Skip to main content
Antimicrobial Agents and Chemotherapy logoLink to Antimicrobial Agents and Chemotherapy
. 2005 Feb;49(2):619–626. doi: 10.1128/AAC.49.2.619-626.2005

Conservation of Amino Acids in Human Rhinovirus 3C Protease Correlates with Broad-Spectrum Antiviral Activity of Rupintrivir, a Novel Human Rhinovirus 3C Protease Inhibitor

S L Binford 1, F Maldonado 1, M A Brothers 1, P T Weady 1, L S Zalman 1, J W Meador III 1, D A Matthews 1, A K Patick 1,*
PMCID: PMC547258  PMID: 15673742

Abstract

The picornavirus 3C protease is required for the majority of proteolytic cleavages that occur during the viral life cycle. Comparisons of published amino acid sequences from 6 human rhinoviruses (HRV) and 20 human enteroviruses (HEV) show considerable variability in the 3C protease-coding region but strict conservation of the catalytic triad residues. Rupintrivir (formerly AG7088) is an irreversible inhibitor of HRV 3C protease with potent in vitro activity against all HRV serotypes (48 of 48), HEV strains (4 of 4), and untyped HRV field isolates (46 of 46) tested. To better understand the relationship between in vitro antiviral activity and 3C protease-rupintrivir binding interactions, we performed nucleotide sequence analyses on an additional 21 HRV serotypes and 11 HRV clinical isolates. Antiviral activity was also determined for 23 HRV clinical isolates and four additional HEV strains. Sequence comparison of 3C proteases (n = 58) show that 13 and 11 of the 14 amino acids that are involved in side chain interactions with rupintrivir are strictly conserved among HRV and HEV, respectively. These sequence analyses are consistent with the comparable in vitro antiviral potencies of rupintrivir against all HRV serotypes, HRV isolates, and HEV strains tested (50% effective concentration range, 3 to 183 nM; n = 125). In summary, the conservation of critical amino acid residues in 3C protease and the observation of potent, broad-spectrum antipicornavirus activity of rupintrivir highlight the advantages of 3C protease as an antiviral target.


The picornavirus family consists of over 200 medically important viruses, including human rhinoviruses (HRV) and human enteroviruses (HEV). HRV, comprising over 100 different serotypes, are a major cause of mild upper respiratory infections (reviewed in references 4, 8, and 40). Although HRV infections are usually mild and self-limiting, they can also be associated with exacerbation of disease in individuals with underlying respiratory disorders (4, 48). The HEV include over 70 viruses that are associated with diverse clinical syndromes ranging from mild, self-limiting infections to fulminant and potentially fatal disease (3, 52, 53). Earlier clinical studies with agents that inhibit virus attachment and/or uncoating (e.g., tremacamra, a soluble intracellular adhesion molecule-1, pirodavir, and pleconaril) and nonspecific antiviral agents such as α-2β interferon have demonstrated that prevention and early treatment of HRV colds could provide clinical benefit (3, 18-23, 53). Recently, a retrospective analysis of two multicenter clinical trials demonstrated that pleconaril, a capsid-function inhibitor (17), significantly reduced the duration and severity of picornavirus-induced colds (24). To date, however, no antiviral agents have been approved for the prevention or treatment of HRV infection.

We have focused our antiviral strategy on the inhibition of 3C protease, a viral enzyme that is absolutely required for the proteolytic cleavage of viral precursor polyproteins to functional proteins (10-14, 42, 46, 47) and, thus, is essential for viral replication. Initial sequence comparisons of the 3C protease-coding regions from six HRV serotypes (15, 28, 38, 57, 58, 63), together with the experimentally derived three-dimensional structure of 3C protease (41), indicate significant homology in the substrate-inhibitor binding site, including strict conservation of the three amino acid residues that comprise the catalytic triad (His 40, Glu 71, and Cys 147). Consistent with this finding, rupintrivir (formerly AG7088), a novel, irreversible inhibitor of 3C protease (10-13, 61) has demonstrated broad-spectrum, potent in vitro antiviral activity against all picornaviruses tested, including 48 HRV serotypes, 4 HEV strains, and 46 untyped field isolates of HRV (34, 46, 64). Furthermore, recent data demonstrating the ability of rupintrivir to moderate illness severity and reduce viral load in human subjects following experimental HRV infection provide proof of concept for the mechanism of 3C protease inhibition (25). To better understand the relationship between the broad-spectrum in vitro antiviral activity of rupintrivir and 3C protease-rupintrivir binding interactions, we performed nucleotide sequence analyses on an additional number of laboratory strains and clinical isolates of HRV and antiviral assays to include HRV clinical isolates and additional HEV strains. Our results indicate that among the HRV and HEV strains evaluated, 13 and 11, respectively, of the 14 amino acid residues previously shown to be critical for 3C protease inhibitor binding (41) are absolutely conserved. This conservation is consistent with the comparable in vitro antiviral activity of rupintrivir against numbered HRV serotypes, HRV isolates, and HEV strains evaluated (50% effective concentration range, 3 to 183 nM).

MATERIALS AND METHODS

Antiviral compound.

Rupintrivir (molecular weight, 598.7; aqueous solubility of 15 μg/ml) was synthesized at Pfizer Global Research and Development (formerly Agouron Pharmaceuticals, Inc.), San Diego, Calif. (11).

Cells and virus strains.

H1-HeLa cells, MRC-5 cells, all numbered HRV serotypes, and HEV were purchased from the American Type Culture Collection (Manassas, Va.). Nasal lavage samples from subjects experimentally infected with challenge strains of HRV 39, HRV Hanks, HRV 23, or HRV 16 were kindly provided by Frederick G. Hayden (Department of Internal Medicine, University of Virginia Health Sciences Center, Charlottesville, Va.) and Ronald B. Turner (Department of Pediatrics, University of Virginia Health Sciences Center). Field isolates from symptomatic individuals were identified as HRV by the optical immunoassay OIA (45) and were generously provided by Rachel Ostroff (Thermo BioStar, Inc., Boulder, Colo.). Clinical viral isolates were obtained by inoculating H1-HeLa cells with nasal lavage samples. All HRV clinical isolates were propagated and antiviral assays were performed in H1-HeLa incubated at 34°C. Cells were grown in minimal essential medium (Invitrogen, Carlsbad, Calif.) with 10% fetal bovine serum (HyClone Lab. Inc., Logan, Utah). Coxsackieviruses B2 and B5 (CVB 2 and CVB 5) were propagated and antiviral assays were performed in H1-HeLa cells incubated at 37°C. Echovirus serotypes 6 and 9 (EV 6 and EV 9) were propagated and antiviral assays were performed in MRC-5 cells incubated at 37°C. Cell-free HRV lysates were either extracted immediately or stored at −80°C for subsequent RNA isolation.

DNA sequence analysis.

RNA was purified from virus by silica-based extraction by using the RNeasy method (QIAGEN, Valencia, Calif.), and then cDNA was synthesized from viral RNA by using a First Strand synthesis kit (Amersham Pharmacia Biotech, Inc., Piscataway, N.J.) and random nanomers. For each strain, the target region (corresponding to bases 5116 to 5827 in HRV 89) was amplified by PCR with Taq 2000 polymerase (Stratagene, San Diego, Calif.). The 3C protease nucleotide sequences from HRV serotypes 3, 9, 10, 11, 13, 15, 17, 19, 23, 24, 31, 36, 39, 52, 56, 59, 62, 78, 81, 87, and HRV Hanks were determined, when possible, by sequencing of PCR products. Otherwise, the PCR amplicons were subcloned into pGEM-T Easy vectors (ProMega Corp., Madison, Wis.) and sequenced by using T7 and SP6 primers. The difficulty in amplifying the 3C protease regions from very diverse HRV serotypes was addressed by using multiple degenerate PCR primers. The sense PCR primers used were 5′-CARGGNCCWTAYTCNGGNWANCC-3′, 5′-CARGGNCCATAYTCNGGNRANCC-NAAR-3′, and 5′-CAGGGACCATACTCAGG-3′. The antisense PCR primers used were 5′-ATCYTYYAANACACTWGGNTGNAYTWNNGTTTT-3′ and 5′-ATCYTYAANACACTWGG-NTGNAYT-3′. Since previous studies identified the sequence flanking the 3C protease of HRV 39, species-specific primers consisting of the sense primer 5′-AAAACCATGGAGCGGAGGGTAG-TTGC-3′ and antisense primer 5′-CACAGGATCCTTATTACATTCACTGGTTTTCTTTGA-3′ were used for HRV 39. The DNA sequences were analyzed, relevant regions were translated, and phylogenetic relationships based on similarities of 3C protease amino acid sequences were determined by DNASTAR analysis software (DNASTAR Inc., Madison, Wis.). For illustration purposes, a phylogenetic tree was generated by Clustal W analysis in MegAlign. The peptide sequences for 3C protease from HRV serotypes 1A (63), 1B (28), 2 (57), 14 (58), 16 (38), and 89 (15) and from all HEV (6, 7, 9, 29-33, 36, 39, 44, 49, 50, 54, 59, 65-69) were obtained from the GenBank database (Bethesda, Md.).

Antiviral assays.

The ability of rupintrivir to protect cells against infection with HRV clinical isolates and HEV was measured in 96-well plates by a modified and standard cytopathic effect (CPE) inhibition assay, respectively, by using the 2,3-bis(2-methoxy-4-nitro-5-sulfophenyl)-5-[(phenylamino) carbonyl]-2H-tetrazolium hydroxide (XTT) dye reduction method (HEV) (34, 56, 62, 64). In the standard CPE assay, H1-HeLa and MRC-5 cells were initially resuspended at a final concentration of 2 × 105 or 5 × 104 cells per ml, respectively, and infected with virus (50 μl) at a constant multiplicity of infection (MOI) determined to yield ≥65 to ≤95% cell death (CPE) for each individual virus (MOI range, 0.004 to 1) or mock infected with medium only. At 1 to 6 days later, cultures were evaluated for CPE microscopically, XTT and phenazine methosulfate were added to the test plates, and the amount of formazan produced was quantified spectrophotometrically at 450 nm with adjustment for background at 650 nm. Data were expressed as the percentage of formazan produced in compound-treated cells compared to formazan produced in wells of uninfected, compound-free cells. The 50 and 90% effective concentrations (EC50 and EC90) were calculated as the concentrations of compound that increased the percent of formazan production in infected, compound-treated cells to 50% and 90% of that produced by uninfected, compound-free cells, respectively. The modified CPE assay, which allows for the simultaneous determination of antiviral activity and optimal CPE, was utilized with HRV clinical isolates. Briefly, each clinical HRV isolate was inoculated at six serial 10-fold dilutions in quadruplicate for a 50% tissue culture infectious dose determination (back titration) and at three 10-fold virus dilutions for compound titration. The EC50 values were determined by utilizing the inoculum dilution for which the calculated 50% tissue culture infectious dose per milliliter was between 32 and 320. XTT and phenazine methosulfate were added to the test plates and processed as above.

RESULTS

In vitro antiviral activity of rupintrivir against HRV clinical isolates and HEV strains.

Previous studies have demonstrated potent, broad-spectrum antiviral activity of rupintrivir against 48 HRV numbered serotypes and 4 related picornaviruses as well as 46 untyped HRV field isolates (34, 46). To extend these findings, we evaluated the antiviral activity of rupintrivir against 23 clinical isolates of HRV derived from subjects during experimental challenge studies (Table 1). Four additional laboratory strains of HEV, selected for their diversity and association with human disease (CVB 2 and 5 and EV 6 and 9), were likewise evaluated (Table 1). Rupintrivir demonstrated potent in vitro antiviral activity against 23 of 23 HRV clinical isolates tested in H1-HeLa cell protection assays, with a mean EC50 of 24 nM (range, 3 to 104 nM) and a mean EC90 of 214 nM (range, 7 to 1078 nM). Potent in vitro activity was also demonstrated against four HEV strains with a mean EC50 of 41 nM (range, 7 to 137 nM) and a mean EC90 of 103 nM (range, 34 to 254 nM). Overall, rupintrivir demonstrated comparable antiviral potencies (61-fold range [3 to 183 nM] in EC50 values) against all the HRV and HEV isolates tested (n = 125; Table 1) (34, 46).

TABLE 1.

In vitro antiviral activity of rupintrivir against HRV clinical isolates and HEVa

Virus Antiviral activity
EC50 (nM) EC90 (nM)
HRV clinical isolateb
    100 22 ± 20 32 ± 8
    103 58, 24 316, 746
    108 43, 19 182, 42
    113 46, 26 464
    119 29 ± 11 145 ± 130
    123 23 ND
    126 23 794
    129 19 39
    142 19 38
    O 54 ± 51 275 ± 258
    G 14, 20 66, 1078
    I 104, 16 604, 29
    W 10, 3 44, 7
    J 18 ± 6 69 ± 90
    R 14 ± 8 204 ± 291
    Q 17 33
    B 12 40
    C 12 28
    T 42 ± 28 123 ± 24
    E 23 562
    15 26 ± 7 262
    28 8 18
    32 7 19
Enterovirus
    CVB 2 22, 137 88, 254
    CVB 5 7, 47 147, 83
    EV 6 51, 34 94, 59
    EV 9 11, 18 65, 34
a

Antiviral activity was determined by measuring XTT dye reduction following 2 to 7 days of infection by using the standard or modified CPE inhibition assays as described in Materials and Methods. Results represent the mean ± standard deviation (three to five experiments) or individual values (one or two experiments). ND, not determined.

b

Clinical isolates were derived from nasal lavage samples collected from subjects infected with HRV serotype 39 (designated 100, 103, 108, 113, 119, 123, 126, 129, 142, O, G, I, W), Hanks (designated J, R, Q, and B), 16 (designated C, T, and E), or 23 (designated 15, 28, and 32).

Amino acid sequence analyses of HRV and HEV 3C protease.

High-resolution crystallographic studies have identified 23 different amino acid residues located in the substrate (inhibitor) binding site of HRV 2 3C protease that interact with rupintrivir (41). Inhibitor binding interactions have been identified, 17 of which occur with side chains of 14 amino acids (including Phe 25, His 40, Glu 71, Ile 125, Leu 127, Ser 128, Asn 130, Thr 142, Cys 147, His 161, Gly 163, Gly 164, Asn 165, and Phe 170), while rupintrivir interactions with nine other amino acids involve polypeptide main chain atoms only (Fig. 1.). To understand the extent of conservation of the 14 amino acid residues that are involved in side chain interactions with rupintrivir and to correlate this with the observed spectrum of rupintrivir in vitro antiviral activity, we performed sequence analysis on the 3C protease-coding regions from 21 HRV serotypes and 11 HRV clinical isolates. The amino acid sequences from the combined 32 HRV isolates were aligned together with the published sequences from 6 HRV serotypes and 20 HEV strains (n = 58) (Fig. 2), and the extent of homology was evaluated. Of the 14 amino acids involved in side chain interactions with rupintrivir, 13 (93%) and 11 (79%) were absolutely conserved (identical) in HRV and HEV strains, respectively. Among the HRV strains, only Asn 130 was not conserved. Among the HEV strains, Asn 130 was not conserved, and Ile 125 and Ser 128 were replaced with conservative (similar in charge and polarity) substitutions of Leu 125 in all HEV analyzed and Gly 128 in 17 of 20 HEV analyzed (3 of 20 were Ser), respectively. Rupintrivir demonstrated comparable in vitro antiviral activity against the HEV (mean EC50 of 65 nM) and HRV (mean EC50 of 22 nM) strains in which matching sequence data were available (n = 41; Table 1) (46). This suggests that the differences in level of conservation (11 and 13 of 14 amino acids conserved for HEV and HRV, respectively) did not have an effect on rupintrivir antiviral activity.

FIG. 1.

FIG. 1.

Interactions between HRV 2 3C protease and rupintrivir. All amino acid residues having side chains interacting directly with the bound inhibitor are included (Phe 25, His 40, Glu 71, Ile 125, Leu 127, Ser 128, Asn 130, Thr 142, Cys 147, His 161, Gly 163, Gly 164, Asn 165, and Phe 170). Amino acid residues (Ser 144, Gly 145, and Val 162) having only polypeptide main chain interactions are shown. Other amino acid residues (Asn 22, Gly 23, Lys 24, Leu 126, Lys 143, and Tyr 146) having only polypeptide main chain interactions are not shown. Hydrogen bonds are indicated by dashed lines. The covalent bond between the sulfur γ atom (SG) of Cys 147 and carbon 19 atom (C19) of the inhibitor is shown by a solid line. The drawing was prepared by using the program LIGPLOT (60).

FIG. 2.

FIG. 2.

FIG. 2.

Amino acid sequences of 3C proteases from HRV and HEV (n = 58) including PV (poliovirus), CVB, coxsackievirus A (CAV), EV (echovirus serotypes 6, 9, 11, 12, or 30 or enterovirus serotypes 70 or 71), HRV clinical isolates (designated 100, 103, 108, 113, 119, 138, J, and Q), and HRV field isolates (M24, M36, and MR). Dashes represent homologous sequences. Fourteen amino acids of 3C protease that have side chain interactions with rupintrivir, including the catalytic triad (His 40, Cys 147, and Glu 71), are boxed in gray.

Despite the strict conservation of amino acids involved in critical binding interactions with rupintrivir, the overall level of diversity among the complete 3C proteases from the picornaviruses evaluated (n = 58) is reflected in the low overall homology at the amino acid level (42 of 183 amino acids, or <23% identity and approximately 80% dissimilarity) (Fig. 3). Furthermore, the levels of variability between the HRV and HEV strains evaluated were comparable to the levels of variability observed among all HRV isolates evaluated (≤41% amino acid identity). Consistent with the two genetic subgroups A and B that have been previously identified (26, 35, 55), HRV isolates also appear to cluster into two groups based on the level of homology of their 3C proteases. Specifically, one cluster contained HRV serotypes 14, 3, 17, and 52 (HRV group B viruses) and, with the exception of HRV 87, all other HRV analyzed belong to a second cluster, identified as HRV group A viruses. Although both subgroups share less than 42% amino acid sequence identity, within each subgroup they share comparable homology (greater than 73 and 77% for groups A and B, respectively). In addition, sequence analysis of 3C proteases from the HRV clinical isolates derived following inoculation of subjects in experimental human clinical studies were identical to the specific challenge virus strain utilized, e.g., HRV Hanks and isolates J and Q; HRV 39 and isolates 119 and 138. In a similar manner, the three field isolates (M24, M36, and MR) demonstrated greater than 95% homology with other known HRV numbered serotypes.

FIG. 3.

FIG. 3.

Phylogenetic tree of HRV and HEV based on homology of 3C protease amino acid residues was generated by MegAlign by using Clustal W analysis as described in Materials and Methods. The length of each pair of branches represents the distance between the sequences, and the horizontal axis indicates the approximate percent dissimilarities of the 3C protease amino acid sequences.

HRV serotype 87 is currently considered an unclassified rhinovirus falling into neither group A nor B. The comparison of the 3C protease amino acid sequences for HRV 87 demonstrated less than 40% amino acid homology with other HRV strains but a greater similarity (53 to 83% identity) with HEV strains. In this analysis, HRV 87 related most closely to HEV 70, sharing 83% identity. To confirm this finding, we performed sequence analyses in the 3C protease-coding region of a second independent lot of HRV 87 as well as the 5′ untranslated region of the genome. Results confirmed our observation, demonstrating that HRV serotype 87 is more closely related to HEV than to HRV (data not shown).

DISCUSSION

Prevention and treatment of picornavirus infection remain a significant unmet medical need. Given the remarkable heterogeneity of HRV (composed of over 200 different serotypes) and HEV (including poliovirus, coxsackievirus, and echovirus), an antiviral agent should at minimum demonstrate significant antiviral potency against the majority of virus serotypes and strains. Although small molecule inhibitors targeting viral attachment and uncoating have demonstrated antiviral activity, these inhibitors are often limited in their spectrum of activity (2, 34, 46). The clinical relevance of this limitation was recently shown for pleconaril; in two pivotal human clinical trials 13% of baseline isolates were not susceptible to pleconaril (16).

We have focused our research efforts on the HRV 3C protease, an enzyme that mediates viral proteolytic processing and is absolutely required for virus replication. Nucleotide sequence comparisons of a limited number of HRV serotypes had revealed considerable variability in the 3C protease-coding regions but strict conservation of critical amino acids located in the enzyme active sites, suggesting that inhibitors against 3C should have broad-spectrum activity. This was demonstrated in studies with rupintrivir, a novel inhibitor of 3C protease. In in vitro antiviral assays, rupintrivir was active against all numbered HRV serotypes, untyped field isolates, and related picornaviruses tested (34, 46). Moreover, rupintrivir demonstrated comparable antiviral potencies against all HRV serotypes tested, exhibiting only a 27-fold range in EC50 values (46). This was in contrast to results observed with pirodavir and pleconaril, inhibitors of capsid function. These compounds demonstrated a wider range in antiviral potencies (1,590- to 2,707-fold range in EC50 values) and were not active against all HRV serotypes tested. In this study, we have further evaluated the in vitro antiviral activity of rupintrivir against HRV clinical isolates that were derived from nasal lavage specimens obtained from human volunteers after experimental challenge as well as four additional HEV strains. Consistent with results from previous studies, rupintrivir inhibited the replication of all isolates tested. Importantly, rupintrivir demonstrated comparable antiviral potencies across all picornaviruses tested (n = 125; 61-fold range in EC50 values) (34, 46).

To better understand the relationship between the broad-spectrum antiviral activity of rupintrivir and the level of homology among different 3C protease-coding regions, we performed nucleotide sequence analyses on 33 HRV numbered serotypes and clinical isolates. Previous studies by Matthews et al. (41) show that extensive interactions occur between rupintrivir and the substrate (inhibitor) binding pocket residues of 3C protease. Specifically, examination of the high-resolution X-ray structure of rupintrivir bound to type 2 rhinovirus 3C protease (41) indicates extensive protein-inhibitor interactions involving 23 different amino acid residues. Individual amino side chains are involved in protein-inhibitor contacts for 14 of these residues, while for the remaining nine interacting residues, only polypeptide main chain atoms directly contact rupintrivir. We anticipate that side chain variation (nonconservation) for residues making only main chain contact with rupintrivir will have a minor effect on inhibitor binding. On the other hand, serotype-dependent substitutions for amino acids having side chain interactions with rupintrivir could be expected to directly affect inhibitor binding. In this regard, it is remarkable that of the 14 residues involved in such side chain-mediated interactions with rupintrivir, 13 are identically conserved in all known HRV 3C protease sequences. The single exception is Asn 130 (type 2) located at the back of the S2 specificity pocket. It is known that other 3C protease inhibitors with bulky substituents that penetrate deeply into the S2 pocket lose potency against certain HRV serotypes (41). It is likely that the high conservation of these extended binding site residues is directly related to stringent requirements for recognition, binding, and cleavage of viral precursor polyprotein sequences. Among HEV strains, Asn 130 was also not conserved, and Ile 125 and Ser 128 were replaced with conservative substitutions. The small variations in the EC50 values are well below the level at which they could be rationalized by structure. Since the MOIs utilized in these experiments did not have a significant effect on virus susceptibility to rupintrivir (different MOIs yield comparable EC50 values), it is likely that these variations are due to assay variability. Overall, these observations provide a compelling molecular explanation for the broad-spectrum potency of rupintrivir across a range of different picornavirus strains.

With the exception of HRV 87, phylogenetic analyses based on 3C protease-deduced amino acid sequences also confirm the segregation of numbered HRV prototype strains into two known genetic groups (5, 26, 43, 55). Specifically, HRV serotypes 3, 14, 17, and 52 appeared to form one cluster (group B), and all others appeared to cluster in the other group (group A). In terms of picornavirus evolution, it is noteworthy that relationships that have been defined for other regions of the genome (e.g., 5′ untranslated region and VP1 and VP4/2 capsid proteins in the P1 region) are also exemplified in 3C protease, which occupies the terminal third or P3 region of the genome. The preservation of these relationships suggests that significant intertypic recombination has not occurred (51) and implies that recombination is not a primary mechanism for diversification in HRV. Consistent with analyses based on other regions of the picornavirus gene regions (1, 27, 37, 55), our phylogenetic analysis of 3C protease amino acid sequences also suggests that HRV 87 is more closely related to other HEV than to other HRV. Although HRV 87 also differs from other HRV in its receptor specificity, HRV 87 exhibits acid lability, a property typical for HRV and not other HEV. Despite these differences and consistent with the level of 3C protease homology, HRV 87 demonstrates susceptibility to rupintrivir comparable to other HRV and HEV.

In this study, we utilized nucleotide sequence and structural analyses to better understand the molecular basis for the broad-spectrum antiviral activity of rupintrivir. In summary, the absolute conservation of the amino acids involved in side chain interactions with rupintrivir explains rupintrivir's broad-spectrum activity and comparable antiviral potencies across diverse picornavirus strains and further highlights 3C protease as a novel antiviral target.

Acknowledgments

We thank George Smith III for his critical review of the manuscript and Jules Beardsley for help in preparation of the manuscript.

REFERENCES

  • 1.Andeweg, A. C., T. M. Bestebroer, M. Huybreghs, T. G. Kimman, and J. C. De Jong. 1999. Improved detection of rhinoviruses in clinical samples by using a newly developed nested reverse transcription-PCR assay. J. Clin. Microbiol. 37:524-530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Andries, K., B. Dewindt, J. Snoeks, L. Wouters, H. Moereels, P. J. Lewi, and P. A. Janssen. 1990. Two groups of rhinoviruses revealed by a panel of antiviral compounds present sequence divergence and differential pathogenicity. J. Virol. 64:1117-1123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Arruda, E., and F. G. Hayden. 1995. Clinical studies of antiviral agents for picornaviral infections, p. 321-355. In D. J. Jeffries and E. De Clerq (ed.), Antiviral chemotherapy. John Wiley & Sons Ltd., Chichester, N.Y.
  • 4.Arruda, E., A. Pitkaranta, T. J. Witek, Jr., C. A. Doyle, and F. G. Hayden. 1997. Frequency and natural history of rhinovirus infections in adults during autumn. J. Clin. Microbiol. 35:2864-2868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Blomqvist, S., C. Savolainen, L. Raman, M. Roivainen, and T. Hovi. 2002. Human rhinovirus 87 and enterovirus 68 represent a unique serotype with rhinovirus and enterovirus features. J. Clin. Microbiol. 40:4218-4223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Brown, B. A., and M. A. Pallansch. 1995. Complete nucleotide sequence of enterovirus 71 is distinct from poliovirus. Virus Res. 39:195-205. [DOI] [PubMed] [Google Scholar]
  • 7.Chang, K. H., P. Auvinen, T. Hyypiä, and G. Stanway. 1989. The nucleotide sequence of coxsackievirus A9; implications for receptor binding and enterovirus classification. J. Gen. Virol. 70:3269-3280. [DOI] [PubMed] [Google Scholar]
  • 8.Couch, R. B. 2001. Rhinoviruses, p. 777-797. In D. M. Knipe, P. M. Howley, D. E. Griffin, R. A. Lamb, M. A. Martin, B. Roizman, and S. E. Straus (ed.), Fields virology, 4th ed. Lippincott Williams & Wilkins, Philadelphia, Pa.
  • 9.Dahllund, L., L. Nissinen, T. Pulli, V. P. Hyttinen, G. Stanway, and T. Hyypiä. 1995. The genome of echovirus 11. Virus Res. 35:215-222. [DOI] [PubMed] [Google Scholar]
  • 10.Dragovich, P. S., T. J. Prins, R. Zhou, S. A. Fuhrman, A. K. Patick, D. A. Matthews, C. E. Ford, J. W. Meador III, R. A. Ferre, and S. T. Worland. 1999. Structure-based design, synthesis, and biological evaluation of irreversible human rhinovirus 3C protease inhibitors. 3. Structure-activity studies of ketomethylene-containing peptidomimetics. J. Med. Chem. 42:1203-1212. [DOI] [PubMed] [Google Scholar]
  • 11.Dragovich, P. S., T. J. Prins, R. Zhou, S. E. Webber, J. T. Marakovits, S. A. Fuhrman, A. K. Patick, D. A. Matthews, C. A. Lee, C. E. Ford, B. J. Burke, P. A. Rejto, T. F. Hendrickson, T. Tuntland, E. L. Brown, J. W. Meador III, R. A. Ferre, J. E. Harr, M. B. Kosa, and S. T. Worland. 1999. Structure-based design, synthesis, and biological evaluation of irreversible human rhinovirus 3C protease inhibitors. 4. Incorporation of P 1 lactam moieties as l-glutamine replacements. J. Med. Chem. 42:1213-1224. [DOI] [PubMed] [Google Scholar]
  • 12.Dragovich, P. S., S. E. Webber, R. E. Babine, S. A. Fuhrman, A. K. Patick, D. A. Matthews, C. A. Lee, S. H. Reich, T. J. Prins, J. T. Marakovits, E. S. Littlefield, R. Zhou, J. Tikhe, C. E. Ford, M. Wallace, J. W. Meador III, R. A. Ferre, E. L. Brown, S. L. Binford, J. E. V. Harr, D. M. DeLisle, and S. T. Worland. 1998. Structure-based design, synthesis, and biological evaluation of irreversible human rhinovirus 3C protease inhibitors. 1. Michael acceptor structure-activity studies. J. Med. Chem. 41:2806-2818. [DOI] [PubMed] [Google Scholar]
  • 13.Dragovich, P. S., S. E. Webber, R. E. Babine, S. A. Fuhrman, A. K. Patick, D. A. Matthews, S. H. Reich, J. T. Marakovits, T. J. Prins, R. Zhou, J. Tikhe, E. S. Littlefield, T. M. Bleckman, M. Wallace, T. Little, C. E. Ford, J. W. Meador III, R. A. Ferre, E. L. Brown, S. L. Binford, D. M. DeLisle, and S. T. Worland. 1998. Structure-based design, synthesis, and biological evaluation of irreversible human rhinovirus 3C protease inhibitors. 2. Peptide structure-activity studies. J. Med. Chem. 41:2819-2834. [DOI] [PubMed] [Google Scholar]
  • 14.Dragovich, P. S., R. Zhou, D. J. Skalitzky, S. A. Fuhrman, A. K. Patick, C. E. Ford, J. W. Meador III, and S. T. Worland. 1999. Solid-phase synthesis of irreversible human rhinovirus 3C protease inhibitors. 1. Optimization of tripeptides incorporating N-terminal amides. Bioorg. Med. Chem. 7:589-598. [DOI] [PubMed] [Google Scholar]
  • 15.Duechler, M., T. Skern, W. Sommergruber, C. Neubauer, P. Gruendler, I. Fogy, D. Blaas, and E. Kuechler. 1987. Evolutionary relationships within the human rhinovirus genus: comparison of serotypes 89, 2, and 14. Proc. Natl. Acad. Sci. USA 84:2605-2609. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Fleischer, R., and K. Laessig. 2003. Safety and efficacy evaluation of pleconaril for treatment of the common cold. Clin. Infect. Dis. 37:1722. [DOI] [PubMed] [Google Scholar]
  • 17.Fromtling, R. A., and J. Castañer. 1997. VP-63843 pleconaril WIN-63843. Drugs Future 22:40-44. [Google Scholar]
  • 18.Gwaltney, J. M., Jr., B. Winther, J. T. Patrie, and J. O. Hendley. 2002. Combined antiviral-antimediator treatment for the common cold. J. Infect. Dis. 186:147-154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Hayden, F. G., K. Andries, and P. A. Janssen. 1992. Safety and efficacy of intranasal pirodavir (R77975) in experimental rhinovirus infection. Antimicrob. Agents Chemother. 36:727-732. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Hayden, F. G., T. Coats, K. Kim, H. A. Hassman, M. M. Blatter, B. Zhang, and S. Liu. 2002. Oral pleconaril treatment of picornavirus-associated viral respiratory illness in adults: efficacy and tolerability in phase II clinical trials. Antivir. Ther. 7:53-65. [PubMed] [Google Scholar]
  • 21.Hayden, F. G., and J. M. Gwaltney, Jr. 1984. Intranasal interferon-alpha 2 treatment of experimental rhinoviral colds. J. Infect. Dis. 150:174-180. [DOI] [PubMed] [Google Scholar]
  • 22.Hayden, F. G., G. J. Hipskind, D. H. Woerner, G. F. Eisen, M. Janssens, P. A. Janssen, and K. Andries. 1995. Intranasal pirodavir (R77,975) treatment of rhinovirus colds. Antimicrob. Agents Chemother. 39:290-294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Hayden, F. G., D. L. Kaiser, and J. K. Albrecht. 1988. Intranasal recombinant alfa-2b interferon treatment of naturally occurring common colds. Antimicrob. Agents Chemother. 32:224-230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Hayden, F. G., D. T. Herrington, T. L. Coats, K. Kim, E. C. Cooper, S. A. Vilano, S. Liu, S. Hudson, D. C. Pevear, M. Collett, M. McKinlay, and the Pleconaril Respiratory Infection Study Group. 2003. Efficacy and safety of oral pleconaril for treatment of colds due to picornaviruses in adults: results of 2 double-blind, randomized, placebo-controlled trials. Clin. Infect. Dis. 36:1523-1532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Hayden, F. G., R. B. Turner, J. M. Gwaltney, K. Chi-Burris, M. Gersten, P. Hsyu, A. K. Patick, G. J. Smith III, and L. S. Zalman. 2003. Phase II, randomized, double-blind, placebo-controlled studies of rupintrivir nasal spray 2-percent suspension for prevention and treatment of experimentally induced rhinovirus colds in healthy volunteers. Antimicrob. Agents Chemother. 47:3907-3916. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Horsnell, C., R. E. Gama, P. J. Hughes, and G. Stanway. 1995. Molecular relationships between 21 rhinovirus serotypes. J. Gen. Virol. 76:2549-2555. [DOI] [PubMed] [Google Scholar]
  • 27.Hovi, T., J. Nokso-Koivisto, S. Blomqvist, C. Savolainen, R. Raty, P. Lukkari, and T. Kilpi. 2001. Phylogenetic analysis of human rhinoviruses calls for revision of taxonomy and sequence based strain identification, abstr. LB1. Abstr. 4th Int. Symp. Influenza Other Respir. Viruses, 29 November to 2 December, 2001, Curacao, Netherlands Antilles.
  • 28.Hughes, P. J., C. North, C. H. Jellis, P. D. Minor, and G. Stanway. 1988. The nucleotide sequence of human rhinovirus 1B: molecular relationships within the rhinovirus genus. J. Gen. Virology. 69:49-50. [DOI] [PubMed] [Google Scholar]
  • 29.Hughes, P. J., C. North, P. D. Minor, and G. Stanway. 1989. The complete nucleotide sequence of coxsackievirus A21. J. Gen. Virol. 70:2943-2952. [DOI] [PubMed] [Google Scholar]
  • 30.Hughes, P. J., D. M. A. Evans, P. D. Minor, G. C. Schild, J. W. Almond, and G. Stanway. 1986. The nucleotide sequence of a type 3 poliovirus isolated during a recent outbreak of poliomyelitis in Finland. J. Gen. Virol. 67:2093-2102. [DOI] [PubMed] [Google Scholar]
  • 31.Iizuka, N., S. Kuge, and A. Nomoto. 1987. Complete nucleotide sequence of the genome of coxsackievirus B1. Virology 156:64-73. [DOI] [PubMed] [Google Scholar]
  • 32.Ishiko, H., N. Takeda, K, Miyamura, N. Kato, M. Tanimura, K. H. Lin, M. Yin-Murphy, J. S. Tam, G. F. Mu, and S. Yamazaki. 1992. Phylogenic analysis of a coxsackievirus A24 variant: the most recent worldwide pandemic was caused by progenies of a virus prevalent around 1981. Virology 187:748-759. [DOI] [PubMed] [Google Scholar]
  • 33.Jenkins, O., J. D. Booth, P. D. Minor, and J. W. Almond. 1987. The complete nucleotide sequence of coxsackievirus B4 and its comparison to other members of the Picornaviridae. J. Gen. Virol. 68:1835-1848. [DOI] [PubMed] [Google Scholar]
  • 34.Kaiser, L., C. E. Crump, and F. G. Hayden. 2000. In vitro activity of pleconaril and AG7088 against selected serotypes and clinical isolates of human rhinoviruses. Antivir. Res. 47:215-220. [DOI] [PubMed] [Google Scholar]
  • 35.King, A. M. Q., F. Brown, P. Christian, T. Hovi, T. Hyypiä, N. J. Knowles, S. M. Lemon, P. D. Minor, A. C. Palmenberg, T. Skern, and G. Stanway. 2000. Picornaviridae, p. 657-678. In M. H. V. van Regenmortel, C. M. Fauquet, D. H. L. Bishop, E. B. Carsten, M. K. Estes, S. M. Lemon, J. Maniloff, M. A. Mayo, D. J. McGeoch, C. R. Pringle, and R. B. Wickner (ed.), Virus taxonomy: classification and nomenclature of viruses. Seventh Report of the International Committee for the Taxonomy of Viruses. Academic Press, San Diego, Calif.
  • 36.Kraus, W., H. Zimmermann, H. J. Eggers, and B. Neson-Salz. 1995. Infectious cDNA clones of echovirus 12 and a variant resistant against the uncoating inhibitor rhodanine differ in seven amino acids. J. Virol. 69:5853-5858. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Ledford, R. M., N. R. Patel, T. M. Demenczuk, A. Watanyar, T. Herbertz, M. S. Collett, and D. C. Pevear. 2004. VP1 sequencing of all human rhinovirus serotypes: insights into genus phylogeny and susceptibility to antiviral capsid-binding compounds. J. Virol. 78:3663-3674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Lee, W. M., W. Wang, and R. R. Rueckert. 1995. Complete sequence of the RNA genome of human rhinovirus 16, a clinically useful common cold virus belonging to the ICAM-1 receptor group. Virus Genes 9:177-181. [DOI] [PubMed] [Google Scholar]
  • 39.Lindberg, A. M., P. O. Stalhandske, and U. Pettersson. 1987. Genome of coxsackievirus B3. Virology 156:50-63. [DOI] [PubMed] [Google Scholar]
  • 40.Makela, M. J., T. Puhakka, O. Ruuskanen, M. Leinonen, P. Saikku, M. Kimpimaki, S. Blomqvist, T. Hyypiä, and P. Arstila. 1998. Viruses and bacteria in the etiology of the common cold. J. Clin. Microbiol. 36:539-542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Matthews, D. A., P. S. Dragovich, S. E. Webber, S. A. Fuhrman, A. K. Patick, L. S. Zalman, T. F. Hendrickson, R. A. Love, T. J. Prins, J. T. Marakovits, R. Zhou, J. Tikhe, C. E. Ford, J. W. Meador III, R. A. Ferre, E. L. Brown, S. L. Binford, M. A. Brothers, D. M. DeLisle, and S. T. Worland. 1999. Structure-assisted design of mechanism-based irreversible inhibitors of human rhinovirus 3C protease with potent antiviral activity against multiple rhinovirus serotypes. Proc. Natl. Acad. Sci. USA 96:11000-11007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Matthews, D. A., W. W. Smith, R. A. Ferre, B. Condon, G. Budahazi, W. Sisson, J. E. Villafranca, C. A. Janson, H. E. McElroy, C. L. Gribskov, and S. Worland. 1994. Structure of human rhinovirus 3C protease reveals a trypsin-like polypeptide fold, RNA-binding site, and means for cleaving precursor polyprotein. Cell 77:761-771. [DOI] [PubMed] [Google Scholar]
  • 43.Mori, J., and J. P. Clewley. 1994. Polymerase chain reaction and sequencing for typing rhinovirus RNA. J. Med. Virol. 44:323-329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Nomoto, A., T. Omata, H. Toyoda, S. Kuge, H. Horie, Y. Kataoka, Y. Genba, Y. Nakano, and N. Imura. 1982. Complete nucleotide sequence of the attenuated poliovirus Sabin 1 strain genome. Proc. Natl. Acad. Sci. USA 79:5793-5797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Ostroff, R., A. Ettinger, H. La, M. Rihanek, L. Zalman, J. Meador III, A. K. Patick, S. Worland, and B. Polisky. 2001. Rapid multiserotype detection of human rhinoviruses on optically coated silicon surfaces. J. Clin. Virol. 21:105-117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Patick, A. K., S. L. Binford, M. A. Brothers, R. L. Jackson, C. E. Ford, M. D. Diem, F. Maldonado, P. S. Dragovich, R. Zhou, T. J. Prins, S. A. Fuhrman, J. W. Meador III, L. S. Zalman, D. A. Matthews and S. T., Worland. 1999. Antiviral activity of AG7088, a potent inhibitor of human rhinovirus 3C protease. Antimicrob. Agents Chemother. 43:2444-2450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Patick, A. K., and K. Potts. 1998. Protease inhibitors as antiviral agents. Clin. Microbiol. Rev. 11:614-627. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Pitkaranta, A., and F. G. Hayden. 1998. Rhinoviruses:important respiratory pathogens. Ann. Med. 30:529-537. [DOI] [PubMed] [Google Scholar]
  • 49.Poyry, T., T. Hyypiä, C. Horsnell, L. Kinnunen, T. Hovi, and G. Stanway. 1994. Molecular analysis of coxsackievirus A16 reveals a new genetic group of enteroviruses. Virology 202:982-987. [DOI] [PubMed] [Google Scholar]
  • 50.Righthand, V. F., and T. E. Gratsch. 1994. Construction of a recombinant cDNA of echovirus 6 that established a persistent in vitro infection. Virology 201:341-348. [DOI] [PubMed] [Google Scholar]
  • 51.Rodrigo, M. J., and J. Dopazo. 1995. Evolutionary analysis of the picornavirus family. J. Mol. Evol. 40:362-371. [DOI] [PubMed] [Google Scholar]
  • 52.Rotbart, H. A. 2002. Treatment of picornavirus infections. Antivir. Res. 53:83-98. [DOI] [PubMed] [Google Scholar]
  • 53.Rotbart, H. A., and A. D. Webster. 2001. Treatment of potentially life-threatening enterovirus infections with pleconaril. Clin. Infect. Dis. 32:228-235. [DOI] [PubMed] [Google Scholar]
  • 54.Ryan, M. D., O. Jenkins, P. J. Hughes, A. Brown, N. J. Knowles, D. Booth, P. D. Minor, and J. W. Almond. 1990. The complete nucleotide sequence of enterovirus type 70: relationships with other members of the picornaviridae. J. Gen. Virol. 71:2291-2299. [DOI] [PubMed] [Google Scholar]
  • 55.Savolainen, C., S. Blomqvist, M. N. Mulders, and T. Hovi. 2002. Genetic clustering of all 102 human rhinovirus prototype strains: serotype 87 is close to human enterovirus 70. J. Gen. Virol. 83:333-342. [DOI] [PubMed] [Google Scholar]
  • 56.Scudiero, D. A., R. H. Shoemaker, K. D. Paull, A. Monks, S. Tierney, T. H. Nofziger, M. J. Currens, D. Seniff, and M. R. Boyd. 1988. XTT evaluation of a soluble tetrazolium/formazan assay for cell growth and drug sensitivity in culture using human and other tumor cell lines. Cancer Res. 48:4827-4833. [PubMed] [Google Scholar]
  • 57.Skern, T., W. Sommergruber, D. Blaas, P. Gruendler, F. Fraundorfer, C. Pieler, I. Fogy, and E. Kuechler. 1985. Human rhinovirus 2: complete nucleotide sequence and proteolytic processing signals in the capsid protein region. Nucleic Acids Res. 13:2111-2126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Stanway, G., P. J. Hughes, R. C. Mountford, P. D. Minor, and J. W. Almond. 1984. The complete nucleotide sequence of a common cold virus: human rhinovirus 14. Nucleic Acids Res. 12:7859-7875. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Toyoda, H., M. Kohara, Y. Kataoka, T. Suganuma, T. Omata, N. Imura, and A. Nomoto. 1984. Complete nucleotide sequences of all three poliovirus serotype genomes. Implication for genetic relationship, gene function and antigenic determinants. J. Mol. Biol. 174:561-585. [DOI] [PubMed] [Google Scholar]
  • 60.Wallace, A. C., R. A. Laskowski, and J. M. Thornton. 1995. LIGPLOT: a program to generate schematic diagrams of protein-ligand interactions. Protein Eng. 8:127-134. [DOI] [PubMed] [Google Scholar]
  • 61.Webber, S. E., K. Okano, T. L. Little, S. Reich, Y. Xin, S. T. Worland, S. A. Fuhrman, D. A. Matthews, T. F. Hendrickson, R. A. Love, A. K. Patick, J. W. Meador III, R. A. Ferre, E. L. Brown, C. E. Ford, and S. L. Binford. 1998. Tripeptide aldehyde inhibitors of human rhinovirus 3C protease: design, synthesis, biological evaluation, and cocrystal structure solution of P1 glutamine isosteric replacements. J. Med. Chem. 41:2786-2805. [DOI] [PubMed] [Google Scholar]
  • 62.Weislow, O. S., R. Kiser, D. L. Fine, J. Bader, R. H. Shoemaker, and M. R. Boyd. 1989. New soluble-formazan assay for HIV-1 cytopathic effects: application to high-flux screening of synthetic and natural products for aids-antiviral activity. J. Natl. Cancer Inst. 81:577-586. [DOI] [PubMed] [Google Scholar]
  • 63.Werner, G., B. Rosenwirth, E. Bauer, J. M. Seifert, F. J. Werner, and J. Besemer. 1986. Molecular cloning and sequence determination of the genomic regions encoding protease and genome-linked protein of three picornaviruses. J. Virol. 57:1084-1093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Zalman, L. S., M. A. Brothers, P. S. Dragovich, R. Zhou, T. J. Prins, S. T. Worland, and A. K. Patick. 2000. Inhibition of human rhinovirus-induced cytokine production by AG7088, a human rhinovirus 3C protease inhibitor. Antimicrob. Agents Chemother. 44:1236-1241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Zell, R. H., E. Birch-Hirschfeld, U. Fortmuller, A. Henke, and A. Stelzner. Unpublished data.
  • 66.Zell, R. H. 1997. Unpublished data.
  • 67.Zell, R. H. 1999. Unpublished data.
  • 68.Zhang, G., G. Wilsden, N. J. Knowles, and J. W. McCauley. 1993. Complete nucleotide sequence of a coxsackie B5 virus and its relationship to swine vesicular disease virus. J. Gen. Virol. 74:845-853. [DOI] [PubMed] [Google Scholar]
  • 69.Zimmermann, H., H. J. Eggers, and B. Nelson-Salz. 1996. Molecular cloning and sequence determination of the complete genome of the virulent echovirus 9 strain Barty. Virus Genes 12:149-154. [DOI] [PubMed] [Google Scholar]

Articles from Antimicrobial Agents and Chemotherapy are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES