Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Oct 1.
Published in final edited form as: Drug Discov Today. 2012 Jun 13;17(19-20):1111–1120. doi: 10.1016/j.drudis.2012.06.002

Recent progress in structure-based anti-influenza drug design

Juan Du 1,3, Timothy A Cross 2,3,4, Huan-Xiang Zhou 1,3,*
PMCID: PMC3459301  NIHMSID: NIHMS387332  PMID: 22704956

Abstract

Seasonal and pandemic influenza have caused high morbidity and mortality worldwide. Recent emergence of influenza A H5N1 and H1N1 strains has heightened concern, especially as a result of their drug resistance. The life cycle of influenza viruses has been well studied and nearly all the viral proteins are becoming potential therapeutic targets. In this review, we present an overview of recent progress in structure-based anti-influenza drug design, paying close attention to the increasing role of computation and strategies for overcoming drug resistance.

Introduction

Seasonal and pandemic influenza represent one of the major threats to public health. The annual influenza epidemic results in 250 000–500 000 deaths worldwide [1]. During the past century, the 1918 Spanish flu, 1957 Asian flu and 1968 Hong Kong flu pandemics caused millions of fatalities [2]. More-recent years have seen the emergence of the 1997 H5N1 virus in Hong Kong, or ‘bird flu’, known for its high fatality rate (although low transmission in humans) [1], and the 2009 H1N1 virus in Mexico, or ‘swine flu’, which contributed to at least 16 000 deaths [1]. These two viruses have heightened concern, especially because they can carry drug-resistant mutations [3,4].

Influenza viruses are negative-sense single-stranded RNA viruses, belonging to the family Orthomyxoviridae. Based on the antigenic difference in their nucleoproteins and matrix proteins, the viruses are classified into three types: A, B and C [5]. Influenza A is the major pathogen for most cases of epidemic influenza, and has thus attracted the most attention. The influenza A genome is composed of eight RNA segments, five of which code for one protein each and the other three code for two proteins each [6]. The proteins (Table 1) are: hemagglutinin (HA), neuraminidase (NA), matrix protein 1 (M1), M2 proton channel, nucleoprotein (NP), non-structural protein 1 (NS1), nuclear export protein (NEP; formerly known as NS2), polymerase acid protein (PA), polymerase basic proteins (PB1 and PB2) and a protein named PB1-F2 which is expressed from a second reading frame (+1) of the PB1 gene [7]. PB1, PB2 and PA form the RNA polymerase. The surface glycoproteins HA and NA provide the viruses with distinct antigenic properties. Influenza A viruses are further organized according to HA and NA subtypes. Sixteen HA subtypes (H1 to H16) and nine NA subtypes (N1 to N9) have been identified. The subtypes of the 1997 bird flu and the 2009 swine flu viruses have been indicated above. Recent seasonal flu epidemics are dominated by the H3N2 and H1N1 subtypes (along with influenza B viruses).

Table 1.

Functions, binding sites and inhibitors of influenza proteins

Proteins Functions Binding sites Inhibitors
HA Virus attachment to sialic acid receptors on host cell surface; fusion of virus and cell membranes Sialic acid binding site; TBHQ binding site TBHQ
NA Cleavage of sialic acid receptors to release progeny viruses from host cells Active site Zanamivir; oseltamivir
M2 Acidification and uncoating of endosome-entrapped virus; virus assembly and budding Inside pore near Ser31 Amantadine; rimantadine
NP Capsidation of viral RNA and binding of three polymerase subunits to form ribonucleoprotein particles Tail-loop binding site; RNA binding site {6. AU: are there none available? If none available please state here and hereafter for absolute clarity in Table 1}
Polymerase Viral RNA transcription and replication PA: endonuclease active site; binding site for PB1
PB1: polymerase active site
PB2: cap binding site; importing binding site
Ml Structural component of virion; nuclear export of ribonucleoprotein particles NEP binding site
NEP Nuclear export of ribonucleoprotein particles from host-cell nucleus Crm1 binding site; M1 binding site
NS1 Protection against host-cell antiviral responses Double-stranded RNA binding site; CPSF30 binding site

Vaccines and drugs are two effective strategies against influenza.{1. AU: Because available drugs don't really provide an effective anti-influenza strategy this sentence should reflect that. Also Vaccination isn't always an option as you go onto say. Please re-word this sentence for clarification. Would something like: Vaccine and drug usage are two strategies for combating influenza infection, be ok here?} Vaccination is not a realistic plan for a rapidly spreading influenza pandemic, because of the substantial lead time for vaccine production. The antiviral drugs provide alternative options to control influenza infections. To date, four antiviral drugs have been approved by the FDA, including two NA inhibitors, oseltamivir (Tamiflu®) and zanamivir (Relenza®), and two M2 channel blockers, amantadine (Symmetrel®) and rimantadine (Flumadine®). Rapid emergence of drug-resistant viral mutations has limited the use of the NA inhibitors [3,8,9] and rendered the M2 blockers ineffective [4,1012]. It is urgent that novel anti-influenza drugs are developed.

The life cycle of influenza viruses has been well studied and nearly all the viral proteins are becoming potential therapeutic targets [13,14]. Here, we present an overview of recent progress in structure-based anti-influenza drug design, with close attention being paid to the increasing role of computation strategies that can be used for overcoming drug resistance.

Hemagglutinin

The surface glycoprotein HA attaches the viral particle to sialic acid receptors on the host cell surface for viral entry and promotes the release of viral ribonucleoprotein complexes through membrane fusion [15,16]. HA is a trimer, commonly divided into a head region and a stem region (Figure 1a); each chain is synthesized as a precursor polypeptide and then cleaved into two fragments, HA1 (328 amino acids) and HA2 (221 amino acids), linked by a disulfide bond. The 16 subtypes of HA are phylogenetically divided into two groups [17]: group 1, composed of H1, H2, H5, H6, H8, H9, H11, H12, H13, H16 subtypes; and group 2, composed of H3, H4, H7, H10, H14, H15 subtypes.

Figure 1.

Figure 1

(a) Hemagglutinin (HA) trimer, bound to a sialic acid, N-acetylneuraminic acid (Neu5Ac); tert-butylhydroquinone (TBHQ); or a designed protein, HB36 (PDB entries 3M5I, 3EYK and 3R2X, respectively). Two of the HA monomers are represented as magenta and yellow surfaces, and the third as a cartoon, with HA1 and HA2 in light and dark blue, respectively. (b) Small molecules that either target the sialic acid binding site (Neu5Ac and compounds 1 and 2) or inhibit membrane fusion (TBHQ and compound 3).

{8. AU: I am unable to edit Fig 1. Please follow DDT house style and use sentence case throughout: i.e. Sialic acid}

Both roles of HA in the viral life cycle have been targeted for therapeutic designs (Figure 1a). The receptor binding site is in the head region, lined by three structural elements of HA1 within each monomer, including the 190-helix (residues 188–194), the 130-loop (residues 134–138) and the 220-loop (residues 221–228) [1821]. Virtual screening of the ZINC database (http://zinc.docking.org/) against an H5-subtype HA by Nandi [22], based on docking and ligand–receptor hydrogen-bonding, identified compounds 1 and 2 (Figure 1b) as potential lead molecules.

Binding of small molecules such as tert-butylhydroquinone (TBHQ; Figure 1b) to HA was found to inhibit membrane fusion [2325]. Recent structure determination [26] identified the TBHQ binding site in a group-2 HA at the prefusion trimeric interface of the stem region. Residues lining this site include L29{2. AU: You have used a combination of amino acid codes in this manuscript I have changed amino acids to single letter code for consistency. Please check that I have done so accurately} of HA1 and L98 and A101 of HA2 in one monomer, and L55 and L99 of HA2 in an adjacent monomer. Apparently, TBHQ binding stabilizes the prefusion trimer conformation and prevents the pH-induced conformational changes of HA1 necessary for membrane fusion. This binding site is blocked in group-1 HAs [26], but fusion inhibitors that target group-1 HAs appear to bind at a neighboring site at the trimeric interface [24,25]. Based on similarity to these earlier ligands, Tang et al. [27] screened the Roche collection of ~1 million compounds. Compound 3 (Figure 1b) was found to be a potent and selective inhibitor to H1-subtype HAs.

Seasonal vaccines elicit neutralizing antibodies that recognize epitopes on the surface of the highly variable head region of HA; vaccine components are varied each year in the hope of matching the circulating viral strains in the forthcoming flu season. Recently, neutralizing antibodies that recognize highly conserved sites in the stem region have been discovered [2830]. These antibodies promise to withstand viral mutations and provide broad neutralization. The same consideration led Fleishman et al. [31] to design proteins computationally that bind to a conserved surface patch in the stem region (Figure 1a). The designed proteins bound HA with nanomolar affinity and inhibited the pH-induced fusogenic conformational changes of HA. The success of the computational design was further demonstrated by the crystal structure of the complex, with a binding interface nearly identical to that in the design.

Neuraminidase

NA is another glycoprotein, a tetramer with relatively independent monomers, expressed at the surface of the influenza virus. It is responsible for releasing progeny viral particles by cleaving the terminal sialic acid from HA receptors on cell membranes [32], and facilitates the mobility of viruses in the respiratory tract [33]. Phylogenetically the nine NA subtypes can also be divided into two groups [34]: group 1, composed of N1, N4, N5 and N8 subtypes; and group 2, composed of N2, N3, N6, N7 and N9 subtypes. Based on the structures of N2 and N9 NAs [35,36] and with 2-deoxy-2,3-didehydro-N-acetylneuraminic acid (Neu5Ac2en; Figure 2a), a naturally occurring NA inhibitor, as the template, two inhibitors, zanamivir and oseltamivir (Figure 2a), were designed [37,38] and subsequently approved by the FDA. They were among the earliest examples of structure-based drug design and represent the most successful effort in structure-based anti-influenza drug discovery [39]. Other NA inhibitors were developed subsequently, including peramivir [40] and A315675 [41] (Figure 2a) as well as oligomeric zanamivir [42].

Figure 2.

Figure 2

(a) Neuraminidase inhibitors. (b) The active site of N1 (light blue) and N9 (yellow) neuraminidases (PDB entries 2HU0 and 2C4A, respectively). The N1 structure has the 150-loop in the open conformation, even with oseltamivir (shown with carbons in green) bound. The 150-loop and 430-loop of N1 and N9 are shown in dark blue and in orange, respectively. N1 sidechains interacting with oseltamivir are also shown.

{9. AU: I am unable to edit Fig 2. Please follow DDT house style and close up in-house nomenclature for compounds: i.e. A315675. Because this is DDT house style please note that I have made this change in your manuscript}

The most common oseltamivir-resistant mutations include H274Y in group-1 NAs and R292K in group-2 NAs [8]. The latter mutation also confers partial resistance to zanamivir [8,43]. Both of the mutated positions are near the active site of the enzymes (Figure 2b). H274Y-carrying viruses and R292K-carrying viruses are susceptible to A315675, but only partially susceptible to peramivir [8].

The residues around the active site of NAs are well-conserved among the nine subtypes, and the two FDA-approved drugs are effective against group-1 and group-2 enzymes, with similar nanomolar IC50s [8]. However, structure determination for group-1 NAs in 2006 [34] showed that, contrary to group-2 NAs, the 150-loop (residues 147–152) adopts an open conformation in the apo form, opening a cavity adjacent to the active site (Figure 2b). Molecular dynamics simulations have further demonstrated considerable conformational flexibility of the 150-loop [44,45]. The 150-cavity provides opportunities for designing N1-selective inhibitors, which were exploited recently by a number of computational design studies [4651]. Notable among these is the study of Rudrawar et al. [51], who introduced allyl and arylallyl substituents at the C3 position of Neu5Ac2en (compound 4, Figure 2a). These derivatives selectively inhibited group-1 NAs over group-2 NAs, and are equally potent against wild-type NA and the H274Y mutant (with micromolar IC50s). In agreement with their molecular modeling prediction, the substituents indeed occupy the 150-cavity in crystal structures of the NA-inhibitor complexes. By extending the moiety at the C4 position of zanamivir, Wen et al. [50] also obtained a compound with a micromolar IC50.

Other computational design studies [5257] have taken advantage of the multitude of NA structures available in the Protein Data Bank (PDB). In all of these studies, including those based on the open conformation of the 150-loop, a database of compounds or a library of derivatives of an established NA inhibitor (oseltamivir and A315675 in particular) is screened using the dock-and-score approach. This approach, although providing the necessary high throughput, has ample room for improvement in accuracy [58]. Validation such as that seen in the study of Rudrawar et al. [51] will continue to be important.

M2 proton channel

M2 forms a tetrameric proton-selective ion channel activated by the low pH of endosomes. Each M2 monomer has only 97 residues. The N-terminal 25 residues are exterior to the{3. AU: please confirm edit here is ok} virus; the next 21 residues form a single transmembrane helix, followed by a 16-residue amphipathic helix residing at the hydrophobic–hydrophilic interface of the viral membrane inner leaflet [59] (Figure 3a); and the C-terminal 35 residues are in the viral interior. While in the endosome, inward proton conductance by M2 mediates acidification of the viral interior and facilitates uncoating of the virus after its endocytosis into the host cells. In addition, the amphipathic helix is associated with virus budding [60] and the C-terminal segment is involved in M1 binding [61].

Figure 3.

Figure 3

(a) The M2 conductance domain, comprising the transmembrane and amphipathic helices (PDB entry 2L0J). The bound amantadine (shown with carbons in magenta) is from PDB entry 2KQT. Sidechains prone to drug-resistant mutations or functionally important are shown. Yellow shade represents the viral membrane. (b) M2 channel blockers.

{10. AU: I am unable to edit Fig 3. Please follow DDT house style and close up in-house compound nomenclature: i.e. BL1743, please also use the correct symbol for the minus sign on Cl}

M2 is the target of the amantadine and rimantadine drugs (Figure 3b). After a brief period of controversy about the drug binding site [62,63], it is now established that these drugs act as channel blockers, binding at a site on{4. AU: Edit ok?} the channel pore where a number of drug-resistant mutations occur [59,6367] (Figure 3a). Amantadine, first synthesized by Prelog and Seiwerth in 1941 [68], was approved as an anti-influenza drug by the FDA in 1966. Its derivative, rimantadine received FDA approval in 1994. Owing to their side effects on the central nervous system, these drugs were not widely used. Other M2 inhibitors, including BL1743 [69] (Figure 3b), have also been identified.

High incidence of drug-resistant mutations in recent years [10,11] has led the Centers for Disease Control and Prevention (CDC) to recommend that these drugs should not be used [70]. The most common drug-resistant mutation is S31N, but V27A and L26F are also often found. Several recent design studies have targeted these mutants, using either amantadine [7174] or BL1743 [7578] as the template. Studies of the former{5. AU: Edit ok?} resulted in compounds that are more-potent inhibitors than amantadine against wild-type M2 but are ineffective against the amantadine-resistant mutants. The latter studies, however, resulted in several compounds, that are potent inhibitors of the V27A and the L26F mutants. Compound 5 is an example [77]. Molecular dynamics simulations suggested that the pore site in the mutants is expanded; the bulky hydrophobic group of compound 5 appears to be well accommodated there. No effective inhibitor for the most common M2 mutant, S31N, has been found.

The M2 pore-lining H37–W41 cluster (Figure 3a), through its unique chemistry, shepherds the permeant proton through the channel [59]. H37 is responsible for channel activation by low exterior pH [79] and for proton selectivity [80], whereas W41 is the channel gate that prevents outward flow of proton current [81]. Given its essential functional role, the H37–W41 cluster seems a promising target for designing inhibitors that will withstand viral mutations.

Nucleoprotein

The primary function of NP is to encapsidate the segmented RNA and bind with the three polymerase subunits, PA, PB1 and PB2, to form ribonucleoprotein particles (RNPs) for RNA transcription, replication and packaging [82]. In each RNP, the viral RNA wraps around individual NP molecules, which are strung together by burying the ‘tail loop’ (residues 402–428) of one NP molecule (498 residues) inside an adjacent NP molecule [8385] (Figure 4). In mature virions and when newly assembled RNPs are ready for export from the host cell nucleus, RNPs are bound with the M1 protein [86] (Figure 4).

Figure 4.

Figure 4

M1 and nuclear export protein (NEP) as adaptor proteins between a ribonucleoprotein particle and Crm1.

By random screening Kao et al. [87] recently discovered that a compound, nucleozin (Figure 5a), induced NP aggregation, thereby inhibiting nucleus accumulation of NP and viral replication. Subsequent structure determination for NP bound with a nucleozin analog identified a NP dimer with a back-to-back arrangement between the monomers, resulting in an interface that harbors two anti-parallel ligand molecules [88]. Nucleozin-resistant mutations around the two ligand binding sites have been found [8789].

Figure 5.

Figure 5

(a) Small molecules that induce nucleoprotein (NP) aggregation (nucleozin) and target the tail-loop binding pocket (compound 6) or the RNA binding groove (compound 7). (b) Structure of an NP monomer, with the tail loop (in magenta) of an adjacent NP molecule buried (PDB entry 2IQH). A salt bridge important for tail-loop binding is highlighted in the inset. Arginine sidechains defining the RNA binding groove are also shown.

{11. AU: I am unable to edit Fig 5. Please follow DDT house style and use sentence case throughout: i.e. Tail-loop binding pocket}

The NP tail-loop binding pocket and the RNA binding groove have recently been targeted for rational drug design (Figure 5b). It has been established that NP oligomerization mediated by tail-loop binding is essential for the transcription and replication activity of RNPs, and that a salt bridge between E339 lining the binding pocket and R416 on the tail loop is essential for tail-loop binding [90]. Now, Shen et al. [91] have shown that the tail-loop peptide can inhibit NP oligomerization and slow down viral replication. By virtual screening, they have also identified compound 6 (Figure 5a) as a potent inhibitor of NP oligomerization and replication of wild-type and nucleozin-resistant strains. The latter finding unequivocally demonstrates that compound 6 and nucleozin target different NP sites.

Targeting the RNA binding groove by the dock-and-score method, Fedichev et al. [92] identified compound 7 (Figure 5a) as a potential NP binder. In vitro and in vivo experiments confirmed the efficacy of this compound as an influenza A virus inhibitor.

Other influenza proteins

Structures for many of the domains of the remaining proteins (Figure 6a) have now been determined. The structures harbor numerous functional sites and protein–protein and protein–RNA binding sites. These provide new targets for structure-based drug design.

Figure 6.

Figure 6

Figure 6

(a) Domain organizations of PA, PB1, PB2, M1, NEP and NS1. (b) Left: endonuclease domain of PA (PDB entry 3HW5). The Mg2+ ion and the surrounding residues defining the active site are shown. Right: C-terminal domain of PA bound with the N-terminal region of PB1 (PDB entry 3CM8). (c) The N-terminal (light blue) and middle (cyan) domains of M1 (PDB entry 1EA3). Cationic residues, including those in the nuclear localization signal motif 101RKLKR105 are displayed. These residues are involved in NEP binding. (d) The C-terminal domain of NEP (PDB entry 1PD3). W78 and surrounding anionic residues involved in M1 binding are shown. (e) Left: N-terminal domain of NS1 bound to double-stranded RNA (PDB entry 2ZKO). Two cationic residues in the binding interface are displayed. Right: effector domain of NS1 bound to two copies of the second and third zinc fingers of CPSF30 (PDB entry 2RHK). The zinc fingers are in yellow and green.

RNA polymerase

The heterotrimer of PA, PB1 and PB2 forms the RNA polymerase of influenza. Once inside the nucleus of the host cell, the polymerase (as part of an RNP) first transcribes and then replicates the viral RNA. The newly synthesized viral mRNA has a 5′-capped fragment cleaved from host pre-mRNA [93]; the cleavage is afforded by the endonuclease activity of the N-terminal domain (residues 1–256) of PA [94] (Figure 6b). The C-terminal domain (residues 257–716) harbors a ‘mouth’ to which the N-terminal region (residues 1–16) of PB1 is inserted [95,96] (Figure 6b). PB1 contains the polymerase domain, but neither the structure nor the precise boundary of this domain is known. The C-terminal region (residues 678–757) of PB1 complexes with the N-terminal region (residues 1–37) of PB2 [97]. For the remainder of PB2, residues 318–483 form the cap-binding domain [98]; residues 535–684 form a putative RNA binding domain [99]; and residues 686–759 form a domain that binds human importin α5, the classical eukaryotic nuclear import adaptor [100].

Recent screening by Iwai et al. [101] of 33 macrocyclic bisbibenzyl compounds identified marchantin E as an inhibitor of PA endonuclease activity. This molecule docked well into the endonuclease active site and inhibited the growth of influenza A viruses. The many protein–protein and protein–RNA interfaces involving polymerase domains provide additional target sites for drug design. In particular, a peptide corresponding to the N-terminal 25 residues of PB1 inhibits the polymerase activity and viral replication, presumably by blocking the assembly of the polymerase trimer [102]. Small-molecule compounds can confer a similar inhibitory effect.

An alternative reading frame within the PB1 gene produces PB1-F2 [7]. This protein (ranging from 80 to 90 residues [6]) is relatively short-lived, localized in mitochondria membranes and induces cell death, but no structural information is yet available.

Matrix protein 1

In the virion, M1 forms an intermediate layer between the membrane-bound HA, NA and M2 proteins and the eight RNPs. In the nucleus of infected cells, binding of M1 to newly assembled RNPs is essential for their export [103]. M1 is also the major driving force in virus budding [104].

The structure of M1 residues 1–164 has been determined [105,106] (Figure 6c). Residues 1–67 and residues 89–164 form four-helix bundles that are stacked side by side; the linker between them also contains a short helix. No structure is available for the C-terminal region but it is known to have a significant helix content [106]. The middle domain contains the basic motif 101RKLKR105, which is a nuclear localization signal and essential for M1 import into the nucleus of infected cells [107]. This basic motif is also involved in the binding of M1 to NEP [108].

Nuclear export protein

NEP, produced by a spliced form of the NS1 gene, is the second adaptor protein between RNPs and the cellular protein Crm1, which mediates the nuclear export of many proteins containing a leucine-rich nuclear export signal (NES) [108] (Figure 4). The N-terminal region (residues 1–54) of NEP bears the NES motif and binds Crm1. The C-terminal domain (residues 63–116) is a helical hairpin (Figure 6d). W78 has been identified as pivotal for the M1 binding site, and is surrounded by glutamate residues. The M1–NEP and NEP–Crm1 interfaces can potentially be targets for drug design.

Non-structural protein 1

The major function of NS1 is to protect viruses against the antiviral responses mediated by interferon α/β in infected cells [109]. NS1 consists of 230 or 237 residues depending on the strain, and forms a dimer. It can be divided into two functional domains: an N-terminal domain (residues 1–73) that binds double-stranded RNA [110] and a C-terminal effector domain (residues 74–230/7) that binds multiple cellular proteins [109], including CPSF30, a protein required for the 3′-end processing of all cellular pre-mRNAs [111] (Figure 6e).

Twu et al. [112] found that a 61-residue fragment comprising the second and third zinc fingers of CPSF30 inhibited viral replication. This fragment binds to NS1 [109]; therefore its antiviral effect suggests that inhibiting the binding of NS1 with CPSF30 opens a route for drug design. Using a yeast-based assay, Basu et al. [113] screened the NCI Diversity Set Library (http://dtp.nci.nih.gov/branches/dscb/diversity_explanation.html) for compounds that inhibited the actions of NS1. Four such compounds were identified, but it was unknown where they bound to NS1, or, indeed, whether they bound to NS1 directly or to cellular factors that regulated NS1.

Recombinant viruses that express mutated or truncated viral proteins can be designed as live-attenuated virus vaccines. This has now been demonstrated for NS1 [114] and NEP [115].

Concluding remarks

The 1997 H5N1 and 2009 H1N1 influenza strains and the rapid emergence of drug-resistant mutations in pandemic and seasonal influenza viruses have heightened the urgency to develop new antivirals. Advances in viral biology and structure determination for nearly all the influenza viral proteins have provided a solid foundation. Taking advantage of that, in the past few years structure-based anti-influenza drug design has been a remarkably active area of research: on the established targets (HA, NA and M2) and on the newly identified potential drug targets (in particular NP and the RNA polymerase).

Computation is now playing an ever-growing role in structure-based drug design. Virtual screening has the throughput to handle libraries containing large numbers of compounds. However, the level of accuracy is still such that experimental validation is essential. Another contribution that computation can make is to generate structural models for viral protein domains and for complexes between viral proteins themselves and between viral proteins and cellular factors.

One can but marvel at the ingenuity with which influenza viruses evade host-cell immune responses and generate drug-resistant mutations. Experience with structure-based drug design during the past three decades has resulted in valuable lessons on overcoming drug resistance. A promising strategy is to target conserved regions of viral protein structures or sites essential for viral functions. Although a note of caution should be considered when reflecting on the experience of oseltamivir and zanamivir, which targeted the well-conserved active site of NA, conserved and essential functional sites should nevertheless be a priority for future drug design.

Winning the war against influenza will probably require combination therapies that target multiple viral protein functions. The observation that in vitro selection of mutant viruses resistant to NA inhibitors identified many mutations in HA demonstrates that these two viral proteins compensate for each other's functions [116]. Correlated mutations in circulating viruses [6] further suggest that such compensation occurs between other viral proteins as well. A combination therapy will render viral functional compensation ineffective, providing a strong motivation to target all the viral proteins and their interactions in future drug design.

  • All flu proteins are becoming drug targets

  • Computation is playing an ever-growing role in structure-based anti-flu drug design

