Next Article in Journal
Phenolic Plant Extracts Versus Penicillin G: In Vitro Susceptibility of Staphylococcus aureus Isolated from Bovine Mastitis
Next Article in Special Issue
Computer-Assisted and Data Driven Approaches for Surveillance, Drug Discovery, and Vaccine Design for the Zika Virus
Previous Article in Journal
Multilevel Impacts of Iron in the Brain: The Cross Talk between Neurophysiological Mechanisms, Cognition, and Social Behavior
Previous Article in Special Issue
Antiviral Agents in Development for Zika Virus Infections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Therapeutic Advances Against ZIKV: A Quick Response, a Long Way to Go

Department of Biotechnology, Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA), 28040 Madrid, Spain
Pharmaceuticals 2019, 12(3), 127; https://0-doi-org.brum.beds.ac.uk/10.3390/ph12030127
Submission received: 3 July 2019 / Revised: 15 August 2019 / Accepted: 16 August 2019 / Published: 30 August 2019
(This article belongs to the Special Issue Zika Virus: Therapeutic Advances)

Abstract

:
Zika virus (ZIKV) is a mosquito-borne flavivirus that spread throughout the American continent in 2015 causing considerable worldwide social and health alarm due to its association with ocular lesions and microcephaly in newborns, and Guillain–Barré syndrome (GBS) cases in adults. Nowadays, no licensed vaccines or antivirals are available against ZIKV, and thus, in this very short time, the scientific community has conducted enormous efforts to develop vaccines and antivirals. So that, different platforms (purified inactivated and live attenuated viruses, DNA and RNA nucleic acid based candidates, virus-like particles, subunit elements, and recombinant viruses) have been evaluated as vaccine candidates. Overall, these vaccines have shown the induction of vigorous humoral and cellular responses, the decrease of viremia and viral RNA levels in natural target organs, the prevention of vertical and sexual transmission, as well as that of ZIKV-associated malformations, and the protection of experimental animal models. Some of these vaccine candidates have already been assayed in clinical trials. Likewise, the search for antivirals have also been the focus of recent investigations, with dozens of compounds tested in cell culture and a few in animal models. Both direct acting antivirals (DAAs), directed to viral structural proteins and enzymes, and host acting antivirals (HAAs), directed to cellular factors affecting all steps of the viral life cycle (binding, entry, fusion, transcription, translation, replication, maturation, and egress), have been evaluated. It is expected that this huge collaborative effort will produce affordable and effective therapeutic and prophylactic tools to combat ZIKV and other related still unknown or nowadays neglected flaviviruses. Here, a comprehensive overview of the advances made in the development of therapeutic measures against ZIKV and the questions that still have to be faced are summarized.

1. Introduction

Human and animal health has to face changes in the ecology of pathogens resulting from climate warming, vector colonization of new geographical niches, global trade, and human behavior. In this sense, emerging viruses represent a worrisome challenge because of their current worldwide spread, as exemplified by the outbreaks of the West Nile virus (WNV) in North America in 1999 [1], Chikungunya virus (CHIKV) in Asia in 2004–2005 [2], or Ebola virus (EBOV) in Western Africa in 2014 [3]. However, the more recent threatening has been caused by the emergence and dissemination of Zika virus (ZIKV) in the Americas in 2015 [4]. The virus has provoked more than 580,000 autochthonous suspected disease cases in the Americas since then [5], causing social and medical alarm mainly due to the evidence of a causal link between ZIKV and several congenital injuries, like microcephaly and ocular lesions, as well as to its association with neurological disorders, such as Guillain–Barré syndrome in adults [6]. All these events led the World Health Organization (WHO) to the declaration of a Public Health Emergency of International Concern (PHEIC) in February 2016 [7]. Since then, and even though the number of reported clinical cases has decreased a lot [8], a great effort has been conducted by the scientific community to understand the virus biology and pathogenesis, to improve diagnostic tools, and to develop therapeutic and prophylactic alternatives.
ZIKV is a mosquito-borne flavivirus (family Flaviviridae), which genome (about 11 kb in length) encodes a single open reading frame translated into a viral polyprotein that, after cotranslational and postranslational proteolytic cleavage, generates 10 mature proteins: Three structural proteins [C (capsid), prM/M (membrane) and E (envelope)] and seven non-structural proteins (NS1, NS2A, NS2B, NS3, NS4A, NS4B and NS5) [4,9].
The natural course of the ZIKV infection is usually asymptomatic or produces a relatively mild illness with a wide range of symptoms that go from fever, rush, and headache, to joint and muscle pain, and that usually ends up in an uneventful recovery [6,10]. Nevertheless, since ZIKV colonized the Americas, it has shown an increased virulence and rapid spread, and it has been associated with an unexpected rise of microcephaly cases in fetuses and newborns, severe ocular lesions, and a remarkable increase in Guillain–Barré syndrome (GBS) cases [10,11]. Additionally, ZIKV can also be sexually transmitted and persist in the male genital tract, which may lead to infertility [12].
Albeit a huge amount of research has been conducted in the last three years, which extensively surpasses that of any other recently emerging pathogen [13], there is still a lack of effective ZIKV vaccines or antivirals. Our current knowledge about ZIKV molecular mechanisms of pathogenesis, including the development of amenable animal models [14,15,16] and the elucidation of the crystal structure of some of the ZIKV proteins (C, NS1, NS2b-NS3, NS3, and NS5) [17], are certainly contributing to develop vaccine prototypes and to identify antiviral candidates; however, we are still far from controlling the virus. Here, a comprehensive overview of the advances made in the development of therapeutic measures against ZIKV is given.

2. Vaccines

Nowadays, measures such as vector control and surveillance programs, and avoiding mosquito bites are the more effective ones against the ZIKV infection. In addition, as pregnant and fertile women are the more at risk population, authorities from some countries recommend to abstain from unprotected sex if they, or their partners, live or travel to endemic regions. Therefore, development of efficient vaccines is more probably the best way to fight the ZIKV disease.
Several questions should be posed before a vaccine can be licensed to use in the clinical practice. In addition to its effectiveness, safety for pregnant women is mandatory [18], as they will be the main target for the ZIKV vaccine administration and other flavivirus vaccines, such as those for the Japanese encephalitis virus (JEV) or yellow fever virus (YFV), are usually contraindicated for them. Similarly, because different flaviviruses co-circulate in the same geographical areas, the possibility that a vaccine could generate cross-reactive antibodies and, thus, induce antibody dependent enhancement (ADE), should be taken into account. Additionally, as the ZIKV vaccine will be used preferably in developing countries, cost is another important point to take into account. In this sense, the WHO has published an updated target product profile (TPP) for a Zika vaccine that could be used in an outbreak response scenario [19].
So far, in less than three years, a large number of candidate vaccines has been evaluated in the early stages, and some of them have even entered into clinical trials (Table 1). Almost all available approaches have been rehearsed: Purified inactivated and live attenuated viruses, nucleic acid (DNA or RNA) based candidates, virus-like particles, subunit elements, and recombinant viruses.

2.1. Purified Inactivated Vaccines

Several purified inactivated ZIKV vaccines have been tested in mice and in non-human primates (NHP). Overall, the production of high levels of neutralizing antibodies, reduction of viremia, and protection of interferon deficient mice challenged with homologous and heterologous viral strains, and passive protection after administration of immunoglobulins from vaccinated animals, which seems to correlate with neutralizing antibody titers, have been reported [20,32]. One of these vaccines [20] entered the phase I clinical trial in India (CTRI/2017/05/008539) [21]. Another purified formalin-inactivated Zika virus vaccine (ZPIV) tested in NHP also induced specific neutralizing antibodies and full protection after viral challenge [22]. This ZPIV vaccine has entered too in phase I, placebo-controlled, double-blind trials (NCT02963909, NCT02952833, NCT02937233, NCT03343626, and NCT03425149). Initial reports indicate that most of the enrolled participants seroconverted, developed antibody titers higher than the threshold recorded in animal studies, and that the passive transfer of purified neutralizing IgG from vaccinated participants protected mice against the ZIKV infection [23].

2.2. Live Attenuated Vaccines

Different approaches have been tested as live attenuated vaccine candidates. A prototype engineered by deleting 10 nucleotides from the 3’ untranslated region of the viral genome (10-del ZIKV) induced high neutralizing antibody titers and a strong cellular response, completely reduced viremia in interferon-receptor deficient mice, and protected CD1 newborn mice from intracranial inoculation of a lethal strain [33]. Even more, the same group demonstrated that a single dose of this vaccine diminishes levels of viral RNA in maternal, placental, and fetal tissues in challenged mice, preventing viral transmission during pregnancy, as well as testes damage and oligospermia in vaccinated male mice [33]. This vaccine prototype also protected NHP from ZIKV infection, but animals showed an anamnestic antibody response when challenged, indicating that the virus did replicate and, thus, there were no sterilizing immunity [34]. Immunization with an attenuated vaccine due to a deletion of nine amino acids in the viral capsid protein also elicited protective immunity that completely prevented viremia, conferred protection after challenge, and avoided virus transmission in immunodeficient mice [35]. Another attenuated prototype encoding a non-glycosylated NS1 protein diminished viral RNA levels in maternal, placental, and fetal tissues, and protected against placental damage and fetal demise when tested in dams early after pregnancy [36].

2.3. Nucleic Acid (DNA or RNA) Vaccines

Several DNA-based vaccines expressing the viral prM/E proteins have been evaluated in animal models, including humanized mice. Overall, these candidates induce the production of specific humoral and cellular responses, reduce viremia, protect against viral challenge, and confer protection after passive transfer of immune serum from both vaccinated mice and NHP, and, in some cases, prevent vertical virus transmission [24,27,37,38,39]. A DNA (prM/E) prototype has been shown to protect mice against ZIKV-associated damage to the testes and sperm after viral challenge, preventing viral persistence in the testicles [39]. These types of vaccines have also been effective in protecting NHP against the ZIKV challenge [32]. Even more, one of the above mentioned DNA vaccine (GLS-5700), which had been previously tested in mice [25], has entered clinical trials phase I (NCT02809443 and NCT02887482) [26]. In these studies over 80% of the vaccinated individuals elicited neutralizing antibodies, and inoculation in immunodeficient mice of their post-vaccination sera protected most of the animals against a lethal viral challenge. Likewise, and based on previous data obtained after mice vaccination [27], preliminary results of two other trials (NCT02840487 and NCT02996461) have been reported, one evaluating a chimeric ZIKV-JEV plasmid and the other a ZIKV wild-type plasmid. Both candidates were safe and well tolerated, and strong specific antibody and cellular responses were achieved with the chimeric one [28]. The one encoding the wild type ZIKV has advanced to the phase II clinical trial (NCT03110770).
Similar approaches have been conducted using RNA-based vaccines encoding also the prM/E region of the viral genome and administered as nanoparticles (NP). An early report showed that a candidate tested in immunocompetent mice induced the production of specific antibodies and a CD8 + response against a unique H-2Db-restricted epitope was identified [40]. Another non-replicating lipid-nanoparticle-encapsulated nucleoside-modified mRNA (mRNA-LNP), encoding the prM/E and expressed in cells to produce virus-like particles (VLPs), induced a potent, sterilizing, and durable neutralizing antibody response in mice and NHP, and protected both of them against the ZIKV infection [29]. Similarly, another mRNA-LNP candidate also protected against the ZIKV infection and conferred sterilizing immunity in mice [36]. Still more, taking into account the possibility that the vaccine could induce cross-reactive antibodies against DENV and, thus, induce antibody dependent enhancement (ADE), the authors introduced mutations in the conserved fusion-loop epitope of the E protein [36]. These mutant vaccines not only protected against the ZIKV challenge, but also diminished the production of antibodies that may enhance the DENV infection in cells and mice [25]. Even more, the same RNA-LNP construct (mRNA-1325) diminished levels of viral RNA in maternal, placental, and fetal tissues [30], and has advanced to clinical trials (NCT03014089).

2.4. Virus-like Particles Vaccines

In addition to the above mentioned modified lipid-encapsulated RNA vaccines encoding the prM/E region that render virus-like particles (VLPs), several other groups have described the immunogenicity of this kind of prototypes. VLPs assembled by co-expressing the C/prM/E and the NS2B/NS3 viral proteins stimulated the production of high virus neutralizing antibodies in mice [41,42]. Likewise, VLPs produced by the expression of just the prM/E proteins in HEK293 cells induced neutralizing antibodies that were protective in infected mice and in those receiving passive transfer of purified IgG [43]. A different approach was conducted by using recombinant VLPs based on the hepatitis B core antigen (HBcAg) that incorporates the ZIKV E protein domain III (DIII) and were produced in Nicotiana benthamiana plants. This candidate elicited potent humoral and cellular responses in mice and did not enhance the infection of DENV in Fc gamma receptor-expressing cells [44].

2.5. Subunit Vaccines

Purified soluble ZIKV E protein, expressed in Drosophila melanogaster cells or E. coli, induced virus-specific neutralizing antibodies, and avoided viremia when inoculated in immunocompetent mice [45,46]. When assayed in NHP, one prototype produced in insect cells elicited high neutralizing antibody titers, conferred sterilizing immunity, and protected mice after passive transfer of plasma from vaccinated NHPs [47]. Likewise, an N-terminal truncated E (∆E) protein, expressed in either of the two systems, elicited ZIKV-specific humoral and cellular response in immunocompetent mice, and passive transfer of sera from immunized animals conferred full protection against the lethal ZIKV challenge in newborns mice [48,49]. A similar ∆E protein has also been shown to reduce ZIKV-infected cells in the brains of fetal or suckling mice, preventing the onset of microcephaly [50].
Purified ZIKV envelope domain III (DIII) has also been expressed in D. melanogaster cells, E. coli, and Nicotiana benthamiana plants. These constructs induced humoral and cellular responses when administered to mice, and passive transfer of sera protected recipient mice against the lethal ZIKV [49,51,52]. In addition, they did not enhance the infection of DENV when tested in Fc gamma receptor (FcγR)-expressing cells [51,52]. Likewise, a fragment (aa 298-409) of the DIII expressed in 293T cells stimulated long-term immunogenicity and protected immunodeficient mice and newborns from vaccinated mothers against the virus challenge, and also pups from naïve mothers that received a passive transfer of specific antibodies [53].

2.6. Recombinant Vaccines

Different viruses have been assayed as a backbone for the development of ZIKV recombinant vaccines. Adenovirus strains of various serotypes have been engineered to include the ZIKV structural proteins. A recombinant chimpanzee adenovirus type 7 (AdC7) expressing ZIKV M/E proteins elicited durable humoral and cellular responses, and protective and sterilizing immunity against heterologous ZIKV infection in mice, and also protected against testes damage [54]. Several human adenoviruses (Ad4, Ad5 and Ad26) encoding either the E alone or the prM/E have shown to induce a durable antibody and cellular responses in mice and NHP, reducing viral titers in tissues and blood, and protecting mice from viral challenge [55,56,57].
Similar approaches have been conducted by using modified vaccinia Ankara (MVA) as a vector. One prototype encoding the ZIKV prM/E elicited neutralizing antibodies, induced a potent and polyfunctional ZIKV-specific CD8 + T cell response, and significantly reduced viremia in mice [58]. Another MVA candidate expressing the ZIKV-NS1 also provided robust humoral and cellular responses, and afforded 100% protection in mice against a lethal intracerebral challenge [59]. This line of attack is interesting because it could circumvent potential ADE, as it does not encode the E protein. By using the recently developed vaccinia-based Sementis Copenhagen Vector (SCV), a multi-pathogen vaccine encoding the structural proteins of chikungunya and Zika viruses was shown to induce neutralizing antibodies against both viruses and to reduce viremia in mice [60]. In addition, it conferred protection against chikungunya-related arthritis, and ZIKV fetal/placental and testis infection.
As in the case of adeno and vaccinia viruses-based vaccines, vesicular stomatitis virus (VSV) has also been evaluated as a platform for ZIKV vaccines. Immunization with a VSV construct expressing the GP of the EBOV and the ZIKV prM/E, or a prM/ΔE, resulted in complete protection of challenged mice [61]. Likewise, when the ZIKV prM-E-NS1 was produced as a VSV recombinant construct, it induced the ZIKV specific humoral and cellular responses that protected against the ZIKV challenge [62]. Furthermore, a VSV-based CHIKV (E3-E2-6K-E1) and ZIKV (M/E) bivalent vaccine induced neutralizing antibody responses to both CHIKV and ZIKV in mice and protected them against infection with either viruses [63].
Recently, results obtained after immunization of mice with a live-attenuated measles virus (MV) strain encoding the ZIKV prM/E have been reported [31]. Specific neutralizing antibodies and cellular responses, reduction of viral titers in sera and organs, and prevention of infection of the fetus were observed in immunized animals. This candidate has entered the phase I clinical trial (NCT02996890).
Finally, advantages have been taken of the already available vaccines against other related flaviviruses. So that, a recombinant chimeric ZIKV vaccine candidate that expresses the ZIKV prM/E proteins using the licensed JEV live-attenuated vaccine (SA14-14-2) as a backbone elicited robust and long-lasting immune responses and conferred complete protection against the ZIKV challenge both in mice and NHP [64]. Even more, the vaccine also protected against the ZIKV placental and fetal damage during pregnancy [64]. Likewise, when the prM/E proteins of the live attenuated yellow fever 17D vaccine virus (YF 17D) were replaced by those of ZIKV, it was shown a significantly induction of neutralizing antibodies and a reduction of viral loads in organs, avoiding neuroinvasion, and protecting against the ZIKV challenge in mice [65]. A ZIKV (prM/E)-YF 17D variant with adaptive mutations in the E protein also drove to seroconversion and to a strong cellular response in vaccinated mice that conferred full protection against homologous and heterologous viral challenge, and protected fetus from intra-placental challenged vaccinated dams [66]. Improvements of this kind of prototypes by modifying the cleavage site between the pM and M, which are compensated by mutations in the E protein, induced high neutralizing antibody titers in mice and protected against the heterologous viral challenge [67].

