Herbert, A.S.; Davidson, C.; Kuehne, A.I.; Bakken, R.; Braigen, S.Z.; Gunn, K.E.; Whelan, S.P.; Brummelkamp, T.R.; Twenhafel, N.A.; Chandran, K.; et al. Niemann-pick C1 is essential for ebolavirus replication and pathogenesis in vivo. MBio 2015, 6, 1–12.
  28. Côté, M.; Misasi, J.; Ren, T.; Bruchez, A.; Lee, K.; Filone, C.M.; Hensley, L.; Li, Q.; Ory, D.; Chandran, K.; et al. Small molecule inhibitors reveal Niemann-Pick C1 is essential for Ebola virus infection. Nature 2011, 477, 344–348.
  29. Carette, J.E.; Raaben, M.; Wong, A.C.; Herbert, A.S.; Obernosterer, G.; Mulherkar, N.; Kuehne, A.I.; Kranzusch, P.J.; Griffin, A.M.; Ruthel, G.; et al. Ebola virus entry requires the cholesterol transporter Niemann-Pick C1. Nature 2011, 477, 340–343.
  30. Miller, E.H.; Obernosterer, G.; Raaben, M.; Herbert, A.S.; Deffieu, M.S.; Krishnan, A.; Ndungo, E.; Sandesara, R.G.; Carette, J.E.; Kuehne, A.I.; et al. Ebola virus entry requires the host-programmed recognition of an intracellular receptor. EMBO J. 2012, 31, 1947–1960.
  31. Coleman, E.M.; Walker, T.N.; Hildreth, J.E.K. Loss of Niemann Pick type C proteins 1 and 2 greatly enhances HIV infectivity and is associated with accumulation of HIV Gag and cholesterol in late endosomes/lysosomes. Virol. J. 2012, 9, 31.
  32. Tang, Y.; Leao, I.C.; Coleman, E.M.; Broughton, R.S.; Hildreth, J.E.K. Deficiency of Niemann-Pick Type C-1 Protein Impairs Release of Human Immunodeficiency Virus Type 1 and Results in Gag Accumulation in Late Endosomal/Lysosomal Compartments. J. Virol. 2009, 83, 7982–7995.
  33. van ′t Wout, A.B.; Swain, J.V.; Schindler, M.; Rao, U.; Pathmajeyan, M.S.; Mullins, J.I.; Kirchhoff, F. Nef Induces Multiple Genes Involved in Cholesterol Synthesis and Uptake in Human Immunodeficiency Virus Type 1-Infected T Cells. J. Virol. 2005, 79, 10053–10058.
  34. Schloer, S.; Brunotte, L.; Mecate-Zambrano, A.; Zheng, S.; Tang, J.; Ludwig, S.; Rescher, U. Drug synergy of combinatory treatment with remdesivir and the repurposed drugs fluoxetine and itraconazole effectively impairs SARS-CoV-2 infection in vitro. Br. J. Pharmacol. 2021, 178, 2339–2350.
  35. Van Damme, E.; De Meyer, S.; Bojkova, D.; Ciesek, S.; Cinatl, J.; De Jonghe, S.; Jochmans, D.; Leyssen, P.; Buyck, C.; Neyts, J.; et al. In vitro activity of itraconazole against SARS-CoV-2. J. Med. Virol. 2021, 93, 4454–4460.
  36. Schloer, S.; Goretzko, J.; Kühnl, A.; Brunotte, L.; Ludwig, S.; Rescher, U. The clinically licensed antifungal drug itraconazole inhibits influenza virus in vitro and in vivo. Emerg. Microbes Infect. 2019, 8, 80–93.
  37. Lu, F.; Liang, Q.; Abi-Mosleh, L.; Das, A.; de Brabander, J.K.; Goldstein, J.L.; Brown, M.S. Identification of NPC1 as the target of U18666A, an inhibitor of lysosomal cholesterol export and Ebola infection. Elife 2015, 4, 1–16.
  38. Liscum, L.; Faust, J.R. The intracellular transport of low density lipoprotein-derived cholesterol is inhibited in Chinese hamster ovary cells cultured with 3-β-androst-5-en-17-one. J. Biol. Chem. 1989, 264, 11796–11806.
  39. Lloyd-Evans, E.; Morgan, A.J.; He, X.; Smith, D.A.; Elliot-Smith, E.; Sillence, D.J.; Churchill, G.C.; Schuchman, E.H.; Galione, A.; Platt, F.M. Niemann-Pick disease type C1 is a sphingosine storage disease that causes deregulation of lysosomal calcium. Nat. Med. 2008, 14, 1247–1255.
