key: cord-0738013-790d7v7q authors: Mottin, Melina; Borba, Joyce V.V.B.; Braga, Rodolpho C.; Torres, Pedro H.M.; Martini, Matheus C.; Proenca-Modena, Jose Luiz; Judice, Carla C.; Costa, Fabio T.M.; Ekins, Sean; Perryman, Alexander L.; Horta Andrade, Carolina title: The A–Z of Zika drug discovery date: 2018-06-20 journal: Drug Discov Today DOI: 10.1016/j.drudis.2018.06.014 sha: d9a9ee2dce090f86cb22d0325e3ec14f2e3f3adb doc_id: 738013 cord_uid: 790d7v7q Despite the recent outbreak of Zika virus (ZIKV), there are still no approved treatments, and early-stage compounds are probably many years away from approval. A comprehensive A–Z review of the recent advances in ZIKV drug discovery efforts is presented, highlighting drug repositioning and computationally guided compounds, including discovered viral and host cell inhibitors. Promising ZIKV molecular targets are also described and discussed, as well as targets belonging to the host cell, as new opportunities for ZIKV drug discovery. All this knowledge is not only crucial to advancing the fight against the Zika virus and other flaviviruses but also helps us prepare for the next emerging virus outbreak to which we will have to respond. Zika virus (ZIKV) remains a global health concern. Originally discovered in Africa in 1947 [1] , ZIKV became an epidemic 60 years later, reaching several tropical regions of the Americas, Africa and Asia. Despite causing mild symptoms such as fever, rashes and conjunctivitis, the major concern about ZIKV regards the severe neurological disorders, such as microcephaly, craniofacial disproportion, spasticity, seizures, irritability and other brainstem dysfunctions [2, 3] . In 2016, infant head computed tomographic findings, of subjects infected during pregnancy, confirmed the causal relationship between microcephaly and Zika infection [3] . A recent study by Yuan and co-workers demonstrated that a single mutation (S139N) in the pre-membrane (prM) structural protein increased ZIKV infectivity in neural progenitor cells (NPCs), making the virus more virulent [4] . This mutation arose in the French Polynesia strain, and it has contributed to the increased incidence of microcephaly and higher mortality in neonates, according to experimental assays [4] . The disorders attributed to Zika infection mainly affect infants but can also impact adults. There have been ZIKV-related cases of Guillain-Barré syndrome [5] , myelitis [6] , uveitis [7] and meningoencephalitis [8] reported in adults. Currently, neither a specific antiviral drug nor a vaccine is available for treating or preventing ZIKV infection. However, there are several promising drug targets encoded by the virus or present in host cells. There have been several reports on compounds found to have activity against ZIKV and its proteins. Here, we present a comprehensive A-Z review of the recent advances in ZIKV drug design, including viral and host cell inhibitors and several experimental and computational techniques that have been applied in these studies. This information will contribute to the design of drugs against ZIKV and related viruses. ZIKV is a spherical, enveloped virus, with an icosahedral-like symmetry [9] (Fig. 1a) . Belonging to the genus Flavivirus of the Scheme of Zika virus (ZIKV) surface, structural and nonstructural proteins. (a) Surface-shaded depth cued representation of mature ZIKV (built using UCSF Chimera package [171] , http://www.rbvi.ucsf.edu/chimera, based on PDB ID 5IRE), showing the icosahedral-like symmetry arrangement of surface proteins. (b) Virion components, highlighting the E, M and C proteins, as well as genomic RNA. ZIKV encodes a large polyprotein, which after processing yields three structural proteins (C, M and E) and seven nonstructural proteins (NS1; NS2A; NS2B; NS3 protease and helicase domains; NS4A; NS4B; NS5 methyltransferase and RNA polymerase domains), built using the VMD program [172] (http://www.ks.uiuc.edu/Research/vmd/). NS5 domains are represented separately, as two distinct targets, but NS5 methyltransferase is attached to the NS5 polymerase domain to form the full-length NS5. (c) ZIKV infectious life cycle: the virus is attached (1) and subsequently internalized (2) by receptor-mediated endocytosis. The virus is then trafficked to early endosomes, where the acidic environment induces fusion (3) between the virus and host membrane resulting in particle disassembly and genome release (4) . RNA is replicated and translated into a single polyprotein, which is processed by host and virus-encoded proteins (5) . Following translation, a replication complex is assembled and associated to virusinduced membranes where viral replication takes place (6) . The progeny RNA (+) strands can either initiate a new translation cycle or be assembled into virions within the endoplasmic reticulum (ER) (7) . The resultant immature virions are transported to the trans-Golgi where the immature virions are transformed into mature infectious particles (8) that are released by exocytosis (9) . In the pink boxes are the names of the compounds that can inhibit the marked steps of the virus lifecycle. Flaviviridae family [10] , ZIKV carries a positive-sense single-stranded RNA genome, encoding a large polyprotein, which after processing by host and viral proteases yields three structural and seven nonstructural (NS) proteins (Fig. 1b) . Envelope protein (E), membrane protein (M), which is expressed as prM, the precursor to M, and capsid (C) are the structural proteins, which form the virion. The NS proteins (NS1, NS2A, NS2B, NS3, NS4A, NS4B, NS5) are essential for genomic replication and modulation of host immunity [10] (Fig. 1b) . Figure 1c presents an overview of the replication process of ZIKV in the infected cell, showing the marked steps of the replication, which can be inhibited by the compounds discussed below in this review. We also present a special section (see supplementary material online: viral entry and replication mechanisms). Initially, before the availability of ZIKV structures in early March 2016, there were several efforts to develop homology models of the ZIKV proteins based on close homologs such as dengue virus (DENV; see Glossary) and other flaviviruses [11] [12] [13] . Since late March 2016, ZIKV protein structures have been determined mostly by X-ray crystallography and have been made available in databases such as the Protein Data Bank (PDB) [14] : NS1, NS2B-NS3 protease, NS3 helicase, NS5 methyltransferase, NS5 polymerase, NS5 full-protein and envelope glycoprotein (Table 1). Among them, NS2B-NS3 protease, NS3 helicase and NS5 methyltransferase structures are available with ligands, ATP or RNA; this is very useful for ligand-binding-site identification in virtual screens versus apo conformations. These protein structures are fundamental when performing experimental and computational studies. Figure 2 shows some ZIKV protein 3D structures and their ligand and or substrate binding pockets. The substrate binding pocket of ZIKV NS2B/NS3 protease (NS2B residues: S81, D83, K54 and NS3pro residues: D75, H51, S135, G133, G151, D129, Y161) ( Fig. 2a) was based on the co-crystallized compound 1H-1,3-benzodiazol-1ylmethanol [15] and a boronate inhibitor [16] , which bind to a similar pocket. NS3 helicase presents two binding sites, one encompasses the RNA strand and the other is where the ATP molecule binds (Fig. 2b) [17] . The full-length NS5 protein contains two domains, representing distinct targets: the NS5 methyltransferase and NS5 polymerase [18] . The NS5 polymerase domain has three adjacent binding sites: the active site, the RNA site and the NTP channel [18] (Fig. 2c) . The NS5 methyltransferase domain has two substrates: SAM, a co-substrate involved in methyl group transfer, and GTP, a substrate for RNA synthesis [18] . Consequently, the accessibility of these two substrates, together with a conserved catalytic tetrad of Lys61, Asp146, Lys182, Glu218, forms the active site [18] (Fig. 2d) . ZIKV envelope protein is composed of three distinct domains: the b-barrel-shaped domain (DI), the finger-like domain (DII) and the immunoglobulin-like domain (DIII) (Fig. 2e ). Domain II is responsible for the dimerization [19] . A binding site between DI and DIII, enclosing a hydrophobic cavity around a flexible linker, was computationally predicted [20] . The protein-protein binding pocket is composed of residues from DI and DIII (Fig. 2e) . The crystallographic structure of the capsid protein was recently solved [21, 22] . It contains four a helices in each monomer, and it forms a dimer (Fig. 2f) . We predicted the capsid pocket druggability using the PockDrug server [23] , which identified three pockets: pocket 1 (between a2 and a3 helices), pocket 2 (between N-terminal and a1 helix of the monomers) and pocket 3 (between a4 helices of the monomers). The viral proteins play an important part in virus infection and replication processes. The E glycoprotein is associated with virus adsorption, internalization and fusion with the host cell, as well as with the development of neutralizing humoral immunity. The prM protein has several roles in the flavivirus lifecycle, such as assisting in the chaperone-mediated