Sunteți pe pagina 1din 21

176

VARGHESE DRUG DEVELOPMENT RESEARCH 46:176196 (1999)

Research Overview

Development of Neuraminidase Inhibitors as Anti-Influenza Virus Drugs


Joseph N. Varghese*
Biomolecular Research Institute, Parkville, Victoria, Australia

Strategy, Management and Health Policy Venture Capital Preclinical Enabling Research Technology Preclinical Development Toxicology, Formulation Drug Delivery, Pharmacokinetics Clinical Development Phases I-III Regulatory, Quality, Manufacturing Postmarketing Phase IV

ABSTRACT Structure-based design and synthesis of potent influenza virus neuraminidase inhibitors are now being evaluated in human trials as anti-influenza virus drugs. The first drug of this class, Relenza (Zanamivir/GG167), is now awaiting pharmaceutical evaluation and registration in Australia, Europe, and North America for both treatment and prophylaxis of influenza. The target for the drug is the active site of neuraminidase, which is a pocket that has been totally conserved in both Type A and B influenza in all known subtypes of influenza (animal and human). Mutations in residues that surround this conserved pocket allow the virus to escape binding to circulating antibodies that recognise the molecular surface around the active site of the wild-type virus. High-affinity neuraminidase inhibitors have been designed that interact only with the conserved active site residues. The design of these sialic acid analogues was based on the crystal structure of influenza virus neuraminidase and its complex with N-acetyl neuraminic acid (sialic acid) and 2-deoxy-2,3-dehydro-N-acetyl neuraminic acid. These novel inhibitors are highly specific for influenza neuraminidase, and have been shown to inhibit influenza virus replication in both cell culture and animal models. The development of drugs against a rapidly mutating organism like influenza has to address to the possibility of emerging drug resistance. This is examined in the light of drug resistant mutants selected after in vitro passaging of virus in the presence of neuraminidase inhibitors. Drug Dev. Res. 46:176 196, 1999. 1999 Wiley-Liss, Inc.
Key words: sialidase; Zanamivir; Relenza

INTRODUCTION

Influenza, a viral infection of the upper respiratory tract in humans, has plagued mankind since the dawn of history. The disease in modern times continues to affect a significant proportion of the population irrespective of age or previous infection history. These periodic epidemics that reinfect otherwise healthy people have devastated communities worldwide. Some pandemics like the 19171919 Spanish flu were responsible for the deaths of tens of millions of people throughout the world. The origins, spread, and severity of influenza epidemics have been a puzzle that has only in the last two decades been adequately addressed. The virus is spread by aerosols produced by infected animals, and the continual production of new strains of the virus results in reinfection [reviewed by Kilbourne, 1987]. There are three types of influenza virus classified
1999 Wiley-Liss, Inc.

on their serological cross-reactivity with viral matrix proteins and soluble nucleoprotein (A, B, and C). Only type A and B are known to cause severe human disease. Type B is only found in humans, while type A has a natural reservoir in birds and some mammals like pigs and horses [Webster et al., 1993]. Influenza, an orthomyxovirus, is a 100 nm lipid-enveloped virus containing an eight-segment negative single-stranded genome [Lamb, 1989]. Two of the segments code for the surfaces glycoproteins hemagglutinin HA (which binds to terminal sialic acid) and neuraminidase NA (which cleaves terminal sialic acid) which appear as spikes protruding out of the viral enve-

*Correspondence to: Joseph N. Varghese, Biomolecular Research Institute, 343 Royal Parade, Parkville, Victoria 3052, Australia. E-mail: jose@bioresi.com.au

ANTI-VIRAL DRUGS FOR INFLUENZA

177

lope. The viral target in humans is the upper respiratory tract epithelial cells. Replication begins with penetration of the virion through the mucin layer covering the epithelial surface, followed by attachment to the viral receptor by the HA. Penetration of the cell is achieved by endocytosis and the virion core is released after the fusion of the virion and vesicle membrane, mediated by the HA. Fusion is enabled by a conformational change in the HA, made possible by lowering the pH of the endosome by the M2 ion channel protein. Following replication, the progeny virions are released by budding off the cell membrane [Murphy and Bang, 1952; Compans and Dimmons, 1969]. The release of virons occurs 8 h postinfection and the onset of infection is sudden, resulting in pyroxia, muscular and joint pain, and a dry cough [Murphy and Webster, 1990]. Virus shedding continues for up to a week, when a rise in virus-specific antibody clears the virus from the host. The vulnerability of the host succumbing to viremea during this week of rising viral titre is dependent on interferon induction [Ennis and Meager, 1981] 48 h postinfection, which attenuates viral replication until the cell-mediated immune response begins to clear the virus. The severity of the illness is thought to depend on the level of cross protection arising from antibodies raised from previous influenza infections [Fox et al., 1982]. The course of the illness can be dehabilitating, and no effective treatment is available at present to halt the progression of the disease. Death can result for susceptible populations (neonate and elderly), primarily as a result of secondary infections [Sprenger et al., 1993]. In this article, a structural basis for the continual emergence of new influenza strains is discussed, and the reasons why current vaccines against influenza fail to protect against all strains of influenza. The discovery in the early 1980s of the molecular structure of influenza virus NA, the detailed atomic description of the active site of the molecule, and the exploitation of its structural conservation is discussed in terms of the design of potent NA inhibitors. The therapeutic use of these inhibitors as anti-viral drugs against influenza virus infections shall be examined and the possibility of drug resistance to NA inhibitors shall be addressed.
Antigenic Variation

host. This mechanism, termed Antigenic Drift, accounts for most of the strain variation within a particular subtype of influenza. However, infrequently a mutation arises by genetic re-assortment of viruses from different animal hosts (Antigenic Shift), whereby an entirely new gene for one of the surface glycoproteins is generated which is significantly different (~50%) in amino acid sequence of the parent virus. This is the mechanism by which new subtypes of influenza arise which are primarily responsible for the major pandemics that occur. Such a great change in the antigenic surface of these glycoproteins results in little or no cross-protection by circulating antibodies to previous influenza infections, leading to the global spread of severe influenza and the high mortality rate such pandemics inflict. Strains of influenza virus are classified by type (A, B, or C), geographic location, date of original isolation, and the subtype of the HA and NA antigens. There exist nine known subtypes (N1 to N9) of NA and 13 known subtypes (H1 to H13) of HA from influenza A in all animal populations. Two NA (N1 and N2) and three HA (H1, H2, and H3) subtypes of influenza A have occurred in strains that have infected humans since 1933, when isolates were first characterised [Smith et al., 1933]. Prior to 1933 there is indirect evidence of antigenic shift occurring in human populations [Beveridge, 1977]. The N1 subtype was associated with virus isolated between 1933 and 1957, after which time the N2 subtype appeared in the Asian influenza. No major change in the structure of NA has occurred since, although the HA subtype has changed from H2 to H3 in 1968 in the Hong Kong pandemic, and H1N1 reappeared in 1978 as the Russian pandemic. Recently (1998) an avian strain, H5N1, caused a brief outbreak in humans in Hong Kong that was quickly contained, as it appears that infections were only transmitted from avian to human and not human to human [Yeun et al., 1998; Subbaroa et al., 1998; Claas et al., 1998]. Influenza B, which infects only human hosts, has only one subtype, but like type A undergoes continual antigenic drift.
Current Therapeutics and Vaccines

The plethora of different strains of the virus that are responsible for continued reinfection of the virus in humans is primarily related to mutations in the viral genes of two surface glycoproteins, HA and NA [Smith and Palase, 1989]. The current paradigm for this genetic variation [Skehel, 1974; Webster and Laver, 1975] is that these mutations arise primarily from incremental changes in the amino acid sequences of these glycoproteins by selection pressure of the immune system of the infected

Amantadine and rimantadine are the only class of drugs that have been approved for therapy [Douglas, 1990; Wintermeyer and Nahata, 1995]. At high concentration (>50 g/mL), amantadine is thought to buffer the pH of the endosome and prevent the conformational change of the HA necessary for fusion. Drug-resistant mutants arise where the hemagglutinin trimers are thought to be less stable than the wild type [Daniels et al., 1985]. At low concentrations (<1 g/mL), amantadine blocks the activity of the viral M2 ion channel protein, which plays a role in virus uncoating and

178

VARGHESE

glycoprotein maturation [Hay et al., 1993]. Amantadine prevents fusion by altering the ability of the virus to change the ion balance [Pinto et al., 1992] within the endosome and the trans-Golgi network. However, amantadine rapidly gives rise to drug-resistant strains by mutations in the HA and the M2 protein that circumvent or block the activity of the drug. Amantidine and rimantadine can cause adverse CNS complaints and other complications [Srange et al., 1991; Monto et al., 1995], although rimantadine can be taken at lower doses, in part related to lower plasma levels, and much less dependence on renal secretion for elimination than amantidine. Rimantadine has been shown to lead to drug resistance in humans 2 days posttreatment [Hayden, 1993]. Influenza virus vaccines [Tyrrell, 1980; Murphy and Webster, 1990] prepared from killed (formalin inactivated) virus are used currently worldwide as the only prophylactic treatment available against the disease. Killed vaccines contain whole virus or fractionated subunits. These vaccines attempt to incorporate antigens from influenza strains that are predicted to circulate in the community during the next season. They also confer immunity to viral strains that are closely related to the strains incorporated into the vaccine. However, the susceptibility of the virus to neutralisation by antibodies raised by immunisation is reduced by antigenic drift. For example, commercial vaccines containing inactivated A/Beijing/353/89 H3N2 strain circulating among humans from 1990 to 1993 provided significantly less protection [Chakraverty et al., 1993] against the antigenic drift [Both et al., 1983] variant A/Georgia/03/93. A more radical vaccination treatment involving the direct injection into muscle of plasmid DNA encoding HA, nuclear protein, and M1 influenza viral proteins has successfully been attempted in animal models [Donnelly et al., 1995]. This is thought to involve the incorporation of the plasmid into muscle cells, and the production of the viral proteins cell-mediated immune (CMI) response elicited offers cross-protection due to reactivity with the more conserved epitopes of the HA and nuclear proteins, thus overcoming antigenic drift in the virus. The success of the method arises from a CMI response to segments of conserved amino acids in these proteins. Problems arising from possible autoimmune response to DNA, and in ensuring the exogenous DNA is not integrated into the cell genome or sensitive cell lines elsewhere in the host, have precluded testing in humans at present. Furthermore, mutations in the nuclear proteins of influenza (which have been relatively stable to date) arising from CMI selection pressure could undermine the efficacy of this technique.
VIRAL NEURAMINIDASE AS AN ANTI-VIRAL TARGET

blood cells once agglutinated by influenza virus could not again be agglutinated by either the eluted virus or fresh virus preparations. This activity is now attributed to NA, which is one of the two integral membrane glycoproteins of influenza virus [for review, see Colman, 1994; Varghese, 1997] Neuraminidase is an exoglycosidase which destroys the hemagglutinin receptor by cleaving the -ketosidic linkage of terminal sialic acid (N-actylneuraminic acid (Neu5Ac)) to an adjacent sugar [Klenk et al., 1955; Gottschalk, 1957]. Viral HA binds specifically to Neu5Accontaining receptors on the surface of susceptible cells [Rogers and Paulson, 1983]. NA, which also removes terminal sialic acid from a range of glycoconjugates, plays an important, but not completely understood, role in the viral replication cycle. Without NA activity, viruses [Burnett, 1947] were thought to be immobilised by mucosal secretions in the upper respiratory tract. By removing terminal sialic acid from the sialic acid-rich mucous layer protecting target cells [Gottschalk, 1957, 1958], NA could facilitate penetration of the virus to the cell surface. It has been shown that neuraminidase-deficient virus [Liu and Air, 1993] can still replicate in vivo, albeit at a much reduced rate [Liu et al., 1995]. This shows that NA does not play an essential role in viral entry, replication, assembly, or budding in mice. It has an important role in facilitating the spread of the infection by preventing aggregation at the cell surface and possible immobilisation in the mucin by HA. During virus replication, the freshly synthesised viral glycoproteins have to be desialylated to prevent self-aggregation at the infected host cell surface by HA binding to terminal sialic acid on these glycoproteins. Finally, on elution of progeny virions from infected cells, NA activity is required to facilitate viral escape from the cell surface. Inactivation or inhibition of NA during budding has been observed to result in aggregation of virons on the cell surface [Palese et al., 1974; Palese and Compans, 1976; Griffin and Compans, 1979]. Inhibition of this glycohydrolase thus provides a means of controlling this disease, while the number of infected cells is low, by slowing the rate of viral attachment and subsequent release of progeny virons. This would allow the host immune system to eliminate the virus.
Molecular Structure of Neuraminidase

Enzyme activity on the surface of influenza virus was first detected by Hirst [1942], who observed that red

There are between 50 to 100 NA spikes per virion [Bucher and Palese, 1975], which are approximately 10% of the visible spikes projecting out of the surface of the virion [White, 1974]. These spikes can be removed from the virus by treatment with detergent [Laver, 1963]. Electron microscopic images of the NA spikes [Laver and Valentine, 1969] reveal a mushroom-shaped molecule made up of a box-like head of about 80 80 40 A, with

ANTI-VIRAL DRUGS FOR INFLUENZA

179

a narrow centrally attached stalk (15 A wide and 100 A long) which terminates in a hydrophobic knob anchored in the viral envelope. The detergent-released spikes can be digested by pronase to release the NA heads, which retain full antigenic and enzyme activity [Drzenick et al., 1968; Laver, 1978]. The molecule was found to be a tetramer of molecular weight 240,000, reducing to 200,000 when treated with pronase [Blok et al., 1982]. The X-ray 3-dimensional molecular structure of NA heads was determined [Varghese et al., 1983] for two N2 subtypes, A/Tokyo/3/67 and A/RI/5+/57. The structure of A/Tokyo/3/67 N2 has since been refined [Varghese and Colman, 1991] to higher resolution (2.2 A), as have the structures of two avian N9 subtypes [Baker et al., 1987; Tulip et al., 1992a] and influenza type B [Burmeister et al., 1991]. They were found to have an identical protein fold with 60 residues (including 16 conserved cysteine residues) being invariant. Bacterial sialidases from Salmonella [Crennell et al., 1993] and cholera [Crennell et al., 1994] have homologous structures to influenza NA, but only a few of the residues in the active site are structurally invariant. The protein fold consists of a symmetric arrangement of six four-stranded antiparallel -sheets arranged as the blades of a propeller, the propeller axis being approximately parallel to but tilted away from the circular 4-fold axis of the tetramer. This tilt angle varies between the known subtypes. There is a calcium-binding site connected to the active site via conserved residues. The functional role of calcium in the structure is unknown, although calcium has been shown to be necessary for NA activity [Baker and Gandhi, 1976] and it is essential for the thermostablity of the molecule [Burmeister et al., 1994].
Carbohydrate Structure

for the 1918 pandemic, selected by serial passaging by mouse brain) binds and sequesters plasminogen. This leads to higher local concentrations of this protease precursor and, thus, increasing cleavage of HA. This is a possible explanation for the pantropism of this strain [Kunkel et al., 1987]. The structural basis for this unusual function of NA appears to be the presence of a carboxylterminus lysine and the absence of the carbohydrate at Asn146. Goto and Kawkaoka [1998] suggest that this carbohydrate obstructs the binding of plasminogen to the terminal lysine, suppressing HA cleavage in cells other than its usual targets. This oligosaccaride is a complex sugar containing N-acetylgalactosamine [Ward et al., 1983] that is not found in any other of the known oligosaccarides of influenza virus glycoproteins, and is the only glycopeptide antigenically related to chick embryo host antigen [Ward et al., 1982]. The oligosaccharide appears as a spike emanating from the top of the monomer, forming a crystal contact with a neighbouring tetramer in crystals of A/Tokyo/3/67 NA. The four carbohydrate spikes of a tetramer form an open barrel structure of eight carbohydrate chains with the neighbouring tetramer, with no apparent inter-carbohydrate contacts [Varghese and Colman 1991]. This oligosaccharide may play an important but as yet unidentified role in NA structure or activity.
A Second Sialic Acid Binding Site

Carbohydrates at four N-linked glycosylation sites were observed in N2 NA at residues 86, 146, 200, and 234 in the X-ray structure. Two N-actylglucosamines were resolved at Asp86 and Asp234, both at the bottom surface of the monomer. The carbohydrate at Asp200 consists of at least eight sugar residues with linkages consistent with known mannose-rich simple N-linked carbohydrate [Wagh and Bahl, 1981]. This oligosaccharide emerges from the side of the monomer and covers a neighbouring subunit (see Fig. 1). The oligosaccharide site at Asn146 is the most conserved of all neuraminidase glycosylation sites, except that of the neurovirulent virus A/WSN/33 (H1N1) [Francis and Moore, 1940; Hitte and Nayak, 1982]. The absence of this glycosylation site in A/WSN/33 has been shown to confer neurovirulence in mice [Li et al., 1993]. It has recently been shown [Goto and Kawaoka, 1998] that the NA of A/WSN/33 (a descendent of the virus responsible

Hemagglutination activity has been reported for the N9 subtype of NA of influenza type A virus [Laver et al., 1984] at 4C, which was not related to aberrant neuraminidase activity, but was associated with a second Neu5Ac binding site on the surface of NA away from the active site [Webster et al., 1987]. The residues on the surface of NA responsible for the haemabsorbing (HB) activity have been identified by monoclonal variants which lost the capacity to bind red blood cells [Webster et al., 1987], and the activity was successfully transferred to the N2 subtype of NA [Nuss and Air, 1991] by sitedirected mutagenesis. Furthermore, it was shown that N9 NA activity did not remove the putative sialic acidrelated moiety that bound red blood cells to this HB site [Air and Laver, 1995], as the agglutination was restored on cooling to 4C. An HB site has also been discovered on the NA of A/FPV/Rostock/34 H7N1 [Hausmann et al., 1995], which appears to have the same location on the NA surface as in N9 NA. Recently, Varghese and co-workers [1997] located the HB site on N9 by X-ray diffraction (Fig. 2). They have shown that six residues on three separate loops of N9 NA interact directly with the sialic acid in the second sialic acid binding site. These are the three serine residues (367, 370, and 372) in the loop containing residues 367-SIASRS-372, Asn400, and Trp403 in the loop con-

180

VARGHESE

Fig. 1. A CPK atomic model of the upper surface of an influenza virus neuraminidase N2 tetrameric head [Varghese and Colman, 1991]. The long spike-like structures emerging from the face of the tetramer towards the right are carbohydrate moieties attached to Asn146. The dark spheres represent those residues that have been totally conserved in all known

influenza viruses, and they cluster near the corners of the molecule where the active site of the enzyme is situated. It is this site to which the neuraminidase inhibitors are targeted. Residues that have mutated since 1933 are represented by lighter shaded spheres, and they surround the conserved active site.

taining 400-NTSW-403, and Lys432 in a third loop. The sequence variation at residues 400 and 432 in NAs with known HB activity indicates that these residues are not essential for HB activity, and the triple serine SxxSxS loop of 367 to 372 and Trp403 are a minimal signature for HB binding in NA.

All known strains from N1 to N9 which carry the signature are avian (or equine), with the exception of the two human N2 strains, RI5+/57 and Leningrad/134/57 [Varghese et al., 1997]. All avian sequences appear to carry this signature, and in general human and swine NAs do not, indicating that the HB site has a functional role in

ANTI-VIRAL DRUGS FOR INFLUENZA

181

Fig. 2. A surface rendered image [Nicholls et al., 1991] of a subunit of influenza neuraminidase N9. Shown in darker grey are residues conserved in all influenza viruses that interact with sialic acid (twist-boat) in the

catalytic site. In dark grey are also residues that are conserved in all avian strains that interact with sialic acid (chair) in haemabsorption (HB) site.

avian influenza. It has been shown [Air and Laver, 1995] that the HB site on A/tern/Australia/G70C/75 NA binds a moiety on red cells that is not cleavable by A/tern/Australia/G70C/75 NA. This is a consequence of the moiety either being Neu5Ac in a noncleavable linkage, or something other than Neu5Ac. The biological significance of this HB site has not been determined, but the above re-

sults would indicate that it may function as a lectin to some avian cell receptor, which is unaffected by influenza NA activity. Influenza is normally an asymptomatic infection in birds [Alexander, 1986], but replicates preferentially in the cells lining the intestinal tract of waterfowl [Slemons et al., 1974; Hinshaw et al., 1980], where all of the different subtypes of influenza A have been iso-

182

VARGHESE

lated. This has led to the proposition that waterfowl are the primary vectors for the global spread of the disease [Webster et al., 1992] through faecal droppings. This avirulent adaptation of the virus to avian species enables it to survive and persist in a vast global reservoir. If this HB signature relates to some as-yet unidentified role in avian influenza, then it could be inferred that the first human 1957 pandemic N2 strains carrying the HB signature (RI5+/57 and Leningrad/134/57) were most probably derived from a genetic reassortment event with an avian strain [Skehel, 1974]. This HB signature has since disappeared under antigenic drift in post-1957 human and swine stains, indicating that it serves no biological function in pathogenesis of influenza in humans and pigs.
Antigenic Variation in Neuraminidase Structures

Comparison of all known sequences of approximately 390 residues of the NA globular head [Varghese and Colman, 1991], indicates that only 54 (excluding 16 conserved cysteine residues) are invariant (Fig. 1). Apart from 21 residues involved in preserving the structural integrity of the molecule [Varghese and Colman, 1991], the main clustering of these invariant residues is within the enzyme active site (Fig. 1), where 17 are in the active site and 16 neighbouring the active site. This is a cavity on the upper surface of the molecule into which sialic acid has been observed to bind [Colman et al., 1983; Varghese et al., 1992; Burmeister et al., 1991]. Excluding the active site pocket, strain variation occurs over the entire surface of the NA heads. The active site was found to be in a pocket of totally conserved (over all animal subtypes) residues [Colman et al., 1983]. In this way the enzyme active site pocket is surrounded by highly variable surface residues that prevent immune recognition by antibody molecules of the active site [Colman, 1994]. X-ray diffraction studies of NAantibody complexes have shown that the footprint of an antibody in the complex is larger than the exposed surface of the conserved region of the active site [Colman et al., 1987, 1989; Malby et al., 1994]. These structural results indicate that antibodies are unable to exert mutational pressure on the conserved active site because they cannot bind there without engaging strain variable residues as part of the binding surface [Colman, 1997]. However, since antibodies bind to strain variable elements of the structure, the virus can overcome host immune pressure by point mutations of these residues that do not have a catalytic or structural role [Varghese, 1988; Tulip et al., 1992b], but are able to disrupt the antigenantibody binding interface. The rapid emergence of these escape mutants explains the failure to produce a universal vaccine for influenza.
Enzyme Active Site

The structures of N-acetyl neuraminic acid (sialic acid Neu5Ac) and the 2-deoxy-2,3-dehydro-N-acetyl

neuraminic acid (Neu5Ac2en) inhibitor complexed with N2 NA [Varghese et al., 1992] revealed the nature of the interactions of the molecules in the active site pocket (Fig. 3). Sialic acid binds in the active site in the -anomer, and in a distorted half-chair conformation through the same face as used in its interaction with HA [Weis et al., 1998]. The carboxylate group of the sugar interacts with three guanidinium groups of arginine residues 118, 292, and 371 and has an equatorial conformation with respect to the sugar ring (this group, axial toward the floor in the undistorted structure, is probably held equatorial by interactions with these arginine residues). The NH group of the 5-NAcetyl side chain interacts with the floor of the active site cavity, via a bound water molecule. The oxygen of the 5-N-Acetyl side chain is hydrogen bonded to the N of Arg152, while the methyl group lies in a hydrophobic pocket near Ile222 and Trp178. The last two hydroxyl groups of the 6-glycerol sidechain are hydrogen bonded to carboxylate oxygens of Glu276 and the 4-hydroxyl is directed to a carboxylate oxygen of Glu119, and the glycosidic oxygen O2 interacts with a carboxylate oxygen of Asp151. Similar binding of sialic acid in the active site was observed in type B [Burmeister et al., 1991]. Comparison of active sites of N2, N9, and type B NA [Varghese et al., 1995] show there are no significant differences between active site orientations, except for some minor displacements of Arg224 and Glu276, where the major interactions with the 6-glycerol group of sialic acid occur. However, there are some differences in the water structure in the active sites of the different subtypes. Comparison of the active sites of influenza neuraminidases and bacterial sialidases [Crennell et al., 1993, 1994] indicates that there is considerable conservation of the catalytic site at the carboxylate-binding end. The residues Asp151, Arg118, Glu277, Arg292, Val or Ile349, Arg371, Try406, and Glu425 are conserved over all known viral and bacterial strains. The argininyl residues 118, 292, and 371 position the 2-carboxylate group and the Val(or Ile)349, Glu425 and Glu 277 are important in positioning the triarginyl cluster. Asp151 and Tyr406 are presumably important in bond cleavage, but the precise mechanism is still unclear. These eight residues (Fig. 4) are thus most likely to be conserved in all neuraminidases. Differences between viral, bacterial, and mammalian NA structures may correspond with the different role these enzymes have in vivo. These differences are likely to be in the interactions of the 6-glycerol, 5-Nacetyl, and 4-OH groups of silaic acid. In influenza, the turnover rate must be balanced against the requirement to maintain sufficient sialic acid at the cell surface to enable attachment via the HA. This balance may require some configuration of residues in the active site not di-

ANTI-VIRAL DRUGS FOR INFLUENZA

183

Fig. 3. A ball and stick representation [Kraulis, 1991] of sialic acid (Neu5Ac, in the centre of the diagram) bound in the active site of neuraminidase N2 [Varghese et al., 1992]. The tube represents the back-

bone of the protein, and black and grey spheres represent oxygen and nitrogen atoms. Thin lines represent bonding interactions between sialic acid and the conserved amino acid side-chains of the active site.

rectly responsible for catalysis, but only involved in the binding and release of sialic acid [Varghese et al., 1995].
NEURAMINIDASE INHIBITOR DESIGN

Earlier screening programs [Edmond et al., 1966] failed to identify potent inhibitors of viral NA. Neu5Ac2en (Fig. 5) was the first inhibitor synthesised [Meindl and Tuppy, 1969] whose Ki was in the micromolar range. This was based on the proposed transition state of the reaction, where the anomeric carbon (C2) bound to the ketosidic oxygen has a trigonal state. Several analogues of Neu5Ac2en were synthesised soon after, and the most potent of these, a trifluoracetyl derivative, had a Ki of only 0.8 M [Meindl et al., 1974]. While this compound

showed that in cell culture it retarded virus shedding [Palase and Compans, 1976; Palese et al., 1974], it failed as an effective antiviral agent in animals [Palese and Schulman, 1977]. The X-ray structure of Neu5Ac2en/neuraminidase complexes have been determined for N2 [Varghese et al., 1992], type B [Burmeister et al., 1993], and N9 [BossartWhitaker et al., 1993]. Neu5Ac2en binds in the active site of NA with the carboxylate oxygen atoms placed in the same location as the carboxylate of sialic acid. The 5N-Acetyl, 4-hydroxy, and 6 Neu5Ac2en-glycerol are positioned isosterically in the two molecules. An alternative approach to developing sialidase inhibitors has been made using synthetic thioglyco-

184

VARGHESE

Fig. 4. A ball-and-stick representation [Kraulis, 1991] of Zanamivir/ GG167 (centre in light grey) in the active site of neuraminidase, showing catalytic residues (black) that are conserved over both bacte-

rial and viral neuraminidases, and recognition residues (grey) conserved only in influenza virus neuraminidase. The tube is a representation of the protein backbone.

side-analogues of gangliosides such as Neu5Ac(2-S6)Glc(1-1)Ceramide [Suzuki et al., 1990], which inhibit different subtypes of human and animal influenza virus with a K i of up to 2.8 M. These metabolically stable ganglioside analogues contain a thioglycosidic linkage to the terminal neuraminic acid that resists cleavage by the enzyme. Several flavonoid neuraminidase inhibitors have been isolated from plant extracts [Nagai et al., 1990], one of which, 5,7,4-trihydroxy-8methoxyflavone was a more potent inhibitor than

Neu5Ac2en. Recently, in vivo anti-influenza virus activity of a Kampo (Japanese herbal medicine) preparation has shown promising results in inhibiting influenza virus replication in mice [Nagai and Yamada, 1994], but the mode of action of these compounds is unclear.
Enzyme Mechanism

The similar positioning of the carboxylate oxygens and ring of Neu5Ac2en and sialic acid suggests that Neu5Ac2en is probably a transition state analogue. As

ANTI-VIRAL DRUGS FOR INFLUENZA

185

Fig. 5. Influenza neuraminidase inhibitors: 1) Sialic Acid, NAcetylneuraminic acid (Neu5Ac); 2) 2-deoxy-2,3-dehydro-Nacetylneuraminic acid (Neu5Ac2en); 3) 4-amino-Neu5Ac2en; 4) Zanamivir, 4-guanidino-Neu5Ac2en; 5) 5-N-acetyl-4-guanidino-6methyl(propyl) carboxamide-4,5-dihydro-2H-pyran-2-carboxylic acid; 6) 5-N-acetyl-4-amino-6-diethyl carboxamide-4,5-dihydro-2H-pyran-2-carboxylic acid; 7) GS4071, 4-N-acetyl-5-amino-3-(1-ethylpropoxy)-1cyclohexene-1-carboxylic acid.

topic effects [Chong et al., 1992] support the existence of a oxycarbonium ion intermediate. However, the structural basis for NA activity is still unclear. It has been suggested [Varghese et al., 1992] that the tyrosyl oxygen of Tyr406 assisted by the sialic acid carboxylate itself could stabilise the developing charge on the oxycarbonium ion intermediate. The reaction is completed by the activation of a water molecule by a deprotonated Asp151, and its attack on the carbonium, resulting in the formation of the anomer of sialic acid [Friebolin et al., 1980]. However, the pH activity profile of neuraminidase [Lentz et al., 1987; Chong et al., 1991; Burmeister et al., 1993] suggest a bell-shaped profile, indicating normal activity from a pH range of about 4.5 to 9, which would indicate that the role of Asp151 as the acid group in the catalysis is unclear. A nonspecific proton donor has been proposed [Taylor and von Itzstein, 1994], probably a water molecule as the acid group, with a deprotonated Tyr406 stabilising the oxycarbonium ion, and a proton transferred from water to the departing aglycon group. It has also been postulated that a proton elimination at C3 leads to the transformation of the oxycarbonium ion into Neu5Ac2en, which is produced irreversibly at low levels from sialic acid by the enzyme [Burmeister et al., 1993]. A SN1-type mechanism has been suggested [Taylor and von Itzstein, 1994] that is facilitated by an activated water molecule that can be expelled upon inhibitor binding. The catalytic mechanism could possibly proceed without an acid group, by the electrostatic potential of the enzyme lowering the barrier preventing the breaking of the ketosidic bond, and the solvent protonating the aglycon after release [Janakiraman et al., 1994]. Clearly, details of the enzyme mechanism have yet to be elucidated definitively, as structural considerations can only indicate that Tyr406 (and possibly Glu277) and the triarginyl cluster are essential in the enzyme mechanism, and that Asp151 could be implicated in it.
Inhibitor Design Principles

Neu5Ac2en has a higher affinity for NA than sialic acid, a mechanism of catalysis was proposed [Miller et al., 1978] that involves the distortion of the substrate by the formation of a oxycarbonium ion intermediate, which has a similar structure to Neu5Ac2en. However, the structural results from sialic acid/NA complexes suggest that the tighter binding of Neu5Ac2en arises more likely from the relaxation of the conformational strain arising from the transition from chair to boat of the pyranose ring of sialic acid in the active site [Varghese et al., 1992]. Evidence for a sialyl cation transition state by iso-

All the nearest neighbour interactions with sialic acid and Neu5Ac2en and the protein are with totally conserved amino acids. Thus, an inhibitor designed to bind only to the conserved active site residues of NA would inhibit NA activity across all strains of influenza. This would enable the development of an anti-viral drug that would affect the spread of viral replication potentially in three ways, i.e., transport through the protective mucosal layer, desialyation of freshly synthesised viral glycoproteins, and elution of progeny virions from infected cells. The development of potent inhibitors was based on the structural information of the N2 NA conserved active site and its complex with sialic acid and Neu5Ac2en. There are no reports of de novo molecules designed to fit into the cavity, and the most useful approach was to

186

VARGHESE

consider molecules that were structurally related to Neu5Ac2en. This involved the design of molecules that would bind isosterically to Neu5Ac2en, but which were modified to increase the number of favourable interactions with the protein. The method of Goodford [1985] enabled the calculation of favourable binding sites for a variety of chemical probes. The validity of this method was indicated by its methods ability to identify the positions of the carboxylate binding site of sialic acid as an energy minima for a carboxylate probe, and the successful prediction of known bound water sites in the active sites by a water probe. Utilising this methodology, predictions of energetically favourable substitutions to Neu5Ac2en were examined [von Itzstein et al., 1993]. A replacement of the hydroxyl at the 4-position of the pyranose ring of Neu5Ac2en by an amino group was identified by this procedure as an energetically favourable substitution. A protonated primary amine probe identified a favourable binding site of -16 kcal mol1 at this location, and in a pocket in the active site near two conserved glutamate residues, Glu119 and Glu227. This suggested that the substitution of the 4-hydroxyl group by an amino group would increase the overall binding interactions by forming a salt link with Glu119. Furthermore, the substitution of the 4-hydroxyl group with a much more bulky guanidinyl group would lead to even tighter binding as a result of lateral interactions of the terminal nitrogens of the guanidinyl group and the carboxylate groups of Glu119 and Glu277. This led to the design and syntheses of 3 and 4 (4amino-Neu5Ac2en and 4-guanidino-Neu5Ac2en in Fig. 5) [von Itzstein et al., 1993] (4 is GG167, Zanamivir, Relenza), which bound to A/Tokyo/3/67 with a Ki of 50 nM and 0.2 nM, respectively. These compounds were later shown by X-ray studies of 3 and 4 complexed with A/Tokyo/3/67 NA to bind close to that predicted by the design studies. However, details of the interactions of the guanidinyl group of 4 with the glutamic acid groups (Glu119 and Glu277) in the floor of the active site were slightly different. This was confirmed on a higher resolution X-ray study (Fig. 4) of 4 complexed with Tern N9 NA [Varghese et al., 1995]. One of the primary guanidinyl nitrogens of 4 is hydrogen bonded to the main chain oxygen at residue 178, a carbxylate oxygen of Glu227, and a water molecule. The other primary guanidinyl nitrogen interacts with the main chain oxygen of residues 178 and 151. The secondary guanidinyl nitrogen interacts with the carboxylate of Glu119 and Asp 151. The interactions with Glu119 lack the hydrogen bonding geometry (postulated in the design study), as the carboxylate group of Glu119 stacks parallel to the guanidinyl group, with its interactions being electrostatic and van der Waals in character. Furthermore, theoretical energy-minimised struc-

tures of the complex using AMBER [Pearlman et al., 1990] converged to the X-ray structure only if the protein nonhydrogen atom were kept rigid in the X-ray structure [Varghese et al., 1995]. Otherwise, this resulted in active site residues showing large distortions in their conformation. This is an example of the difficulty in correctly modelling even modest changes in the interactions of an inhibitor/active site complex. Zanamivir shows potent inhibition of NA activity in all known wild strains of influenza. Furthermore, it is very specific to influenza, as it shows weak inhibition to bacterial, para influenza, and mammalian neuraminidases [von Itzstein et al., 1993; Woods et al., 1993]. This is possibly due to the specific interactions of the 4-guanidino group within the subpocket of the active site of influenza NA that is not conserved in other neuraminidases. In bacterial neuraminidases [Crennell et al., 1993, 1994], this pocket is much smaller, and would prevent the binding of the 4-guanidino group in this region of the active site. This is consistent with the proposition that the interactions of sialic acid with the active site of NA are function-specific at the C4,C5, and C6 position of sialic acid, and that modification at these positions confers specificity to the target enzyme [Varghese et al., 1995]. Recently, a new class of potent NA inhibitors, exemplified by 5 and 6 (Fig. 5) has been reported in which hydrophobic substituents have replaced the glycerol moiety at the 6-position [Sollis et al., 1996; Smith et al., 1996; Taylor et al., 1998]. A small conformational change in the active site of NA occurs to enable these inhibitors to be accommodated (Fig. 6). Glu276 changes its position to form a salt link with Arg224, and thereby creates the necessary hydrophobic pocket for the carboxamide substituents. It has been proposed that the unexpected strong binding of these inhibitors is a result of the burial of hydrophobic surface area and salt-bridge formation in an environment of low dielectric. Taylor and co-workers [1998] have shown that there is a greater degree of distortion of the active site residues in influenza B NA than influenza A NA on binding with these carboxamide derivatives. This correlates with the decreased affinity for influenza B NA when compared with influenza A NA on binding to ligand. They have also shown by molecular dynamics calculations that the tendency for salt-bridge formation is greater in influenza A NA than influenza B NA, and this is a useful descriptor for the prediction of inhibitor potency. A carbocyclic analogue of sialic acid (GS4071, Ro640796) [Kim et al., 1997] (7 in Fig. 5), which has a hydrophobic group attached to the 6 position via an ether link, has also been shown to inhibit NA and virus replication in vitro. This inhibitor binds in a similar mode as the 6-carboximide derivatives of Neu5Ac2en [Varghese et al., 1998], by forming a hydrophobic pocket created

ANTI-VIRAL DRUGS FOR INFLUENZA

187

Fig. 6. A surface rendered image [Nicholls et al., 1991] of neuraminidase N9 with a ball and stick representation of (a) Zanamivir [Varghese et al., 1993] and (b) a carboxamide derivative of Neu5Ac2en in the active site [Taylor et al., 1998]. The change in conformation of Glu276 (grey

patch at the upper centre of diagram) from H-bond interactions with the glycerol group of GG167 to formation of a salt-link with Arg224 is shown by the creation of a hydrophic binding pocket in (b). Black spheres represent bound water molecules.

by the salt link of Glu276 and Arg224. It could be inferred, as with the carboxamide derivatives, that this compound would have a differential binding to influenza A and B neuraminidases. GS4071 has a similar low bioavailablity compared to GG167 [Li et al., 1998]; however, an ethyl ester linked to one of the carboxylate oxygens of GS4071 improves the oral bioavailablity following rapid conversion to the active form during gastrointestinal absorption. This prodrug, GS4104, when administered orally results in high and sustained systemic absorption in animal tests [Li et al., 1998] with a half-life of 5 h in most tissues. Whole body autoradiography in rats has shown a twofold higher concentration in the lungs compared to plasma 6 h postdose, and a 30-fold higher concentration at 24 h postdose [Eisenberg et al., 1997]. Furthermore, distribution in brain tissue was minimal. This would indicate that the GS4104 inhibitor is orally active and the inhibitor is currently undergoing human clinical trials for use as an anti-influenza therapeutic. Other approaches [Luo et al., 1995] to structurebased design of inhibitors have to date produced only mM inhibition of neuraminidase activity.
ANTIVIRAL ACTIVITY OF NEURAMINIDASE INHIBITORS

activity in vivo was not demonstrated [Palese and Schulman, 1977]. As a consequence of this, efforts were directed towards HA, which was then considered a better target for anti-influenza drugs. The interest in neuraminidase inhibitors as anti-influenza drugs has only been revived with the success of 4-guanidino-Neu5Ac2en and its analogues in attenuating viral titres in mice when administered directly into the lungs [von Itzstein et al., 1993; Ryan et al., 1994].
Inhibition In Vitro

In vitro inhibition of viral replication in tissue culture was demonstrated [Palese and Compans, 1976] for the tri-fluro derivative of Neu5Ac2en, but its antiviral

von Izstein and co-workers [1993] have shown that the 4-amino- and 4-guanidino-Neu5Ac2en inhibit influenza strains A/Singapore/1/57 and B/Victoria/102/95 in MDCK cells with IC50 values (the concentration required to inhibit plaque formation in MDCK cells by 50%) of 1.5 M and 0.065 M (4-amino) and 0.014 M and 0.005 M (4-guanidino), respectively. These IC50 values, in particular for the 4-guanidino compound, are well below those found for amantadine, ribavirin, and Neu5Ac2en. Furthermore, in comparison to Neu5Ac2en the 4guanidino-Neu5Ac2en inhibitor was 100-fold less active against human lysosomal sialidase and over 1,000-fold more active against a wide range of clinical isolates of influenza A and B, including amantidine- and rimantadine-resistant variants [Woods et al., 1993]. The inhibition of replication of virus in MDCK cells has been confirmed [Thomas et al., 1994] and Hayden

188

VARGHESE

and co-workers [1994] extended the inhibition studies to human respiratory epithelium cells in vitro, indicating high antiviral activity for strains of A(H1N1) and A(N3N2) isolates. They also found that delayed administration of the drug after viral replication was well established was associated with inhibition of virus replication. However, the viral titre was higher (1.3 log10 compared to 4.0 log10 at 10 g/ml concentration of the drug) for the delayed administration compared with the viral titre when the drug was present throughout the period of viral exposure. The clinical significance of this in the treatment of established infections has been explored in detail in clinical trails.
Administration and Inhibition In Vivo

The earlier work of Palese and Schulman [1977] indicated that the failure in inhibiting viral replication after intranasal and subcutaneous treatment by Neu5aC2en in mice would produce similar results for Neu5Ac2en analogues. This failure of Neu5Ac2en as an anti-viral treatment in animal models can now be ascribed to the rapid excretion of the compound [Nohle et al., 1982], thereby not delivering sufficient concentration of the inhibitor to the infected tissue. It has been shown [von Itzstein et al., 1993] that when Zanamivir is administrated intranasally in mice it is considerably more effective than when the drug is injected intra-peritoneally. This can be attributed to the localisation of sufficiently high concentrations of inhibitor in the lining of the nasal and respiratory epithelia, where influenza virus replication is believed to occur. Animal trials with ferrets challenged with influenza virus have shown that Zanamivir is effective in studies involving prophylactic administration of the drug [von Itzstein et al., 1993]. When the drug is administrated intranasally, 50 g/kg, twice daily, 1 day before infection with the virus and the succeeding 6 days, it substantially reduced virus titre in nasal washings and abolished fever that usually appears 3 days after infection. Results from a double-blind, randomised, placebocontrolled trial using this compound have been successful both for early treatment and prophylaxis of experimental inoculation of 166 adult human volunteers [Hayden et al., 1996] with 105 TCID50 of influenza A/Texas/91 (H1N1). For all dose groups combined, Zanamivir was 82% effective in preventing laboratory evidence of infection, and 95% effective in preventing febrile illness. These results indicate that NA is important for viral replication in humans. In separate randomised, double-blind studies in 38 centres in North America and 32 centres in Europe during the 19941995 influenza season, a total of 417 adults with influenza-like illness of less than 2 days duration were randomly assigned to the following groups: 6.4 mg of Zanamivir by intranasal spray plus 10 mg by inhala-

tion; 10 mg of Zanamivir plus placebo spray; or placebo by both routes [Hayden et al., 1997]. Treatments were self-administrated for 5 days. This study showed that in adult infections of both influenza A or B, Zanamivir directly administered to the respiratory tract was safe and reduced symptoms if begun early. Administration of Zanamivir by both groups reduced from 7 to 4 days the median time to the alleviation of major symptoms when compared with the placebo group, if treatment commenced within 30 h after onset of symptoms. Similar results were obtained in a double-blind randomised, placebo-controlled study completed in the southern hemisphere winter of 1997, on healthy and high risk patients, showing Zanamivir was well tolerated and alleviated symptoms of influenza 1.5 to 2.5 days earlier than placebo, and high risk patients had fewer complications and a reduction in the use of antibiotics with these complications [Silagy et al., 1998]. In a double-blind placebo controlled prophylaxis trail, Zanamivir (10 mg inhaled daily) was shown to be 67% (P < 0.001) effective in preventing laboratory-confirmed clinical influenza illness [Monto et al., 1998], and when confirmed illness was restricted to those with fever, the efficacy rose to 84% (P = 0.001). Efficacy in prevention of total influenza infection was 31% (P = 0.034), indicating subclinical infection, which allows immunity to develop in subsequent infections. Reduced single daily dosage of 6.4mg inhaled Zanamivir has also been shown to be protective against experimental infections in humans [Calfee et al., 1998]. Similar results have been obtained with the orally active prodrug GS4104. The parent drug, GS4071, has low bioavailability like GG167 (24%) compared to the high bioavailability of GS4104 [Li et al., 1998]. GS4104 undergoes rapid enzymatic conversion to GS4101 following gastrointestinal absorption. The oral administration of GS4104 in mice produced dose-dependent protective effects against the neurotropic influenza strain A/NWS/ 33 (H1N1), where 1 mg/kg/day afforded 100% protection [Mendel et al., 1998]. Increasing the dose to 10 mg/ kg/day afforded similar protection against A/Victoria/3/ 75 (H3N2) and B/HongKong/5/72. Similar results were obtained with the ferret model [Oxford et al., 1998]. No significant drug-related toxicity was observed after the administration of oral dosages of GS4104 of up to 800 mg/kg/day for 14 days in nonclinical toxicology studies with rats [Mendel et al., 1998]. In human clinical trials in the US, following oral treatment (75 mg or 150 mg) of GS4104 for 5 days, after febrile respiratory naturally acquired illness of 36 h or less, in a placebo-controlled, double-blind study of healthy non-immunized adults ages 1865 have shown the efficacy of the drug [Treanor et al., 1998]. The results showed a 30% reduction in both 75 mg (P = 0.0001) and 150 mg (P = 0.001) groups in duration of symptoms, 40%

ANTI-VIRAL DRUGS FOR INFLUENZA

189

reduction (both 75 mg, P = 0.0001; and 150 mg, P = 0.001) in severity of symptoms, and a 50% reduction (P = 0.017) of secondary complications. Similar results were obtained for treatment of naturally acquired influenza during the 19971998 influenza season in Canada, Europe, and China [Aoki et al., 1998], and no serious adverse events were reported. A study of the long-term prophylaxis of natural influenza [Hayden et al., 1998] showed an overall protective efficacy of 74% against influenza illness when 75 mg was taken once or twice daily during the period of local influenza activity. Long-term administration of once daily GS4104 was found to be safe and well tolerated.
DRUG RESISTANCE

Although the active site of influenza virus has been conserved in all known field strains of the virus, the possibility of drug resistance needs to be addressed. NA inhibitors, if used extensively as anti-influenza drugs, will apply selection pressure on the active site residues for the first time in its evolutionary history. Experience with influenza and other viruses, in particular HIV, have shown [Kimberlin et al., 1995] that drug-resistant mutants arise very rapidly, resulting in the effectiveness of antiviral drugs being short-lived. One attempted solution to the problem is the use of several drugs during therapy [Madren et al., 1995], making it more difficult for the virus to develop resistance. In the case of influenza virus, to date there have been no reports of drug resistance to Zanamivir from field strains [Osterhaus et al., 1998], except in an immunocompromised host [Gubereva et al., 1998]. However, it has recently been reported [McKimmBreschkin et al., 1994; Gubareva et al., 1995] that 4guanidino-Neu5Ac2en-resistant mutants can arise from multiple serial passages of virus in MDCK cells in the presence of the inhibitor. It has been demonstrated [McKimm-Breschkin et al., 1996] that almost all of the mutations arise in the HA receptor binding site and not on the NA. These altered HA variants, which have weaker binding to HA receptors, appear to arise as a result of increased inhibition of the NA by the drug. This is consistent with the earlier proposition that the rate of desialylation of receptor is critically related to the rate of attachment to receptor for the virus infection and elution [Varghese et al., 1995]. The decreased activity of the NA by the drug selects HA mutants with decreased binding to receptors.
E119G Neuraminidase Variant

cine in Tern N9. The crystal structure of this mutant and its complex (Fig. 7) with 4-guanidino-Neu5Ac2en has been determined [Blick et al., 1995]. The structures suggest that the decrease in inhibitor binding arises from the loss of stabilising interaction of the 4-guanidino group of Zanamivir [Varghese et al., 1995] and alterations in the solvent structure of the active site. This alteration arises from a water molecule that binds near the location of one of the carboxylate oxygens of the glutamic acid in the wild-type molecule. The location of the 4-guanidinoNeu5Ac2en drug in the complex with the mutant enzyme is isosteric compared to the drug/wild-type complex [Varghese et al., 1995], the only differences being the interactions with residue 119. This variant is, however, inherently thermally unstable, as measured both by enzyme activity and NC10 monoclonal antibody reactivity [Sahasrabudhe et al., 1998]. Substrate, inhibitor, or monoclonal antibodies stabilised the NA against all methods of denaturation. These results suggests that the instability of the variant is primarily due to the level of polypeptide chain folding rather than the level of association of monomers and tetramers. Furthermore, the presence of high level of substrate, either cell- or virus-associated, may be sufficient to stabilise the NA during virus replication.
R276K Neuraminidase Variant

A Zanamivir-resistant NA mutation has been isolated [Blick et al., 1995; Gubareva et al., 1995] that results in a single active-site residue mutation, with apparently unaltered activity, of glutamic acid 119 to gly-

Resistance to the carboxamide inhibitor 5 (Fig. 5) has been investigated in type A influenza with the N9 subtype NA, and an active site mutant, R292K, has been isolated [McKimm-Breschkin et al., 1998]. The specific activity of the R292K variant is only 20% of that of wildtype, and virus bearing this mutant NA shows decreased sensitivity to all NA inhibitors [McKimm-Breschkin et al., 1998]. The R292K mutant has also been reported to arise in an avian N2 background after passaging in Zanamivir [Gubareva et al., 1997]. Varghese and co-workers [1998] have shown by Xray diffraction that the mutation of Arg292 to Lys results in a very local change of structure when compared to the wild-type enzyme. In the wild-type NA Arg292, together with Arg118 and Arg371, engages the 2-carboxylate group of sialic acid in the active site and is partly responsible for distorting the pyranose ring from a chair to a boat conformation [Varghese et al., 1992]. The distal primary guanidinyl amide of Arg292 interacts with Glu277 and the tyrosyl oxygen of Tyr406 (Fig. 2), and the other two guanidinyl nitrogens interact with Asn294. The guanidinyl group of Arg292 is parallel with the peptide plane of mainchain atoms of residues 348349. Two water molecules in the active site cavity have H-bond interactions with the primary amides of the Arg292 prior to the entry of substrate into the active site. The R292K variant of influenza NA affects the bind-

190

VARGHESE

Fig. 7. A ball-and-stick representation of the complex of Zanamivir in the active site of wild-type N9 neuraminidase and the Glu199 to Gly mutation superimposed on each other. Spheres represent oxygen, nitrogen, and carbon atoms. A black sphere (Wat) represents a solvent molecule that occupies the position of the carboxylate group of Glu119 when

it mutates to a Gly. Thin lines represent H-bond interactions of the guanidinium group of Zanamivir and the enzyme. The loss of the interactions with Glu119 in the mutant is thought to be responsible for the reduction of binding of Zanamivir to the mutant. The tubes represent the backbone of the protein.

ing of substrate by modification of the interaction with the substrate carboxylate. This may be one of the structural correlates of the reduced enzyme activity of the variant. Inhibitors which have replacements for the glycerol at the 6-position are further affected in the R292K variant because of structural changes in the binding site which apparently raise the energy barrier for the conformational change in the enzyme required to accommodate such inhibitors. It has been shown [Varghese et al., 1998] that there

is a correlation between the effect of the Arg292 to Lys substitution on the structure of complexes with inhibitors and on the binding properties of the inhibitors. There is also a correlation between the degree of resistance of the variant over wild-type and the degree of structural dissimilarity of the inhibitor to the transition state analogue 2. The largest structural consequences occur with the inhibitors whose binding is most severely compromised, i.e., 7 and then 6 and 5. While 6 and 5 are able to form the hydrophobic pocket to accommodate the sub-

ANTI-VIRAL DRUGS FOR INFLUENZA

191

stitutions in the 6-position of the ring, 7 is unable to achieve this in the Arg292 to Lys mutation (Fig. 8). Smaller structural alterations are associated with inhibitors that retain the glycerol sidechain, and these compounds are only marginally impaired as inhibitors as a consequence of the Arg292 replacement by lysine. The R292K variant has retained only 20% of wildtype enzyme activity, but its high resistance to 7, 6, and 5 suggests that this mutant will be more successful against

these inhibitors than against those which retain a glycerol moiety at C6. Varghese and co-workers [1998] concluded that both the structural and binding effects of the Arg292 to Lys mutation are most pronounced on inhibitors which least resemble the natural ligand (for example, compound 7). They suggested that inhibitor design, at least for targets that have high mutation frequency, should aim to retain as many features as practicable of the natural ligand. The principle behind such a design strategy is

Fig. 8. A ball-and-stick representation of the complex of GS4071 in the active site of wild-type N9 neuraminidase and the Arg292 to Lys mutation superimposed on each other. Spheres represent oxygen, nitrogen, and carbon atoms, and the tube is a representation of the protein back-

bone. GS4071 fails to induce the formation of a salt-link of Glu276 with Arg224 and binds in a different conformation to that of the wild-type enzyme. The thin lines represent the H-bond interaction of GS4071 with the mutant enzyme.

192

VARGHESE

simple variants in the natural ligand binding site must retain some binding to that ligand to preserve the protein function.
CONCLUSION

Jenny McKimm-Breschkin for drug resistance, Brian Smith for enzyme mechanisms.
REFERENCES
Air GM, Laver WG. 1995. Red cells bound to influenza virus N9 neuraminidase are not released by the N9 neuraminidase activity. Virology 211:278284. Alexander DJ. 1986. Avian diseaseshistorical aspects. In: Proceedings of the 2nd International Symposium on Avian Influenza. Richmond, VA: U.S. Animal Heath Assosiation. p 413. Aoki F Osterhaus A, Rimmelzwaan, Kinnersley N, Ward P 1998. Oral , . GS4104 successfully reduces duration and severity of naturally acquired influenza. San Diego: ICAAC 2427 Sept. (Abstr. LB-5). Baker NJ, Gandhi SS. 1976. Effect of Ca++ on the stability of influenza virus neuraminidase. Arch Virol 52:718. Baker AT, Varghese JN, Laver WG, Air GM, Colman PM. 1987. The three-dimensional structure of neuraminidase of subtype N9 from an avian influenza virus. Proteins 1:111117. Beveridge WI. 1977. Influenza: the last great plague. London: Heinemann. Blick TJ, Tiong T, Sahasrabudhe A, Varghese JN, Colman PM, Hart GJ, Bethell RC, McKimm-Breschkin JL. 1995. Generation and characterization of an influenza virus variant with decreased sensitivity to the neuraminidase specific inhibitor 4-guanidino-Neu5Ac2en. Virology 214:475484. Blok J, Air GM, Laver WG, Ward CW, Lilley GG, Woods EF Roxburgh , CM, Inglis AS. 1982. Studies on the size, chemical composition and partial sequence of the neuraminidase (NA) from type A influenza virus show that the N-terminal region of the NA is not processed and serves to anchor the NA in the viral membrane. Virology 119:109121. Bossart-Whitaker P Carson M, Babu YS, Smith CD, Laver WG, Air , GM. 1993. Three dimensional structure of influenza A N9 neuraminidase and its complex with the inhibitor 2-deoxy 2,3-dehydro-Nacetyl neuraminic acid J Mol Biol 232:10691083. Both GW, Sleigh MJ, Cox NJ, Kendal AP 1983. Antigenic drift in influ. enza virus H3 haemagglutinin from 1968 to 1980: multiple evolutionary pathways and sequential amino acid changes at key antigenic sites. J Virol 48:5260. Bucher DJ, Palese P 1975. The biologically active proteins of influ. enza virus neuraminidase. In: Kilbourne ED, editor. Influenza virus and influenza. New York: Academic Press. p 83123. Burmeister WP Ruigrok RWH, Cusack S. 1991. The 2.2 resolution , crystal structure of influenza B neuraminidase and its complex to sialic acid. EMBO J 11:4956. Burmeister WP Henrissat B, Bosso C, Cusack S, Ruigrok RWH. 1993. , Influenza B virus neuraminidase can synthesize its own inhibitor. Structure 1:1926. Burmeister WP Cusack S, Ruigrok RWH. 1994. Calcium is needed , for the thermostability of influeza B virus neuraminidase. J Gen Virol 75:381388.87. Burnett FM. 1947. Mucins and mucoids in relation to influenza virus action. IV. Inhibition by purified mucoid of infection and haemagglutinin with the virus strain WSE. Aust J Exp Biol Med Sci 26:381387. Calfee DP Peng AW, Hussey EK, Lobo M, Hayden FG. 1998. Protec, tive efficacy of reduced frequency dosing of intranasal zanamivir in experimental human influenza. San Diego: ICAAC 2427 Sept. (Abstr. H-68).

It has now been 15 years since the structure of influenza virus neuraminidase was determined [Varghese et al., 1983; Colman et al., 1983]. The development of the 4-subtituted Neu5Ac2en analogue inhibitors dates from 1987, when the structure of sialic acid and Neu5Ac2en complexed with neuraminidase were determined to sufficient accuracy to permit modelling of potential inhibitors [Varghese et al., 1992]. The development was a multidisciplinary collaboration of biochemists, crystallographers, molecular modellers, and synthetic chemists, and culminated in the synthesis [von Itzstein et al., 1994] and biological testing of the compounds [von Itzstein et al., 1993]. Similar results have been obtained by the orally active Gilead carbocyclic analogue GS4104. Preliminary data on the efficacy of potent neuraminidase inhibitors on humans indicate its effectiveness in both prophylaxis and treatment of the disease [Hayden, 1998]. Zanamivir has completed stage III clinical trials in Australia, Europe, and North America, and awaits clearance as a pharmaceutical by regulatory authorities for use as a pharmaceutical both for prophylaxis and treatment of all influenza A and B infections. GS4104 is still undergoing clinical trials and offers an alternative oral therapy for influenza virus infections. While drug resistance of the virus has been observed in vitro, the question of whether this translates into resistance in field strains (as has occurred in the amantidine class of drugs) infecting human populations will be a cause for concern in the long-term viability of this class of antivirals. It has recently been reported that the Arg292 to Lys mutation in neuraminidase has emerged in two immunocompetent patients in clinical trials of GS4104 [Hayden, 1998]. The significance of these isolates in terms of infectivity and spread of resistant variants is yet to be determined. The structural studies on drug-resistant variants of neuraminidase have suggested a minimalist approach [Varghese et al., 1998] to drug design against shifting targets like influenza virus proteins. The further removed the drug is from its natural ligand the greater the potential of the organism developing resistance. This is one of the first structure-based design compounds to be a useful anti-viral drug and portends well for this methodology to be used as an additional weapon in controlling the many pathogens that have plagued humanity for so long.
ACKNOWLEDGMENTS

The author would like to acknowledge Peter Colman for crystallographic and drug design strategies,

ANTI-VIRAL DRUGS FOR INFLUENZA


Chakraverty et al., 1993. Influenza activity United States and worldwide, and composition of the 199394 influenza vaccine. JAMA 269:17781779. Chong AKJ, Pegg MS, von Itzstein M. 1991. Characterisation of an ionisable group involved in the binding and catalysis by sialidase from influenza virus. Biochem Int 24:165171. Chong AKJ, Pegg MS, Taylor NR, von Itzstein M. 1992. Evidence for a sialyl cation transition state complex in the reaction of sialidase from influenza. Eur J Biochem 207:335343. Claas EC, Osterhaus AD, van Beek R, De Jong JC, Rimmelzwaan GF , Krauss S, Shortridge KF Webster RG. 1998. Human influenza A , H5N1 virus related to a highly pathogenic avian influenza virus. Lancet 351:472477 Colman PM. 1989. Influenza virus neuraminidase: Enzyme and antigen. In: Krug RM, editor. The influenza viruses. New York: Plenum Press. p 175218. Colman PM. 1994. Influenza virus neuraminidase: structure, antibodies, and inhibitors. Protein Sci 3:16871696. Colman PM. 1997. Virus versus antibody. Structure 5:591593. Colman PM, Laver WG. 1981. The structure of influenza virus neuraminidase at 5A resolution. In: Structural aspects of recognition and assembly in biological macromolecules. Rehovot, Israel: I.S.S. p 869872. Colman PM, Varghese JN, Laver WG. 1983. Structure of the catalytic and antigenic sites in influenza virus neuraminidase. Nature 303:4144. Colman PM, Laver WG, Varghese JN, Baker AT, Tullock PA, Air GM, Webster RG. 1987. Three-dimensional structure of a complex of antibody with influenza virus neuraminidase. Nature 326: 358363. Colman PM, Tulip WR, Varghese JN, Tulloch PA, Baker AT, Laver WG, Air GM. 1989. 3-D structures of influenza virus neuraminidase-antibody complexes. Proc R Soc Lond B323:511518. Compans RW, Dimmock NJ. 1969. An electron microscopic study of single-cycle infection of chick embryo fibroblasts by influenza virus. Virology 39:499. Crennell SJ, Garman EF Laver WG, Vimr ER, Taylor GL. 1993. Crys, tal structure of a bacterial sialidase (from Salmonella typhimurium LT2) shows the same fold as an influenza virus neuraminidase. Proc Natl Acad Sci USA 90:98529856. Crennell SJ, Garman E, Laver G, Vimr E, Taylor GL. 1994. Crystal structure of Vibro Cholerae neuraminidase reveals dual lectin-like domains in addition to the catalytic domain. Structure 2:535544. Daniels RS, Downie JC, Hay AJ, Knossow M, Skehel JJ, Wang ML, Wiley DC. 1985. Fusion mutants of the influenza virus haemagglutinin glycoprotein. Cell 40:431439. Donnelly JJ, Friedman A, Martinez D, Montgomery DL, Shiver JW, Motzel SL, Ulmer JB, Liu MA. 1995. Preclinical efficacy of a prototype DNA vaccine: enhanced protection against antigenic drift in influenza virus. Nat Med 1:583586. Douglas RG Jr. 1990. Drug therapy, prophylaxis and treatment of influenza. N Engl J Med 322:443450. Drzenick R, Frank H, Rott R. 1968. Electron microscopy of purified influenza virus neuraminidase. Virology 36:703707. Edmond JD, Johnston RG, Kidd D, Rylance HJ, Sommerville RG. 1966. The inhibition of neuraminidase and antiviral action. Br J Pharmacol Chemother 27:415426. Eisenberg EJ, Bidgood A, Cundy KC. 1997. Penetration of GS4071, a novel influenza neuraminidase inhibitor, into rat bronchoalveolar

193

lining fluid following oral administration of prodrug GS4104, Antimicrob Agents Chemother 41:19491952. Ennis FA, Meager A. 1981. Immune interferon produced to high levels antigenic stimulation of human lymphocytes with influenza virus. J Exp Med 154:12791289. Fox JP, Cooney MK, Hall CE, Foy HM. 1982. Influenza virus infections in Seattle families, 19751979. II. Pattern of infection of invaded households and relation of age and prior antibody to occurrence of infection by time and age. Am J Epidemiol 116: 228242. Francis T Jr, Moore AE. 1940. A study of the neurotropic tendency in strains of the virus of epidemic influenza. J Exp Med 72:717728. Friebolin H, Brossmer R, Keilich G, Ziegler D. 1980. Supp M. 1HNMR spektroskopischer nachweis der N-acetyl-a-D-neuraminsaure als primares spaltprodukt der neuraminidasen. Hoppe Seylers Z Physiol Chem 361:697702. Goodford PJ. 1985. A computational procedure for determining energetically favourable binding sites on biologically important macromolecules. J Med Chem 28:849857. Gottschalk A. 1957. Neuraminidase: the specific enzyme of influenza virus and vibrio cholerae. Biochem Biophys Acta 23:645646. Gottschalk A. 1958. Neuraminidase: its substrate and mode of action. Adv Enzymol 20:135145. Goto H, Kawaoka Y. 1998. A novel mechanism for the acquisition of virulence by a human influenza A virus. Proc Natl Acad Sci USA 95:1022410228. Griffin JA, Compans RW. 1979. Effect of cytochalsin B on the maturation of enveloped viruses. J Exp Med 150:379391. Gubareva LV, Bethell R, Hart GJ, Penn CR, Webster RG. 1995. Characterization of mutants of influenza virus selected with 4-guanidinoNeu5Ac2en. Abstracts, 14th Annual Meeting, Am Soc Virol, Austin, Texas W44-1. Gubareva LV, Robinson MJ, Bethell R, Webster RG. 1997. Catalytic framework mutations in the neuraminidase active site of influenza viruses that are resistant to 4-guanidino-Neu5Ac2en. J Virol 71:3385 3390. Gubareva LV, Matrosovich MN, Brenner MK, Bethel RC, Webster RG. 1998. Evidence for Zanamivir resistence in an immunocompromised child infected with influenza B virus. J Infect Diseases 178:12571262. Hausmann J, Kretzschmar E, Garten W, Klenk H-D. 1995. N1 neuraminidase of influenza virus A/FPV/Rostock/34 has haemadsorbing activity. J Gen Virol 76:17191728. Hay AJ, Thompson CA, Geraghty A, Hayhurst S, Grambas S, Bennett MS. 1993. The role of the M2 protein in influenza virus infection. In: Hannoun C et al., editors. Options for the control of influenza virus II. Amsterdam: Excerpta Medica. p 281288. Hayden FG. 1993. Update on antiviral agents and viral drug resistance. In: Mandell GL, Douglas RG, Bennett JE, editors. Principles and practice of infectious disease. New York: Churchill Livingston. p 2(2):315. Hayden FG. 1996. Amantidine and rimantidine clinical aspects. In: Richmond DD, editor. Antiviral drug resistance. Chichester, UK: John Wiley. p 6977. Hayden FG, Rollins BS, Madren LK. 1994. Anti-influenza activity of the neuraminidase inhibitor 4-guanidino-Neu5Ac2en in cell culture and in human respiratory epithelium. Antiviral Res 25:123131. Hayden FG, Treanor JJ, Betts RF Lobo M, Esinhart JD, Hussey EK. , 1996. Safety and efficacy of the neuraminidase inhibitor GG167 in experimental human influenza. JAMA 275:295299.

194

VARGHESE
Li W, Escarpe PA, Eisenberg EJ, Cundy KC, Sweet C, Jakeman KJ, Merson J, Lew W, Williams M, Zhang L, Kim CU, Bischofberger N, Chen MS, Mendel DB. 1998. Identification of GS4104 as an orally bioavailable prodrug of the influenza virus neuraminidase inhibitor GS4071. Antimicrob Agents Chemother 42:647653. Liu C, Air GM. 1993. Selection and characterization of a neuraminidase-minus mutant of influenza virus and its rescue by cloned neuraminidase genes. Virology 194:403407. Liu C, Eichelberger MC, Compans RW, Air GM. 1995. Influenza type A viurs neuraminidase does not play a role in viral entry, replication, assembly, or budding. J Virol 69:10991106. Luo M, Jedrzejas MJ, Singh S, White CL, Brouillette WJ, Air GM, Laver WG. 1995. Benzoic acid inhibitors of influenza virus neuraminidase. Acta Crystallogr D51:504510. Madren LK, Shipman C, Hayden FG. 1995. In vitro inhibitory effects of combinations of anti-influenza agents. Antiviral Chem Chemother 6(2):109113. Malby RL, Tulip WR, Harley VR, McKimm-Breschkin JL, Laver WG, Webster RG, Colman PM. 1994. The structure of a complex between the NC10 antibody and influenza virus neuraminidase and comparison with the overlapping binding site of the NC41 antibody. Structure 2:733746. McKimm-Breschkin JL, Marshall D, Penn CR. 1994. Phenotypic changes observed in influenza viruses passaged in 4-amino or 4guanidino-Neu5Ac2en in vitro. Abstracts, 9th International Conference on Negative Strand Viruses, Estoril, Portugal, p 260. McKimm-Breschkin JL, Blick TJ, Sarasrabudhe AV, Tiong T, Marshall D, Hart GJ, Bethell RC, Penn CR. 1996. Generation and characterisation of variants of the NWS/G70C influenza virus after in vitro passage in 4-amino-Neu5Ac2en and 4-guanidinoNeu5Ac2en. Antimicrob Agents Chemother 40:4046. McKimm-Breschkin JL, Sahasrabudhe A, Blick TJ, McDonald M, Colman PM, Hart GJ, Bethell RC, Varghese JN. 1998. Mutations in a conserved residue in the influenza virus neuraminidase active site decreases sensitivity to Neu5Ac2en derivatives. J Virol 72:2456 2462. Meindl P Tuppy H. 1969. 2-Deoxy-2,3-dehydrosialic acids. I. Synthe, sis and properties of 2-deoxy-2,3-dehydro-N-acetylneuraminic acids and their methyl esters. Monatsh Chem. 100:12951306. Meindl P Bodo G, Palase P Schulman J, Tuppy H. 1974. Inhibition of , , neuraminidase activity by derivatives of 2-deoxy-2,3-dehydro-Nacetylneuraminic. Virology 58:457463. Mendel DB, Tai CY, Escarpe PA, Sidwell RW, Huffman JH, Sweet C, Jakeman KJ, Merson J, Lacy SA, Lew W, Williams MA, Zhang L, Chen MS, Bischofberger N, Kim CU. 1998. Oral administration of a prodrug of the influenza virus neuraminidase inhibitor GS4071 protects mice and ferrets against influenza infection. Antimicrob Agents Chemother 42:640646 Miller CA, Wang P Flashner M. 1978. Mechanism of Arthro-bacter , sialophilus neuraminidase: the binding of substrates and transition state analogues. Biochem Biophys Res Commun 83:14791487. Monto AS, Ohmir Se, Hornbuckle K, Pearce CL. 1995. Safety and efficacy of long term use of rimantadine for prophylaxis of type 1 influenza in nursing homes. Antimicrob Agents Chemother 39:2224 2228. Monto AS, Robinson DP, Herlocher L, Hinson JM, Elliott M, Keene O. 1998. Efficacy and safety of zanamivir in prevention of influenza among healthy adults. San Diego: ICAAC 2427 Sept. (Abstr. LB-7). Murphy JS, Bang FB. 1952. Observations with the electron micro-

Hayden FG, Osterhaus ADME, Treanor JJ, Fleming DM, Aoki FY, Nicholson KG, Bohnen AM, Hirst HM, Keene O, Whiteman K. 1997. Efficacy and safety of the neuramimnidase inhibitor Zanamivir in the treatment of influenza virus infections. N Engl J Med 337:874880. Hayden FG, Atmar R, Schilling M, Johnson C, Poretz D, Parr D, Huson L, Ward P, Mills R. 1998. Safety and efficacy of oral GS4104 in longterm prophylaxis of natural influenza. San Diego: ICAAC 24 27 Sept. (Abstr. LB-6). Hinshaw VS, Webster RG, Turner B. 1980. The perpetuation of orthomyxoviruses and paramyxoviruses in Canadian waterfowl. Can J Microbiol 26:622629. Hirst GK. 1942. Adsorption of influenza hemagglutinins and virus by red blood cells. J Exp Med 76:17401743. Hitte AL, Nayak DP 1982. Complete nucleotide sequences of the . neuraminidase gene of human influenza virus A/WSN/33. J Virol 41:730734. Janakiraman MN, White CL, Laver WG, Air MA, Luo M. 1994. Structure of influenza virus neuraminidase B/Lee/40 complexed with sialic acid and a dehydro analog at 1.8A-resolution: implications for the catalytic mechanism. Biochemstry 33:81728179. Kilbourne ED. 1987. Influenza. New York: Plenum Press. Kim CU, Lew W, Williams MA, Liu H, Zhang L, Swaminathan S, Bischofberger N, Chen MS, Mendel DB, Tai CY, Laver WG, Stevens RC. 1997. 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 119:681690. Kimberlin DW, Crumpacker CS, Straus SE, Biron KK, Drew WL, Hayden FG, McKinlay M, Richman DD, Whitley RJ. 1995. Antiviral resistance in clinical practice. Antiviral Res 26:423438. Klenk E, Faillard H, Lempfrid H. 1955. Uber die enzymatishe Wirkung von Ifluenza. Z Physiol Chem 301:235246. Kraulis PJ. 1991. MOLSCRIPT: A program to produce both detailed and schematic plots of protein structures. J Appl Cryst D50:869873. Kunkel TA, Roberts JD, Zakaur RA. 1987. Rapid and efficient sitespecific mutagenesis without phenotypic selection. Methods Enzymol 154:369382. Lamb RA. 1989. Genes and proteins of the influenza virus. In: Krug RM, editor. The influenza viruses. New York: Plenum Press. p 187. Laver WG. 1963. The structure of influenza virus 3. Disruption of the virus particle and separation of neuraminidase activity. Virology 20:251262. Laver WG. 1978. Crystallization and peptide maps of neuraminidase heads from H2N2 and H3N2 influenza virus strains. Virology 86:7887. Laver WG, Valentine RC. 1969. Morphology of the isolated hemagglutinin and neuraminidase subunits of influenza virus. Virology 38:105119. Laver WG, Colman PM, Webster RG, Hinshaw VS, Air GM. 1984. Influenza virus neuraminidase with hemagglutinin activity. Virology 137:314323. Lentz MR, Webster RG, Air GM. 1987. Site-directed mutation of the active site of influenza neuraminidase and implications for the catalytic mechanism. Biochemistry 26:53515358. Li S, Schulman J, Itamura S, Palase P 1993. Glycosylation of neuramini. dase determines the neurovirulence of influenza A/WSN/33 virus. J Virol 67:66676673.

ANTI-VIRAL DRUGS FOR INFLUENZA


scope on cells of chick chorio-allantoic membrane infected with influenza virus. J Exp Med 95:259. Murphy BR, Webster RG. 1990. Orthomyxovirus. In: Fields BN, Knipe DM, editors. Virology. New York: Raven Press. p 10911152. Nagai T, Yamada H. 1994. In vivo anti-influenza virus activity of Kampo (Japanese herbal) medicine Sho-seiryu-to and its mode of action. Int J Immunopharmacol 16(8):605613. Nagai T, Miyaichi Y, Tomimori T, Suzuki Y, Yamada H. 1990. Inhibition of influenza virus sialidase and anti-influenza virus activity by plant flavonoids. Chem Pharm Bull 38(5):13291332. Nicholls A, Sharp K, Honig B. 1991. Protein folding and association: insights from the interfacial and thermodynamic properties of hydrocarbons. Proteins 11:281296. Nohle U, Beau J-M, Schauer R. 1982. Uptake, metabolism and excretion of orally and intravenously administered, double-labeled N-glycoloylneuraminic and single-labeled 2-deoxy-2,3dehydro-N-acetylneuraminic acid in mouse and rat. Eur J Biochem 126:543548. Nuss JM, Air GM. 1991. Transfer of the hemagglutinin activity of influenza virus neuraminidase subtype N9 into N2 neuraminidase background. Virology 183:496504. Osterhaus ADME, Tisdale M, Elliot M. 1998. Double blind randomised trial of zanamivir in the treatment of acute influenza clinical and virological efficacy results. San Diego: ICAAC 2427 Sept. (Abstr. H-67). Palese P Compans RW. 1976. Inhibition of influenza virus replication , in tissue culture 2-deoxy-2,3-dehydro-N-trifluroacetylneuraminic acid. FANA: mechanism of action. J Gen Virol 33:159163. Palese P Schulman JL. 1977. Inhibitors of viral neuraminidase as po, tential antiviral drugs. In: Chemoprophylaxis and virus infections of the upper respiratory tract, vol 1. Boca Raton, FL: CRC Press. p 189205. Palese P Tobita K, Ueda M, Compans RW. 1974a. Characterisation of , temperature sensitive influenza virus mutants defective in neuraminidase. Virology 61:397410. Palese P Schulman JL, Bodo G, Meidl P 1974b. Inhibition of influ, . enza and parainfluenza virus replication in tissue culture by 2-deoxy2,3-dehydroNtrifluroacetylneuraminic acid. FANA Virol 59:490498. Pearlman DA, Case DA, Caldwell J, Seibel G, Singh UC, Weiner PK, Kollman PA. 1990. AMBER 4.0 San Franscisco, CA: University of California. Pinto LH, Holsinger LJ, Lamb RA. 1992. Influenza virus M2 protein has ion channel activity. Cell 69:517528. Rogers GN, Paulson JC. 1983. Receptor determinants of human and animal influenza virus isolates: difference in receptor specificity of the H3 haemagglutinin based on species of origin. Virology 127:361373. Ryan DM, Ticehurst J, Dempsey MH, Penn CR. 1994. Inhibition of influenza virus replication in mice by GG167 (4-guanidino-2,4dideoxy-2,3-dehydro-N-acetylneuraminic acid) is consistent with extracellular activity of viral neuraminidase. Antimicrob Chem Chemother 38(10):22702275. Sahasrabudhe A, Lawrence L, Epa VC, Varghese JN, Colman PM, McKimm-Breschkin JL. 1998. Substrate, inhibitor, or antibody stabilises the Glu 119 Gly mutant influenza virus neuraminidase. Virology 247:1421. Skehel JJ. 1974. The origin of pandemic influenza virus. Symp Soc Gen Microbiol 24:321342.

195

Slemons RD, Johnson DC, Osborn JS, Hayes F 1974. Type A influ. enza viruses iosloated from wild free-flying ducks in California. Avian Dis 18:119125. Smith FI, Palese P. 1989. Variation in influenza virus genes. Epidemiological, pathogenic, and evolutionary consequences. In: Krug RM, editor. The influenza viruses. New York: Plenum Press. p 319359. Smith W, Andrewes CH, Laidlaw PP 1933. A virus obtained from in. fluenza patients. Lancet 2:6668. Smith PW, Sollis SL, Howes PD, Cherry PC, Cobley KN, Taylor H, Whittington AR, Scicinski J, Bethell RC, Taylor N, Skarzynski T, Cleasby A, Singh O, Wonacott A, Varghese J, Colman P 1996. Novel . inhibitors of influenza sialidases related to GG167. Structure-activity, crystallographic and molecular dynamics studies with 4H-pyran-2-carboxylic acid 6-carboxamides. Bioorg Med Chem Lett 6(24):29312936. Sollis SL, Smith PW, Howes PD, Cherry PC, Bethell RC. 1996. Novel inhibitors of influenza sialidase related to GG167. Synthesis of 4amino and 4-guanidino-4H-pyran-2-carboxylic acid-6-propylamides; selective inhibitors of influenza A virus sialidase. Bioorg Med Chem Lett 6(15):18051808. Sprenger MJW, Beyer WEP Kempen BM, Mulder PGH, Masurel N. , 1993. Risk factors for influenza mortality. In: Hannoun C et al., editors. Options for the control of influenza virus II. Amsterdam: Excerpta Medica. p 1523. Srange KC, Little DW, Blarnik B. 1991. Adverse reactions to amantadine prophylaxis of influenza in a retirement home. J Am Geriatr Soc 33:700705. Subbarao K, Klimov A, Katz J, Regnery H, Lim W, Hall H, Perdue M, Swayne D, Bender C, Huang J, Hemphill M, Rowe T, Shaw M, Xu X, Fukuda K, Cox N. 1998. Characterisation of an avian influenza A (H5N1) virus isolated from a child with fatal respiratory illness. Science 279:393396. Suzuki Y, Sato K, Kiso M, Hasegawa A. 1990. New ganglioside analogs that inhibit influenza virus sialidase. Glycoconjugate J 7:349356. Taylor NR, von Itzstein M. 1994. Molecular modeling studies on ligand binding to sialidase from influenza virus and the mechanism of catalysis. J Med Chem 37:616624. Taylor NR, Cleasby A, Singh O, Skarzynski T, Wanacott A, Smith PW, Sollis SL, Howes PD, Cherry PC, Bethell R, Colman P Varghese J. , 1998. Dihydropyran carboxamides related to Zanamivir. A new series of inhibitors of influenza virus sialidases. Part 2: A crystallographic and molecular modeling study of complexes of 4-amino-4H-pyran-6carboxamides and sialidase from influenza types A and B. J Med Chem 41:798807. Thomas GP Forsyth M, Penn CR, McCauley JW. 1994. Inhibition of , growth of influenza virus in vitro by 4-guanidino-2,4-dideoxy-2,3dehydro-N-acetylneuraminic acid. Antiviral Res 24:351356. Treanor JJ, Vrooman PS, Hayden FG, Kinnersley N, Ward P Mills RG. , 1998. Efficacy of oral GS4104 in treating acute influenza. San Diego: ICAAC 2427 Sept. (Abstr. LB-4). Tulip WG, Varghese JN, Baker AT, van Donkelaar A, Laver WG, Webster RG, Colman PM. 1992a. Refined atomic structures of N9 subtype influenza virus neuraminidase and escape mutants. J Mol Biol 221:487497. Tulip WG, Varghese JN, Laver WG, Webster RG, Colman PM. 1992b. Refined crystal structure of the influenza virus N9 neuraminidaseNC41 Fab complex. J Mol Biol 227:122148. Tyrrell DAJ. 1980. Influenza vaccines. Phil Trans R Soc Lond B288:449460.

196

VARGHESE
Ward CW, Elleman TC, Azad AA. 1982. Amino acid sequence of the pronase-released heads of neuraminidase subtype N2 from the Asian strain A/Tokyo/3/67 of influenza virus. Biochem J 207:9195. Ward CW, Murry JM, Roxburgh CM, Jackson DC. 1983. Chemical and antigenic characterisation of the carbohydrate side chains of an Asian (N2) influenza virus neuraminidase. Virology 126:370375. Webster RG, Laver WG. 1975. Antigenic variation in influenza viruses. In: Kilbourne ED, editor. Influenza virus and influenza. New York: Academic Press. p 269314. Webster RG, Air GM, Metzger DW, Colman PM, Varghese JN, Baker AT, Laver WG. 1987. Antigenic structure and variation in an influenza N9 neuraminidase. J Virol 61:29102916. Webster RG, Bean WJ, Gorman OT, Chambers TM, Kawaoka Y. 1992. Evolution and ecology of influenza A viruses. Microbiol Rev 56:152179. Webster RG, Schafer JR, Suss J, Bean WJ Jr, Kawaoko Y. 1993. Evolution and ecology of influenza viruses. In: Hannoun C et al., editors. Options for the control of influenza virus II. Amsterdam: Excerpta Medica. p 177185. Weis W, Brown JH, Cusack S, Paulson JC, Skehel JJ, Wiley DC. 1988. Structure of the influenza virus haemagglutinin complexed with its receptor, sialic acid. Nature 333:426431. White DO. 1974. Influenza viral proteins: identification and synthesis. Curr Top Microbiol Immunol 63:148. Wintermeyer SM, Nahata MC. 1995. Rimantadine: a clinical perspective. Ann Pharmacother 29:299310. Woods JM, Bethell RC, Coates JAV, Healy N, Hiscox SA, Pearson BA, Ryan DM, Ticehurst J, Tilling J, Walcott SM, Penn CR. 1993. 4Guanidino-2-4-dideoxy-2,3-dehydro-n-acetylneuraminic acid is a highly effective inhibitor both of the Sialidase (Neuraminidase) and growth of a wide range of influenza A and B viruses in vitro. Antimicrob Agents Chemother 37:14731479. Wright CE, Laver WG. 1978. Preliminary crystallographic data for influenza virus neuraminidase heads. J Mol Biol 120: 133136. Yeun KY, Chan PK, Peiris M, Tsang DN, Que TL, Shortridge KF , Cheung PT, To WK, Ho ET, Sung R, Cheng AF 1998. Clinical fea. tures and rapid viral diagnosis of human disease associated with avian influenza A H5N1 virus. Lancet 351:467471.

Varghese JN. 1997. The design of anti-influenza virus drugs from the X-ray molecular structure of influenza virus neuraminidase. In: Veerapandian P editor. Structure based drug design: diseases, tar, gets, techniques and developments, vol 1. New York: Marcel Dekker. p 459486. Varghese JN, Colman PM. 1991. Three-dimensional structure of the neuraminidase of influenza virus A/Tokyo/3/67 at 2.2 resolution. J Mol Biol 221:473486. Varghese JN, Laver WG, Colman PM. 1983. Structure of the influenza virus glycoprotein antigen neuraminidase at 2.9 resolution. Nature 303:3540. Varghese JN, Webster RG, Laver WG, Colman PM. 1988. The threedimensional structure of an escape mutant of the neuraminidase of influenza virus A/Tokyo/3/67. J Mol Biol 200:201203. Varghese JN, McKimm-Breschkin JL, Caldwell JB, Kortt AA, Colman PM. 1992. The structure of the complex between influenza virus neuraminidase and sialic acid, the viral receptor. Proteins 14:327332. Varghese JN, Epa VC, Colman PM. 1995. Three-dimensional structure of 4-guanidino-Neu5Ac2en and influenza virus neuraminidase. Protein Sci 4:10811087. Varghese JN, Colman PM, van Donkelaar A, Blick TJ, Sahasrabudhe A, McKimm-Breschkin JL. 1997. Structural evidence for a second sialic acid binding site in avian influenza virus neuraminidase. Proc Natl Acad Sci USA 94:1180811812. Varghese JN, Smith PW, Sollis SL, Blick TJ, Sahasrabudhe A, McKimmBreschkin JL, Colman PM. 1998. Drug design against a shifting target: a structural basis for resistance to inhibitors in a variant of influenza virus neuraminidase. Structure 6:735746. von Itzstein M, Wu W-Y, Kok GB, Pegg MS, Dyson JC, Jin B, VanPhan T, Symthe ML, White HF Oliver SW, Colman PM, Varghese JN, , Ryan DM, Woods JM, Bethell RC, Hotham VJ, Cameron JM, Penn CR. 1993. Rational design of potent sialidase-based inhibitors of influenza virus replication. Nature 363:418423. von Itzstein M, Wu W-Y, Jin B. 1994. The synthesis 2,3-didehydro2,4-dideoxy-4-guanidinyl-N-acetylneuraminic acid: a potent influenza virus inhibitor. Carbohydr Res 259:301305. Wagh PV, Bahl OP 1981. Sugar residues on proteins. Crit Rev Biochem . 307377. Ward CW. 1981. Structure of influenza virus hemagglutinin. Curr Top Microbiol Immunol 94/95:174.

S-ar putea să vă placă și