  40. Cenedella, R.J.; Jacob, R.; Borchman, D.; Tang, D.; Neely, A.R.; Samadi, A.; Mason, R.P.; Sexton, P. Direct perturbation of lens membrane structure may contribute to cataracts caused by U18666A, an oxidosqualene cyclase inhibitor. J. Lipid Res. 2004, 45, 1232–1241.
  41. Kummer, S.; Lander, A.; Goretzko, J.; Kirchoff, N.; Rescher, U.; Schloer, S. Pharmacologically induced endolysosomal cholesterol imbalance through clinically licensed drugs itraconazole and fluoxetine impairs Ebola virus infection in vitro. Emerg. Microbes Infect. 2022, 11, 195–207.
  42. Liesenborghs, L.; Spriet, I.; Jochmans, D.; Belmans, A.; Gyselinck, I.; Teuwen, L.-A.; ter Horst, S.; Dreesen, E.; Geukens, T.; Engelen, M.M.; et al. Itraconazole for COVID-19: Preclinical studies and a proof-of-concept randomized clinical trial. EBioMedicine 2021, 66, 103288.
  43. Ridgway, N.D.; Dawson, P.A.; Ho, Y.K.; Brown, M.S.; Goldstein, J.L. Translocation of oxysterol binding protein to Golgi apparatus triggered by ligand binding. J. Cell Biol. 1992, 116, 307–319.
  44. Mesmin, B.; Bigay, J.; Moser von Filseck, J.; Lacas-Gervais, S.; Drin, G.; Antonny, B. A Four-Step Cycle Driven by PI(4)P Hydrolysis Directs Sterol/PI(4)P Exchange by the ER-Golgi Tether OSBP. Cell 2013, 155, 830–843.
  45. Weber-Boyvat, M.; Zhong, W.; Yan, D.; Olkkonen, V.M. Oxysterol-binding proteins: Functions in cell regulation beyond lipid metabolism. Biochem. Pharmacol. 2013, 86, 89–95.
  46. Raychaudhuri, S.; Prinz, W.A. The diverse functions of oxysterol-binding proteins. Annu. Rev. Cell Dev. Biol. 2010, 26, 157–177.
  47. Meutiawati, F.; Bezemer, B.; Strating, J.R.P.M.; Overheul, G.J.; Žusinaite, E.; van Kuppeveld, F.J.M.; van Cleef, K.W.R.; van Rij, R.P. Posaconazole inhibits dengue virus replication by targeting oxysterol-binding protein. Antiviral Res. 2018, 157, 68–79.
  48. Hsu, N.Y.; Ilnytska, O.; Belov, G.; Santiana, M.; Chen, Y.H.; Takvorian, P.M.; Pau, C.; van der Schaar, H.; Kaushik-Basu, N.; Balla, T.; et al. Viral reorganization of the secretory pathway generates distinct organelles for RNA replication. Cell 2010, 141, 799–811.
  49. 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.
  50. Arita, M. Phosphatidylinositol-4 kinase III beta and oxysterol-binding protein accumulate unesterified cholesterol on poliovirus-induced membrane structure. Microbiol. Immunol. 2014, 58, 239–256.
  51. Roulin, P.S.; Lötzerich, M.; Torta, F.; Tanner, L.B.; Van Kuppeveld, F.J.M.; Wenk, M.R.; Greber, U.F. Rhinovirus uses a phosphatidylinositol 4-phosphate/cholesterol counter-current for the formation of replication compartments at the ER-Golgi interface. Cell Host Microbe 2014, 16, 677–690.
  52. Betz, C.; Hall, M.N. Where is mTOR and what is it doing there? J. Cell Biol. 2013, 203, 563–574.
  53. Castellano, B.M.; Thelen, A.M.; Moldavski, O.; Feltes, M.; Van Der Welle, R.E.N.; Mydock-McGrane, L.; Jiang, X.; Van Eijkeren, R.J.; Davis, O.B.; Louie, S.M.; et al. Lysosomal cholesterol activates mTORC1 via an SLC38A9-Niemann-Pick C1 signaling complex. Science 2017, 355, 1306–1311.