folding of the E protein and preventing premature fusion during virion egress [24] . The main role of the C protein is the assembly and packaging of the viral RNA genome to form the viral nucleocapsid, in addition to acting as elements of viral particle assembly when associated with other proteins on lipid droplets and the endoplasmic reticulum (ER) [25] . The function of NS flavivirus proteins is still not completely understood but can be summarized as follows. NS1 has several roles, including flaviviral replication and virion maturation [26] . The NS2B subdomain is required for the proper formation of the substrate recognition site of the NS3 protease [27] . The NS3 helicase (NS3h) protein promotes the separation of RNA strands during viral replication and unwinds the RNA secondary structure in the 3 0 nontranslated region [28] . Together with NS2B, NS3 protease is responsible for cleavage and posttranslational modification of the viral polyprotein [16, 29] . Flavivirus NS4A and NS4B proteins compose the ER membrane-associated replication complex [30] [31] [32] . DENV NS4B has been shown to interact with NS3h and dissociate it from single-stranded RNA to modulate RNA synthesis [33] . ZIKV NS4A and NS4B cooperatively suppress the host's Akt/mammalian target of rapamycin (mTOR) pathway, inhibiting neurogenesis and inducing autophagy [34] . For these reasons, NS4B is an important target and its inhibition could impair viral propagation. Unfortunately, no crystal structures of ZIKV NS4B are available (as of 25 th June 2018). NS5 contains a methyltransferase domain, which methylates the RNA cap structure, and an RNA-dependent RNA polymerase domain, which synthesizes the viral RNA and is thus essential to ZIKV survival and establishment of the infection in host cells [18] . The innate immune response is the initial line of host defense, where type I interferon (IFN) production and signaling have a central role. Regarding this, NS5 proteins have been shown to inhibit the IFN signaling to evade host antiviral defense [35] . The NS5 protein uses different mechanisms (depending on the specific flavivirus) to target this signaling pathway. For Tick-borne encephalitis virus (TBEV) and West Nile virus (WNV), NS5 proteins inhibit signal transducer and activator of transcription (STAT1) phosphorylation and nuclear translocation [36] , this action was related to Ribbon-representation of some of the ZIKV protein 3D structures. (a) NS2B-NS3 protease, highlighting the substrate/ligand-binding pocket occupied by a boronate inhibitor [16] and the compound 1H-1,3-benzodiazol-1-ylmethanol [15] . (b) NS3 helicase, highlighting the RNA and ATP binding sites [17] . [20] . This predicted binding pocket encloses a hydrophobic cavity around the flexible linker joining DI and DIII. (f) Capsid protein [21, 22] and the predicted ligand-binding pockets. All these 3D structures are available in PDB and the figures were built using the VMD program [172] (http://www.ks.uiuc.edu/Research/vmd/). the polymerase domain [37] . For Japanese encephalitis virus (JEV), NS5 blocks tyrosine phosphorylation of tyrosine protein kinase Tyk2 [38] . For yellow fever virus (YFV), NS5 can bind to STAT2, but only after IFN treatment, and this prevents STAT2 binding to IFNstimulated responsive promoter elements present on IFN target genes [39] . A recent study showed that ZIKV NS5 expression resulted in proteasomal degradation of the IFN-regulated transcriptional activator STAT2 from humans, but not mice, resembling the DENV NS5 mode of action [40, 41] . Particularly for ZIKV, a docking study was carried out to predict the protein-protein interactions between NS5, seven in absentia homolog (SIAH2) protein and STAT2 proteins. The study suggests that: (i) NS5 recruits SIAH2 for the ubiquitination-dependent degradation of STAT2; and (ii) the NS5 amino acid residues involved in interaction with SIAH2 and/or STAT2 were found to be conserved across related flaviviruses [42] . Another structural study of ZIKV NS5 indicated that the smallmolecule inhibitor-binding site of DENV3 NS5 is structurally conserved in ZIKV NS5, indicating a potential mechanism for functional inhibition of ZIKV NS5 [43] . In principle, small molecules interfering with the function of any ZIKV viral proteins have potential in restricting virus replication and preventing the progress of ZIKV-infection-related pathogenesis and diseases. Here, we summarize several inhibitors targeting viral proteins that have been identified from high-throughput cell-based screening, in silico docking or compound library screening. Some of them have even undergone in vivo testing. A polyphenol present in green tea, epigallocatechingallate (EGCG), has been shown to have antiviral activity for many viruses, such as HIV, herpes simplex virus (HSV), influenza virus and hepatitis C virus (HCV) [44] . Recently, Carneiro and co-workers, through in vitro assays (in Vero E6 cells), showed that EGCG also inhibited the entry of ZIKV into the host cell by at least 1 log (>90%) at higher concentrations (>100 mM). The authors proposed that the inhibition is probably related to the direct interaction of EGCG with the lipid envelope, leading to a subsequent destruction of the virus particle [44] . Conversely, EGCG contains the catechol group, which is a well-known PAIN substructure (panassay interference compound) that promiscuously or nonspecifically inhibits many different targets [45] [46] [47] . A subsequent computational study by Sharma and co-workers, using induced-fit docking, molecular dynamics (MD) and druglike property calculations, revealed that EGCG binds to a hydrophobic site located between DI and DIII, close to a flexible linker (Fig. 1e ). Using MD simulations, they showed that EGCG blocks the major conformational change during the membrane fusion process [20] . Another natural product, nanchangmycin, from Streptomyces nanchangensis, proved to be a potent inhibitor of ZIKV in HTS assays using human cells [48] , with an IC 50 0.1 mM (Fig. 3a) . To determine whether nanchangmycin was blocking a pre-or post-entry step in the viral lifecycle, the authors treated cells with the drug and removed the drug 4 h post-infection (hpi) and monitored the level of infection (24 h later). Nanchangmycin inhibited infection, which suggests that it blocks an early step in the viral lifecycle [48] . Unfortunately, nanchangmycin contains a reactive Michael acceptor, which is also a well-known PAINS substructure [49] . Fernando and co-workers, through homology modeling, docking and MD simulation approaches, predicted two compounds: ZINC33683341 and ZINC49605556, could inhibit ZIKV E protein. These two small molecules preferentially bind to the glycan-binding domain, related to the virulence of the virus. The antiviral activity of ZINC33683341 was confirmed by cell culture assay (in Vero cells) [50] . Although the capsid protein has emerged as a promising target for antiflaviviral agents, few capsid inhibitors have been identified to date. The compound ST-148 has been shown to interact with the capsid protein and was identified as a potent inhibitor of all four serotypes of DENV in vitro and in vivo, reducing viremia and viral load in vital organs [24] . Characterization of the mode-of-action of ST-148 showed that it directly targeted the capsid protein and presented bimodal antiviral activity affecting assembly and/or release and entry of infectious DENV particles [51] . Another example of a flaviviral capsid inhibitor is the biotinylated derivative of compound SL209, which inhibited HCV infection, blocking the core dimerization with an IC 50 of 2.80 mM [52] . Both studies support the hypothesis that inhibitors of viral capsid formation might constitute a new class of antiviral agents, and therefore could also be applicable to ZIKV. Natural products like polyphenols are known to have antiviral activity against influenza virus, coronavirus (CoV), DENV and others [53] [54] [55] [56] . In this way, many natural products have been tested against NS2B-NS3 protease, and some of them were able to inhibit ZIKV protease activity. In one study, 22 compounds were tested at (the relatively high concentration of) 100 mM, and seven compounds were able to inhibit >50% of the protease activity [57] . The IC 50 s of these compounds were determined and ranged from 22 to 112 mM [57] . In another study, five flavonoids and one natural phenol were tested [58] . The flavonoids were myricetin (IC 50 = 1.3 mM), quercetin (IC 50 = 2.4 mM), luteolin (IC 50 = 2.7 mM), isorhamnetin (IC 50 = 15.5 mM), apigenin (IC 50 = 56.3 mM) and the natural phenol curcumin (IC 50 = 3.5 mM). These polyphenols inhibit the enzyme in a noncompetitive mode, which means that they could perhaps be allosteric inhibitors. Docking studies showed that the compounds are predicted to bind to a pocket on the back of the active site of ZIKV NS2B-NS3 [58] . However, curcumin, quercetin and other flavonoids have been shown to be promiscuous inhibitors, for example via colloidal aggregation [59] [60] [61] ; curcumin also contains reactive Michael acceptors and quercetin has a catechol, a well-known PAINS substructure, which might make these compounds less favorable. An HTS assay was developed to test compounds that inhibit the interaction between NS3 and the NS2B N-terminal fragment [62] . Then, a library of FDA-approved drugs and investigational drugs (2816 drugs) was screened using this assay, and 23 compounds produced an IC 50 below 15 mM, of which 12 were considered PAINS. The remaining 11 compounds were tested for their protease inhibition activities, and three could inhibit with IC 50 values ranging from 1.1 to 15.9 mM. The three compounds were tested against ZIKV, DENV, WNV, JEV and YFV. They presented inhibitory activity close to the nanomolar level, whereas temoporfin (Fig. 3b ) displayed an EC 50 (half-maximal effective concentration) for all flaviviruses tested at low nanomolar concentrations. The therapeutic indices for all three compounds were high, because their CC 50 (concentration that inhibits 50% of mammalian cell proliferation) was considerably higher than their EC 50 . The com-pounds were also able to inhibit ZIKV replication in placental epithelial and neuronal cells and to inhibit viral polyprotein cleavage. Temoporfin was tested in a viremia mouse model and a lethal mouse model and was able to inhibit viremia and protect Chemical structures of selected Zika virus (ZIKV) protein inhibitors. (a) Envelope glycoprotein inhibitor: nanchangmycin (IC 50 = 0.1 mM) [48] . (b) NS2B-NS3 protease inhibitors: temoporfin (IC 50 = 1.1 mM) [62] and NSC157058 (IC 50 = 0.82 mM) [66] . (c) NS3 helicase inhibitor: suramin (EC 50 = 0.42 mM), which was tested only in a cell-based assay in ZIKV [75] . (d) NS5 polymerase inhibitors: sofosbuvir (IC 50 = 7.3 mM) [105] , 2-C-ethynyl-UTP (IC 50 = 0.46 mM) [99] and DMB213 (IC 50 = 5.2 mM) [105] . (e) NS5 methyltransferase inhibitor: sinefungin (IC 50 = 1.18 mM) [90, 91] . 83% of the mice; the mice that survived did not present any signs of neurological disorder [62] . A similar study was done using structure-based virtual screening (VS) of 8277 compounds from the DrugBank database, and the top 100 candidates were identified [63] . From these, eight clinically approved compounds belonging to different drug classes were selected for further validation studies. From the eight selected compounds, five were validated as NS2B-NS3 protease inhibitors. Then, the compounds novobiocin, lopinavir, ritonavir and rifampicin which had favorable safety profiles and could be administrated to pregnant women were selected for phenotypic screening. Two compounds were able to inhibit ZIKV replication. Because the compound novobiocin had a higher selectivity index, it was tested in vivo. Dexamethasone immunosuppressed mice with disseminated ZIKV infection and novobiocin treatment had a significantly higher survival rate, lower mean blood and tissue viral loads, as well as less severe histopathological changes than untreated controls [63] . A similar strategy is to appropriate other flavivirus active compounds to inhibit ZIKV, because they are closely related in terms of the sequence and structure of their proteins [64] . Based upon a literature review, 11 drugs were selected to be tested in ZIKV if they could match the following criteria: (i) known toxicity profile in humans and pregnancy; (ii) known broad-spectrum antiviral activity; (iii) known antiviral activity against flavivirus or another RNA virus; (iv) coverage of a broad range of indications and drug classes. Bromocriptine presented inhibitory activity in a cytopathic-effect inhibition assay, virus-yield-reduction assay and plaqueyield-reduction assay. The drug was predicted to bind to NS2B-NS3 protease through molecular docking studies and also presented inhibition of the proteolytic activity of the enzyme [64] . Twentyseven HCV NS3-NS4A protease inhibitors were tested in ZIKV NS2B-NS3; ten presented IC 50 values below 50 mM. The best two inhibitors were tested to determine their inhibition type, and they presented competitive inhibition profiles, meaning that they might bind at the active site [65] . Aprotinin, a 58 amino acid bovine trypsin inhibitor, previously tested against WNV protease, also inhibits ZIKV protease (IC 50 = 70 nM), and molecular modeling studies predicted that the inhibitor probably blocks the interactions of NS3 and NS2B [66] . By contrast, in our analysis of crystal structures of WNV protease with bovine pancreatic trypsin inhibitor (PDB ID 2IJO) [67] and of DENV protease bound to aprotinin (PDB ID 3U1J) [68] , these types of inhibitors bind to and occlude the substrate site. A focused library of protease inhibitors that bind to WNV exosites was tested, and $700 structurally similar compounds were screened [66] . The best inhibitors were NSC157058 (IC 50 = 0.82 mM) (Fig. 3b) , NSC86314 (IC 50 = 0.97 mM) and NSC716903 (IC 50 = 1.12 mM). They were also tested against furin, a human serine protease with a cleavage sequence preference similar to those of flaviviral proteases, and they showed no effect at 100 mM. These compounds decreased the viral yield in neuronal precursor cells. The best inhibitor, NSC157058, was tested using an in vivo mouse model and decreased the viremia by tenfold; unfortunately, it has a very unfavorable pharmacokinetics profile [66] . Brecher et al. developed an assay to analyze the conformational changes in DENV NS2B-NS3 protease using luciferase [69] . They performed a VS assay using an allosteric site conformation that is present only in the enzyme's inactive state. Twenty-nine compounds were selected from a VS pipeline to be tested in a protease inhibition trial, of which three were subsequently selected to be tested in the allosteric assay. Only NSC135618 was able to inhibit the protease. The same compound also inhibited viral replication of DENV, ZIKV, WNV and YFV [69] . Crystal structure elucidations and MD studies have been undertaken to further understand the interactions and conformational changes of the protein bound to its inhibitors. The inhibitor complex seen in the crystal structures could provide a model for assemblies formed at the site of polyprotein processing, guiding future drug discovery processes against ZIKV protease [15, 16, 70] . Helicase displays ATPase and RNA triphosphatase (RTPase) activities. There are few studies regarding NS3h inhibitors, mainly using VS, but they are not experimentally validated. Although we have recently had some promising preliminary results in our own virtual screens against ZIKV NS3h, developing drugs that target ZIKV NS3h could involve significant hurdles. For instance, the accessible pocket(s) within the active site of the NS3h-RNA complex is relatively shallow, and previous efforts to discover specific inhibitors of DENV NS3h have been challenging [71] . A potential strategy to deal with these issues could involve targeting other sites of ZIKV NS3h, such as the putative allosteric site that we discovered in our MD simulations [72] . We are currently performing new VS against snapshots of ZIKV NS3h that we generated with MD to test this strategy. Other challenges are related to: (i) the development of potent inhibitors, as a result of the lack of the specific amino acid interaction with ATP nucleobase or ribose ring; and (ii) the discovery of specific inhibitors, because the host helicase is within the same SF2 superfamily as NS3h [71] . In an ATPase activity analysis of NS3 helicase, the ATPase inhibitors resveratrol and quercetin were tested at 150 mM; their inhibitory effects were 51% and 15% ATPase activity inhibition, respectively [73] . Suramin (Fig. 3c) , an antiparasitic drug, was previously characterized as a DENV NS3 helicase inhibitor with a noncompetitive mode of action [74] , and recently it was shown to inhibit ZIKV replication with EC 50 value of 0.42 mM [75] . Albulescu and co-workers suggested through experimental assays that suramin affected ZIKV binding and/or entry, virion biogenesis and attachment to host cells [75] . The authors also suggest that suramin might inhibit the NS3h activity or might affect packaging by binding to positively charged residues on the capsid protein [75] . MD simulations of NS3h where performed to study the proteins flexibility and conformations preferences [72] . This work showed that the RNA binding loop is influenced by the presence of the RNA strand, being more stable and in a closed conformation, when RNA is bound to the protein [72] . The different conformations observed during MD showed a distinct region beneath or behind the ATP site, which could perhaps be an allosteric site, and could help guide the discovery or design of new inhibitors. These conformations and strategies are guiding new VS experiments in the OpenZika project, which is being performed on IBM's World Community Grid (WCG). OpenZika has been virtually screening millions of compounds against all of the ZIKV protein structures (and targets from related flaviviruses) using molecular docking (on WCG) and QSAR modeling filtering (performed in-house) [76] . This project is ongoing and several compounds are currently being assayed that were identified in these virtual screens. Other flaviviral NS3 helicase inhibitors such as ivermectin [77] and benzoxazole have activity against YFV NS3h and DENV NS3h, respectively [78] . Ivermectin was also reported to inhibit DENV NS5 interaction with its nuclear transporter importin (IMP) a/b in vitro [79, 80] . All these NS3h agents are also interesting compounds to be tested against ZIKV NS3h. The NS4B protein is known to have transmembrane domains and is placed in the ER [31] . The biological function of this protein is not fully established, but it is part of the replication complex (RC), and some studies have demonstrated that NS4B induces the formation of a membrane web that forms a scaffold for the formation of the RC [81] . NS4B also interacts with other viral proteins, such as NS1, NS2B, NS3 and NS4A [33, [82] [83] [84] . For those reasons, NS4B is a good target to develop inhibitors. Unfortunately, the search for NS4B inhibitors faces many challenges, such as poor pharmaceutical properties of the inhibitors, difficulties to find pan-flaviviral inhibitors and also a high risk of developing resistant viruses [85] . There are few studies regarding ZIKV NS4B protein inhibition thus far. An aminothiazole derivative (NITD-618) was identified as a specific DENV inhibitor, capable of inhibiting all four serotypes (but not closely related flaviviruses); it was identified in an HTS of 1.8 million compounds using a DENV replicon containing the luciferase gene [86] . NITD-618 faced problems with pharmacokinetics in in vivo studies because of its high lipophilicity, and medicinal chemistry attempts to overcome this issue resulted in loss of antiviral selectivity against DENV [86] . In another screen using the NIH clinical collection, a naltrindole analog (SDM25N) was found to inhibit DENV, and specific NS4B point mutations (F164L and P104L) conferred resistance against the compound, indicating that NS4B is probably the targeted protein [87] . A spiropyrazolopyridone compound was found to inhibit DENV serotypes 2 and 3 but not 1 and 4. This fact is correlated with the low identity between NS4B of flaviviruses and even between different serotypes of DENV [88] . The compound was able to suppress viremia in a mouse model of DENV serotype 2 infection. Two compounds (CCG-3394 and CCG-4088) were hits from a screen against YFV NS4B, and mutations on the K128 residue conferred resistance to YFV against both of these compounds [89] . Different types of DENV methyltransferase (MTase) inhibitors were also tested for ZIKV MTase. The classes of inhibitors tested were S-adenosyl-L-methionine (SAM) analogs, RNA Cap analogs and compounds targeting an allosteric site of DENV MTase. Sadenosyl-L-homocysteine (SAH) and the SAM analog sinefungin [90] . The structure of ZIKV MTase with the SAM analog sinefungin was recently elucidated, and the construction of an inhibitor connecting sinefungin with a Cap analog attached by a linker was proposed, which could perhaps grant higher affinity to the protein [91] . VS using a hydrophobic site close to the SAM pocket was performed with >20 000 compounds [92] . The ten compounds with the best scores were selected for experimental screening: four were able to inhibit viral growth in concentrations below 20 mM and the best inhibitor presented an IC 50 value of 4.8 mM [92] . Jain and co-workers developed a SAM analog that encroaches into the guanosine triphosphate (GTP)/Cap pocket of ZIKV MTase [93] . They started from the crystallographic structure bound to SAM and developed the analog, attaching a 4-fluorophenyl group that gave the compound the capacity to bind to a part of the RNA tunnel and to occupy the Cap methylation site [93] . Designing ZIKV-specific MTase inhibitors is a challenge, as is the case for DENV and other flaviruses, mainly owing to the selectivity of these compounds to SAM-binding or SAM-utilizing proteins present in the host cell, because of the similarity to human RNA and DNA MTases [29] . Moreover, another issue with the design of competitive MTase inhibitors is the high cellular SAM concentration [94] . Still, these issues have been overcome in the field of kinases [95] , indicating that it could also be feasible for MTases [94] . NS5 RdRp remains one of the most promising targets, as demonstrated by the large number of marketed nucleoside inhibitor (NI) and non-nucleoside inhibitor (NNI) drugs available for the treatment of HIV-1, cytomegalovirus and hepatitis B and C viruses [96, 97] . Nucleoside analogs, such as the prodrug sofosbuvir, have been used as part of certain combination therapies that can treat HCV. Some NIs, like 2 0 -C-and 2 0 -O-methyl-substituted nucleosides, 2 0 -C-fluoro-2 0 -C-methyl-substituted nucleosides, 3 0 -O-methyl-substituted nucleosides, 3 0 -deoxynucleosides, derivatives with a 4 0 -C-azido substitution, heterobase-modified nucleosides and neplanocins, were tested against ZIKV, and the most promising inhibitors were the 2-C-methylated nucleosides, with IC 50 values <10 mM [98] . A similar observation was found when ribonucleoside triphosphate analogs were tested against NS5 polymerase: the 2 0 -C-methyl-and 2 0 -C-ethinyl-substituted ribonucleoside triphosphates were the best inhibitors. The compounds were tested for their abilities to be incorporated into the RNA chain and to terminate its elongation. The analogs showing both abilities were 2 0 -F-2-C-ME-UTP (IC 50 = 90.76 mM), 2 0 -C-ME-UTP (IC 50 = 5.78 mM) and 2 0 -C-ethynyl_UTP (IC 50 = 0.46 mM) (Fig. 3e) [99] . ATP analogs were tested in NS5 polymerase in another study, and the compounds that led to the strongest enzyme inhibition were the 2 0 -Cmethylated ATPs [100] . The best inhibitors presented IC 50 values of 5.6 mM and 7.9 mM [100] . An adenosine analog called BCX4430 inhibited ZIKV in cell culture, as well as in a mouse model [101] . This compound already presented broad-spectrum activity against a wide range of RNA viruses: West Nile, tick-borne yellow fever, Marburg and Ebola [102, 103] . Previous studies suggested that BCX4430 acts on NS5 polymerase, promoting chain termination of viral RNA synthesis [102] . In a recent study, another adenosine analog and a known flaviviral NS5 polymerase inhibitor was tested on cell-based assays in ZIKV. The compound 7-deaza-2 0 -C-methyladenosine (7DMA) inhibited ZIKV proliferation in Vero cells and in an animal model, it decreased viremia and delayed morbidity and mortality caused by ZIKV [104] . Although the authors have not tested this compound against the ZIKV protein, it represents a very promising candidate [104] . Drug Discovery Today Volume 23, Number 11 November 2018 Nucleoside inhibitors against the polymerase have made great progress; however, the major challenges for the NI-based antiviral approach include toxicity and the added complication of being a prodrug (which requires host kinases to convert it to the triphosphate form of NI) [94] . Using a drug repurposing strategy, Xu et al. tested compounds already used for HCV NS5 polymerase: sofosbuvir (Fig. 3e) , a NI, and DMB213 (Fig. 3e) , a NNI [105] . Both compounds inhibited ZIKV NS5 polymerase with IC 50 values of 7.3 mM and 5.2 mM, respectively. Mutations that confer resistance to nucleoside analog inhibitors in HCV also led to resistance to sofosbuvir in ZIKV, which was not the case for DMB213 [105] . Mumtaz and co-workers tested sofosbuvir against ZIKV using Vero, A549 and Huh7 cells, measuring the level of the active sofosbuvir metabolite by mass spectrometry [106] . Sofosbuvir showed a cell-type-dependent activity, inhibiting ZIKV with an IC 50 of $4 mM (but only in Huh7 cells) [106] . This correlated with differences in the intracellular concentration of the active metabolite of sofosbuvir, which was higher in Huh7 cells compared with the other cells. These results highlight the importance of a careful selection of cell system for repurposing trials of prodrugs to evaluate antiviral activity [106] . A pharmacophore-based strategy was used to search for NS5 inhibitors [12] . The researchers constructed models for NS5 polymerase and NS5 methyltransferase using the molecules ribavirin and BG323. They used the models to conduct VS of the ZINC database and found 23 candidates for NS5 polymerase and 18 candidates for methyltransferase. These candidates were docked to their respective targets, and three potential leads were selected for each protein, based on the docking scores. The compounds with the best docking scores were ZINC39563464 for NS5 polymerase and ZINC64717952 for NS5 MTase [12] . These results remain to be confirmed through validation assays. Lim et al. performed a fragment-based screening by X-ray crystallography targeting the DENV NS5 polymerase and elucidated an allosteric binding pocket at the base of the thumb subdomain, close to the enzyme active site [107] . Promising allosteric inhibitors were developed through structure-guided design, and they were active in DENV polymerase biochemical and cell-based assays, inhibiting the enzyme with IC 50 values of 1-2 mM [107] . The same strategy could probably be used for ZIKV NS5 allosteric inhibitor design, because the viral proteins have conserved sequences and similar structures. ZIKV, as well as other flaviviridae members, has a small genome and requires the host cell machinery to carry out several core functions that are essential to viral replication. In addition to the inhibition of the function of viral proteins, an attractive broad-spectrum strategy is to target host cell processes, because they are often employed by multiple viruses and are less prone to the development of drug resistance [108] . The recently described inhibitors of the most commonly targeted cellular functions in ZIKV can be found in Table S1 (see supplementary material online) and are reviewed below. Moreover, we also present a special section regarding the antiviral effects of the type I and II interferons in the supplementary material. Viruses rely on the supply of nucleosides from the host cell to maintain proper RNA replication. Furthermore, there is evidence that inhibition of nucleoside biosynthesis triggers the activation of antiviral interferon-stimulated genes in human cells [109] . Thus, host enzymes involved in the de novo biosynthesis of nucleosides, such as inosine monophosphate dehydrogenase (IMPDH) and dihydroorotate dehydrogenase (DHODH), are interesting targets for broad-spectrum antiviral therapy. IMPDH catalyzes the oxidative conversion of inosine 5 0 -monophosphate into xanthosine 5 0monophosphate, which is the first committed and rate-limiting step of the guanine nucleotide biosynthetic pathway [110] . Known inhibitors of IMPDH include ribavirin, 5-ethynyl-1-beta-D-ribofuranosylimidazole-4-carboxamide (EICAR) and mycophenolic acid (MPA). Ribavirin, one of the first clinically used broad-spectrum antivirals, is commonly employed in combination therapies to treat HCV and is thought to have multiple mechanisms of action, including the inhibition of the viral polymerase and host IMPDH [111] . It was found to inhibit the virus-induced cytopathic effects (CPE) of several flaviviruses, including ZIKV (EC 50 = 142.9 mg/ml) [112] . A recent study tested the activity of known broad-spectrum antivirals, including ribavirin, which yielded a poor inhibition of virus-induced CPE (EC 50 > 50 mM for MR766 strain), concluding it was not a suitable candidate as a ZIKV therapeutic [113] . Other nucleoside biosynthesis inhibitors were also tested in the same study, among which were MPA (EC 50 = 0.11 mM), brequinar (EC 50 = 0.08 mM) and 6-azauridine (EC 50 = 0.98 mM). The latter two compounds inhibit dihidroorotate dehydrogenase (DHODH) and orotidylic acid (OMP) decarboxylase, respectively, which are two enzymes involved in pyrimidine biosynthesis [113] . It is important to highlight that several nucleotide biosynthesis inhibitors suppress viral growth through innate immunity [109] . In a study by Lucas-Hourani and co-workers, the authors showed that pyrimidine deprivation, using brequinar and another compound DD264, is not directly responsible for the antiviral activity of pyrimidine biosynthesis inhibitors but it rather involves the induction of a metabolic stress and the subsequent triggering of the cellular immune response [109] . Barrows and co-workers also tested MPA, which presented potent anti-ZIKV activity, inhibiting ZIKV infection in HuH-7, HeLa, JEG3, hNSC and HAEC cells [114] . Their screen also detected other nucleoside biosynthesis inhibitors, such as azathiopurine, mercaptopurine hydrate, mycophenolate mofetil (a prodrug of MPA) and thioguanine. Another cell-based screen carried out by Pascoalino et al. [115] identified 6-azauridine (EC 50 = 2.3 AE 0.1 mM) and another pyrimidine biosynthesis inhibitor: 5-fluorouracil (EC 50 = 14.3 AE 8.6 mM), which inhibits thymidylate synthase, the enzyme that catalyzes the final step of thymidine biosynthesis. 5-Fluorouracil and floxuridine also showed dose-dependent inhibition of ZIKV replication in a study by Tiwari et al. [116] . Although the 'pregnancy categories' labeling system is being replaced [117, 118] , it is worth noting that these compounds were generally classified in pregnancy category D by the FDA [114] , and thus have presented 'positive evidence of human fetal risk'. This is not unexpected, because they deplete the cellular pool of nucleotides, which certainly affects proper development of the fetus. Flaviviral infection has been associated with alterations in the lipid homeostasis [119] and membrane structure of infected cells [120] . DENV infection, for example, is known to induce dramatic relocalization of the fatty acid synthase (FAS) to the sites of viral replication [121, 122] . Cholesterol has been identified as an important modulator of the host response to several flaviviruses, but the exact mechanism by which this modulation occurs is not yet fully understood. Nevertheless, inhibition of the cholesterol biosynthesis pathway represents an attractive therapeutic approach. Several enzymes involved in cholesterol biosynthesis such as mevalonate decarboxylase, 3-hydroxy-3-methylglutaryl-CoA (HMG-CoA) synthase and squalene synthetase were found to be important for efficient replication of DENV in A549 and K562 cell lines [123] . Lovastatin, an HMG-CoA reductase inhibitor, was tested against ZIKV as part of a cell-based screen of 725 FDA-approved drugs, and its activity was confirmed through a dose-response assay (EC 50 = 20.7 AE 8.6 mM) [115] . Its antiflaviviral activity was previously reported in HCV [124] and DENV [123] . Moreover, Sarkey and co-workers demonstrated that lovastatin attenuated nervous system injury in an animal model and could be used in inflammatory peripheral nerve diseases, including Guillain-Barré syndrome, which can be a consequence of ZIKV infection [125] . Nevertheless, although it is considered safe, treatment with lovastatin showed no evidence of beneficial impact on dengue infections in a randomized, double-blind, placebo-controlled trial [126] . Mevastatin also had anti-ZIKV activity at concentrations of 1-5 mM, but it was not dose-responsive at concentrations above 5 mM [113] . Mevastatin is known to induce apoptosis [127] , which could perhaps explain the lack of antiviral efficacy at higher concentrations. Nordihydroguaiaretic acid (NDGA) and its derivative tetra-Omethyl nordihydroguaiaretic acid (M 4 N) were tested against ZIKV and showed inhibition in the low micromolar range (IC 50 values of 9.1 and 5.7 mM, respectively) [128] . They are polyphenols, whose mechanism of action is not fully elucidated (and could involve promiscuous inhibition; see the aforementioned section on PAINS in the NS2B-NS3 protease section), but NDGA has been shown to affect HCV replication through the reduction of the amount of lipid droplets, thought to be mediated by the sterol regulatory element-binding protein (SREBP) pathway [129] . Early studies have established NDGA as a 5-lipoxygenase inhibitor [130] , but it has been shown to bind several other molecular partners, such as glucose transporter 1 (GLUT1) [131] , tyrosine kinases [132, 133] and even transthyretin [134] . The molecular mechanisms and clinical applications of these compounds have been thoroughly reviewed by Lü and co-workers [135] . Protein kinases catalyze the addition of phosphate groups on several molecular entities, such as proteins, lipids and carbohydrates, thus controlling many cellular processes. Because viral replication requires the hijacking of several cellular mechanisms, it is expected to be hindered by the modulation of kinase activity. Indeed, host cell kinases have been implicated in the replication of several RNA virus families [136] . Tang and co-workers have carried out a transcriptome analysis of human neural progenitor cells (hNPCs) and enrichment analysis of their supplementary data indicates that several of the upregulated genes are related to protein kinase activity [137] . Xu et al. performed a two-step drug repurposing screen by initially measuring the caspase-3 activity and subsequently measuring cell viability with the primary hits. Using this approach, they were able to detect PHA-690509, an investigational cyclindependent kinase inhibitor (CDKi), which inhibited ZIKV infection with an EC 50 value of 1.72 mM [138] . They then tested an additional 27 CDKis and identified nine that could inhibit ZIKV replication. Among them, seliciclib (a purine analog, also called roscovitine) and RGB-286147 inhibited ZIKV infection at submicromolar concentrations [138] . The authors concluded that these results suggest that one or more host CDKs might be important for ZIKV replication, because flaviviruses are not known to encode any CDK [138] . AXL (from the Greek word anexelekto, or uncontrolled) is a tyrosine kinase receptor (TKR) thought to mediate viral attachment to the host cells [139, 140] . Its function is also responsible for a downregulation of interferon production [141] . Although this recognition might be unrelated to its kinase activity, AXL inhibitors were found to inhibit ZIKV infection rates. Rausch et al. screened a library of $2000 compounds on human osteosarcoma cells (U2OS) and found 19 ZIKV inhibitors, eight of which had protein kinases as targets, including five TKRs and, among these, two AXL inhibitors [48] . The same AXL inhibitors did not exhibit the same activity on different cell lines (HBMEC and Jeg-3), indicating that the effect is cell-type specific. The authors also point out that Jeg-3 cells are highly permissive to ZIKV infection, despite showing no detectable AXL expression levels, which suggests that this receptor is not essential for infection [48] . This hypothesis is corroborated by an in vivo study [142] . Intracellular membrane traffic is a mechanism that is also exploited by viruses and depends heavily on the enzymatic activity of protein kinases, which regulate the vesicle traffic via the phosphorylation of a specific subunit of the associated adaptor proteins (APs) [143] . Bekerman and co-workers tested the antiviral activity of erlotinib and sunitinib for several different flaviviruses (and other genera), including ZIKV (EC 50 values of 6.28 and 0.51 mM, respectively) [143] . They present several lines of evidence that indicate that the likely targets are AP2-associated kinase (AAK1) and G-associated kinase (GAK), with other possible candidates being AXL, KIT proto-oncogene receptor tyrosine kinase (KIT) and proto-oncogene RET [143] . Monel et al. described in-depth the vacuolization of ZIKVinfected cells and suggested that they undergo a paraptosis-like death [144] , which is associated with the activity of phophoinositide 3-kinase (PI3K) [145] and membrane-associated protein kinase (MAPK) [146] . They therefore tested the activity of several kinase inhibitors and verified that they could prevent the onset of vacuoles, particularly the specific class-1 PI3K (ZSTK474) and AKT (triciribine) inhibitors. However, none of the tested inhibitors blocked ZIKV infection [144] . Viral replication extensively hijacks the ER functions. ER is intimately associated with the intracellular membrane network remodeling, and it is the framework where the viral polyprotein is expressed and processed. These events generate a considerable amount of stress on this organelle [147] . Correct expression and processing of nascent proteins are paramount for efficient viral replication. Several host Drug Discovery Today Volume 23, Number 11 November 2018 proteins are responsible for the monitoring of proper protein synthesis, folding and degradation. Impairment of these functions results in reduced viral assembly and budding. Examples are the ER membrane complex (EMC), a-glucosidase, cyclophilin and proteasome elements. Some inhibitors have been investigated under this context. EMC, an ensemble of nine components that are thought to assist protein folding, has been shown to be upregulated in ZIKV-infected cells. Several proteins of this complex are multipass membrane proteins [148] . Another key player related to protein metabolism is a-glucosidase, an ER-resident enzyme that catalyzes the removal of glucose units from N-linked oligosaccharides. This processing is crucial for the nascent glycoprotein maturation and subsequent correct folding. Glycoproteins that fail to be processed are subjected to abnormal accumulation in the ER and ER-associated degradation (ERAD), in which nascent misfolded proteins are retrotranslocated back to the cytosol, ubiquitinated and degraded by the proteasome [149] . a-Glucosidase inhibitors exhibit broad antiviral activity against multiple genera [150, 151] , but celgosvir (6-O-butanoyl castanospermine) showed poor activity against ZIKV in a CPEbased assay in Vero 76 cells infected with either MR766 or PRVABC59 strains (EC 50 > 50 mM in both cases) [113] . Xin and co-workers have carried out a quantitative proteomic analysis of C6/36 cells infected with ZIKV and found upregulated genes from two protein-related pathways: the ubiquitin-proteasome system (UPS) and the unfolded protein response (UPR) [152] . They subsequently tested the dose-dependent effect of bortezomib (EC 50 = 5.525 nM) and MG132 (EC 50 = 1.151 mM), two proteasome inhibitors, in ZIKV-infected Vero cells. They also found that bortezomib can reduce viral load and signs of pathology in ZIKVinfected mice [152] . Bortezomib was also detected in one of the previous screens, but it showed moderate toxicity in the hNSC cell line, as expected based on its mechanism and primary use as an anticancer agent [114] . Cyclophilins are cytoplasmic proteins responsible for the isomerization of proline peptide bonds from the trans to cis conformation, thus facilitating protein folding [153] . Their activities were shown to be essential for HIV and HCV replication. Cyclosporine A targets cyclophilins and was found to inhibit HCV [154] , WNV, DENV and YFV infections [155] . Barrows et al. detected cyclosporine as a potential hit but it displayed controversial results, even enhancing the infection at 1 mM compared with the control [114] . Most flaviviruses, including ZIKV, are endocytosed by a clathrinmediated mechanism and undergo pH-dependent fusion processes, in which the endosomal membrane fuses with the viral envelope through the action of the envelope protein. So that these steps can occur, the E protein must undergo conformational changes and expose its fusional loop [156] . Li and co-workers identified that 25-hydroxycholesterol (25HC) is probably capable of blocking viral entry by modifying host-cell membrane properties, with a calculated EC 50 of 188 nM. This compound was also able to (i) reduce viremia in mice and non-human primates and (ii) protect mouse embryos from microcephaly [157] . The antimalarial drug chloroquine is known to exhibit broadspectrum antiviral activity [158] . It is a weak base (pK 1 = 8.1, pK 2 = 10.1), and the pH of lysosomes in the presence of chloro-quine increases from $4 to $6 [159] . This increase is thought to be responsible for the inhibition of the pH-dependent fusion step of viral entry. Chloroquine was tested against ZIKV, showing activity in Vero cells (9.82 AE 2.79 mM), human brain microvascular endothelial cells (14.20 AE 0.18 mM) and human neural stem cells (12.36 AE 2.76 mM). It reduced the number of infected cells in vitro, inhibited virus production and cell death promoted by ZIKV infection and displayed relatively minor cytotoxic effects (CC 50 ranging from 94.95 to 134.54 mM). Moreover, chloroquine partially reversed morphological changes induced by ZIKV in mouse neurospheres [160] . Chloroquine was found to reduce viral burden in the placenta of ZIKV-infected C57BL/6 pregnant mice [161] . However, it also showed controversial results in another study [113] . Quinacrine (EC 50 = 2.27 AE 0.14 mM), mefloquine (EC 50 3.95 AE 0.21 mM) and GSK369796 (EC 50 = 2.57 AE 0.09 mM), other antimalarial drugs with a similar mechanism of action, were also recently tested against ZIKV [162] . The anticancer drug obatoclax promotes rapid neutralization of lysosomal pH, showing activity against YFV (EC 50 < 0.125 mM), WNV (EC 50 = 0.10 AE 0.04 mM) and ZIKV (EC 50 = 0.13 AE 0.01 mM). The authors suggest that the anticancer and antiviral activities are independent and rely on different mechanisms [163] . Obatoclax was also tested in retinal pigment epithelial (RPE) cells and displayed an EC 50 value of 0.04 AE 0.01 mM in the recovery of infected-cell viability [164] . In addition, it was also detected in a drug screen in Jeg-3 cells with an EC 50 value of 0.08 mM [48] . Saliphenylhalamide (SaliPhe) is another known viral entry blocker, which acts through the inhibition of the vacuolar ATPase [164, 165] . It was tested in Vero 76 cells, yielding an EC 50 value of 0.62 mM for MR766 strain and 0.49 mM for the PRVABC59 strain. It was also tested by Kuivanen et al. in RPE cells, yielding an EC 50 value of 0.05 AE 0.02 mM [164] . Niclosamide blocks the acidification of endosomes, albeit using a different mechanism that is not yet fully elucidated [166, 167] . It was detected in a screen against ZIKV in SNB-19 cells showing an EC 50 value of 0.37 mM based on the measurement of intracellular viral RNA [138] . Niclosamide is an FDA-approved drug, formerly designated in pregnancy category B and broadly used in the treatment of intestinal helminthiasis. Drugs in pregnancy category B have not undergone controlled studies in pregnant women but they have failed to demonstrate a risk to the fetus in animal reproduction studies. Inhibitors of endosomal sorting complexes required for transport (ESCRT) machinery could be promising targets to ZIKV drug discovery (see supplementary material online). Moreover, ZIKV inhibitors with an unclassified or unconfirmed mechanism of action are also discussed (see supplementary material online). One of the key steps in the drug discovery pipeline is the development of screening assays to assess the antiviral activity of compounds [168] . Viruses depend on cell machinery to replicate, and for this reason in vitro assays are developed using host cells for culture and viral replication. ZIKV has recently been shown to infect different cells in multiple species. These findings show that a diversity of cell lines can be used to study ZIKV infection, providing a good framework for the drug discovery process [169] . We also present and discuss the available screening assays for anti-ZIKV hit discovery (see supplementary material online). Although the scientific community has devoted considerable efforts to the search for a vaccine and antiviral drugs to prevent and treat ZIKV infection, there are still no approved treatments for this flavivirus, nor vaccines to prevent it. In a relatively short time we have gone from having no structures of proteins for this virus to a wealth of data. Owing to the current state of known anti-ZIKV compounds, there is probably some way to go until clinical trials of the discovered candidates are undertaken, especially when considering that some patients, including immunosuppressed and pregnant women, will have underlying medical conditions. In this regard, one strategy to overcome the risk of teratogenesis is the development of new drugs to treat Zika based on two-stage Phase II clinical trials: first-stage focusing on testing the efficacy in nonpregnant patients and the second-stage focusing on pregnant women. These will possibly reduce the overall costs of trials, facilitating their logistical set up and enhancing the safety level for the patients [170] . Many of the compounds are clearly unsuitable for these patients. Moreover, one of the frequent issues we see with many of the published screens is that they used unrealistic concentrations (e.g., 100 mM) or compounds (e.g., cytotoxic anticancer compounds) that could never be used in pregnant women. Much of the work described herein would have benefited from having experienced medicinal chemists involved to advise on the utility of compounds selected. This would also help to avoid compounds like PAINS [47] , which were reported as hits from several screens. Ultimately, the recent advances in the discovery of anti-ZIKV agents, ZIKV protein structures and host target protein structures, and our understanding of the disease itself, are not only crucial to advancing the fight against ZIKV but they can also be useful for the next emerging virus outbreak to which we will have to respond. We should also heed the lessons learned from ZIKV drug discovery so we can be more successful and avoid dead-ends in future. All authors except S.E. (CEO of Collaborations Pharmaceuticals) do not have any conflicts of interest to declare. These views and reflections are those of the authors. Molecular evolution of Zika virus during its emergence in the 20th century Defining the syndrome associated with congenital Zika virus infection Computed tomographic findings in microcephaly associated with Zika virus A single mutation in the prM protein of Zika virus contributes to fetal microcephaly Guillain-Barré Syndrome outbreak associated with Zika virus infection in French Polynesia: a case-control study Case report: acute myelitis due to Zika virus infection Uveitis associated with Zika virus infection Zika virus associated with meningoencephalitis The 3.8 Å resolution cryo-EM structure of Zika virus Flaviviridae: the viruses and their replication Illustrating and homology modeling the proteins of the Zika virus Zika virus NS5 protein potential inhibitors: an enhanced in silico approach in drug discovery Virtual screening for potential inhibitors of NS3 protein of Zika virus The Protein Data Bank Crystal structure of unlinked NS2B-NS3 protease from Zika virus Crystal structure of Zika virus NS2B-NS3 protease in complex with a boronate inhibitor Structure of the NS3 helicase from Zika virus Structure and function of the Zika virus full-length NS5 protein Structures of the Zika virus envelope protein and its complex with a flavivirus broadly protective antibody Epigallocatechin gallate, an active green tea compound inhibits the Zika virus entry into host cells via binding the envelope protein Crystal structure of the capsid protein from Zika virus Structural insight into the Zika virus capsid encapsulating the viral genome PockDrug-Server: a new web server for predicting pocket druggability on holo and apo proteins A novel inhibitor of dengue virus replication that targets the capsid protein Dengue virus capsid protein usurps lipid droplets for viral particle formation The many faces of the flavivirus NS1 protein offer a multitude of options for inhibitor design Crystal structure of the NS3 protease-helicase from dengue virus Crystal structure and activity of Kunjin virus NS3 helicase; protease and helicase domain assembly in the full length NS3 protein The medicinal chemistry of dengue fever Genetic interaction between NS4A and NS4B for replication of Japanese encephalitis virus Subcellular localization and membrane topology of the dengue virus type 2 non-structural protein 4B The non-structural protein 4A of dengue virus is an integral membrane protein inducing membrane alterations in a 2K-regulated manner Dengue virus NS4B interacts with NS3 and dissociates it from single-stranded RNA Zika virus NS4A and NS4B proteins deregulate Akt-mTOR signaling in human fetal neural stem cells to inhibit neurogenesis and induce autophagy Viral tricks to grid-lock the type I interferon system Inhibition of interferon-stimulated JAK-STAT signaling by a Tick-borne flavivirus and identification of NS5 as an interferon antagonist Identification of residues critical for the interferon antagonist function of Langat virus NS5 reveals a role for the RNA-dependent RNA polymerase domain Blocking of interferon-induced Jak-Stat signaling by Japanese encephalitis virus NS5 through a protein tyrosine phosphatase-mediated mechanism The interferon signaling antagonist function of yellow fever virus NS5 protein is activated by type I interferon Zika virus targets human STAT2 to inhibit type I interferon signaling Mouse STAT2 restricts early dengue virus replication Structural analysis and insight into Zika virus NS5 mediated interferon inhibition The structure of Zika virus NS5 reveals a conserved domain conformation The green tea molecule EGCG inhibits Zika virus entry Enhancement of chemical rules for predicting compound reactivity towards protein thiol groups New substructure filters for removal of pan assay interference compounds (PAINS) from screening libraries and for their exclusion in bioassays Pan assay interference compounds (PAINS) and other promiscuous compounds in antifungal research Screening bioactives reveals nanchangmycin as a broad spectrum antiviral active against Zika virus Nonleadlikeness and leadlikeness in biochemical screening An approach for Zika virus inhibition using homology structure of the envelope protein Characterization of the mode of action of a potent dengue virus capsid inhibitor Direct binding of a hepatitis C virus inhibitor to the viral capsid protein The preventive and therapeutic potential of natural polyphenols on influenza Naturally derived anti-hepatitis B virus agents and their mechanism of action Polyphenols: a diverse class of multi-target anti-HIV-1 agents Advances and challenges in the synthesis of highly oxidised natural phenols with antiviral, antioxidant and cytotoxic activities Inhibitory effect of flavonoids against NS2B-NS3 protease of ZIKA virus and their structure activity relationship Solution conformations of Zika NS2B-NS3pro and its inhibition by natural products from edible plants A common mechanism underlying promiscuous inhibitors from virtual and high-throughput screening Flavonoids: true or promiscuous inhibitors of enzyme? The case of deoxyxylulose phosphate reductoisomerase Colloidal aggregation and the in vitro activity of traditional Chinese medicines Existing drugs as broad-spectrum and potent inhibitors for Zika virus by targeting NS2B-NS3 interaction Structure-based discovery of clinically approved drugs as Zika virus NS2B-NS3 protease inhibitors that potently inhibit Zika virus infection in vitro and in vivo Novel antiviral activity and mechanism of bromocriptine as a Zika virus NS2B-NS3 protease inhibitor Identification of novel small molecule inhibitors against NS2B/ NS3 serine protease from Zika virus Characterization of the Zika virus two-component NS2B-NS3 protease and structure-assisted identification of allosteric smallmolecule antagonists Structural evidence for regulation and specificity of flaviviral proteases and evolution of the Flaviviridae fold Ligand-bound structures of the dengue virus protease reveal the active conformation A conformational switch high-throughput screening assay and allosteric inhibition of the flavivirus NS2B-NS3 protease Structural dynamics of Zika virus NS2B-NS3 protease binding to dipeptide inhibitors Ten years of dengue drug discovery: progress and prospects Molecular dynamics simulations of Zika virus NS3 helicase: insights into RNA binding site activity Structural features of Zika virus non-structural proteins 3 and -5 and its individual domains in solution as well as insights into NS3 inhibition Suramin inhibits helicase activity of NS3 protein of dengue virus in a fluorescence-based high throughput assay format Suramin inhibits Zika virus replication by interfering with virus attachment and release of infectious particles OpenZika: an IBM world community grid project to accelerate Zika virus drug discovery Ivermectin is a potent inhibitor of flavivirus replication specifically targeting NS3 helicase activity: new prospects for an old drug Novel benzoxazole inhibitor of dengue virus replication that targets the NS3 helicase Ivermectin is a specific inhibitor of importin a/b-mediated nuclear import able to inhibit replication of HIV-1 and dengue virus Nuclear localization of dengue virus (DENV) 1-4 nonstructural protein 5; protection against all 4 DENV serotypes by the inhibitor Ivermectin Expression of hepatitis C virus proteins induces distinct membrane alterations including a candidate viral replication complex expression of hepatitis C virus proteins induces distinct membrane alterations including a candidate viral replication complex Evidence for a genetic and physical interaction between nonstructural proteins NS1 and NS4B that modulates replication of West Nile virus Palmitoylation and polymerization of hepatitis C virus NS4B protein Characterization of dengue virus NS4A and NS4B protein interaction Targeting dengue virus NS4B protein for drug discovery Inhibition of dengue virus by targeting viral NS4B protein Identification of a new dengue virus inhibitor that targets the viral NS4B protein and restricts genomic RNA replication Discovery of dengue virus NS4B inhibitors Identification of inhibitors of yellow fever virus replication using a replicon-based high-throughput assay Zika virus methyltransferase: structure and functions for drug design perspective Structural basis of Zika virus methyltransferase inhibition by sinefungin Structural insight into NS5 of Zika virus leading to the discovery of MTase inhibitors Development of a S-adenosylmethionine analog that intrudes the RNA-cap binding site of Zika methyltransferase The dengue virus NS5 protein as a target for drug discovery Targeting histone lysine methylation in cancer Nucleoside and nucleotide HIV reverse transcriptase inhibitors: 25 years after zidovudine A 40-year journey in search of selective antiviral chemotherapy Nucleoside inhibitors of zika virus Analysis of ribonucleotide inhibitors of Zika virus RNAdependent RNA polymerase by using nonradioactive polymerase assays Adenosine triphosphate analogs can efficiently inhibit the Zika virus RNA-dependent RNA polymerase Efficacy of the broad-spectrum antiviral compound BCX4430 against Zika virus in cell culture and in a mouse model Antiviral activity of the adenosine analogue BCX4430 against West Nile virus and tick-borne flaviviruses Protection against filovirus diseases by a novel broadspectrum nucleoside analogue BCX4430 The viral polymerase inhibitor 7-deaza-2-Cmethyladenosine is a potent inhibitor of in vitro Zika virus replication and delays disease progression in a robust mouse infection model Purification of Zika virus RNA-dependent RNA polymerase and its use to identify small-molecule Zika inhibitors Cell-line dependent antiviral activity of sofosbuvir against Zika virus Potent allosteric dengue virus NS5 polymerase inhibitors: mechanism of action and resistance profiling Combating emerging viral threats Inhibition of pyrimidine biosynthesis pathway suppresses viral growth through innate immunity IMP dehydrogenase: structure, mechanism and inhibition Broad-spectrum agents for flaviviral infections: dengue, Zika and beyond Interferon, ribavirin, 6-azauridine and glycyrrhizin: antiviral compounds active against pathogenic flaviviruses Evaluation of anti-Zika virus activities of broad-spectrum antivirals and NIH clinical collection compounds using a cell-based, highthroughput screen assay A screen of FDA-approved drugs for inhibitors of Zika virus infection Zika antiviral chemotherapy: identification of drugs and promising starting points for drug discovery from an FDA-approved library Zika virus infection reprograms global transcription of host cells to allow sustained infection The new pregnancy and lactation labeling rule Content and format of labeling for human prescription drug and biological products; requirements for pregnancy and lactation labeling. Final rule Dengue virus infection perturbs lipid homeostasis in infected mosquito cells Leptospira interrogans endostatin-like outer membrane proteins bind host fibronectin, laminin and regulators of complement Dengue virus nonstructural protein 3 redistributes fatty acid synthase to sites of viral replication and increases cellular fatty acid synthesis Rab18 facilitates dengue virus infection by targeting fatty acid synthase to sites of viral replication Cholesterol biosynthesis modulation regulates dengue viral replication Different anti-HCV profiles of statins and their potential for combination therapy with interferon Lovastatin attenuates nerve injury in an animal model of Guillain-Barré syndrome Lovastatin for the treatment of adult patients with dengue: a randomized, double-blind, placebo-controlled trial Mevastatin, an inhibitor of HMG-CoA reductase, induces apoptosis, differentiation and Rap1 expression in HL-60 cells Antiviral activity of nordihydroguaiaretic acid and its derivative tetra-O-methyl nordihydroguaiaretic acid against West Nile virus and Zika virus Effects of hypolipidemic agent nordihydroguaiaretic acid on lipid droplets and hepatitis C virus Comparative effects of indomethacin, acetylenic acids, 15-HETE, nordihydroguaiaretic acid and BW755C on the metabolism of arachidonic acid in human leukocytes and platelets Effect of nordihydroguaiaretic acid on cell viability and glucose transport in human leukemic cell lines Nordihydroguaiaretic acid (NDGA), an inhibitor of the HER2 and IGF-1 receptor tyrosine kinases, blocks the growth of HER2-overexpressing human breast cancer cells Inhibition of IGF-1R and lipoxygenase by nordihydroguaiaretic acid (NDGA) analogs Inhibitory activities of propolis and its promising component, caffeic acid phenethyl ester, against amyloidogenesis of human transthyretin Molecular mechanisms and clinical applications of nordihydroguaiaretic acid (NDGA) and its derivatives: an update Inhibitors of cellular kinases with broad-spectrum antiviral activity for hemorrhagic fever viruses Zika virus infects human cortical neural progenitors and attenuates their growth Identification of small-molecule inhibitors of Zika virus infection and induced neural cell death via a drug repurposing screen The TIM and TAM families of phosphatidylserine receptors mediate dengue virus entry Axl mediates ZIKA virus entry in human glial cells and modulates innate immune responses Enveloped viruses disable innate immune responses in dendritic cells by direct activation of TAM receptors Axl is not an indispensable factor for zika virus infection in mice Anticancer kinase inhibitors impair intracellular viral trafficking and exert broad-spectrum antiviral effects Zika virus induces massive cytoplasmic vacuolization and paraptosis-like death in infected cells Endoplasmic reticulum vacuolation and unfolded protein response leading to paraptosis like cell death in cyclosporine A treated cancer cervix cells is mediated by cyclophilin B inhibition Paraptosis: mediation by MAP kinases and inhibition by AIP-1/Alix Stress responses in flavivirus-infected cells: activation of unfolded protein response and autophagy Identification of Zika virus and dengue virus dependency factors using functional genomics Roles of ubiquitin in endoplasmic reticulum-associated protein degradation (ERAD) Antiviral therapies targeting host ER alpha-glucosidases: current status and future directions Inhibitors of endoplasmic reticulum a-glucosidases potently suppress hepatitis C virus virion assembly and release Quantitative proteomic analysis of mosquito C6/36 cells reveals host proteins involved in Zika virus infection Design, modification and 3D QSAR studies of novel naphthalin-containing pyrazoline derivatives with/without thiourea skeleton as anticancer agents Cyclosporine A inhibits hepatitis C virus nonstructural protein 2 through cyclophilin A Cyclosporine inhibits flavivirus replication through blocking the interaction between host cyclophilins and viral NS5 protein Flavivirus cell entry and membrane fusion 25-Hydroxycholesterol protects host against Zika virus infection and its associated microcephaly in a mouse model Targeting endosomal acidification by chloroquine analogs as a promising strategy for the treatment of emerging viral diseases Hydroxychloroquine and Chloroquine Retinopathy Chloroquine, an endocytosis blocking agent, inhibits Zika virus infection in different cell models Inhibition of autophagy limits vertical transmission of Zika virus in pregnant mice Antiviral activities of selected antimalarials against dengue virus type 2 and Zika virus Obatoclax inhibits alphavirus membrane fusion by neutralizing the acidic environment of endocytic compartments Obatoclax, saliphenylhalamide and gemcitabine inhibit Zika virus infection in vitro and differentially affect cellular signaling, transcription and metabolism Obatoclax, saliphenylhalamide, and gemcitabine inhibit influenza A virus infection Structure-activity analysis of niclosamide reveals potential role for cytoplasmic pH in control of mammalian target of rapamycin complex 1 (mTORC1) signaling Niclosamide is a proton carrier and targets acidic endosomes with broad antiviral effects Open drug discovery for the Zika virus Working with Zika and Usutu viruses in vitro Small molecules and antibodies for Zika therapy UCSF Chimera-a visualization system for exploratory research and analysis VMD -visual molecular dynamics Supplementary data associated with this article can be found, in the online version, at https://doi.org/10.1016/j.drudis.2018.06.014.