  • Strategies for overcoming drug-resistant flu virus mutations are emerging

Acknowledgments

This work was supported in part by Grants GM58187 and AI23007 from the National Institutes of Health.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.World Health Organization Website http://www.who.int/
  • 2.Cox NJ, Subbarao K. Global epidemiology of influenza: past and present. Annu. Rev. Med. 2000;51:407–421. doi: 10.1146/annurev.med.51.1.407. [DOI] [PubMed] [Google Scholar]
  • 3.Gubareva LV, et al. Selection of influenza virus mutants in experimentally infected volunteers treated with oseltamivir. J. Infect. Dis. 2001;183:523–531. doi: 10.1086/318537. [DOI] [PubMed] [Google Scholar]
  • 4.Gubareva L, et al. Update: drug susceptibility of swine-origin influenza A (H1N1) viruses, April 2009. Morb. Mortal. Wkly Rep. 2009;58:433–435. [PubMed] [Google Scholar]
  • 5.Palese P, Young JF. Variation of influenza A, B, and C viruses. Science. 1982;215:1468–1474. doi: 10.1126/science.7038875. [DOI] [PubMed] [Google Scholar]
  • 6.Ghedin E, et al. Large-scale sequencing of human influenza reveals the dynamic nature of viral genome evolution. Nature. 2005;437:1162–1166. doi: 10.1038/nature04239. [DOI] [PubMed] [Google Scholar]
  • 7.Chen W, et al. A novel influenza A virus mitochondrial protein that induces cell death. Nat. Med. 2001;7:1306–1312. doi: 10.1038/nm1201-1306. [DOI] [PubMed] [Google Scholar]
  • 8.Mishin VP, et al. Susceptibilities of antiviral-resistant influenza viruses to novel neuraminidase inhibitors. Antimicrob. Agents Chemother. 2005;49:4515–4520. doi: 10.1128/AAC.49.11.4515-4520.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Meijer A, et al. Oseltamivir-resistant influenza virus A (H1N1), Europe, 2007-08 season. Emerg. Infect. Dis. 2009;15:552–560. doi: 10.3201/eid1504.081280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Bright RA, et al. Adamantane resistance among influenza A viruses isolated early during the 2005-2006 influenza season in the United States. J. Am. Med. Assoc. 2006;295:891–894. doi: 10.1001/jama.295.8.joc60020. [DOI] [PubMed] [Google Scholar]
  • 11.Deyde VM, et al. Surveillance of resistance to adamantanes among influenza A(H3N2) and A(H1N1) viruses isolated worldwide. J. Infect. Dis. 2007;196:249–257. doi: 10.1086/518936. [DOI] [PubMed] [Google Scholar]
  • 12.Furuse Y, et al. Large-scale sequence analysis of M gene of influenza A viruses from different species: mechanisms for emergence and spread of amantadine resistance. Antimicrob. Agents Chemother. 2009;53:4457–4463. doi: 10.1128/AAC.00650-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Krug RM, Aramini JM. Emerging antiviral targets for influenza A virus. Trends Pharmacol. Sci. 2009;30:269–277. doi: 10.1016/j.tips.2009.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Das K, et al. Structures of influenza A proteins and insights into antiviral drug targets. Nat. Struct. Mol. Biol. 2010;17:530–538. doi: 10.1038/nsmb.1779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Skehel JJ, Wiley DC. Receptor binding and membrane fusion in virus entry: the influenza hemagglutinin. Annu. Rev. Biochem. 2000;69:531–569. doi: 10.1146/annurev.biochem.69.1.531. [DOI] [PubMed] [Google Scholar]
  • 16.Harrison SC. Viral membrane fusion. Nat. Struct. Mol. Biol. 2008;15:690–698. doi: 10.1038/nsmb.1456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Nobusawa E, et al. Comparison of complete amino acid sequences and receptor-binding properties among 13 serotypes of hemagglutinins of influenza A viruses. Virology. 1991;182:475–485. doi: 10.1016/0042-6822(91)90588-3. [DOI] [PubMed] [Google Scholar]
  • 18.Ha Y, et al. X-ray structures of H5 avian and H9 swine influenza virus hemagglutinins bound to avian and human receptor analogs. Proc. Natl Acad. Sci. U S A. 2001;98:11181–11186. doi: 10.1073/pnas.201401198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Gamblin SJ, et al. The structure and receptor binding properties of the 1918 influenza hemagglutinin. Science. 2004;303:1838–1842. doi: 10.1126/science.1093155. [DOI] [PubMed] [Google Scholar]
  • 20.Stevens J, et al. Structure and receptor specificity of the hemagglutinin from an H5N1 influenza virus. Science. 2006;312:404–410. doi: 10.1126/science.1124513. [DOI] [PubMed] [Google Scholar]
  • 21.Yang H, et al. Structures of receptor complexes of a North American H7N2 influenza hemagglutinin with a loop deletion in the receptor binding site. PLoS Pathog. 2010;6:e1001081. doi: 10.1371/journal.ppat.1001081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Nandi T. Proposed lead molecules against Hemagglutinin of avian influenza virus (H5N1). Bioinformation. 2008;2:240–244. doi: 10.6026/97320630002240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Bodian DL, et al. Inhibition of the fusion-inducing conformational change of influenza hemagglutinin by benzoquinones and hydroquinones. Biochemistry. 1993;32:2967–2978. doi: 10.1021/bi00063a007. [DOI] [PubMed] [Google Scholar]
  • 24.Staschke KA, et al. Inhibition of influenza virus hemagglutinin-mediated membrane fusion by a compound related to podocarpic acid. Virology. 1998;248:264–274. doi: 10.1006/viro.1998.9273. [DOI] [PubMed] [Google Scholar]
  • 25.Deshpande MS, et al. An approach to the identification of potent inhibitors of influenza virus fusion using parallel synthesis methodology. Bioorg. Med. Chem. Lett. 2001;11:2393–2396. doi: 10.1016/s0960-894x(01)00459-0. [DOI] [PubMed] [Google Scholar]
  • 26.Russell RJ, et al. Structure of influenza hemagglutinin in complex with an inhibitor of membrane fusion. Proc. Natl Acad. Sci. U S A. 2008;105:17736–17741. doi: 10.1073/pnas.0807142105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Tang G, et al. Discovery of novel 1-phenyl-cycloalkane carbamides as potent and selective influenza fusion inhibitors. Bioorg. Med. Chem. Lett. 2010;20:3507–3510. doi: 10.1016/j.bmcl.2010.04.141. [DOI] [PubMed] [Google Scholar]
  • 28.Ekiert DC, et al. Antibody recognition of a highly conserved influenza virus epitope. Science. 2009;324:246–251. doi: 10.1126/science.1171491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Sui J, et al. Structural and functional bases for broad-spectrum neutralization of avian and human influenza A viruses. Nat. Struct. Mol. Biol. 2009;16:265–273. doi: 10.1038/nsmb.1566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Corti D, et al. A neutralizing antibody selected from plasma cells that binds to group 1 and group 2 influenza A hemagglutinins. Science. 2011;333:850–856. doi: 10.1126/science.1205669. [DOI] [PubMed] [Google Scholar]
  • 31.Fleishman SJ, et al. Computational design of proteins targeting the conserved stem region of influenza hemagglutinin. Science. 2011;332:816–821. doi: 10.1126/science.1202617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Palese P, Compans RW. Inhibition of influenza virus replication in tissue culture by 2-deoxy-2,3-dehydro-N-trifluoroacetylneuraminic acid (FANA): mechanism of action. J. Gen. Virol. 1976;33:159–163. doi: 10.1099/0022-1317-33-1-159. [DOI] [PubMed] [Google Scholar]
  • 33.Klenk HD, Rott R. The molecular biology of influenza virus pathogenicity. Adv. Virus Res. 1988;34:247–281. doi: 10.1016/S0065-3527(08)60520-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Russell RJ, et al. The structure of H5N1 avian influenza neuraminidase suggests new opportunities for drug design. Nature. 2006;443:45–49. doi: 10.1038/nature05114. [DOI] [PubMed] [Google Scholar]
  • 35.Varghese JN, et al. The structure of the complex between influenza virus neuraminidase and sialic acid, the viral receptor. Proteins. 1992;14:327–332. doi: 10.1002/prot.340140302. [DOI] [PubMed] [Google Scholar]
  • 36.White CL, et al. A sialic acid-derived phosphonate analog inhibits different strains of influenza virus neuraminidase with different efficiencies. J. Mol. Biol. 1995;245:623–634. [PubMed] [Google Scholar]
  • 37.von Itzstein M, et al. Rational design of potent sialidase-based inhibitors of influenza virus replication. Nature. 1993;363:418–423. doi: 10.1038/363418a0. [DOI] [PubMed] [Google Scholar]
  • 38.Kim CU, et al. Influenza neuraminidase inhibitors possessing a novel hydrophobic interaction in the enzyme active site: design, synthesis, and structural analysis of carbocyclic sialic acid analogues with potent anti-influenza activity. J. Am. Chem. Soc. 1997;119:681–690. doi: 10.1021/ja963036t. [DOI] [PubMed] [Google Scholar]
  • 39.von Itzstein M. The war against influenza: discovery and development of sialidase inhibitors. Nat. Rev. Drug Discov. 2007;6:967–974. doi: 10.1038/nrd2400. [DOI] [PubMed] [Google Scholar]
  • 40.Babu YS, et al. BCX-1812 (RWJ-270201): discovery of a novel, highly potent, orally active, and selective influenza neuraminidase inhibitor through structure-based drug design. J. Med. Chem. 2000;43:3482–3486. doi: 10.1021/jm0002679. [DOI] [PubMed] [Google Scholar]
  • 41.Kati WM, et al. In vitro characterization of A-315675, a highly potent inhibitor of A and B strain influenza virus neuraminidases and influenza virus replication. Antimicrob. Agents Chemother. 2002;46:1014–1021. doi: 10.1128/AAC.46.4.1014-1021.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Watson KG, et al. Highly potent and long-acting trimeric and tetrameric inhibitors of influenza virus neuraminidase. Bioorg. Med. Chem. Lett. 2004;14:1589–1592. doi: 10.1016/j.bmcl.2003.09.102. [DOI] [PubMed] [Google Scholar]
  • 43.Yen HL, et al. Neuraminidase inhibitor-resistant influenza viruses may differ substantially in fitness and transmissibility. Antimicrob. Agents Chemother. 2005;49:4075–4084. doi: 10.1128/AAC.49.10.4075-4084.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Amaro RE, et al. Remarkable loop flexibility in avian influenza N1 and its implications for antiviral drug design. J. Am. Chem. Soc. 2007;129:7764–7765. doi: 10.1021/ja0723535. [DOI] [PubMed] [Google Scholar]
  • 45.Amaro RE, et al. Mechanism of 150-cavity formation in influenza neuraminidase. Nat. Commun. 2011;2:388. doi: 10.1038/ncomms1390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.García-Sosa AT, et al. Design of multi-binding-site inhibitors, ligand efficiency, and consensus screening of avian influenza H5N1 wild-type neuraminidase and of the oseltamivir-resistant H274Y variant. J. Chem. Inf. Model. 2008;48:2074–2080. doi: 10.1021/ci800242z. [DOI] [PubMed] [Google Scholar]
  • 47.Cheng LS, et al. Ensemble-based virtual screening reveals potential novel antiviral compounds for avian influenza neuraminidase. J. Med. Chem. 2008;51:3878–3894. doi: 10.1021/jm8001197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Li Y, et al. Rational design of Tamiflu derivatives targeting at the open conformation of neuraminidase subtype 1. J. Mol. Graph. Model. 2009;28:203–219. doi: 10.1016/j.jmgm.2009.07.001. [DOI] [PubMed] [Google Scholar]
  • 49.Mitrasinovic PM. On the structure-based design of novel inhibitors of H5N1 influenza A virus neuraminidase (NA). Biophys. Chem. 2009;140:35–38. doi: 10.1016/j.bpc.2008.11.004. [DOI] [PubMed] [Google Scholar]
  • 50.Wen W-H, et al. Analogs of zanamivir with modified C4-substituents as the inhibitors against the group-1 neuraminidases of influenza viruses. Bioorg. Med. Chem. 2010;18:4074–4084. doi: 10.1016/j.bmc.2010.04.010. [DOI] [PubMed] [Google Scholar]
  • 51.Rudrawar S, et al. Novel sialic acid derivatives lock open the 150-loop of an influenza A virus group-1 sialidase. Nat. Commun. 2010;1:113. doi: 10.1038/ncomms1114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.An J, et al. A novel small-molecule inhibitor of the avian influenza H5N1 virus determined through computational screening against the neuraminidase. J. Med. Chem. 2009;52:2667–2672. doi: 10.1021/jm800455g. [DOI] [PubMed] [Google Scholar]
  • 53.Rungrotmongkol T, et al. Design of oseltamivir analogs inhibiting neuraminidase of avian influenza virus H5N1. Antivir. Res. 2009;82:51–58. doi: 10.1016/j.antiviral.2009.01.008. [DOI] [PubMed] [Google Scholar]
  • 54.Park JW, Jo WH. Computational design of novel, high-affinity neuraminidase inhibitors for H5N1 avian influenza virus. Eur. J. Med. Chem. 2010;45:536–541. doi: 10.1016/j.ejmech.2009.10.040. [DOI] [PubMed] [Google Scholar]
  • 55.Wang S-Q, et al. Three new powerful oseltamivir derivatives for inhibiting the neuraminidase of influenza virus. Biochem. Biophys. Res. Comm. 2010;401:188–191. doi: 10.1016/j.bbrc.2010.09.020. [DOI] [PubMed] [Google Scholar]
  • 56.Durrant JD, McCammon JA. Potential drug-like inhibitors of Group 1 influenza neuraminidase identified through computer-aided drug design. Comput. Biol. Chem. 2010;34:97–105. doi: 10.1016/j.compbiolchem.2010.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Rungrotmongkol T, et al. Combinatorial design of avian influenza neuraminidase inhibitors containing pyrrolidine core with a reduced susceptibility to viral drug resistance. Comb. Chem. High Throughput Screening. 2010;13:268–277. doi: 10.2174/138620710790980504. [DOI] [PubMed] [Google Scholar]
  • 58.Gilson MK, Zhou HX. Calculation of protein-ligand binding affinities. Annu. Rev. Biophys. Biomol. Struct. 2007;36:21–42. doi: 10.1146/annurev.biophys.36.040306.132550. [DOI] [PubMed] [Google Scholar]
  • 59.Sharma M, et al. Insight into the mechanism of the influenza A proton channel from a structure in a lipid bilayer. Science. 2010;330:509–512. doi: 10.1126/science.1191750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Rossman JS, Lamb RA. Influenza virus assembly and budding. Virology. 2011;411:229–236. doi: 10.1016/j.virol.2010.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Chen BJ, et al. The influenza virus M2 protein cytoplasmic tail interacts with the M1 protein and influences virus assembly at the site of virus budding. J. Virol. 2008;82:10059–10070. doi: 10.1128/JVI.01184-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Schnell JR, Chou JJ. Structure and mechanism of the M2 proton channel of influenza A virus. Nature. 2008;451:591–595. doi: 10.1038/nature06531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Stouffer AL, et al. Structural basis for the function and inhibition of an influenza virus proton channel. Nature. 2008;451:596–599. doi: 10.1038/nature06528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Hu J, et al. Backbone structure of the amantadine-blocked trans-membrane domain M2 proton channel from influenza A virus. Biophys. J. 2007;92:4335–4343. doi: 10.1529/biophysj.106.090183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Yi M, et al. A secondary gate as a mechanism for inhibition of the M2 proton channel by amantadine. J. Phys. Chem. B. 2008;112:7977–7979. doi: 10.1021/jp800171m. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Cady SD, et al. Structure of the amantadine binding site of influenza M2 proton channels in lipid bilayers. Nature. 2010;463:689–692. doi: 10.1038/nature08722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Cross TA, et al. M2 protein from influenza A: from multiple structures to biophysical and functional insights. Curr. Opin Virol. 2012;2:128–133. doi: 10.1016/j.coviro.2012.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Prelog V, Seiwerth R. Über eine neue, ergiebigere Darstellung des Adamantans. Berichte der Deutschen Chemischen Gesellschaft (A and B Series) 1941;74:1769–1772. [Google Scholar]
  • 69.Kurtz S, et al. Growth impairment resulting from expression of influenza virus M2 protein in Saccharomyces cerevisiae: identification of a novel inhibitor of influenza virus. Antimicrob. Agents Chemother. 1995;39:2204–2209. doi: 10.1128/aac.39.10.2204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Bright RA, et al. High levels of adamantane resistance among influenza A (H3N2) viruses and interim guidelines for use of antiviral agents—United States, 2005-06 influenza season. Morb. Mortal. Wkly Rep. 2006;55:44–46. [PubMed] [Google Scholar]
  • 71.Hu W, et al. Identification of hits as Matrix-2 protein inhibitors through the focused screening of a small primary amine library. J. Med. Chem. 2010;53:3831–3834. doi: 10.1021/jm901664a. [DOI] [PubMed] [Google Scholar]
  • 72.Zhao X, et al. Discovery of highly potent agents against influenza A virus. Eur. J. Med. Chem. 2011;46:52–57. doi: 10.1016/j.ejmech.2010.10.010. [DOI] [PubMed] [Google Scholar]
  • 73.Wang J, et al. Exploring the requirements for the hydrophobic scaffold and polar amine in inhibitors of M2 from influenza A virus. ACS Med. Chem. Lett. 2011;2:307–312. doi: 10.1021/ml100297w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Duque MD, et al. Exploring the size limit of templates for inhibitors of the M2 ion channel of influenza A virus. J. Med. Chem. 2011;54:2646–2657. doi: 10.1021/jm101334y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Wang J, et al. Discovery of spiro-piperidine inhibitors and their modulation of the dynamics of the M2 proton channel from influenza A virus. J. Am. Chem. Soc. 2009;131:8066–8076. doi: 10.1021/ja900063s. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Balannik V, et al. Design and pharmacological characterization of inhibitors of amantadine-resistant mutants of the M2 ion channel of influenza A virus. Biochemistry. 2009;48:11872–11882. doi: 10.1021/bi9014488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Wang J, et al. Molecular dynamics simulation directed rational design of inhibitors targeting drug-resistant mutants of influenza A virus M2. J. Am. Chem. Soc. 2011;133:12834–12841. doi: 10.1021/ja204969m. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Wang J, et al. Exploring organosilane amines as potent inhibitors and structural probes of influenza A virus M2 proton channel. J. Am. Chem. Soc. 2011;133:13844–13847. doi: 10.1021/ja2050666. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Hu J, et al. Histidines, heart of the hydrogen ion channel from influenza A virus: toward an understanding of conductance and proton selectivity. Proc. Natl Acad. Sci. U S A. 2006;103:6865–6870. doi: 10.1073/pnas.0601944103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Venkataraman P, et al. Chemical rescue of histidine selectivity filter mutants of the M2 ion channel of influenza A virus. J. Biol. Chem. 2005;280:21463–21472. doi: 10.1074/jbc.M412406200. [DOI] [PubMed] [Google Scholar]
  • 81.Tang Y, et al. The gate of the influenza virus M2 proton channel is formed by a single tryptophan residue. J. Biol. Chem. 2002;277:39880–39886. doi: 10.1074/jbc.M206582200. [DOI] [PubMed] [Google Scholar]
  • 82.Portela A, Digard P. The influenza virus nucleoprotein: a multifunctional RNA-binding protein pivotal to virus replication. J. Gen. Virol. 2002;83:723–734. doi: 10.1099/0022-1317-83-4-723. [DOI] [PubMed] [Google Scholar]
  • 83.Ye Q, et al. The mechanism by which influenza A virus nucleoprotein forms oligomers and binds RNA. Nature. 2006;444:1078–1082. doi: 10.1038/nature05379. [DOI] [PubMed] [Google Scholar]
  • 84.Ng AK-L, et al. Structure of the influenza virus A H5N1 nucleoprotein: implications for RNA binding, oligomerization, and vaccine design. FASEB J. 2008;22:3638–3647. doi: 10.1096/fj.08-112110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Coloma R, et al. The structure of a biologically active influenza virus ribonucleoprotein complex. PLoS Pathog. 2009;5:e1000491. doi: 10.1371/journal.ppat.1000491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Boulo S, et al. Nuclear traffic of influenza virus proteins and ribonucleoprotein complexes. Virus Res. 2007;124:12–21. doi: 10.1016/j.virusres.2006.09.013. [DOI] [PubMed] [Google Scholar]
  • 87.Kao RY, et al. Identification of influenza A nucleoprotein as an antiviral target. Nat. Biotechnol. 2010;28:600–605. doi: 10.1038/nbt.1638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Gerritz SW, et al. Inhibition of influenza virus replication via small molecules that induce the formation of higher-order nucleoprotein oligomers. Proc. Natl Acad. Sci. U S A. 2011;108:15366–15371. doi: 10.1073/pnas.1107906108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Su C-Y, et al. High-throughput identification of compounds targeting influenza RNA-dependent RNA polymerase activity. Proc. Natl Acad. Sci. U S A. 2010;107:19151–19156. doi: 10.1073/pnas.1013592107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Chan WH, et al. Functional analysis of the influenza virus H5N1 nucleoprotein tail loop reveals amino acids that are crucial for oligomerization and ribonucleoprotein activities. J. Virol. 2010;84:7337–7345. doi: 10.1128/JVI.02474-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Shen Y-F, et al. E339...R416 salt bridge of nucleoprotein as a feasible target for influenza virus inhibitors. Proc. Natl Acad. Sci. U S A. 2011;108:16515–16520. doi: 10.1073/pnas.1113107108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Fedichev P, et al. Structure-based drug design of a new chemical class of small molecules active against influenza A nucleoprotein in vitro and in vivo. PLoS Curr. 2011:RRN1253. doi: 10.1371/currents.RRN1253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Krug RM, et al. Are the 5' ends of influenza viral mRNAs synthesized in vivo donated by host mRNAs? Cell. 1979;18:329–334. doi: 10.1016/0092-8674(79)90052-7. [DOI] [PubMed] [Google Scholar]
  • 94.Zhao C, et al. Nucleoside monophosphate complex structures of the endonuclease domain from the influenza virus polymerase PA subunit reveal the substrate binding site inside the catalytic center. J. Virol. 2009;83:9024–9030. doi: 10.1128/JVI.00911-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.He X, et al. Crystal structure of the polymerase PA(C)-PB1(N) complex from an avian influenza H5N1 virus. Nature. 2008;454:1123–1126. doi: 10.1038/nature07120. [DOI] [PubMed] [Google Scholar]
  • 96.Obayashi E, et al. The structural basis for an essential subunit interaction in influenza virus RNA polymerase. Nature. 2008;454:1127–1131. doi: 10.1038/nature07225. [DOI] [PubMed] [Google Scholar]
  • 97.Sugiyama K, et al. Structural insight into the essential PB1-PB2 subunit contact of the influenza virus RNA polymerase. EMBO J. 2009;28:1803–1811. doi: 10.1038/emboj.2009.138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Guilligay D, et al. The structural basis for cap binding by influenza virus polymerase subunit PB2. Nat. Struct. Mol. Biol. 2008;15:500–506. doi: 10.1038/nsmb.1421. [DOI] [PubMed] [Google Scholar]
  • 99.Kuzuhara T, et al. Structural basis of the influenza A virus RNA polymerase PB2 RNA-binding domain containing the pathogenicity-determinant lysine 627 residue. J. Biol. Chem. 2009;284:6855–6860. doi: 10.1074/jbc.C800224200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Tarendeau F, et al. Structure and nuclear import function of the C-terminal domain of influenza virus polymerase PB2 subunit. Nat. Struct. Mol. Biol. 2007;14:229–233. doi: 10.1038/nsmb1212. [DOI] [PubMed] [Google Scholar]
  • 101.Iwai Y, et al. Anti-influenza activity of marchantins, macrocyclic bisbibenzyls contained in liverworts. PLoS One. 2011;6:e19825. doi: 10.1371/journal.pone.0019825. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Ghanem A, et al. Peptide-mediated interference with influenza A virus polymerase. J. Virol. 2007;81:7801–7804. doi: 10.1128/JVI.00724-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Martin K, Heleniust A. Nuclear transport of influenza virus ribonucleoproteins: the viral matrix protein (M1) promotes export and inhibits import. Cell. 1991;67:117–130. doi: 10.1016/0092-8674(91)90576-k. [DOI] [PubMed] [Google Scholar]
  • 104.Gómez-Puertas P, et al. Influenza virus matrix protein is the major driving force in virus budding. J. Virol. 2000;74:11538–11547. doi: 10.1128/jvi.74.24.11538-11547.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Sha B, Luo M. Structure of a bifunctional membrane-RNA binding protein, influenza virus matrix protein M1. Nat. Struct. Biol. 1997;4:239–244. doi: 10.1038/nsb0397-239. [DOI] [PubMed] [Google Scholar]
  • 106.Arzt S, et al. Combined results from solution studies on intact influenza virus M1 protein and from a new crystal form of its N-terminal domain show that M1 is an elongated monomer. Virology. 2001;279:439–446. doi: 10.1006/viro.2000.0727. [DOI] [PubMed] [Google Scholar]
  • 107.Liu T, Ye Z. Restriction of viral replication by mutation of the influenza virus matrix protein. J. Virol. 2002;76:13055–13061. doi: 10.1128/JVI.76.24.13055-13061.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Akarsu H, et al. Crystal structure of the M1 protein-binding domain of the influenza A virus nuclear export protein (NEP/NS2). EMBO J. 2003;22:4646–4655. doi: 10.1093/emboj/cdg449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Hale BG, et al. The multifunctional NS1 protein of influenza A viruses. J. Gen. Virol. 2008;89:2359–2376. doi: 10.1099/vir.0.2008/004606-0. [DOI] [PubMed] [Google Scholar]
  • 110.Cheng A, et al. Structural basis for dsRNA recognition by NS1 protein of influenza A virus. Cell Res. 2009;19:187–195. doi: 10.1038/cr.2008.288. [DOI] [PubMed] [Google Scholar]
  • 111.Das K, et al. Structural basis for suppression of a host antiviral response by influenza A virus. Proc. Natl Acad. Sci. U S A. 2008;105:13093–13098. doi: 10.1073/pnas.0805213105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Twu KY, et al. The CPSF30 binding site on the NS1A protein of influenza A virus is a potential antiviral target. J. Virol. 2006;80:3957–3965. doi: 10.1128/JVI.80.8.3957-3965.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Basu D, et al. Novel influenza virus NS1 antagonists block replication and restore innate immune function. J. Virol. 2009;83:1881–1891. doi: 10.1128/JVI.01805-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Talon J, et al. Influenza A and B viruses expressing altered NS1 proteins: a vaccine approach. Proc. Natl Acad. Sci. U S A. 2000;97:4309–4314. doi: 10.1073/pnas.070525997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Akarsu H, et al. Structure-based design of NS2 mutants for attenuated influenza A virus vaccines. Virus Res. 2011;155:240–248. doi: 10.1016/j.virusres.2010.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.McKimm-Breschkin JL. Resistance of influenza viruses to neuraminidase inhibitors—a review. Antiviral. Res. 2000;47:1–17. doi: 10.1016/s0166-3542(00)00103-0. [DOI] [PubMed] [Google Scholar]

RESOURCES