  54. Lim, C.Y.; Davis, O.B.; Shin, H.R.; Zhang, J.; Berdan, C.A.; Jiang, X.; Counihan, J.L.; Ory, D.S.; Nomura, D.K.; Zoncu, R. ER–lysosome contacts enable cholesterol sensing by mTORC1 and drive aberrant growth signalling in Niemann–Pick type C. Nat. Cell Biol. 2019, 21, 1206–1218.
  55. Karam, B.S.; Morris, R.S.; Bramante, C.T.; Puskarich, M.; Zolfaghari, E.J.; Lotfi-Emran, S.; Ingraham, N.E.; Charles, A.; Odde, D.J.; Tignanelli, C.J. mTOR inhibition in COVID-19: A commentary and review of efficacy in RNA viruses. J. Med. Virol. 2021, 93, 1843–1846.
  56. Shin, Y.K.; Liu, Q.; Tikoo, S.K.; Babiuk, L.A.; Zhou, Y. Effect of the phosphatidylinositol 3-kinase/Akt pathway on influenza A virus propagation. J. Gen. Virol. 2007, 88, 942–950.
  57. Hale, B.G.; Jackson, D.; Chen, Y.H.; Lamb, R.A.; Randall, R.E. Influenza A virus NS1 protein binds p85β and activates phosphatidylinositol-3-kinase signaling. Proc. Natl. Acad. Sci. USA 2006, 103, 14194–14199.
  58. Shives, K.D.; Beatman, E.L.; Chamanian, M.; O’Brien, C.; Hobson-Peters, J.; Beckham, J.D. West Nile Virus-Induced Activation of Mammalian Target of Rapamycin Complex 1 Supports Viral Growth and Viral Protein Expression. J. Virol. 2014, 88, 9458–9471.
  59. Bose, S.K.; Shrivastava, S.; Meyer, K.; Ray, R.B.; Ray, R. Hepatitis C Virus Activates the mTOR/S6K1 Signaling Pathway in Inhibiting IRS-1 Function for Insulin Resistance. J. Virol. 2012, 86, 6315–6322.
  60. Joubert, P.E.; Stapleford, K.; Guivel-Benhassine, F.; Vignuzzi, M.; Schwartz, O.; Albert, M.L. Inhibition of mTORC1 Enhances the Translation of Chikungunya Proteins via the Activation of the MnK/eIF4E Pathway. PLoS Pathog. 2015, 11, e1005091.
  61. Das, I.; Basantray, I.; Mamidi, P.; Nayak, T.K.; Pratheek, B.M.; Chattopadhyay, S.; Chattopadhyay, S. Heat shock protein 90 positively regulates Chikungunya virus replication by stabilizing viral non-structural protein nsP2 during infection. PLoS One 2014, 9, e100531.
  62. Shi, Y.; He, X.; Zhu, G.; Tu, H.; Liu, Z.; Li, W.; Han, S.; Yin, J.; Peng, B.; Liu, W. Coxsackievirus A16 elicits incomplete autophagy involving the mTOR and ERK pathways. PLoS One 2015, 10, e0122109.
  63. Kuss-Duerkop, S.K.; Wang, J.; Mena, I.; White, K.; Metreveli, G.; Sakthivel, R.; Mata, M.A.; Muñoz-Moreno, R.; Chen, X.; Krammer, F.; et al. Influenza Virus Differentially Activates mTORC1 and mTORC2 Signaling to Maximize Late Stage Replication. PLoS Pathog. 2017, 13, e1006635.
  64. Bakar, F.A.; Ng, L.F.P. Nonstructural proteins of alphavirus—potential targets for drug development. Viruses 2018, 10, 71.
  65. Mohankumar, V.; Dhanushkodi, N.R.; Raju, R. Sindbis virus replication, is insensitive to rapamycin and torin1, and suppresses Akt/mTOR pathway late during infection in HEK cells. Biochem. Biophys. Res. Commun. 2011, 406, 262–267.
  66. Joubert, P.E.; Werneke, S.W.; de la Calle, C.; Guivel-Benhassine, F.; Giodini, A.; Peduto, L.; Levine, B.; Schwartz, O.; Lenschow, D.J.; Albert, M.L. Chikungunya virus-induced autophagy delays caspase-dependent cell death. J. Exp. Med. 2012, 209, 1029–1047.
  67. Thaa, B.; Biasiotto, R.; Eng, K.; Neuvonen, M.; Götte, B.; Rheinemann, L.; Mutso, M.; Utt, A.; Varghese, F.; Balistreri, G.; et al. Differential Phosphatidylinositol-3-Kinase-Akt-mTOR Activation by Semliki Forest and Chikungunya Viruses Is Dependent on nsP3 and Connected to Replication Complex Internalization. J. Virol. 2015, 89, 11420–11437.
  68. Klaitong, P.; Smith, D.R. Roles of non-structural protein 4a in flavivirus infection. Viruses 2021, 13, 2077.
  69. Lahon, A.; Arya, R.P.; Banerjea, A.C. Dengue Virus Dysregulates Master Transcription Factors and PI3K/AKT/mTOR Signaling Pathway in Megakaryocytes. Front. Cell. Infect. Microbiol. 2021, 11, 715208.
  70. Albentosa-González, L.; de Oya, N.J.; Arias, A.; Clemente-Casares, P.; Martin-Acebes, M.Á.; Saiz, J.C.; Sabariegos, R.; Mas, A. Akt kinase intervenes in flavivirus replication by interacting with viral protein ns5. Viruses 2021, 13, 896.
  71. Urbanowski, M.D.; Hobman, T.C. The West Nile Virus Capsid Protein Blocks Apoptosis through a Phosphatidylinositol 3-Kinase-Dependent Mechanism. J. Virol. 2013, 87, 872–881.
  72. Lee, C.-J.; Liao, C.-L.; Lin, Y.-L. Flavivirus Activates Phosphatidylinositol 3-Kinase Signaling To Block Caspase-Dependent Apoptotic Cell Death at the Early Stage of Virus Infection. J. Virol. 2005, 79, 8388–8399.
  73. George, A.; Panda, S.; Kudmulwar, D.; Chhatbar, S.P.; Nayak, S.C.; Krishnan, H.H. Hepatitis C virus NS5A binds to the mRNA cap-binding eukaryotic translation initiation 4F (elF4F) complex and up-regulates host translation initiation machinery through elF4E-binding protein 1 inactivation. J. Biol. Chem. 2012, 287, 5042–5058.
  74. He, Y.; Nakao, H.; Tan, S.-L.; Polyak, S.J.; Neddermann, P.; Vijaysri, S.; Jacobs, B.L.; Katze, M.G. Subversion of Cell Signaling Pathways by Hepatitis C Virus Nonstructural 5A Protein via Interaction with Grb2 and P85 Phosphatidylinositol 3-Kinase. J. Virol. 2002, 76, 9207–9217.
  75. Street, A.; Macdonald, A.; Crowder, K.; Harris, M. The Hepatitis C Virus NS5A Protein Activates a Phosphoinositide 3-Kinase-dependent Survival Signaling Cascade. J. Biol. Chem. 2004, 279, 12232–12241.
  76. Le Sage, V.; Cinti, A.; Amorim, R.; Mouland, A.J. Adapting the stress response: Viral subversion of the mTOR signaling pathway. Viruses 2016, 8, 152.
  77. Mannová, P.; Beretta, L. Activation of the N-Ras–PI3K–Akt-mTOR Pathway by Hepatitis C Virus: Control of Cell Survival and Viral Replication. J. Virol. 2005, 79, 8742–8749.
  78. Clippinger, A.J.; Maguire, T.G.; Alwine, J.C. Human Cytomegalovirus Infection Maintains mTOR Activity and Its Perinuclear Localization during Amino Acid Deprivation. J. Virol. 2011, 85, 9369–9376.
  79. Kudchodkar, S.B.; Del Prete, G.Q.; Maguire, T.G.; Alwine, J.C. AMPK-Mediated Inhibition of mTOR Kinase Is Circumvented during Immediate-Early Times of Human Cytomegalovirus Infection. J. Virol. 2007, 81, 3649–3651.
  80. Yu, Y.; Alwine, J.C. Human Cytomegalovirus Major Immediate-Early Proteins and Simian Virus 40 Large T Antigen Can Inhibit Apoptosis through Activation of the Phosphatidylinositide 3′-OH Kinase Pathway and the Cellular Kinase Akt. J. Virol. 2002, 76, 3731–3738.
  81. Blanchet, F.P.; Moris, A.; Nikolic, D.S.; Lehmann, M.; Cardinaud, S.; Stalder, R.; Garcia, E.; Dinkins, C.; Leuba, F.; Wu, L.; et al. Human immunodeficiency virus-1 inhibition of immunoamphisomes in dendritic cells impairs early innate and adaptive immune responses. Immunity 2010, 32, 654–669.
  82. Campbell, G.R.; Rawat, P.; Bruckman, R.S.; Spector, S.A. Human Immunodeficiency Virus Type 1 Nef Inhibits Autophagy through Transcription Factor EB Sequestration. PLoS Pathog. 2015, 11, e1005018.
  83. Cinti, A.; Le Sage, V.; Milev, M.P.; Valiente-Echeverría, F.; Crossie, C.; Miron, M.-J.; Panté, N.; Olivier, M.; Mouland, A.J. HIV-1 enhances mTORC1 activity and repositions lysosomes to the periphery by co-opting Rag GTPases. Sci. Rep. 2017, 7, 5515.
  84. Heredia, A.; Le, N.; Gartenhaus, R.B.; Sausville, E.; Medina-Moreno, S.; Zapata, J.C.; Davis, C.; Gallo, R.C.; Redfield, R.R. Targeting of mTOR catalytic site inhibits multiple steps of the HIV-1 lifecycle and suppresses HIV-1 viremia in humanized mice. Proc. Natl. Acad. Sci. USA 2015, 112, 9412–9417.
  85. Donia, M.; McCubrey, J.A.; Bendtzen, K.; Nicoletti, F. Potential use of rapamycin in HIV infection. Br. J. Clin. Pharmacol. 2010, 70, 784–793.
  86. Zhou, Z.; Jiang, X.; Liu, D.; Fan, Z.; Hu, X.; Yan, J.; Wang, M.; Gao, G.F. Autophagy is involved in influenza A virus replication. Autophagy 2009, 5, 321–328.
  87. Zhirnov, O.P.; Konakova, T.E.; Garten, W.; Klenk, H.-D. Caspase-Dependent N-Terminal Cleavage of Influenza Virus Nucleocapsid Protein in Infected Cells. J. Virol. 1999, 73, 10158–10163.
  88. Tsubamoto, H.; Inoue, K.; Sakata, K.; Ueda, T.; Takeyama, R.; Shibahara, H.; Sonoda, T. Itraconazole inhibits AKT/mTOR signaling and proliferation in endometrial cancer cells. Anticancer Res. 2017, 37, 515–520.
  89. Garcia, D.; Shaw, R.J. AMPK: Mechanisms of Cellular Energy Sensing and Restoration of Metabolic Balance. Mol. Cell 2017, 66, 789–800.
  90. Cork, G.K.; Thompson, J.; Slawson, C. Real Talk: The Inter-play Between the mTOR, AMPK, and Hexosamine Biosynthetic Pathways in Cell Signaling. Front. Endocrinol. (Lausanne) 2018, 9, 522.
  91. Head, S.A.; Shi, W.Q.; Yang, E.J.; Nacev, B.A.; Hong, S.Y.; Pasunooti, K.K.; Li, R.J.; Shim, J.S.; Liu, J.O. Simultaneous targeting of NPC1 and VDAC1 by itraconazole leads to synergistic inhibition of MTOR signaling and angiogenesis. ACS Chem. Biol. 2017, 12, 174–182.
  92. Nacev, B.A.; Grassi, P.; Dell, A.; Haslam, S.M.; Liu, J.O. The antifungal drug itraconazole inhibits Vascular Endothelial Growth Factor Receptor 2 (VEGFR2) glycosylation, trafficking, and signaling in endothelial cells. J. Biol. Chem. 2011, 286, 44045–44056.
  93. Vrancken, K.; Paeshuyse, J.; Liekens, S. Angiogenic activity of hepatitis B and C viruses. Antivir. Chem. Chemother. 2012, 22, 159–170.
  94. Li, G.; He, L.; Zhang, E.; Shi, J.; Zhang, Q.; Le, A.D.; Zhou, K.; Tang, X. Overexpression of human papillomavirus (HPV) type 16 oncoproteins promotes angiogenesis via enhancing HIF-1α and VEGF expression in non-small cell lung cancer cells. Cancer Lett. 2011, 311, 160–170.
  95. Hassan, M.; Selimovic, D.; El-Khattouti, A.; Soell, M.; Ghozlan, H.; Haikel, Y.; Abdelkader, O.; Megahed, M. Hepatitis C virus-mediated angiogenesis: Molecular mechanisms and therapeutic strategies. World J. Gastroenterol. 2014, 20, 15467–15475.
More
Information
Subjects: Respiratory System
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 340
Revisions: 2 times (View History)
Update Date: 01 Mar 2022
1000/1000