2.7. Therapeutic Vaccination

As mentioned above, the success of passive transfer of sera from vaccinated mice, NHP, and humans in protecting mice against the ZIKV infection has been extensively reported. In this regard, an early study demonstrated that an EDIII-specific antibody protected mice from lethal ZIKV infection [68]. A DENV-specific antibody against an E-epitope has also been shown to cross-neutralize ZIKV and protect against the ZIKV infection when administered to NHP [69]. Later on, it was reported that neutralizing the convalescent serum from a donor who recovered from the ZIKV infection prevented virus replication and ZIKV-induced microcephaly when inoculated in pregnant mice [70]. Therapeutic human monoclonal antibodies (mAbs) have also been evaluated in mice. So that, a subset of mAbs from patients recovered after the ZIKV infection were shown to neutralize homologous and heterologous viral strains, and administration of one of them in mice reduced tissue pathology, and placental and fetal infection [71]. Similarly, a single injection of ZIKV-specific human neutralizing mAbs induced different levels of protection against the ZIKV infection in adult mice and in utero [72]. A cocktail of three ZIKV-neutralizing mAbs from a ZIKV-infected patient, which were engineered to the abrogate Fcγ receptor binding to eliminate a potential ADE effect, completely prevented viremia when administered to NHP one day before the ZIKV challenge [73]. Likewise, another report showed that four non-neutralizing mAbs from an infected patient that target the non-structural protein NS1 can engage FcγR without inducing ADE in vitro, and that one of them was protective against lethal homologous and heterologous challenges in mice, an effect that was Fc-dependent [74].

2.8. Future Perspectives for ZIKV Vaccines.

As described above, in as little as three years a great number of vaccine candidates has been tested in animal models using many different approaches that, overall, have led to hopeful results. Induction of strong specific humoral and cellular responses, decrease or undetectability of viremia and viral RNA in organs (including placenta, foetal tissues, and testes), prevention of viral transmission during pregnancy and of the onset of microcephaly, and protection in experimental animal models, which seems to correlate with the development of high titers of neutralizing antibodies, have been achieved. Additionally, experiments with passive transfer of immune sera from vaccinated individuals and human mAbs point to therapeutic vaccination as a possible option for the ZIKV control. In fact, several candidates have already entered in clinical trials (Table 1) [75].
In any case, and even though much progress has been accomplished, there is still a long way to go before a prophylactic vaccine can be licensed for use in clinical practice. So that, several considerations should be kept in mind. Although there are evidences pointing that neutralizing antibodies are likely a surrogate for protection, many preclinical tested vaccines did not induce sterilizing immunity and, thus, this point remains to be clarified. In fact, it seems that, in comparison with other flaviviruses, higher levels of neutralizing antibodies are required to protect animals, and probably humans, against the ZIKV infection. Likewise, the durability of the induced immunity has still to be more deeply evaluated. On the other hand, as ZIKV circulates in regions endemic for other related flaviviruses (DENV, YFV, WNV, etc.) another important point is to assure that vaccine implementation would not induce undesirable results like the enhancement of a secondary infection due, for instance, to ADE. In addition, ideally, a vaccine must also prevent infection of reproductive tissues to avoid sexual transmission. Even more, as the main target population that will benefit ZIKV vaccines are fertile and pregnant women and, since other licensed flavivirus vaccines (JEV or YFV) are usually contraindicated for them, safety should be fully evaluated in this population. Moreover, and albeit positive results have been obtained in mice, this is not an adequate model for women pregnancy, and therefore, this point must be carefully taken into account before vaccine campaigns implementation. On top of that, since the number of new ZIKV cases recorded has drastically dropped and other related flaviviruses co-circulate in the same geographical regions, further clinical endpoint trials efficacy will be difficult to undertake. In any case, ZIKV is still circulating at low levels in some areas and its re-emergence in naïve populations or in areas where population immunity wanes remains a real possibility; therefore, the public health need for a ZIKV vaccine persists.

3. Antivirals

Nowadays there is no licensed specific antiviral against ZIKV [76], or any other flavivirus [77], and treatments are palliative and directed to symptoms relief. Hence, as in the case of vaccines development, in a short time many efforts have been conducted to assay dozens of possible antiviral candidates by using many different approaches and tools (in vitro and in silico assays, diverse viral and cell lines, including primary and pluripotent stem-cells, infectious clones, and replicons) [15]. Two major lines of investigation are being evaluated: Search for direct acting antivirals (DAAs) based on interference with viral components, and search for host-acting antivirals (HAAs) based on the inhibition of host factors co-opted for the virus to complete its infectious cycle. In both approaches two important points that ZIKV antivirals have to face is their safety for pregnant women and their ability to cross the blood-brain barrier (BBB) to combat the infection in nervous tissues [78]. With this in mind, efforts have been done by testing of natural compounds, screening of libraries from different sources, and repurposing of drugs with the known antiviral activity that have allowed the identification of several antiviral candidates directed to viral structural proteins and enzymes and to cellular factors.

3.1. Direct Acting Antivirals (DAAs)

DAAs (Table 2) are those compounds that directly interfere with viral components, either structural proteins or enzymes, for which many of their crystal structures (C, NS1, NS2b-NS3, NS3, and NS5) have been resolved [17], what is undoubtedly helping to find more precise antivirals.

3.1.1. Compounds Acting Before Viral Binding

Administration of convalescent sera and neutralizing antibodies suppress ZIKV multiplication and inhibits cell death in infected fetal brains, preventing microcephaly in mice and protecting NHP [68,69,70,71,72,73,74,79]. In this line, as commented above regarding vaccines development, many authors have reported that the passive transfer of immune sera from vaccinated individuals protects mice against the ZIKV challenge, thus pointing to the use of therapeutic antibodies to combat the ZIKV infection. However, as in the case of vaccines implementation, the possibility of undesirable side effects due to cross-reactive antibodies against other flaviviruses, such as ADE in the case of dengue infections, need to be carefully contemplated.
Different compounds impair viral fusion to the cellular membrane, mostly by directly binding to the viral envelope E protein and preventing E-mediated membrane fusion. So that, the synthetic peptide Z2, which is able to penetrate in fetal tissues, avoids vertical transmission and protects against the ZIKV challenge in mice [80]. Porphyrins like Co-protoporphyrin IX (CoPPIX) and Sn-protoporphyrin IX (SnPPIX) induce viral envelope protein loss, affecting viral morphology and entry into target cells [81]. Benzoxazine monomer derived carbon dots (BZM-CDs) [82], polyoxometalates (POMs) [83], cyanohydrazones [84], synthetic carbohydrate receptors (SCRs) [85], and some other small molecules [86,87,88] also inhibit ZIKV infection in cell culture, and reduce viral replication in mice [89].
Compounds exerting a viricidal activity, such as crotoxin, from the Crotalus durissus terrificus venom [90], Aphloia theiformis and Psiloxylon mauritianum extracts [91,92], and polyphenols like epigallocatechin gallate or delphinidin, present in many natural products, also exhibit anti-ZIKV activity [93,94,95], as does the natural flavonoid pinocembrin that impairs the ZIKV infection in culture cells, although it seems to act on post-entry processes [96].

3.1.2. Antivirals Targeting the ZIKV NS5 Polymerase and Methyltransferase Domains

The NS5 protein is the RNA dependent RNA polymerase (RdRp) responsible of viral genome replication. Nucleoside analogs/derivatives target viral but not cellular polymerases and, after their incorporation into the viral nascent RNA chain, abrogate viral RNA replication. These compounds are usually safe for use in humans 3 [143] and, thus, they have been widely evaluated as ZIKV antivirals in cell culture and, in a few cases, in animal models.
Several inhibitors of the inosine monophosphate dehydrogenase (IMPDH), such as ribavirin, a well-known antiviral against many RNA viruses [144], and favipiravir (or T-705) inhibits ZIKV multiplication in different cell lines, including human ones [97,98,99,100,101,102], and abrogate viremia in ZIKV-infected susceptible mice [99], as do merimepodib [103], and mycophenolic acid [104,105,106,107], although the later showed significant cell toxicit [104]. Azathioprine, another purine synthesis inhibitor and immunosuppressant, impairs ZIKV replication in cell culture [108], but it is not recommended for use during pregnancy. Methotrexate (MTX), an anti-cancer chemotherapy and anti-rheumatoid agent, exerts its antiviral activity by inhibiting the dihydrofolate reductase (DHFR) in different cell lines, restricting the synthesis of adenosine triphosphate and disrupting ZIKV replication [109]. Atovaquone, a FDA pregnancy category C drug used to prevent parasitic infections, impairs ZIKV virus production in human cells by inhibiting pyrimidine biosynthesis, and limits ZIKV infection in an ex vivo human placental tissue model [110]. Likewise, 6-methylmercaptopurine riboside [111], 2′-C-methylguanidine, 2′-C-methyladenosine, and 7-deaza-2′-C-methyladenosine inhibit ZIKV replication in cell cultures [105,112,113], and the later one reduces viremia and retards disease signs in mice [113]. Finally, NITD008 [114] and BCX4430 [115] protects immunosuppressed mice, although the later one does not reduce viral loads.
ZIKV replication in cell culture is also affected by a variety of inhibitors of pyrimidine biosynthesis, such as brequinar and CID 91632869 [108], which effect may be due not to pyrimidine privation but to cellular immune response induction [116,145]. Gemcitabine [107,117], although its use may result in damage to the fetus, 2′-2′-C-methylcitosine, 2′-C-methyluridine [118], and their proTides [118], 5-fluorouracil, an anticancer drug, 6-azauridine, an antineoplastic, and finasteride, used in patients with enlarged prostate [97,108,119], also reduce virus multiplication. However, the clinical use of these last two may be impaired by their low solubility. Likewise, P12-23 and P12-34 derivatives of AR-12, an anticancer agent, suppress ZIKV replication in vitro and significantly improves the survival of infected mice [120]. Sofosbuvir, approved by the FDA to treat HCV infected patients, inhibits ZIKV replication in human cell lines of different origin origin [121,122] and improves survival rates in infected susceptible mice, origin [121], though it is not recommended for use in pregnant women. Viperin, an interferon-inducible protein, catalyzes the conversion of cytidine triphosphate (CTP) to 3’-deoxy-3’,4’-didehydro-CTP (ddhCTP), which acts as a chain terminator for the RNA-dependent RNA polymerase and, thus, inhibits replication of Zika virus in mice [123,124]. Aurintricarboxylic acid (ATA), a polyanionic aromatic compound and a potent ribonuclease inhibitor with apoptotic activity that inhibits several bacteria and viruses in vitro, has also been shown to reduce ZIKV multiplication in pre- and post-infection settings [125]. Finally, the methyltransferase domain of the NS5 protein that is responsible for transferring the mRNA cap to the 5’ end of the viral genomic RNA, and which structure is already known, may also serve as a target for the design of antivirals against ZIKV, as it has been demonstrated by the inhibitory activity of several compounds, F3043-0013, F0922-0796, F1609-0442, and F1750-0048 [126,127,128].

3.1.3. Protease Inhibitors.

The N2B-NS3 trypsin-like serine-protease and the NS3 helicase contribute to viral polyprotein processing and, thus, have an important role on virus replication. As a result, several reports have searched for antiviral against these viral enzymes [146,147,148].
An early report identified seven natural flavonoids that inhibited ZIKV multiplication [129]. Additional natural flavonoids like myricitin, quercetin, luteolin, isorhamnetin, and apigerin, and the phenol curcumin inhibit the N2B-NS3 noncompetitively [130], although some of them seem to be promiscuous drugs, as it has been reported that they affect viral binding [130,131,132]. A similar case is suramin, an anti-parasitic, that inhibits NS2B/NS3 complex [133], but also viral binding to the cell membrane [131,132]. On the other hand, by screening a battery of almost 3000 approved and investigational drugs, three compounds (temoporfin, niclosamide, and nitazoxanide) were identified that presented inhibitory activity against ZIKV and also DENV, JEV, YFV, and WNV [134]. Even more, temoprofin inhibited viremia and protected mice against lethal infection [134] In this line, by using a structure-based assay over 8000 approved and investigational drugs, various potent non-competitive NS2B-NS3 inhibitors were identified [135], among them, novobiocin, which can be administered to pregnant women, significantly increases survival rates of infected mice, and reduces blood and tissue viral levels, as well as histopathological damages [135]. By molecular docking studies it was predicted that bromocriptine bind to the NS2B-NS3, inhibiting its proteolytic activity, which was further confirmed in cell culture assays [136]. An additional study tested several HCV NS3-NS4A protease inhibitors and found that two inhibited ZIKV NS2B-NS3 probably binding to the protein active site [137]. Later on, it was reported that aprotinin, a bovine trypsin inhibitor with activity against WNV, also reduces ZIKV multiplication [138]. However, although molecular modeling predicted that it probably blocks the NS2B/NS3 interaction, others study with WNV and DENV indicated that it binds and occludes the enzyme substrate site [139,140]. In this line, another study screened hundreds of protease inhibitors and found three (NSC157058, NSC86314, and NSC716903) that inhibited the enzyme [138]. In fact, NSC157058 significantly decreased viremia in infected mice, but it had an unfavorable pharmacokinetic profile [138]. Another DENV inhibitor, NSC135618, was shown to also inhibit ZIKV as well as WNV and YFV, although the mechanism has not yet been elucidated [141]. Later on, it has been reported that erythrosin B, an FDA-approved food additive, inhibits the ZIKV and DENV2 NS2B-NS3 proteases by a non-competitive mechanism [142]. Finally, viperin was found to interact and colocalize with the ZIKV nonstructural proteins NS2A, NS2B, and NS3, reducing NS3 expression by induction of its proteasome-dependent degradation, inhibiting viral multiplication, an effect that was not observed for JEV or YFV [123], although it has also been reported that acts as pyrimidine inhibitor [124].
On the other hand, searching for inhibitors of the NS3 helicase and the NS4B presents several drawbacks that are complicating its development, although preliminary studies with DENV may open a door to find compounds acting against these ZIKV proteins [149].

3.2. Host-Acting Antivirals (HAAs)

ZIKV, as a flavivirus, has a small RNA genome (~ 10.7 Kb in length), which implies that it requires different host factors and the use of cellular metabolic pathways to propagate efficiently [150]. Thus, this may be used in search of host targets as therapeutic tools. Even more, since different members of the family Flaviviridae often share these host factors, these compounds have the potential of being pan-flaviviral antivirals, and its use would be less prone to the emergence of escape mutants, as often occurs with the DAAs [151,152]. Even though interfering with host factors and metabolism has raised concerns, drugs as ibuprofen and aspirin (COX-2 ciclooxigenase inhibitors) or statins (3-hydroxy-3-methyl-glutaryl-coenzyme A, HMG-CoA reductase inhibitors) are nowadays widely used in the clinical practice. Therefore, HAAs interfering with almost all steps of ZIKV life cycle (binding, entry, fusion, translation, transcription, replication, maturation, and egress) have been evaluated (Table 3).

3.2.1. Early Steps

As in any viral infection, the initial step is the binding of ZIKV to the receptor anchored in the cell membrane and its internalization into the cytoplasm. So far, several molecules have been suggested to be ZIKV receptors (AXL, DC-SIGN, Tyro3, TIM and TAM). These receptors are expressed in different permissive cell types, and, in some instances, abrogation of a given receptor can drive to the use of a different one [153,210]. Inhibition of viral binding has been achieved in different cell types with various molecules, such as R428 and cabozantinib (AXL kinase inhibitors) [153,211], MYD1 (an AXL decoy) [154], curcumin (a food additive) [155], and suramin (an anti-parasitic) [131,132], although it has also been described that it inhibits the NS2B/NS3 complex [133]. Similarly, nonsteroidal anti-inflammatory drugs (NSAIDs), such as ibuprofen, aspirin, acetaminophen, lornoxicam, and naproxen have been reported to prevent the entry of the Zika virus, in some instances by reducing the expression of AXL, the entry cofactor of ZIKV; [156] however, NSAIDs are usually contraindicated as in other flavivirus infection they increase the risk of the hemorrhagic syndrome [212]. On the other hand, since lipids are essential components of ZIKV virions, it has been described that the small molecule CLR01 [157] and an amphipathic α-helical peptide [158] both disrupt virion envelope, reducing ZIKV-infectivity in cell cultures. Even more, the latter showed therapeutic effect on infected mice.
After binding, ZIKV is internalized by endocytosis to reach the endosomes [154] where the viral genome is released into the cytoplasm after endosomal acidic pH triggers the fusion of the viral envelope with endosomal membranes [213], a process dependent of the presence of cholesterol and specific lipids in the membrane [214] These processes have been inhibited by different compounds such as nanchangmycin, an insecticide and antibacterial [107], the broad-spectrum antivirals arbidol [159] and compound 16, a pyridobenzothiazolone that does not reduce ZIKV RNA synthesis, but prevents a second round of infection [160], chlorpromazine, an antipsychotic [161], and daptomycin, an antibiotic [104].
Obatoclax (or GX15-070), an inhibitor of Bcl-2, reduces endosomal pH and inhibits ZIKV fusion to the membranes [117,162], but because of low solubility it was not effective when tested for hematological and myeloid diseases in clinical trials. Saliphenylhalamide (SaliPhe) also blocks endosomes acidification and inhibits ZIKV multiplication in virus natural human cells targets [108,117] .Likewise, amodiaquine, an antimalarial drug [163], inhibits ZIKV mutiplication in cell culture, as does the scorpion Euscorpiops validus venom peptide Ev37 that alkalizes acidic organelles preventing low pH-dependent fusion of the viral membrane-endosomal membrane [164]. Chloroquine (CQ), a FDA-approved anti-inflammatory widely used as an anti-malarian drug, inhibits the fusion to the endosomal membrane by raising the endosomal pH [166,168]. CQ also reduces placental and fetal ZIKV infection [167], and attenuates ZIKV-associated morbidity and mortality, protecting mice foetus from microcephaly [165], and vertical transmission in ZIKV-infected pregnant mice, significantly reducing foetal brain viral loads [138]. Other lysosomotropic agents (ammonium chloride, mefloquine, quinacrine, bafilomycin A1, and GSK369796) that neutralize the endosomal acidic pH also block ZIKV infection of different cell types [161,168]. Similarly, different molecules that interfere with the endosomal activity, like the small inhibitor K22 that produces severe alterations of ZIKV-induced intracellular replication compartments [169], and iron salt ferric ammonium citrate (FAC) [170], tenovin 1 [107], and niclosamide [171], a category B antihelmintic drug approved by FDA, also inhibit ZIKV multiplication. Niclosamide also decreases ZIKV production and prevents apoptosis in human cells, and partially rescues ZIKV-induced microcephaly and attenuates infection in a developed humanized ZIKV-infected embryo model in vivo [172]. Overall, all these molecules have demonstrated its potential for targeting ZIKV entry and internalization in cell culture.

3.2.2. Transcription/Translation

Once ZIKV-RNA is in the cytoplasm, a negative strand RNA is synthesized to direct the production of the new positive strand RNA molecules for their further encapsidation [4]. Different compounds targeting these processes have been assayed as ZIKV antivirals. Silvestrol, a natural compound that inhibits eIF4A, and fenretinide (or 4-HPR), an inhibitor of cancer proliferation that induces apoptosis, limit ZIKV replication in human cell lines [173,174], and the later one significantly reduces viremia and brain viral burden in mice [174]. Likewise, drugs interfering with the polyamine biosynthetic pathway needed for ZIKV translation and transcription [215], as difluoromethylornithine (DFMO or eflornithine), an FDA-approved drug to treat trypanosomiasis and some cancers, and diethylnorspermine (DENSpm) impair viral multiplication [216]. On the other hand, the small molecule GW5074 inhibits the interaction between the NS5 and the host importin IMP α/β1 and, thus, NS5 nuclear localisation [175].

3.2.3. Replication

ZIKV replication and particle morphogenesis are carried out in the ER and then, once packaged, inmature virions traffic thru the trans-Golgi before cleavage of the prM protein to render mature virions that are released from the cell [4]. Thereby, different drugs that affect these processes have been shown to inhibit ZIKV multiplication. Brefeldin A [176], a Penicillium sp inhibitor of protein transport from the ER to the Golgi, emetine [177], an anti-protozoal that disrupts lysosomal function, NGI-1 [178], a modulator of the oligosaccharyltransferase (OST) complex, the host ER-associated signal peptidase (SPase) [179], cavinafungin [180], a fungal compound, and nitazoxanide [181], an FDA approved broad-spectrum antiviral, all impair ZIKV infection in different cell types.
Cell death during ZIKV infection is mediated by the induction of host caspases-3 and neuronal apoptosis [217]. Therefore, caspases inhibitors as bithionol [182], and emricasan [171], currently in phase 2 clinical trials for chronic HCV, inhibit ZIKV in cells of different origin. Likewise, phloretin reduces ZIKV infection by decreasing apoptotic caspase-3 and -7 activities and by reducing AKT/mTOR phosphorylation pathways, which, together with the fact that 2-deoxy-D-glucose disruption of cellular glucose availability inhibits ZIKV propagation, remarks the importance of glucose pathways for ZIKV propagation [183]. Additionally, Bortezomib [104], a proteasome inhibitor approved for treatment myeloma and non-Hodgkin’s lymphoma, and PHA-690509 [171], a cyclin-dependent kinase (CDK) inhibitor, also reduce ZIKV-infection and propagation, although the latter is not adequate for use during pregnancy.
Host lipids are essential for flavivirus replication and particle formation and, thus, they are potential candidates for antiviral intervention [214,218]. Accordingly, two drugs that interfere the sterol regulatory element-binding proteins (SREBP) pathway and disturb the lipid metabolism, nordihydroguaiaretic acid (NDGA) and its derivative the tetra-O-methyl nordihydroguaiaretic acid (M4N), which is currently in Phase I/II clinical trials in patients with advanced cancer, inhibit the infection of ZIKV, as do other structurally unrelated inhibitors of the SREBP pathway, such as PF-429242 and fatostatin [184]. Similarly, three acetyl CoA carboxylase (ACC) inhibitors (PF-05175157, PF-05206574, and PF-06256254) also impair ZIKV replication in cellular culture [185] In addition, activation of adenosine monophosphate-activated protein kinase (AMPK), a regulator of lipid metabolism and ACC, by PF-06409577, metformin, and 5-aminoimidazole-4-carboxamide ribonucleotide (AICAR) were shown to reduce ZIKV replication [186,187].
Cholesterol is also a key player during flavivirus infection. So that, the cholesterol derivative 25-hydroxycholesterol (25-HC) [188,189] reduces mortality and prevents microcephaly in ZIKV-infected mice, and decreases viral loads in the urine and serum of NHP [189]. Likewise, 7-ketocholesterol (7-KC) that probably acts by inducing cellular autophagy, also reduces ZIKV replication [190], as do imipramine, an FDA-approved antidepressant, and benzamil that interfere with intracellular cholesterol transport [191,192]. Two other HMG-CoA inhibitors widely used to control cholesterol in the clinical practice, lovastatin, which attenuates nervous injury in animal model of Guillain–Barré syndrome, Sarkey [193], and mevastatin have been proposed as ZIKV antivirals [120]. In addition, it has been shown that the neutral sphingomyelinase inhibitor GW4869 reduces ZIKV production by affecting viral morphogenesis [194].

3.2.4. Other Compounds Acting Against ZIKV Infection

In addition to the above mentioned antivirals, some other drugs have been reported as effective against ZIKV infection in cell culture, although the specific mechanism by which they act is currently unknown. In this way, the antibiotics kitasamycin [120], a natural product from Streptomyces narbonensis, and azithromycin [195] prevent ZIKV infection in cell culture, as do cyclosporine [104], an immune depressor currently in clinical trials for its possible use in ameliorate neuronal cellular damage, and the antidepressant sertraline [120], a selective serotonin reuptake inhibitor. Similarly, the antiparasitics ivermectin, used to treat worms infections, pyrimethamine, used for toxoplasmosis and cystoisosporiasis [104], and amodiaquine dihydrochloride dihydrate (AQ) [196], an FDA approved drug for treatment of malaria, also inhibit ZIKV infection of human cells, and the last one also in adult mouse brain even after infection has progressed. Palonosetron, a FDA approved drug to prevent chemotherapy-induced nausea and vomiting, has also been proposed for treatment of ZIKV infection [120]. In addition, Hsp70 isoforms are recruited for ZIKV entry, replication, and assembly, and, thus, compounds targeting them reduce viral multiplication in several cell lines, and protect mice, reducing viremia, mortality, and disease symptoms [197].
On top of HDAs, some compounds have been shown to be effective in preventing side effects linked to ZIKV infection. So that, blockers of the N-methyl-D-aspartate receptor (NMDAR), implicated in neuronal damage that often occurs during ZIKV infection [198,219,220], such as ifenprodil, dizocilpine agmatine sulfate, and memantine, used to treat patients with Alzheimer’s disease, prevent neuronal damage in infected mice, but do not inhibit viral replication [199]. Ebselen (EBS), an antioxidant, improves testicular injury in mice by reducing oxidative stress, leucocyte infiltration, and production of pro-inflammatory response, and also prevents virus sexual transmission, but it neither reduces viremia nor improves survival rate [200]. IL-1 receptor antagonist (IRA; Kineret, or anakinra) can preserve placental, increase fetal viability, attenuate fetal neuroinflammation, behavioral deficits, and improve perinatal outcomes [201].

3.2.5. Modulators of the Innate Immunity

ZIKV infection, as many other viruses, activates the innate immune response by activation of interferons (IFNs) stimulated genes to confer resistance to viral infections [16,221,222]. Therefore, IFNs inhibit ZIKV replication in cell culture [202,203] and in mice [204], as did IFN-induced transmembrane proteins (IFITM1 and IFITM3) [205], interferon-activating small molecule AVC [206], and the scorpion venom peptide (Smp76) from Scorpio maurus palmatus that suppresses ongoing viral infection by upregulating the expression of IFN-β activating interferon regulatory transcription factor 3 (IRF3) phosphorylation [207]. Additionally, several small human noncoding microRNAs (miRNAs) that affect gene expression and regulate many different cellular processes have been shown to inhibit ZIKV by activating the IFN-based innate immune response [208]. Nonetheless, ZIKV can also evade type I IFN response, [223,224,225,226] which may lead to spontaneous abortions and growth restriction during pregnancy [227]. Finally, it has been described that inhibitors of epigenetics modulators such as GSK-126 that inhibits histone H3K27 methyltransferases, a suppressor of gene transcriptions, reduces ZIKV multiplication in cell culture by activating the cellular antiviral and immune responses [209].

3.3. Future Perspectives for ZIKV Antivirals.

As in the case of vaccine development, a great effort is being made to find compounds to fight ZIKV infection by applying different approaches, from the screening of bioactive molecules from different libraries to repurposing drugs with known antiviral activity and the use of natural products. However, it should be noted that many different methodologies, viral strains and cell types have been used, which sometimes has driven to contradictory results with the same compound. On top of that, most of the antivirals candidates have been tested in vitro, and only a few have been assayed in vivo, which hampers the interpretation of the results, as extrapolation of in vitro results to in vivo is usually difficult. In fact, in the occasions in which in vivo experiments have been conducted in animal models they have often failed to give the expected outcomes, and thus, this points to the complication for most of them to complete the entire drug development pipeline. In addition, as many of the tested candidates may have untoward effects, careful evaluation should be conducted before they can enter in the clinical practice, mainly when the main target population will be pregnant women and patients with other medical complications.

4. Conclusions

In as little as three years, a great number of vaccine and antiviral candidates has been tested in cell culture and animal models. Most of the assayed vaccines have produced hopeful results (potent specific humoral and cellular responses, decrease of viremia and viral RNA in natural target organs, prevention of vertical viral transmission, development of microcephaly, and protection in experimental animal models), and even some of them have entered in clinical trials. However, several questions remain to be deeply analyzed and resolved before any of these candidates can be implemented in the field. For instance, as other flaviviruses co-circulate with ZIKV in the same regions, avoiding adverse effects such as exacerbation of disease due to, for example, the induction of ADE is required. Likewise, and based on previous experiences with other flaviviruses vaccines, safety for use in pregnant women is mandatory. Even more, vaccines should not only protect from the ZIKV infection, but ideally also prevents replication in sexual organs and viral spread. The cost-effectivity of vaccine campaign implementation should also be considered. Similarly, and even though dozens of antiviral candidates (DAAs and HAAs) have been tested that affect different aspects of the viral life cycle, few of them have been assayed in animal models, and safety and untoward effects should be carefully evaluated before they can be used in clinical practice. Therefore, a long way is still ahead before therapeutic measures against ZIKV can be applied. However, the public health need for ZIKV vaccines and antivirals persists. Although the number of human cases have dropped a lot since 2017, investigations should go on, since, as flaviviruses share many biological characteristics, their implementation should not only be pertinent for ZIKV infections, but also for that of other currently circulating flaviviruses, and for those possible yet unrecognized, or almost neglected, flaviviruses that can emerge in the future.

Author Contributions

Conceptualization, writing and editing, J.-C.S.

Funding

This research was funded by the Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria, INIA (E-RTA2017-00003-C02-01).

Acknowledgments

The author thanks A.B. Blázquez and M.A. Martín-Acebes for critical reading.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Martin-Acebes, M.A.; Saiz, J.C. West nile virus: A re-emerging pathogen revisited. World J. Virol. 2012, 1, 51–70. [Google Scholar] [CrossRef]
  2. Weaver, S.C.; Forrester, N.L. Chikungunya: Evolutionary history and recent epidemic spread. Antivir. Res. 2015, 120, 32–39. [Google Scholar] [CrossRef]
  3. Kaner, J.; Schaack, S. Understanding ebola: The 2014 epidemic. Global Health 2016, 12, 53. [Google Scholar] [CrossRef]
  4. Saiz, J.C.; Vazquez-Calvo, A.; Blazquez, A.B.; Merino-Ramos, T.; Escribano-Romero, E.; Martin-Acebes, M.A. Zika virus: The latest newcomer. Front. Microbiol. 2016, 7, 496. [Google Scholar] [CrossRef]
  5. WHO. The History of Zika Virus. Available online: http://www.who.int/emergencies/zika-virus/history/en/ (accessed on 20 June 2019).
  6. Saiz, J.C.; Martin-Acebes, M.A.; Bueno-Mari, R.; Salomon, O.D.; Villamil-Jimenez, L.C.; Heukelbach, J.; Alencar, C.H.; Armstrong, P.K.; Ortiga-Carvalho, T.M.; Mendez-Otero, R.; et al. Zika virus: What have we learnt since the start of the recent epidemic? Front. Microbiol. 2017, 8, 1554. [Google Scholar] [CrossRef]
  7. Available online: http://www.who.int/emergencies/zika-virus/history/en/ (accessed on 20 June 2019).
  8. Available online: http://www.paho.org/hq/index.php?option=com_content&view=article&id=11599&Itemid=41691&lang=en (accessed on 20 June 2019).
  9. Kuno, G.; Chang, G.J. Full-length sequencing and genomic characterization of bagaza, kedougou, and zika viruses. Arch. Virol. 2007, 152, 687–696. [Google Scholar] [CrossRef]
  10. Duffy, M.R.; Chen, T.H.; Hancock, W.T.; Powers, A.M.; Kool, J.L.; Lanciotti, R.S.; Pretrick, M.; Marfel, M.; Holzbauer, S.; Dubray, C.; et al. Zika virus outbreak on yap island, federated states of micronesia. N. Engl. J. Med. 2009, 360, 2536–2543. [Google Scholar] [CrossRef]
  11. Blazquez, A.B.; Saiz, J.C. Neurological manifestations of zika virus infection. World J. Virol. 2016, 5, 135–143. [Google Scholar] [CrossRef]
  12. Kurscheidt, F.A.; Mesquita, C.S.S.; Damke, G.; Damke, E.; Carvalho, A.; Suehiro, T.T.; Teixeira, J.J.V.; da Silva, V.R.S.; Souza, R.P.; Consolaro, M.E.L. Persistence and clinical relevance of zika virus in the male genital tract. Nat. Rev. Urol. 2018, 16, 211–230. [Google Scholar] [CrossRef]
  13. Martin-Acebes, M.A.; Saiz, J.C. The scientific response to zika virus. J. Clin. Med. 2019, 8, 369. [Google Scholar] [CrossRef]
  14. Nazerai, L.; Pravsgaard Christensen, J.; Randrup Thomsen, A. A ‘furry-tale’ of zika virus infection: What have we learned from animal models? Viruses 2019, 11, 29. [Google Scholar] [CrossRef]
  15. Alves, M.P.; Vielle, N.J.; Thiel, V.; Pfaender, S. Research models and tools for the identification of antivirals and therapeutics against zika virus infection. Viruses 2018, 10, 593. [Google Scholar] [CrossRef]
  16. McArthur, M.A. Zika virus: Recent advances towards the development of vaccines and therapeutics. Viruses 2017, 9, 143. [Google Scholar] [CrossRef]
  17. Hasan, S.S.; Sevvana, M.; Kuhn, R.J.; Rossmann, M.G. Structural biology of zika virus and other flaviviruses. Nat. Struct. Mol. Biol. 2018, 25, 13–20. [Google Scholar] [CrossRef]
  18. Durham, D.P.; Fitzpatrick, M.C.; Ndeffo-Mbah, M.L.; Parpia, A.S.; Michael, N.L.; Galvani, A.P. Evaluating vaccination strategies for zika virus in the americas. Ann. Intern. Med. 2018, 168, 621–630. [Google Scholar] [CrossRef]
  19. Available online: https://www.who.int/immunization/research/development/WHO_UNICEF_Zikavac_TPP_Feb2017.pdf?ua=1 (accessed on 20 June 2019).
  20. Sumathy, K.; Kulkarni, B.; Gondu, R.K.; Ponnuru, S.K.; Bonguram, N.; Eligeti, R.; Gadiyaram, S.; Praturi, U.; Chougule, B.; Karunakaran, L.; et al. Protective efficacy of zika vaccine in ag129 mouse model. Sci. Rep. 2017, 7, 46375. [Google Scholar] [CrossRef]
  21. Available online: http://ctri.nic.in/Clinicaltrials/showallp.php?mid1=16798&EncHid=&userName=CTRI/2017/05/008539 (accessed on 20 June 2019).
  22. Abbink, P.; Larocca, R.A.; De La Barrera, R.A.; Bricault, C.A.; Moseley, E.T.; Boyd, M.; Kirilova, M.; Li, Z.; Ng’ang’a, D.; Nanayakkara, O.; et al. Protective efficacy of multiple vaccine platforms against zika virus challenge in rhesus monkeys. Science 2016, 353, 1129–1132. [Google Scholar] [CrossRef]
  23. Modjarrad, K.; Lin, L.; George, S.L.; Stephenson, K.E.; Eckels, K.H.; De La Barrera, R.A.; Jarman, R.G.; Sondergaard, E.; Tennant, J.; Ansel, J.L.; et al. Preliminary aggregate safety and immunogenicity results from three trials of a purified inactivated zika virus vaccine candidate: Phase 1, randomised, double-blind, placebo-controlled clinical trials. Lancet 2017, 391, 563–571. [Google Scholar] [CrossRef]
  24. Larocca, R.A.; Abbink, P.; Peron, J.P.; Zanotto, P.M.; Iampietro, M.J.; Badamchi-Zadeh, A.; Boyd, M.; Ng’ang’a, D.; Kirilova, M.; Nityanandam, R.; et al. Vaccine protection against zika virus from brazil. Nature 2016, 536, 474–478. [Google Scholar] [CrossRef]
  25. Muthumani, K.; Griffin, B.D.; Agarwal, S.; Kudchodkar, S.B.; Reuschel, E.L.; Choi, H.; Kraynyak, K.A.; Duperret, E.K.; Keaton, A.A.; Chung, C.; et al. In vivo protection against zikv infection and pathogenesis through passive antibody transfer and active immunisation with a prmenv DNA vaccine. NPJ Vaccines 2016, 1, 16021. [Google Scholar] [CrossRef]
  26. Tebas, P.; Roberts, C.C.; Muthumani, K.; Reuschel, E.L.; Kudchodkar, S.B.; Zaidi, F.I.; White, S.; Khan, A.S.; Racine, T.; Choi, H.; et al. Safety and immunogenicity of an anti-zika virus DNA vaccine—Preliminary report. N. Engl. J. Med. 2017. [Google Scholar] [CrossRef]
  27. Dowd, K.A.; Ko, S.Y.; Morabito, K.M.; Yang, E.S.; Pelc, R.S.; DeMaso, C.R.; Castilho, L.R.; Abbink, P.; Boyd, M.; Nityanandam, R.; et al. Rapid development of a DNA vaccine for zika virus. Science 2016, 354, 237–240. [Google Scholar] [CrossRef]
  28. Gaudinski, M.R.; Houser, K.V.; Morabito, K.M.; Hu, Z.; Yamshchikov, G.; Rothwell, R.S.; Berkowitz, N.; Mendoza, F.; Saunders, J.G.; Novik, L.; et al. Safety, tolerability, and immunogenicity of two zika virus DNA vaccine candidates in healthy adults: Randomised, open-label, phase 1 clinical trials. Lancet 2018, 391, 552–562. [Google Scholar] [CrossRef]
  29. Pardi, N.; Hogan, M.J.; Pelc, R.S.; Muramatsu, H.; Andersen, H.; DeMaso, C.R.; Dowd, K.A.; Sutherland, L.L.; Scearce, R.M.; Parks, R.; et al. Zika virus protection by a single low-dose nucleoside-modified mrna vaccination. Nature 2017, 543, 248–251. [Google Scholar] [CrossRef]
  30. Richner, J.M.; Jagger, B.W.; Shan, C.; Fontes, C.R.; Dowd, K.A.; Cao, B.; Himansu, S.; Caine, E.A.; Nunes, B.T.D.; Medeiros, D.B.A.; et al. Vaccine mediated protection against zika virus-induced congenital disease. Cell 2017, 170, 273–283. [Google Scholar] [CrossRef]
  31. Nurnberger, C.; Bodmer, B.S.; Fiedler, A.H.; Gabriel, G.; Muhlebach, M.D. A measles virus-based vaccine candidate mediates protection against zika virus in an allogeneic mouse pregnancy model. J. Virol. 2019, 93. [Google Scholar] [CrossRef]
  32. Baldwin, W.R.; Livengood, J.A.; Giebler, H.A.; Stovall, J.L.; Boroughs, K.L.; Sonnberg, S.; Bohning, K.J.; Dietrich, E.A.; Ong, Y.T.; Danh, H.K.; et al. Purified inactivated zika vaccine candidates afford protection against lethal challenge in mice. Sci. Rep. 2018, 8, 16509. [Google Scholar] [CrossRef]
  33. Shan, C.; Muruato, A.E.; Nunes, B.T.D.; Luo, H.; Xie, X.; Medeiros, D.B.A.; Wakamiya, M.; Tesh, R.B.; Barrett, A.D.; Wang, T.; et al. A live-attenuated zika virus vaccine candidate induces sterilizing immunity in mouse models. Nat. Med. 2017, 23, 763–767. [Google Scholar] [CrossRef]
  34. Shan, C.; Muruato, A.E.; Jagger, B.W.; Richner, J.; Nunes, B.T.D.; Medeiros, D.B.A.; Xie, X.; Nunes, J.G.C.; Morabito, K.M.; Kong, W.P.; et al. A single-dose live-attenuated vaccine prevents zika virus pregnancy transmission and testis damage. Nat. Commun. 2017, 8, 676. [Google Scholar] [CrossRef]
  35. Xie, X.; Kum, D.B.; Xia, H.; Luo, H.; Shan, C.; Zou, J.; Muruato, A.E.; Medeiros, D.B.A.; Nunes, B.T.D.; Dallmeier, K.; et al. A single-dose live-attenuated zika virus vaccine with controlled infection rounds that protects against vertical transmission. Cell Host Microbe 2018, 24, 487–499. [Google Scholar] [CrossRef]
  36. Richner, J.M.; Himansu, S.; Dowd, K.A.; Butler, S.L.; Salazar, V.; Fox, J.M.; Julander, J.G.; Tang, W.W.; Shresta, S.; Pierson, T.C.; et al. Modified mrna vaccines protect against zika virus infection. Cell 2017, 169, 176. [Google Scholar] [CrossRef]
  37. Wang, R.; Liao, X.; Fan, D.; Wang, L.; Song, J.; Feng, K.; Li, M.; Wang, P.; Chen, H.; An, J. Maternal immunization with a DNA vaccine candidate elicits specific passive protection against post-natal zika virus infection in immunocompetent balb/c mice. Vaccine 2018, 36, 3522–3532. [Google Scholar] [CrossRef]
  38. Yi, G.; Xu, X.; Abraham, S.; Petersen, S.; Guo, H.; Ortega, N.; Shankar, P.; Manjunath, N. A DNA vaccine protects human immune cells against zika virus infection in humanized mice. EBioMedicine 2017, 25, 87–94. [Google Scholar] [CrossRef]
  39. Zou, J.; Xie, X.; Luo, H.; Shan, C.; Muruato, A.E.; Weaver, S.C.; Wang, T.; Shi, P.Y. A single-dose plasmid-launched live-attenuated zika vaccine induces protective immunity. EBioMedicine 2018, 36, 92–102. [Google Scholar] [CrossRef]
  40. Chahal, J.S.; Fang, T.; Woodham, A.W.; Khan, O.F.; Ling, J.; Anderson, D.G.; Ploegh, H.L. An rna nanoparticle vaccine against zika virus elicits antibody and cd8+ t cell responses in a mouse model. Sci. Rep. 2017, 7, 252. [Google Scholar] [CrossRef]
  41. Boigard, H.; Alimova, A.; Martin, G.R.; Katz, A.; Gottlieb, P.; Galarza, J.M. Zika virus-like particle (vlp) based vaccine. PLoS Negl. Trop. Dis. 2017, 11, e0005608. [Google Scholar] [CrossRef]
  42. Garg, H.; Sedano, M.; Plata, G.; Punke, E.B.; Joshi, A. Development of virus-like-particle vaccine and reporter assay for zika virus. J. Virol. 2017, 91, e00834-17. [Google Scholar] [CrossRef]
  43. Salvo, M.A.; Kingstad-Bakke, B.; Salas-Quinchucua, C.; Camacho, E.; Osorio, J.E. Zika virus like particles elicit protective antibodies in mice. PLoS Negl. Trop. Dis. 2018, 12, e0006210. [Google Scholar] [CrossRef]
  44. Yang, M.; Lai, H.; Sun, H.; Chen, Q. Virus-like particles that display zika virus envelope protein domain iii induce potent neutralizing immune responses in mice. Sci. Rep. 2017, 7, 7679. [Google Scholar] [CrossRef]
  45. Liang, H.; Yang, R.; Liu, Z.; Li, M.; Liu, H.; Jin, X. Recombinant zika virus envelope protein elicited protective immunity against zika virus in immunocompetent mice. PLoS ONE 2018, 13, e0194860. [Google Scholar] [CrossRef]
  46. To, A.; Medina, L.O.; Mfuh, K.O.; Lieberman, M.M.; Wong, T.A.S.; Namekar, M.; Nakano, E.; Lai, C.Y.; Kumar, M.; Nerurkar, V.R.; et al. Recombinant zika virus subunits are immunogenic and efficacious in mice. mSphere 2018, 3. [Google Scholar] [CrossRef]
  47. Medina, L.O.; To, A.; Lieberman, M.M.; Wong, T.A.S.; Namekar, M.; Nakano, E.; Andersen, H.; Yalley-Ogunro, J.; Greenhouse, J.; Higgs, S.; et al. A recombinant subunit based zika virus vaccine is efficacious in non-human primates. Front. Immunol. 2018, 9, 2464. [Google Scholar] [CrossRef]
  48. Han, J.F.; Qiu, Y.; Yu, J.Y.; Wang, H.J.; Deng, Y.Q.; Li, X.F.; Zhao, H.; Sun, H.X.; Qin, C.F. Immunization with truncated envelope protein of zika virus induces protective immune response in mice. Sci. Rep. 2017, 7, 10047. [Google Scholar] [CrossRef]
  49. Qu, P.; Zhang, W.; Li, D.; Zhang, C.; Liu, Q.; Zhang, X.; Wang, X.; Dai, W.; Xu, Y.; Leng, Q.; et al. Insect cell-produced recombinant protein subunit vaccines protect against zika virus infection. Antivir. Res. 2018, 154, 97–103. [Google Scholar] [CrossRef]
  50. Zhu, X.; Li, C.; Afridi, S.K.; Zu, S.; Xu, J.W.; Quanquin, N.; Yang, H.; Cheng, G.; Xu, Z. E90 subunit vaccine protects mice from zika virus infection and microcephaly. Acta Neuropathol. Commun. 2018, 6, 77. [Google Scholar] [CrossRef]
  51. Yang, M.; Dent, M.; Lai, H.; Sun, H.; Chen, Q. Immunization of zika virus envelope protein domain iii induces specific and neutralizing immune responses against zika virus. Vaccine 2017, 35, 4287–4294. [Google Scholar] [CrossRef]
  52. Yang, M.; Sun, H.; Lai, H.; Hurtado, J.; Chen, Q. Plant-produced zika virus envelope protein elicits neutralizing immune responses that correlate with protective immunity against zika virus in mice. Plant Biotechnol. J. 2017, 16, 572–580. [Google Scholar] [CrossRef]
  53. Tai, W.; He, L.; Wang, Y.; Sun, S.; Zhao, G.; Luo, C.; Li, P.; Zhao, H.; Fremont, D.H.; Li, F.; et al. Critical neutralizing fragment of zika virus ediii elicits cross-neutralization and protection against divergent zika viruses. Emerg. Microbes Infect. 2018, 7, 7. [Google Scholar] [CrossRef]
  54. Xu, K.; Song, Y.; Dai, L.; Zhang, Y.; Lu, X.; Xie, Y.; Zhang, H.; Cheng, T.; Wang, Q.; Huang, Q.; et al. Recombinant chimpanzee adenovirus vaccine adc7-m/e protects against zika virus infection and testis damage. J. Virol. 2018, 92, e01722-17. [Google Scholar] [CrossRef]
  55. Bullard, B.L.; Corder, B.N.; Gorman, M.J.; Diamond, M.S.; Weaver, E.A. Efficacy of a t cell-biased adenovirus vector as a zika virus vaccine. Sci. Rep. 2018, 8, 18017. [Google Scholar] [CrossRef]
  56. Cox, F.; van der Fits, L.; Abbink, P.; Larocca, R.A.; van Huizen, E.; Saeland, E.; Verhagen, J.; Peterson, R.; Tolboom, J.; Kaufmann, B.; et al. Adenoviral vector type 26 encoding zika virus (zikv) m-env antigen induces humoral and cellular immune responses and protects mice and nonhuman primates against zikv challenge. PLoS ONE 2018, 13, e0202820. [Google Scholar] [CrossRef]
  57. Guo, Q.; Chan, J.F.; Poon, V.K.; Wu, S.; Chan, C.C.; Hou, L.; Yip, C.C.; Ren, C.; Cai, J.P.; Zhao, M.; et al. Immunization with a novel human type 5 adenovirus-vectored vaccine expressing the premembrane and envelope proteins of zika virus provides consistent and sterilizing protection in multiple immunocompetent and immunocompromised animal models. J. Infect. Dis. 2018, 218, 365–377. [Google Scholar] [CrossRef]
  58. Perez, P.; Marín, Q.M.; Lazaro-Frias, A.; Jimenez de Oya, N.; Blazquez, A.B.; Escribano-Romero, E.; CO, S.S.; Ortego, J.; Saiz, J.C.; Esteban, M.; et al. A vaccine based on a modified vaccinia virus ankara vector expressing zika virus structural proteins controls zika virus replication in mice. Sci. Rep. 2018, 8, 17385. [Google Scholar] [CrossRef]
  59. Brault, A.C.; Domi, A.; McDonald, E.M.; Talmi-Frank, D.; McCurley, N.; Basu, R.; Robinson, H.L.; Hellerstein, M.; Duggal, N.K.; Bowen, R.A.; et al. A zika vaccine targeting ns1 protein protects immunocompetent adult mice in a lethal challenge model. Sci. Rep. 2017, 7, 14769. [Google Scholar] [CrossRef]
  60. Prow, N.A.; Liu, L.; Nakayama, E.; Cooper, T.H.; Yan, K.; Eldi, P.; Hazlewood, J.E.; Tang, B.; Le, T.T.; Setoh, Y.X.; et al. A vaccinia-based single vector construct multi-pathogen vaccine protects against both zika and chikungunya viruses. Nat. Commun. 2018, 9, 1230. [Google Scholar] [CrossRef]
  61. Emanuel, J.; Callison, J.; Dowd, K.A.; Pierson, T.C.; Feldmann, H.; Marzi, A. A vsv-based zika virus vaccine protects mice from lethal challenge. Sci. Rep. 2018, 8, 11043. [Google Scholar] [CrossRef]
  62. Li, A.; Yu, J.; Lu, M.; Ma, Y.; Attia, Z.; Shan, C.; Xue, M.; Liang, X.; Craig, K.; Makadiya, N.; et al. A zika virus vaccine expressing premembrane-envelope-ns1 polyprotein. Nat. Commun. 2018, 9, 3067. [Google Scholar] [CrossRef]
  63. Chattopadhyay, A.; Aguilar, P.V.; Bopp, N.E.; Yarovinsky, T.O.; Weaver, S.C.; Rose, J.K. A recombinant virus vaccine that protects against both chikungunya and zika virus infections. Vaccine 2018, 36, 3894–3900. [Google Scholar] [CrossRef]
  64. Li, X.F.; Dong, H.L.; Wang, H.J.; Huang, X.Y.; Qiu, Y.F.; Ji, X.; Ye, Q.; Li, C.; Liu, Y.; Deng, Y.Q.; et al. Development of a chimeric zika vaccine using a licensed live-attenuated flavivirus vaccine as backbone. Nat. Commun. 2018, 9, 673. [Google Scholar] [CrossRef]
  65. Giel-Moloney, M.; Goncalvez, A.P.; Catalan, J.; Lecouturier, V.; Girerd-Chambaz, Y.; Diaz, F.; Maldonado-Arocho, F.; Gomila, R.C.; Bernard, M.C.; Oomen, R.; et al. Chimeric yellow fever 17d-zika virus (chimerivax-zika) as a live-attenuated zika virus vaccine. Sci. Rep. 2018, 8, 13206. [Google Scholar] [CrossRef]
  66. Kum, D.B.; Mishra, N.; Boudewijns, R.; Gladwyn-Ng, I.; Alfano, C.; Ma, J.; Schmid, M.A.; Marques, R.E.; Schols, D.; Kaptein, S.; et al. A yellow fever-zika chimeric virus vaccine candidate protects against zika infection and congenital malformations in mice. NPJ Vaccines 2018, 3, 56. [Google Scholar] [CrossRef]
  67. Touret, F.; Gilles, M.; Klitting, R.; Aubry, F.; de Lamballerie, X.; Nougairede, A. Live zika virus chimeric vaccine candidate based on a yellow fever 17-d attenuated backbone. Emerg. Microbes Infect. 2018, 7, 161. [Google Scholar] [CrossRef]
  68. Stettler, K.; Beltramello, M.; Espinosa, D.A.; Graham, V.; Cassotta, A.; Bianchi, S.; Vanzetta, F.; Minola, A.; Jaconi, S.; Mele, F.; et al. Specificity, cross-reactivity, and function of antibodies elicited by zika virus infection. Science 2016, 353, 823–826. [Google Scholar] [CrossRef]
  69. Abbink, P.; Larocca, R.A.; Dejnirattisai, W.; Peterson, R.; Nkolola, J.P.; Borducchi, E.N.; Supasa, P.; Mongkolsapaya, J.; Screaton, G.R.; Barouch, D.H. Therapeutic and protective efficacy of a dengue antibody against zika infection in rhesus monkeys. Nat. Med. 2018, 24, 721–723. [Google Scholar] [CrossRef]
  70. Wang, S.; Hong, S.; Deng, Y.Q.; Ye, Q.; Zhao, L.Z.; Zhang, F.C.; Qin, C.F.; Xu, Z. Transfer of convalescent serum to pregnant mice prevents zika virus infection and microcephaly in offspring. Cell Res. 2017, 27, 158–160. [Google Scholar] [CrossRef]
  71. Sapparapu, G.; Fernandez, E.; Kose, N.; Bin, C.; Fox, J.M.; Bombardi, R.G.; Zhao, H.; Nelson, C.A.; Bryan, A.L.; Barnes, T.; et al. Neutralizing human antibodies prevent zika virus replication and fetal disease in mice. Nature 2016, 540, 443–447. [Google Scholar] [CrossRef]
  72. Li, C.; Gao, F.; Yu, L.; Wang, R.; Jiang, Y.; Shi, X.; Yin, C.; Tang, X.; Zhang, F.; Xu, Z.; et al. A single injection of human neutralizing antibody protects against zika virus infection and microcephaly in developing mouse embryos. Cell Rep. 2018, 23, 1424–1434. [Google Scholar] [CrossRef]
  73. Magnani, D.M.; Rogers, T.F.; Beutler, N.; Ricciardi, M.J.; Bailey, V.K.; Gonzalez-Nieto, L.; Briney, B.; Sok, D.; Le, K.; Strubel, A.; et al. Neutralizing human monoclonal antibodies prevent zika virus infection in macaques. Sci. Transl. Med. 2017, 9. [Google Scholar] [CrossRef]
  74. Bailey, M.J.; Duehr, J.; Dulin, H.; Broecker, F.; Brown, J.A.; Arumemi, F.O.; Bermudez Gonzalez, M.C.; Leyva-Grado, V.H.; Evans, M.J.; Simon, V.; et al. Human antibodies targeting zika virus ns1 provide protection against disease in a mouse model. Nat. Commun. 2018, 9, 4560. [Google Scholar] [CrossRef]
  75. Diamond, M.S.; Ledgerwood, J.E.; Pierson, T.C. Zika virus vaccine development: Progress in the face of new challenges. Annu. Rev. Med. 2019, 70, 121–135. [Google Scholar] [CrossRef]
  76. Saiz, J.C.; Martin-Acebes, M.A. The race to find antivirals for zika virus. Antimicrob. Agents Chemother. 2017, 61. [Google Scholar] [CrossRef]
  77. Menendez-Arias, L.; Richman, D.D. Editorial overview: Antivirals and resistance: Advances and challenges ahead. Curr. Opin. Virol. 2014, 8, iv–vii. [Google Scholar] [CrossRef]
  78. Aliota, M.T.; Bassit, L.; Bradrick, S.S.; Cox, B.; Garcia-Blanco, M.A.; Gavegnano, C.; Friedrich, T.C.; Golos, T.G.; Griffin, D.E.; Haddow, A.D.; et al. Zika in the americas, year 2: What have we learned? What gaps remain? A report from the global virus network. Antivir. Res. 2017, 144, 223–246. [Google Scholar] [CrossRef]
  79. Wang, J.; Bardelli, M.; Espinosa, D.A.; Pedotti, M.; Ng, T.S.; Bianchi, S.; Simonelli, L.; Lim, E.X.Y.; Foglierini, M.; Zatta, F.; et al. A human bi-specific antibody against zika virus with high therapeutic potential. Cell 2017, 171, 229–241. [Google Scholar] [CrossRef]
  80. Yu, Y.; Deng, Y.Q.; Zou, P.; Wang, Q.; Dai, Y.; Yu, F.; Du, L.; Zhang, N.N.; Tian, M.; Hao, J.N.; et al. A peptide-based viral inactivator inhibits zika virus infection in pregnant mice and fetuses. Nat. Commun. 2017, 8, 15672. [Google Scholar] [CrossRef]
  81. Neris, R.L.S.; Figueiredo, C.M.; Higa, L.M.; Araujo, D.F.; Carvalho, C.A.M.; Vercoza, B.R.F.; Silva, M.O.L.; Carneiro, F.A.; Tanuri, A.; Gomes, A.M.O.; et al. Co-protoporphyrin ix and sn-protoporphyrin ix inactivate zika, chikungunya and other arboviruses by targeting the viral envelope. Sci. Rep. 2018, 8, 9805. [Google Scholar] [CrossRef]
  82. Huang, S.; Gu, J.; Ye, J.; Fang, B.; Wan, S.; Wang, C.; Ashraf, U.; Li, Q.; Wang, X.; Shao, L.; et al. Benzoxazine monomer derived carbon dots as a broad-spectrum agent to block viral infectivity. J. Colloid Interface Sci. 2019, 542, 198–206. [Google Scholar] [CrossRef]
  83. Francese, R.; Civra, A.; Ritta, M.; Donalisio, M.; Argenziano, M.; Cavalli, R.; Mougharbel, A.S.; Kortz, U.; Lembo, D. Anti-zika virus activity of polyoxometalates. Antivir. Res. 2019, 163, 29–33. [Google Scholar] [CrossRef]
  84. Li, P.C.; Jang, J.; Hsia, C.Y.; Groomes, P.V.; Lian, W.; de Wispelaere, M.; Pitts, J.D.; Wang, J.; Kwiatkowski, N.; Gray, N.S.; et al. Small molecules targeting the flavivirus e protein with broad-spectrum activity and antiviral efficacy in vivo. ACS Infect. Dis. 2019, 5, 460–472. [Google Scholar] [CrossRef]
  85. Palanichamy, K.; Joshi, A.; Mehmetoglu-Gurbuz, T.; Bravo, M.F.; Shlain, M.A.; Schiro, F.; Naeem, Y.; Garg, H.; Braunschweig, A.B. Anti-zika activity of a library of synthetic carbohydrate receptors. J. Med. Chem. 2019, 62, 4110–4119. [Google Scholar] [CrossRef]
  86. de Wispelaere, M.; Lian, W.; Potisopon, S.; Li, P.C.; Jang, J.; Ficarro, S.B.; Clark, M.J.; Zhu, X.; Kaplan, J.B.; Pitts, J.D.; et al. Inhibition of flaviviruses by targeting a conserved pocket on the viral envelope protein. Cell Chem. Biol. 2018, 25, 1006–1016. [Google Scholar] [CrossRef]
  87. Lian, W.; Jang, J.; Potisopon, S.; Li, P.C.; Rahmeh, A.; Wang, J.; Kwiatkowski, N.P.; Gray, N.S.; Yang, P.L. Discovery of immunologically inspired small molecules that target the viral envelope protein. ACS Infect. Dis. 2018, 4, 1395–1406. [Google Scholar] [CrossRef]
  88. Pitts, J.; Hsia, C.Y.; Lian, W.; Wang, J.; Pfeil, M.P.; Kwiatkowski, N.; Li, Z.; Jang, J.; Gray, N.S.; Yang, P.L. Identification of small molecule inhibitors targeting the zika virus envelope protein. Antivir. Res. 2019, 164, 147–153. [Google Scholar] [CrossRef]
  89. Mazzon, M.; Ortega-Prieto, A.M.; Imrie, D.; Luft, C.; Hess, L.; Czieso, S.; Grove, J.; Skelton, J.K.; Farleigh, L.; Bugert, J.J.; et al. Identification of broad-spectrum antiviral compounds by targeting viral entry. Viruses 2019, 11, 176. [Google Scholar] [CrossRef]
  90. Russo, R.R.; Dos Santos Junior, N.N.; Cintra, A.C.O.; Figueiredo, L.T.M.; Sampaio, S.V.; Aquino, V.H. Expression, purification and virucidal activity of two recombinant isoforms of phospholipase a2 from crotalus durissus terrificus venom. Arch. Virol. 2019, 164, 1159–1171. [Google Scholar] [CrossRef]
  91. Clain, E.; Haddad, J.G.; Koishi, A.C.; Sinigaglia, L.; Rachidi, W.; Despres, P.; Duarte Dos Santos, C.N.; Guiraud, P.; Jouvenet, N.; El Kalamouni, C. The polyphenol-rich extract from psiloxylon mauritianum, an endemic medicinal plant from reunion island, inhibits the early stages of dengue and zika virus infection. Int. J. Mol. Sci. 2019, 20, 1860. [Google Scholar] [CrossRef]
  92. Clain, E.; Sinigaglia, L.; Koishi, A.C.; Gorgette, O.; Gadea, G.; Viranaicken, W.; Krejbich-Trotot, P.; Mavingui, P.; Despres, P.; Nunes Duarte Dos Santos, C.; et al. Extract from aphloia theiformis, an edible indigenous plant from reunion island, impairs zika virus attachment to the host cell surface. Sci. Rep. 2018, 8, 10856. [Google Scholar] [CrossRef]
  93. Carneiro, B.M.; Batista, M.N.; Braga, A.C.S.; Nogueira, M.L.; Rahal, P. The green tea molecule egcg inhibits zika virus entry. Virology 2016, 496, 215–218. [Google Scholar] [CrossRef]
  94. Sharma, N.; Murali, A.; Singh, S.K.; Giri, R. Epigallocatechin gallate, an active green tea compound inhibits the zika virus entry into host cells via binding the envelope protein. Int. J. Biol. Macromol. 2017, 104, 1046–1054. [Google Scholar] [CrossRef]
  95. Vazquez-Calvo, A.; Jimenez de Oya, N.; Martin-Acebes, M.A.; Garcia-Moruno, E.; Saiz, J.C. Antiviral properties of the natural polyphenols delphinidin and epigallocatechin gallate against the flaviviruses west nile virus, zika virus, and dengue virus. Front. Microbiol. 2017, 8, 1314. [Google Scholar] [CrossRef]
  96. Lee, J.L.; Loe, M.W.C.; Lee, R.C.H.; Chu, J.J.H. Antiviral activity of pinocembrin against zika virus replication. Antivir. Res. 2019, 167, 13–24. [Google Scholar] [CrossRef]
  97. Bassi, M.R.; Sempere, R.N.; Meyn, P.; Polacek, C.; Arias, A. Extinction of zika virus and usutu virus by lethal mutagenesis reveals different patterns of sensitivity to three mutagenic drugs. Antimicrob. Agents Chemother. 2018, 62, e00380-18. [Google Scholar] [CrossRef]
  98. Baz, M.; Goyette, N.; Griffin, B.D.; Kobinger, G.P.; Boivin, G. In vitro susceptibility of geographically and temporally distinct zika viruses to favipiravir and ribavirin. Antivir. Ther. 2017, 22, 613–618. [Google Scholar] [CrossRef]
  99. Kamiyama, N.; Soma, R.; Hidano, S.; Watanabe, K.; Umekita, H.; Fukuda, C.; Noguchi, K.; Gendo, Y.; Ozaki, T.; Sonoda, A.; et al. Ribavirin inhibits zika virus (zikv) replication in vitro and suppresses viremia in zikv-infected stat1-deficient mice. Antivir. Res. 2017, 146, 1–11. [Google Scholar] [CrossRef]
  100. Kim, J.A.; Seong, R.K.; Kumar, M.; Shin, O.S. Favipiravir and ribavirin inhibit replication of asian and african strains of zika virus in different cell models. Viruses 2018, 10, 72. [Google Scholar] [CrossRef]
  101. Cai, L.; Sun, Y.; Song, Y.; Xu, L.; Bei, Z.; Zhang, D.; Dou, Y.; Wang, H. Viral polymerase inhibitors t-705 and t-1105 are potential inhibitors of zika virus replication. Arch. Virol. 2017, 162, 2847–2853. [Google Scholar] [CrossRef]
  102. Pires de Mello, C.P.; Tao, X.; Kim, T.H.; Bulitta, J.B.; Rodriquez, J.L.; Pomeroy, J.J.; Brown, A.N. Zika virus replication is substantially inhibited by novel favipiravir and interferon alpha combination regimens. Antimicrob. Agents Chemother. 2017, 62, e01983-17. [Google Scholar] [CrossRef]
  103. Tong, X.; Smith, J.; Bukreyeva, N.; Koma, T.; Manning, J.T.; Kalkeri, R.; Kwong, A.D.; Paessler, S. Merimepodib, an impdh inhibitor, suppresses replication of zika virus and other emerging viral pathogens. Antivir. Res. 2018, 149, 34–40. [Google Scholar] [CrossRef]
  104. Barrows, N.J.; Campos, R.K.; Powell, S.T.; Prasanth, K.R.; Schott-Lerner, G.; Soto-Acosta, R.; Galarza-Munoz, G.; McGrath, E.L.; Urrabaz-Garza, R.; Gao, J.; et al. A screen of fda-approved drugs for inhibitors of zika virus infection. Cell Host Microbe 2016, 20, 259–270. [Google Scholar] [CrossRef]
  105. Eyer, L.; Nencka, R.; Huvarova, I.; Palus, M.; Joao Alves, M.; Gould, E.A.; De Clercq, E.; Ruzek, D. Nucleoside inhibitors of zika virus. J. Infect. Dis. 2016, 214, 707–711. [Google Scholar] [CrossRef]
  106. Goebel, S.; Snyder, B.; Sellati, T.; Saeed, M.; Ptak, R.; Murray, M.; Bostwick, R.; Rayner, J.; Koide, F.; Kalkeri, R. A sensitive virus yield assay for evaluation of antivirals against zika virus. J. Virol. Methods 2016, 238, 13–20. [Google Scholar] [CrossRef]
  107. Rausch, K.; Hackett, B.A.; Weinbren, N.L.; Reeder, S.M.; Sadovsky, Y.; Hunter, C.A.; Schultz, D.C.; Coyne, C.B.; Cherry, S. Screening bioactives reveals nanchangmycin as a broad spectrum antiviral active against zika virus. Cell Rep. 2017, 18, 804–815. [Google Scholar] [CrossRef]
  108. Adcock, R.S.; Chu, Y.K.; Golden, J.E.; Chung, D.H. Evaluation of anti-zika virus activities of broad-spectrum antivirals and nih clinical collection compounds using a cell-based, high-throughput screen assay. Antivir. Res. 2017, 138, 47–56. [Google Scholar] [CrossRef]
  109. Beck, S.; Zhu, Z.; Oliveira, M.F.; Smith, D.M.; Rich, J.N.; Bernatchez, J.A.; Siqueira-Neto, J.L. Mechanism of action of methotrexate against zika virus. Viruses 2019, 11, 338. [Google Scholar] [CrossRef]
  110. Cifuentes Kottkamp, A.; De Jesus, E.; Grande, R.; Brown, J.A.; Jacobs, A.R.; Lim, J.K.; Stapleford, K.A. Atovaquone inhibits arbovirus replication through the depletion of intracellular nucleotides. J. Virol. 2019, 93, e00389-19. [Google Scholar] [CrossRef]
  111. de Carvalho, O.V.; Felix, D.M.; de Mendonca, L.R.; de Araujo, C.; de Oliveira Franca, R.F.; Cordeiro, M.T.; Silva Junior, A.; Pena, L.J. The thiopurine nucleoside analogue 6-methylmercaptopurine riboside (6mmpr) effectively blocks zika virus replication. Int. J. Antimicrob. Agents 2017, 50, 718–725. [Google Scholar] [CrossRef]
  112. Hercik, K.; Kozak, J.; Sala, M.; Dejmek, M.; Hrebabecky, H.; Zbornikova, E.; Smola, M.; Ruzek, D.; Nencka, R.; Boura, E. Adenosine triphosphate analogs can efficiently inhibit the zika virus rna-dependent rna polymerase. Antivir. Res. 2016, 137, 131–133. [Google Scholar] [CrossRef]
  113. Zmurko, J.; Marques, R.E.; Schols, D.; Verbeken, E.; Kaptein, S.J.; Neyts, J. The viral polymerase inhibitor 7-deaza-2’-c-methyladenosine is a potent inhibitor of in vitro zika virus replication and delays disease progression in a robust mouse infection model. PLoS Negl. Trop. Dis. 2016, 10, e0004695. [Google Scholar] [CrossRef]
  114. Deng, Y.Q.; Zhang, N.N.; Li, C.F.; Tian, M.; Hao, J.N.; Xie, X.P.; Shi, P.Y.; Qin, C.F. Adenosine analog nitd008 is a potent inhibitor of zika virus. Open Forum. Infect. Dis. 2016, 3, ofw175. [Google Scholar] [CrossRef]
  115. Julander, J.G.; Siddharthan, V.; Evans, J.; Taylor, R.; Tolbert, K.; Apuli, C.; Stewart, J.; Collins, P.; Gebre, M.; Neilson, S.; et al. Efficacy of the broad-spectrum antiviral compound bcx4430 against zika virus in cell culture and in a mouse model. Antivir. Res. 2017, 137, 14–22. [Google Scholar] [CrossRef]
  116. Chung, D.H.; Golden, J.E.; Adcock, R.S.; Schroeder, C.E.; Chu, Y.K.; Sotsky, J.B.; Cramer, D.E.; Chilton, P.M.; Song, C.; Anantpadma, M.; et al. Discovery of a broad-spectrum antiviral compound that inhibits pyrimidine biosynthesis and establishes a type 1 interferon-independent antiviral state. Antimicrob. Agents Chemother. 2016, 60, 4552–4562. [Google Scholar] [CrossRef]
  117. Kuivanen, S.; Bespalov, M.M.; Nandania, J.; Ianevski, A.; Velagapudi, V.; De Brabander, J.K.; Kainov, D.E.; Vapalahti, O. Obatoclax, saliphenylhalamide and gemcitabine inhibit zika virus infection in vitro and differentially affect cellular signaling, transcription and metabolism. Antivir. Res. 2017, 139, 117–128. [Google Scholar] [CrossRef]
  118. Bernatchez, J.A.; Coste, M.; Beck, S.; Wells, G.A.; Luna, L.A.; Clark, A.E.; Zhu, Z.; Hecht, D.; Rich, J.N.; Sohl, C.D.; et al. Activity of selected nucleoside analogue protides against zika virus in human neural stem cells. Viruses 2019, 11, 365. [Google Scholar] [CrossRef]
  119. Pascoalino, B.S.; Courtemanche, G.; Cordeiro, M.T.; Gil, L.H.; Freitas-Junior, L. Zika antiviral chemotherapy: Identification of drugs and promising starting points for drug discovery from an fda-approved library. F1000Research 2016, 5, 2523. [Google Scholar] [CrossRef]
  120. Yang, C.F.; Gopula, B.; Liang, J.J.; Li, J.K.; Chen, S.Y.; Lee, Y.L.; Chen, C.S.; Lin, Y.L. Novel ar-12 derivatives, p12-23 and p12-34, inhibit flavivirus replication by blocking host de novo pyrimidine biosynthesis. Emerg. Microbes Infect. 2018, 7, 187. [Google Scholar] [CrossRef]
  121. Bullard-Feibelman, K.M.; Govero, J.; Zhu, Z.; Salazar, V.; Veselinovic, M.; Diamond, M.S.; Geiss, B.J. The fda-approved drug sofosbuvir inhibits zika virus infection. Antivir. Res. 2017, 137, 134–140. [Google Scholar] [CrossRef]
  122. Sacramento, C.Q.; de Melo, G.R.; de Freitas, C.S.; Rocha, N.; Hoelz, L.V.; Miranda, M.; Fintelman-Rodrigues, N.; Marttorelli, A.; Ferreira, A.C.; Barbosa-Lima, G.; et al. The clinically approved antiviral drug sofosbuvir inhibits zika virus replication. Sci. Rep. 2017, 7, 40920. [Google Scholar] [CrossRef]
  123. Panayiotou, C.; Lindqvist, R.; Kurhade, C.; Vonderstein, K.; Pasto, J.; Edlund, K.; Upadhyay, A.S.; Overby, A.K. Viperin restricts zika virus and tick-borne encephalitis virus replication by targeting ns3 for proteasomal degradation. J. Virol. 2018, 92, e02054-17. [Google Scholar] [CrossRef]
  124. Gizzi, A.S.; Grove, T.L.; Arnold, J.J.; Jose, J.; Jangra, R.K.; Garforth, S.J.; Du, Q.; Cahill, S.M.; Dulyaninova, N.G.; Love, J.D.; et al. A naturally occurring antiviral ribonucleotide encoded by the human genome. Nature 2018, 558, 610–614. [Google Scholar] [CrossRef]
  125. Park, J.G.; Avila-Perez, G.; Madere, F.; Hilimire, T.A.; Nogales, A.; Almazan, F.; Martinez-Sobrido, L. Potent inhibition of zika virus replication by aurintricarboxylic acid. Front. Microbiol. 2019, 10, 718. [Google Scholar] [CrossRef]
  126. Stephen, P.; Baz, M.; Boivin, G.; Lin, S.X. Structural insight into ns5 of zika virus leading to the discovery of mtase inhibitors. J. Am. Chem. Soc. 2016, 138, 16212–16215. [Google Scholar] [CrossRef]
  127. Zhang, C.; Feng, T.; Cheng, J.; Li, Y.; Yin, X.; Zeng, W.; Jin, X.; Guo, F.; Jin, T. Structure of the ns5 methyltransferase from zika virus and implications in inhibitor design. Biochem. Biophys. Res. Commun. 2017, 492, 624–630. [Google Scholar] [CrossRef]
  128. Coutard, B.; Barral, K.; Lichiere, J.; Selisko, B.; Martin, B.; Aouadi, W.; Lombardia, M.O.; Debart, F.; Vasseur, J.J.; Guillemot, J.C.; et al. Zika virus methyltransferase: Structure and functions for drug design perspectives. J. Virol. 2017, 91, e02202-16. [Google Scholar] [CrossRef]
  129. Lim, H.J.; Nguyen, T.T.; Kim, N.M.; Park, J.S.; Jang, T.S.; Kim, D. Inhibitory effect of flavonoids against ns2b-ns3 protease of zika virus and their structure activity relationship. Biotechnol. Lett. 2016, 39, 415–421. [Google Scholar] [CrossRef]
  130. Roy, A.; Lim, L.; Srivastava, S.; Lu, Y.; Song, J. Solution conformations of zika ns2b-ns3pro and its inhibition by natural products from edible plants. PLoS ONE 2017, 12, e0180632. [Google Scholar] [CrossRef]
  131. Albulescu, I.C.; Kovacikova, K.; Tas, A.; Snijder, E.J.; van Hemert, M.J. Suramin inhibits zika virus replication by interfering with virus attachment and release of infectious particles. Antivir. Res. 2017, 143, 230–236. [Google Scholar] [CrossRef]
  132. Tan, C.W.; Sam, I.C.; Chong, W.L.; Lee, V.S.; Chan, Y.F. Polysulfonate suramin inhibits zika virus infection. Antivir. Res. 2017, 143, 186–194. [Google Scholar] [CrossRef]
  133. Coronado, M.A.; Eberle, R.J.; Bleffert, N.; Feuerstein, S.; Olivier, D.S.; de Moraes, F.R.; Willbold, D.; Arni, R.K. Zika virus ns2b/ns3 proteinase: A new target for an old drug - suramin a lead compound for ns2b/ns3 proteinase inhibition. Antivir. Res. 2018, 160, 118–125. [Google Scholar] [CrossRef]
  134. Li, Z.; Brecher, M.; Deng, Y.Q.; Zhang, J.; Sakamuru, S.; Liu, B.; Huang, R.; Koetzner, C.A.; Allen, C.A.; Jones, S.A.; et al. Existing drugs as broad-spectrum and potent inhibitors for zika virus by targeting ns2b-ns3 interaction. Cell Res. 2017, 27, 1046–1064. [Google Scholar] [CrossRef]
  135. Yuan, S.; Chan, J.F.; den-Haan, H.; Chik, K.K.; Zhang, A.J.; Chan, C.C.; Poon, V.K.; Yip, C.C.; Mak, W.W.; Zhu, Z.; et al. 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. Antivir. Res. 2017, 145, 33–43. [Google Scholar] [CrossRef]
  136. Chan, J.F.; Chik, K.K.; Yuan, S.; Yip, C.C.; Zhu, Z.; Tee, K.M.; Tsang, J.O.; Chan, C.C.; Poon, V.K.; Lu, G.; et al. Novel antiviral activity and mechanism of bromocriptine as a zika virus ns2b-ns3 protease inhibitor. Antivir. Res. 2017, 141, 29–37. [Google Scholar] [CrossRef]
  137. Lee, H.; Ren, J.; Nocadello, S.; Rice, A.J.; Ojeda, I.; Light, S.; Minasov, G.; Vargas, J.; Nagarathnam, D.; Anderson, W.F.; et al. Identification of novel small molecule inhibitors against ns2b/ns3 serine protease from zika virus. Antivir. Res. 2016, 139, 49–58. [Google Scholar] [CrossRef]
  138. Shiryaev, S.A.; Mesci, P.; Pinto, A.; Fernandes, I.; Sheets, N.; Shresta, S.; Farhy, C.; Huang, C.T.; Strongin, A.Y.; Muotri, A.R.; et al. Repurposing of the anti-malaria drug chloroquine for zika virus treatment and prophylaxis. Sci. Rep. 2017, 7, 15771. [Google Scholar] [CrossRef]
  139. Aleshin, A.E.; Shiryaev, S.A.; Strongin, A.Y.; Liddington, R.C. Structural evidence for regulation and specificity of flaviviral proteases and evolution of the flaviviridae fold. Protein Sci. 2007, 16, 795–806. [Google Scholar] [CrossRef]
  140. Noble, C.G.; Seh, C.C.; Chao, A.T.; Shi, P.Y. Ligand-bound structures of the dengue virus protease reveal the active conformation. J. Virol. 2011, 86, 438–446. [Google Scholar] [CrossRef]
  141. Brecher, M.; Li, Z.; Liu, B.; Zhang, J.; Koetzner, C.A.; Alifarag, A.; Jones, S.A.; Lin, Q.; Kramer, L.D.; Li, H. A conformational switch high-throughput screening assay and allosteric inhibition of the flavivirus ns2b-ns3 protease. PLoS Pathog. 2017, 13, e1006411. [Google Scholar] [CrossRef]
  142. Li, Z.; Sakamuru, S.; Huang, R.; Brecher, M.; Koetzner, C.A.; Zhang, J.; Chen, H.; Qin, C.F.; Zhang, Q.Y.; Zhou, J.; et al. Erythrosin b is a potent and broad-spectrum orthosteric inhibitor of the flavivirus ns2b-ns3 protease. Antivir. Res. 2017, 150, 217–225. [Google Scholar] [CrossRef]
  143. De Clercq, E.; Neyts, J. Antiviral agents acting as DNA or rna chain terminators. Handb. Exp. Pharmacol. 2009, 53–84. [Google Scholar] [CrossRef]
  144. Snell, N.J. Ribavirin--current status of a broad spectrum antiviral agent. Expert Opin. Pharmacother. 2001, 2, 1317–1324. [Google Scholar] [CrossRef]
  145. Lucas-Hourani, M.; Dauzonne, D.; Jorda, P.; Cousin, G.; Lupan, A.; Helynck, O.; Caignard, G.; Janvier, G.; Andre-Leroux, G.; Khiar, S.; et al. Inhibition of pyrimidine biosynthesis pathway suppresses viral growth through innate immunity. PLoS Pathog. 2013, 9, e1003678. [Google Scholar] [CrossRef]
  146. Cao, X.; Li, Y.; Jin, X.; Guo, F.; Jin, T. Molecular mechanism of divalent-metal-induced activation of ns3 helicase and insights into zika virus inhibitor design. Nucleic Acids Res. 2016, 44, 10505–10514. [Google Scholar] [CrossRef]
  147. Rut, W.; Zhang, L.; Kasperkiewicz, P.; Poreba, M.; Hilgenfeld, R.; Drag, M. Extended substrate specificity and first potent irreversible inhibitor/activity-based probe design for zika virus ns2b-ns3 protease. Antivir. Res. 2016, 139, 88–94. [Google Scholar] [CrossRef]
  148. Sahoo, M.; Jena, L.; Daf, S.; Kumar, S. Virtual screening for potential inhibitors of ns3 protein of zika virus. Genom. Inform. 2016, 14, 104–111. [Google Scholar] [CrossRef]
  149. Mottin, M.; Borba, J.; Braga, R.C.; Torres, P.H.M.; Martini, M.C.; Proenca-Modena, J.L.; Judice, C.C.; Costa, F.T.M.; Ekins, S.; Perryman, A.L.; et al. The a-z of zika drug discovery. Drug Discov. Today 2018, 23, 1833–1847. [Google Scholar] [CrossRef]
  150. Fernandez-Garcia, M.D.; Mazzon, M.; Jacobs, M.; Amara, A. Pathogenesis of flavivirus infections: Using and abusing the host cell. Cell Host Microbe 2009, 5, 318–328. [Google Scholar] [CrossRef]
  151. Boldescu, V.; Behnam, M.A.M.; Vasilakis, N.; Klein, C.D. Broad-spectrum agents for flaviviral infections: Dengue, zika and beyond. Nat. Rev. Drug Discov. 2017, 16, 565–586. [Google Scholar] [CrossRef]
  152. Saiz, J.C.; Oya, N.J.; Blazquez, A.B.; Escribano-Romero, E.; Martin-Acebes, M.A. Host-directed antivirals: A realistic alternative to fight zika virus. Viruses 2018, 10, 453. [Google Scholar] [CrossRef]
  153. Wells, M.F.; Salick, M.R.; Wiskow, O.; Ho, D.J.; Worringer, K.A.; Ihry, R.J.; Kommineni, S.; Bilican, B.; Klim, J.R.; Hill, E.J.; et al. Genetic ablation of axl does not protect human neural progenitor cells and cerebral organoids from zika virus infection. Cell Stem. Cell 2016, 19, 703–708. [Google Scholar] [CrossRef]
  154. Meertens, L.; Labeau, A.; Dejarnac, O.; Cipriani, S.; Sinigaglia, L.; Bonnet-Madin, L.; Le Charpentier, T.; Hafirassou, M.L.; Zamborlini, A.; Cao-Lormeau, V.M.; et al. Axl mediates zika virus entry in human glial cells and modulates innate immune responses. Cell Rep. 2017, 18, 324–333. [Google Scholar] [CrossRef]
  155. Mounce, B.C.; Cesaro, T.; Carrau, L.; Vallet, T.; Vignuzzi, M. Curcumin inhibits zika and chikungunya virus infection by inhibiting cell binding. Antivir. Res. 2017, 142, 148–157. [Google Scholar] [CrossRef]
  156. Pan, T.; Peng, Z.; Tan, L.; Zou, F.; Zhou, N.; Liu, B.; Liang, L.; Chen, C.; Liu, J.; Wu, L.; et al. Nonsteroidal anti-inflammatory drugs potently inhibit the replication of zika viruses by inducing the degradation of axl. J. Virol. 2018, 92, e01018-18. [Google Scholar] [CrossRef]
  157. Rocker, A.E.; Muller, J.A.; Dietzel, E.; Harms, M.; Kruger, F.; Heid, C.; Sowislok, A.; Riber, C.F.; Kupke, A.; Lippold, S.; et al. The molecular tweezer clr01 inhibits ebola and zika virus infection. Antivir. Res. 2018, 152, 26–35. [Google Scholar] [CrossRef]
  158. Jackman, J.A.; Costa, V.V.; Park, S.; Real, A.; Park, J.H.; Cardozo, P.L.; Ferhan, A.R.; Olmo, I.G.; Moreira, T.P.; Bambirra, J.L.; et al. Therapeutic treatment of zika virus infection using a brain-penetrating antiviral peptide. Nat. Mater. 2018, 17, 971–977. [Google Scholar] [CrossRef]
  159. Haviernik, J.; Stefanik, M.; Fojtikova, M.; Kali, S.; Tordo, N.; Rudolf, I.; Hubalek, Z.; Eyer, L.; Ruzek, D. Arbidol (umifenovir): A broad-spectrum antiviral drug that inhibits medically important arthropod-borne flaviviruses. Viruses 2018, 10, 184. [Google Scholar] [CrossRef]
  160. Cannalire, R.; Tarantino, D.; Piorkowski, G.; Carletti, T.; Massari, S.; Felicetti, T.; Barreca, M.L.; Sabatini, S.; Tabarrini, O.; Marcello, A.; et al. Broad spectrum anti-flavivirus pyridobenzothiazolones leading to less infective virions. Antivir. Res. 2019, 167, 6–12. [Google Scholar] [CrossRef]
  161. Persaud, M.; Martinez-Lopez, A.; Buffone, C.; Porcelli, S.A.; Diaz-Griffero, F. Infection by zika viruses requires the transmembrane protein axl, endocytosis and low ph. Virology 2018, 518, 301–312. [Google Scholar] [CrossRef]
  162. Varghese, F.S.; Rausalu, K.; Hakanen, M.; Saul, S.; Kummerer, B.M.; Susi, P.; Merits, A.; Ahola, T. Obatoclax inhibits alphavirus membrane fusion by neutralizing the acidic environment of endocytic compartments. Antimicrob. Agents Chemother. 2017, 61, e02227-16. [Google Scholar] [CrossRef]
  163. Han, Y.; Mesplede, T.; Xu, H.; Quan, Y.; Wainberg, M.A. The antimalarial drug amodiaquine possesses anti-zika virus activities. J. Med. Virol. 2018, 90, 796–802. [Google Scholar] [CrossRef]
  164. Li, F.; Lang, Y.; Ji, Z.; Xia, Z.; Han, Y.; Cheng, Y.; Liu, G.; Sun, F.; Zhao, Y.; Gao, M.; et al. A scorpion venom peptide ev37 restricts viral late entry by alkalizing acidic organelles. J. Biol. Chem. 2019, 294, 182–194. [Google Scholar] [CrossRef]
  165. Li, C.; Zhu, X.; Ji, X.; Quanquin, N.; Deng, Y.Q.; Tian, M.; Aliyari, R.; Zuo, X.; Yuan, L.; Afridi, S.K.; et al. Chloroquine, a fda-approved drug, prevents zika virus infection and its associated congenital microcephaly in mice. EBioMedicine 2017, 24, 189–194. [Google Scholar] [CrossRef]
  166. Delvecchio, R.; Higa, L.M.; Pezzuto, P.; Valadao, A.L.; Garcez, P.P.; Monteiro, F.L.; Loiola, E.C.; Dias, A.A.; Silva, F.J.; Aliota, M.T.; et al. Chloroquine, an endocytosis blocking agent, inhibits zika virus infection in different cell models. Viruses 2016, 8, 322. [Google Scholar] [CrossRef]
  167. Cao, B.; Parnell, L.A.; Diamond, M.S.; Mysorekar, I.U. Inhibition of autophagy limits vertical transmission of zika virus in pregnant mice. J. Exp. Med. 2017, 214, 2303–2313. [Google Scholar] [CrossRef]
  168. Balasubramanian, A.; Teramoto, T.; Kulkarni, A.A.; Bhattacharjee, A.K.; Padmanabhan, R. Antiviral activities of selected antimalarials against dengue virus type 2 and zika virus. Antivir. Res. 2017, 137, 141–150. [Google Scholar] [CrossRef]
  169. Garcia-Nicolas, O.; V’Kovski, P.; Vielle, N.J.; Ebert, N.; Zust, R.; Portmann, J.; Stalder, H.; Gaschen, V.; Vieyres, G.; Stoffel, M.; et al. The small-compound inhibitor k22 displays broad antiviral activity against different members of the family flaviviridae and offers potential as a panviral inhibitor. Antimicrob. Agents Chemother. 2018, 62, e01206-18. [Google Scholar] [CrossRef]
  170. Wang, H.; Li, Z.; Niu, J.; Xu, Y.; Ma, L.; Lu, A.; Wang, X.; Qian, Z.; Huang, Z.; Jin, X.; et al. Antiviral effects of ferric ammonium citrate. Cell Discov. 2018, 4, 14. [Google Scholar] [CrossRef]
  171. Xu, M.; Lee, E.M.; Wen, Z.; Cheng, Y.; Huang, W.K.; Qian, X.; Tcw, J.; Kouznetsova, J.; Ogden, S.C.; Hammack, C.; et al. Identification of small-molecule inhibitors of zika virus infection and induced neural cell death via a drug repurposing screen. Nat. Med. 2016, 22, 1101–1107. [Google Scholar] [CrossRef]
  172. Cairns, D.M.; Boorgu, D.; Levin, M.; Kaplan, D.L. Niclosamide rescues microcephaly in a humanized in vivo model of zika infection using human induced neural stem cells. Biol. Open 2018, 7. [Google Scholar] [CrossRef]
  173. Elgner, F.; Sabino, C.; Basic, M.; Ploen, D.; Grunweller, A.; Hildt, E. Inhibition of zika virus replication by silvestrol. Viruses 2018, 10, 149. [Google Scholar] [CrossRef]
  174. Pitts, J.D.; Li, P.C.; de Wispelaere, M.; Yang, P.L. Antiviral activity of n-(4-hydroxyphenyl) retinamide (4-hpr) against zika virus. Antivir. Res. 2017, 147, 124–130. [Google Scholar] [CrossRef]
  175. Yang, S.N.Y.; Atkinson, S.C.; Fraser, J.E.; Wang, C.; Maher, B.; Roman, N.; Forwood, J.K.; Wagstaff, K.M.; Borg, N.A.; Jans, D.A. Novel flavivirus antiviral that targets the host nuclear transport importin alpha/beta1 heterodimer. Cells 2019, 8, 281. [Google Scholar] [CrossRef]
  176. Raekiansyah, M.; Mori, M.; Nonaka, K.; Agoh, M.; Shiomi, K.; Matsumoto, A.; Morita, K. Identification of novel antiviral of fungus-derived brefeldin a against dengue viruses. Trop. Med. Health 2017, 45, 32. [Google Scholar] [CrossRef]
  177. Yang, S.; Xu, M.; Lee, E.M.; Gorshkov, K.; Shiryaev, S.A.; He, S.; Sun, W.; Cheng, Y.S.; Hu, X.; Tharappel, A.M.; et al. Emetine inhibits zika and ebola virus infections through two molecular mechanisms: Inhibiting viral replication and decreasing viral entry. Cell Discov. 2018, 4, 31. [Google Scholar] [CrossRef]
  178. Puschnik, A.S.; Marceau, C.D.; Ooi, Y.S.; Majzoub, K.; Rinis, N.; Contessa, J.N.; Carette, J.E. A small-molecule oligosaccharyltransferase inhibitor with pan-flaviviral activity. Cell Rep. 2017, 21, 3032–3039. [Google Scholar] [CrossRef]
  179. Zhang, R.; Miner, J.J.; Gorman, M.J.; Rausch, K.; Ramage, H.; White, J.P.; Zuiani, A.; Zhang, P.; Fernandez, E.; Zhang, Q.; et al. A crispr screen defines a signal peptide processing pathway required by flaviviruses. Nature 2016, 535, 164–168. [Google Scholar] [CrossRef]
  180. Estoppey, D.; Lee, C.M.; Janoschke, M.; Lee, B.H.; Wan, K.F.; Dong, H.; Mathys, P.; Filipuzzi, I.; Schuhmann, T.; Riedl, R.; et al. The natural product cavinafungin selectively interferes with zika and dengue virus replication by inhibition of the host signal peptidase. Cell Rep. 2017, 19, 451–460. [Google Scholar] [CrossRef]
  181. Cao, R.Y.; Xu, Y.F.; Zhang, T.H.; Yang, J.J.; Yuan, Y.; Hao, P.; Shi, Y.; Zhong, J.; Zhong, W. Pediatric drug nitazoxanide: A potential choice for control of zika. Open Forum. Infect. Dis. 2017, 4, ofx009. [Google Scholar] [CrossRef]
  182. Leonardi, W.; Zilbermintz, L.; Cheng, L.W.; Zozaya, J.; Tran, S.H.; Elliott, J.H.; Polukhina, K.; Manasherob, R.; Li, A.; Chi, X.; et al. Bithionol blocks pathogenicity of bacterial toxins, ricin, and zika virus. Sci. Rep. 2016, 6, 34475. [Google Scholar] [CrossRef]
  183. Lin, S.C.; Chen, M.C.; Liu, S.; Callahan, V.M.; Bracci, N.R.; Woodson, C.M.; Dahal, B.; de la Fuente, C.L.; Lin, C.C.; Wang, T.T.; et al. Phloretin inhibits zika virus infection by interfering with the cellular glucose utilization. Int. J. Antimicrob. Agents 2019. [Google Scholar] [CrossRef]
  184. Merino-Ramos, T.; Jimenez de Oya, N.; Saiz, J.C.; Martin-Acebes, M.A. Antiviral activity of nordihydroguaiaretic acid and its derivative tetra-o-methyl nordihydroguaiaretic acid against west nile virus and zika virus. Antimicrob. Agents Chemother. 2017, 61, e00376-17. [Google Scholar] [CrossRef]
  185. Jimenez de Oya, N.; Esler, W.P.; Huard, K.; El-Kattan, A.F.; Karamanlidis, G.; Blazquez, A.B.; Ramos-Ibeas, P.; Escribano-Romero, E.; Louloudes-Lazaro, A.; Casas, J.; et al. Targeting host metabolism by inhibition of acetyl-coenzyme a carboxylase reduces flavivirus infection in mouse models. Emerg. Microbes Infect. 2019, 8, 624–636. [Google Scholar] [CrossRef]
  186. Jimenez de Oya, N.; Blazquez, A.B.; Casas, J.; Saiz, J.C.; Martin-Acebes, M.A. Direct activation of adenosine monophosphate-activated protein kinase (ampk) by pf-06409577 inhibits flavivirus infection through modification of host cell lipid metabolism. Antimicrob. Agents Chemother. 2018, 62, e00360-18. [Google Scholar] [CrossRef]
  187. Cheng, F.; Ramos da Silva, S.; Huang, I.C.; Jung, J.U.; Gao, S.J. Suppression of zika virus infection and replication in endothelial cells and astrocytes by pka inhibitor pki 14-22. J. Virol. 2018, 92, e02019-17. [Google Scholar] [CrossRef]
  188. Li, C.; Deng, Y.Q.; Wang, S.; Ma, F.; Aliyari, R.; Huang, X.Y.; Zhang, N.N.; Watanabe, M.; Dong, H.L.; Liu, P.; et al. 25-hydroxycholesterol protects host against zika virus infection and its associated microcephaly in a mouse model. Immunity 2017, 46, 446–456. [Google Scholar] [CrossRef]
  189. Tricarico, P.M.; Caracciolo, I.; Gratton, R.; D’Agaro, P.; Crovella, S. 25-hydroxycholesterol reduces inflammation, viral load and cell death in zikv-infected u-87 mg glial cell line. Inflammopharmacology 2018. [Google Scholar] [CrossRef]
  190. Willard, K.A.; Elling, C.L.; Stice, S.L.; Brindley, M.A. The oxysterol 7-ketocholesterol reduces zika virus titers in vero cells and human neurons. Viruses 2018, 11, 20. [Google Scholar] [CrossRef]
  191. Wichit, S.; Hamel, R.; Bernard, E.; Talignani, L.; Diop, F.; Ferraris, P.; Liegeois, F.; Ekchariyawat, P.; Luplertlop, N.; Surasombatpattana, P.; et al. Imipramine inhibits chikungunya virus replication in human skin fibroblasts through interference with intracellular cholesterol trafficking. Sci. Rep. 2017, 7, 3145. [Google Scholar] [CrossRef]
  192. Singh, P.K.; Khatri, I.; Jha, A.; Pretto, C.D.; Spindler, K.R.; Arumugaswami, V.; Giri, S.; Kumar, A.; Bhasin, M.K. Determination of system level alterations in host transcriptome due to zika virus (zikv) infection in retinal pigment epithelium. Sci. Rep. 2018, 8, 11209. [Google Scholar] [CrossRef]
  193. Sarkey, J.P.; Richards, M.P.; Stubbs, E.B., Jr. Lovastatin attenuates nerve injury in an animal model of guillain-barre syndrome. J. Neurochem. 2007, 100, 1265–1277. [Google Scholar] [CrossRef]
  194. Huang, Y.; Li, Y.; Zhang, H.; Zhao, R.; Jing, R.; Xu, Y.; He, M.; Peer, J.; Kim, Y.C.; Luo, J.; et al. Zika virus propagation and release in human fetal astrocytes can be suppressed by neutral sphingomyelinase-2 inhibitor gw4869. Cell Discov. 2018, 4, 19. [Google Scholar] [CrossRef]
  195. Retallack, H.; Di Lullo, E.; Arias, C.; Knopp, K.A.; Laurie, M.T.; Sandoval-Espinosa, C.; Mancia Leon, W.R.; Krencik, R.; Ullian, E.M.; Spatazza, J.; et al. Zika virus cell tropism in the developing human brain and inhibition by azithromycin. Proc. Natl. Acad. Sci. USA 2016, 113, 14408–14413. [Google Scholar] [CrossRef] [Green Version]
  196. Zhou, T.; Tan, L.; Cederquist, G.Y.; Fan, Y.; Hartley, B.J.; Mukherjee, S.; Tomishima, M.; Brennand, K.J.; Zhang, Q.; Schwartz, R.E.; et al. High-content screening in hpsc-neural progenitors identifies drug candidates that inhibit zika virus infection in fetal-like organoids and adult brain. Cell Stem. Cell 2017, 21, 274–283. [Google Scholar] [CrossRef]
  197. Taguwa, S.; Yeh, M.T.; Rainbolt, T.K.; Nayak, A.; Shao, H.; Gestwicki, J.E.; Andino, R.; Frydman, J. Zika virus dependence on host hsp70 provides a protective strategy against infection and disease. Cell Rep. 2019, 26, 906–920. [Google Scholar] [CrossRef]
  198. Li, C.; Xu, D.; Ye, Q.; Hong, S.; Jiang, Y.; Liu, X.; Zhang, N.; Shi, L.; Qin, C.F.; Xu, Z. Zika virus disrupts neural progenitor development and leads to microcephaly in mice. Cell Stem. Cell 2016, 19, 120–126. [Google Scholar] [CrossRef]
  199. Costa, V.V.; Del Sarto, J.L.; Rocha, R.F.; Silva, F.R.; Doria, J.G.; Olmo, I.G.; Marques, R.E.; Queiroz-Junior, C.M.; Foureaux, G.; Araujo, J.M.S.; et al. N-methyl-d-aspartate (nmda) receptor blockade prevents neuronal death induced by zika virus infection. MBio 2017, 8, e00350-17. [Google Scholar] [CrossRef]
  200. Simanjuntak, Y.; Liang, J.J.; Chen, S.Y.; Li, J.K.; Lee, Y.L.; Wu, H.C.; Lin, Y.L. Ebselen alleviates testicular pathology in mice with zika virus infection and prevents its sexual transmission. PLoS Pathog. 2018, 14, e1006854. [Google Scholar] [CrossRef]
  201. Lei, J.; Vermillion, M.S.; Jia, B.; Xie, H.; Xie, L.; McLane, M.W.; Sheffield, J.S.; Pekosz, A.; Brown, A.; Klein, S.L.; et al. Il-1 receptor antagonist therapy mitigates placental dysfunction and perinatal injury following zika virus infection. JCI Insight. 2019, 4. [Google Scholar] [CrossRef]
  202. Bayer, A.; Lennemann, N.J.; Ouyang, Y.; Bramley, J.C.; Morosky, S.; Marques, E.T., Jr.; Cherry, S.; Sadovsky, Y.; Coyne, C.B. Type iii interferons produced by human placental trophoblasts confer protection against zika virus infection. Cell Host Microbe 2016, 19, 705–712. [Google Scholar] [CrossRef]
  203. Contreras, D.; Arumugaswami, V. Zika virus infectious cell culture system and the in vitro prophylactic effect of interferons. J. Vis. Exp. 2016. [Google Scholar] [CrossRef]
  204. Jagger, B.W.; Miner, J.J.; Cao, B.; Arora, N.; Smith, A.M.; Kovacs, A.; Mysorekar, I.U.; Coyne, C.B.; Diamond, M.S. Gestational stage and ifn-lambda signaling regulate zikv infection in utero. Cell Host Microbe 2017, 22, 366–376. [Google Scholar] [CrossRef]
  205. Savidis, G.; Perreira, J.M.; Portmann, J.M.; Meraner, P.; Guo, Z.; Green, S.; Brass, A.L. The ifitms inhibit zika virus replication. Cell Rep. 2016, 15, 2323–2330. [Google Scholar] [CrossRef]
  206. Pryke, K.M.; Abraham, J.; Sali, T.M.; Gall, B.J.; Archer, I.; Liu, A.; Bambina, S.; Baird, J.; Gough, M.; Chakhtoura, M.; et al. A novel agonist of the trif pathway induces a cellular state refractory to replication of zika, chikungunya, and dengue viruses. MBio 2017, 8, e00452-17. [Google Scholar] [CrossRef]
  207. Ji, Z.; Li, F.; Xia, Z.; Guo, X.; Gao, M.; Sun, F.; Cheng, Y.; Wu, Y.; Li, W.; Ali, S.A.; et al. The scorpion venom peptide smp76 inhibits viral infection by regulating type-i interferon response. Virol. Sin. 2018, 33, 545–556. [Google Scholar] [CrossRef]
  208. Smith, J.L.; Jeng, S.; McWeeney, S.K.; Hirsch, A.J. A microrna screen identifies the wnt signaling pathway as a regulator of the interferon response during flavivirus infection. J. Virol. 2017, 91, e02388-16. [Google Scholar] [CrossRef]
  209. Arbuckle, J.H.; Gardina, P.J.; Gordon, D.N.; Hickman, H.D.; Yewdell, J.W.; Pierson, T.C.; Myers, T.G.; Kristie, T.M. Inhibitors of the histone methyltransferases ezh2/1 induce a potent antiviral state and suppress infection by diverse viral pathogens. MBio 2017, 8, e01141-17. [Google Scholar] [CrossRef]
  210. Wang, Z.Y.; Wang, Z.; Zhen, Z.D.; Feng, K.H.; Guo, J.; Gao, N.; Fan, D.Y.; Han, D.S.; Wang, P.G.; An, J. Axl is not an indispensable factor for zika virus infection in mice. J. Gen. Virol. 2017, 98, 2061–2068. [Google Scholar] [CrossRef]
  211. Liu, S.; DeLalio, L.J.; Isakson, B.E.; Wang, T.T. Axl-mediated productive infection of human endothelial cells by zika virus. Circ. Res. 2016, 119, 1183–1189. [Google Scholar] [CrossRef]
  212. Musso, D.; Gubler, D.J. Zika virus. Clin. Microbiol. Rev. 2016, 29, 487–524. [Google Scholar] [CrossRef]
  213. Vazquez-Calvo, A.; Saiz, J.C.; McCullough, K.C.; Sobrino, F.; Martin-Acebes, M.A. Acid-dependent viral entry. Virus Res. 2012, 167, 125–137. [Google Scholar] [CrossRef]
  214. Martin-Acebes, M.A.; Vazquez-Calvo, A.; Saiz, J.C. Lipids and flaviviruses, present and future perspectives for the control of dengue, zika, and west nile viruses. Prog. Lipid. Res. 2016, 64, 123–137. [Google Scholar] [CrossRef]
  215. Mounce, B.C.; Poirier, E.Z.; Passoni, G.; Simon-Loriere, E.; Cesaro, T.; Prot, M.; Stapleford, K.A.; Moratorio, G.; Sakuntabhai, A.; Levraud, J.P.; et al. Interferon-induced spermidine-spermine acetyltransferase and polyamine depletion restrict zika and chikungunya viruses. Cell Host Microbe 2016, 20, 167–177. [Google Scholar] [CrossRef]
  216. Mounce, B.C.; Cesaro, T.; Moratorio, G.; Hooikaas, P.J.; Yakovleva, A.; Werneke, S.W.; Smith, E.C.; Poirier, E.Z.; Simon-Loriere, E.; Prot, M.; et al. Inhibition of polyamine biosynthesis is a broad-spectrum strategy against rna viruses. J. Virol. 2016, 90, 9683–9692. [Google Scholar] [CrossRef]
  217. Tang, H.; Hammack, C.; Ogden, S.C.; Wen, Z.; Qian, X.; Li, Y.; Yao, B.; Shin, J.; Zhang, F.; Lee, E.M.; et al. Zika virus infects human cortical neural progenitors and attenuates their growth. Cell Stem. Cell 2016, 18, 587–590. [Google Scholar] [CrossRef]
  218. Martin-Acebes, M.A.; Jimenez de Oya, N.; Saiz, J.C. Lipid metabolism as a source of druggable targets for antiviral discovery against zika and other flaviviruses. Pharmaceuticals 2019, 12, 97. [Google Scholar] [CrossRef]
  219. Cugola, F.R.; Fernandes, I.R.; Russo, F.B.; Freitas, B.C.; Dias, J.L.; Guimaraes, K.P.; Benazzato, C.; Almeida, N.; Pignatari, G.C.; Romero, S.; et al. The brazilian zika virus strain causes birth defects in experimental models. Nature 2016, 534, 267–271. [Google Scholar] [CrossRef]
  220. Garcez, P.P.; Loiola, E.C.; Madeiro da Costa, R.; Higa, L.M.; Trindade, P.; Delvecchio, R.; Nascimento, J.M.; Brindeiro, R.; Tanuri, A.; Rehen, S.K. Zika virus impairs growth in human neurospheres and brain organoids. Science 2016, 352, 816–818. [Google Scholar] [CrossRef] [Green Version]
  221. Lee, I.; Bos, S.; Li, G.; Wang, S.; Gadea, G.; Despres, P.; Zhao, R.Y. Probing molecular insights into zika virus (-) host interactions. Viruses 2018, 10, 233. [Google Scholar] [CrossRef]
  222. Asif, A.; Manzoor, S.; Tuz-Zahra, F.; Saalim, M.; Ashraf, M.; Ishtiyaq, J.; Khalid, M. Zika virus: Immune evasion mechanisms, currently available therapeutic regimens, and vaccines. Viral. Immunol. 2017, 30, 682–690. [Google Scholar] [CrossRef]
  223. Bowen, J.R.; Quicke, K.M.; Maddur, M.S.; O’Neal, J.T.; McDonald, C.E.; Fedorova, N.B.; Puri, V.; Shabman, R.S.; Pulendran, B.; Suthar, M.S. Zika virus antagonizes type i interferon responses during infection of human dendritic cells. PLoS Pathog. 2017, 13, e1006164. [Google Scholar] [CrossRef]
  224. Chen, J.; Yang, Y.F.; Yang, Y.; Zou, P.; He, Y.; Shui, S.L.; Cui, Y.R.; Bai, R.; Liang, Y.J.; Hu, Y.; et al. Axl promotes zika virus infection in astrocytes by antagonizing type i interferon signalling. Nat. Microbiol. 2018, 3, 302–309. [Google Scholar] [CrossRef]
  225. Grant, A.; Ponia, S.S.; Tripathi, S.; Balasubramaniam, V.; Miorin, L.; Sourisseau, M.; Schwarz, M.C.; Sanchez-Seco, M.P.; Evans, M.J.; Best, S.M.; et al. Zika virus targets human stat2 to inhibit type i interferon signaling. Cell Host Microbe 2016, 19, 882–890. [Google Scholar] [CrossRef]
  226. Kumar, A.; Hou, S.; Airo, A.M.; Limonta, D.; Mancinelli, V.; Branton, W.; Power, C.; Hobman, T.C. Zika virus inhibits type-i interferon production and downstream signaling. EMBO Rep. 2016, 17, 1766–1775. [Google Scholar] [CrossRef]
  227. Yockey, L.J.; Jurado, K.A.; Arora, N.; Millet, A.; Rakib, T.; Milano, K.M.; Hastings, A.K.; Fikrig, E.; Kong, Y.; Horvath, T.L.; et al. Type i interferons instigate fetal demise after zika virus infection. Sci. Immunol. 2018, 3. [Google Scholar] [CrossRef]
Table 1. Zika virus (ZIKV) vaccines clinical trials.
Table 1. Zika virus (ZIKV) vaccines clinical trials.
PlatformVaccineAntigenAdjuvantRegimen (Dose, Route, Day/Week)Clinical PhaseTrial NumberReferences
InactivatedBBV121virionAlum2.5 vs. 5 vs. 10 mg 2× (0, + 30)ICTRI/2017/05/008539[20,21]
ZPIVvirionAlum5 mg/IM 2× (+ 1, + 29)INCT02963909[22,23]
ZPIVvirionAlum2.5 vs. 5 vs. 10 mg/IM 2× (0, + 29)INCT02952833[22,23]
ZPIVvirionAlum5 mg/IM 2× (0, + 7: 0, + 14; 0, + 29)INCT02937233[22,23]
ZPIVvirionAlumYF-VAX/IXIARO-5 mg/IM 2× (+ 1, + 29)INCT03008122[22,23]
PIZVvirionAlum2 vs. 5 vs. 10 mg/IM 2× (0, + 30)INCT03343626[22,23]
VLA1601virionAlum3 vs. 6 AgU 2× (0, + 1 vs. 0, + 4)INCT03425149[22,23]
DNA-basedGLS-5700pM/ENone1 mg/IDINCT02809443[24,25,26]
GLS-5700pM/ENone2 mg/IDINCT02887482[24,25,26]
VRC5283pM/ENone4 mg/IM both arms, 2× (0, + 8; 0, + 12) 3× (0, + 4, + 8; 0, + 4, + 20)INCT02840487[27,28]
VRC5283pM/ENone4 mg/IM both arms, 2× (0, + 8) (0, + 12); 3× (0, + 4, + 8) (0, + 4, + 20), needle/needle-freeINCT02996461[27,28]
VRC5283pM/ENone4 mg vs. 8 mg/IM both arms, 3× (0, + 4, + 8) needle-freeII NCT03110770[27,28]
RNA-basedmRNA-1325pM/ENonemRNA-1325INCT03014089[29,30]
RecombinantMVpM/ENonelow/high dose (0 vs. 0, + 30)INCT02996890[31]
Table 2. Direct acting antivirals (DAAs) against ZIKV.
Table 2. Direct acting antivirals (DAAs) against ZIKV.
StepProposed TargetCompounds System Used to Test the DrugsReferences
BindingPreventing E binding Neutralizing antibodiesCell culture, biochemical assays, and animal models[68,69,70,71,72,73,74,79]
Virion morphology and E-mediated membrane fusion & entryZ2, Porphyrins, BZM-CDs, POMs, SCRs, small molecules, [80,81,82,83,84,85,86,87,88,89]
ViricidalCroton, Aphloia theiformis and Psiloxylon mauritianum extracts, epigallocatechin gallate, delphinidin, pinocembrin[90,91,92,93,94,95,96]
ReplicationNS5 polymerase inhibitors (Purine synthesis inhibitors)Ribavirin, merimepodib, favipiravir, mycophenolic acid, azathioprine, methotrexate, avotaquone, 6-methylmercaptopurine riboside, atovaquone, 2’-CMG, 2’-CMA 7-deaza-2’-CMA, NITD008, BCX4430Cell culture, biochemical assays, and animal models[97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115]
NS5 polymerase inhibitors (Pyrimidine synthesis inhibitors)Brequinar, CID91632869, gemcitabine, 2’-CMC, 2’-CMU, 5’-fluorouracil, 6-azauridine, finastenide, P12-23 and P12-34, sofosbuvir, viperin, aurintricarboxylic acidCell culture, biochemical assays, and animal models[97,107,108,110,116,117,118,119,120,121,122,123,124,125]
NS5 methyltransferase inhibitorsF3043-0013, F0922-0796, F1609-0442, and F1750-0048, Compound 10 Biochemical assay, docking [126,127,128]
Viral protease inhibitorsmyricitin, quercetin, luteolin, isorhamnetin, apigerin, compound 2, compound 3, curcumin, suramin, temoporfin, niclosamide, nitazoxanide, novobiocin, bromocriptine, aprotinin, NSC157058, NSC86314, NSC716903, NSC135618, erythrosin B, viperinCell culture, biochemical assays, and animal models[129,130,131,132,133,134,135,136,137,138,139,140,141,142]
Table 3. Host acting antivirals (HAAs) against ZIKV.
Table 3. Host acting antivirals (HAAs) against ZIKV.
StepProposed TargetCompoundsSystem Used to Test the DrugsReferences
Early StepsReceptor binding inhibitionR448, cabozantinib, MYD1 curcumin, suramin, NSAIDs (ibuprofen, aspirin, acetaminophen, lornoxicam, naproxen), CLR01, amphipathic α-helical peptideCell culture[131,132,153,154,155,156,157,158]
Internalizationfusion inhibitors/endosome acidifcationNanchangmycin, arbidol, compound 16, chlorpromazine, daptomycin, Obatoclax, SaliPhe, amodaquine, peptide Ev7, CQ, ammonium chloride, mefloquine, quinacrine, bafilomycin A1, and GSK369796, K22, iron salt ferric ammonium citrate, tenovin 1, niclosamideCell culture, human organoids, animal models[104,107,108,117,138,159,160,161,162,163,164,165,166,167,168,169,170,171,172]
Translation/Transcription Silvestrol, fenretinide, DFMO, DENSpm, GW5074Cell culture[155,173,174,175]
ReplicationIntracellular transportBrefeldin A, emetine, NG-1, SPase, cavinafungin, nitazoxanideCell culture[176,177,178,179,180,181]
Caspases/CDK inhibitorsBithionol, emricasan, phloretin, bortezomib, PHA-690509Cell culture[104,107,171,182,183]
Lipids metabolismNDGA, M4N, PF-429242, fatostatin, metformin, AICAR, PF-05175157, PF-05206574, PF-06256254, PF-06409577 Cell culture[184,185,186,187]
Cholesterol metabolism25-HC, 7-KC, Imipramine, benzamil, lovastatin, mevastatin, GSW4869, Cell culture, animal models[120,188,189,190,191,192,193,194]
Unknown Kitasamycin, azithromycin, cyclosporine, sertraline, ivermectin, pyrimethamine, AQ, palonosetron, Hsp70-NEF interaction inhibitorsCell culture, animal models[104,120,195,196,197]
Side effects ifenprodil, dizocilpine agmatine sulfate, memantine, ebelsen, IRACell culture, animal models[198,199,200,201]
Innate immunity modulation IFNs, IFITM1 and IFITM3, AVC, Smp76, miRNAsCell culture, animal models[202,203,204,205,206,207,208]
Epigenetics GSK-126Cell culture, animal models[209]

Share and Cite

MDPI and ACS Style

Saiz, J.-C. Therapeutic Advances Against ZIKV: A Quick Response, a Long Way to Go. Pharmaceuticals 2019, 12, 127. https://0-doi-org.brum.beds.ac.uk/10.3390/ph12030127

AMA Style

Saiz J-C. Therapeutic Advances Against ZIKV: A Quick Response, a Long Way to Go. Pharmaceuticals. 2019; 12(3):127. https://0-doi-org.brum.beds.ac.uk/10.3390/ph12030127

Chicago/Turabian Style

Saiz, Juan-Carlos. 2019. "Therapeutic Advances Against ZIKV: A Quick Response, a Long Way to Go" Pharmaceuticals 12, no. 3: 127. https://0-doi-org.brum.beds.ac.uk/10.3390/ph12030127

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop