Skip to main content
Journal of Virology logoLink to Journal of Virology
. 2006 Nov 1;81(7):3051–3057. doi: 10.1128/JVI.01466-06

The Novel Human Coronaviruses NL63 and HKU1

Krzysztof Pyrc 1, Ben Berkhout 1, Lia van der Hoek 1,*
PMCID: PMC1866027  PMID: 17079323

In the mid-sixties of the previous century, the first two human coronaviruses (HCoV) were identified: HCoV-229E and HCoV-OC43 (29, 50, 70). These two human coronaviruses were studied extensively from approximately 1965 to the mid-1980s (7, 37, 49, 50, 57, 70). HCoV-229E is a member of the group I coronaviruses, and HCoV-OC43 is a member of group II. Besides the human coronaviruses, there are several group I and group II animal coronaviruses that infect cattle, pigs, cats, dogs, mice, and other animals (33). There is one additional branch, the group III coronaviruses, which are found exclusively in birds (33).

By infecting healthy volunteers, researchers learned that infection with HCoV-229E or HCoV-OC43 results in a common cold (7, 8, 29), and since then, HCoVs have been considered to be relatively harmless respiratory pathogens. This image was roughly disturbed when severe acute respiratory syndrome (SARS)-CoV was introduced into the human population in the winter of 2002 to 2003 in China. SARS-CoV causes a severe respiratory illness with high morbidity and mortality (16, 40). The virus originated from a wild-animal reservoir, most likely bats (43, 46), and was transmitted to humans via infected civet cats. The epidemic was halted in 2003 by a highly effective global public health response, and SARS-CoV is not currently circulating in humans. However, the SARS outbreak brought coronaviruses back into the limelight, and a renewed interest in this virus family resulted in the identification of two more human coronaviruses. We discovered HCoV-NL63, a novel member of group I, in a child with bronchiolitis in The Netherlands (74). HCoV-HKU1, a novel group II virus from an adult with chronic pulmonary disease in Hong Kong, was described in 2005 (81). An animal model is currently lacking for these previously unknown viruses. Nevertheless, since the discovery of HCoV-NL63 and HCoV-HKU1, much knowledge about these viruses has been gained. Several groups have studied their worldwide spread, association with human disease, replication characteristics, genome organization, and genetic diversity.

CORONAVIRUS GENOME ARCHITECTURE

Coronaviruses are positive-stranded RNA viruses, with the largest viral genome of the RNA viruses (27 to 33 kb) (reviewed in reference 47). The single-stranded genome is capped and polyadenylated. The virus particle is enveloped and carries extended spike proteins on the membrane surface, providing the typical crown-like structure (crown = corona) seen by electron microscopy. The coronaviruses share a conserved organization of their RNA genomes (Fig. 1). The 5′ two-thirds of the genome contains the large 1a and 1b open reading frames (ORFs) encoding the proteins necessary for RNA replication (the nonstructural proteins), whereas the 3′ one-third contains the ORFs coding for the structural proteins: hemagglutinin esterase (HE) (only for group IIa), spike protein (S), envelope protein (E), membrane protein (M), and nucleocapsid protein (N) (27, 56, 62). Interspersed among the structural protein genes one can find accessory protein genes that differ among species in number and position.

FIG. 1.

FIG. 1.

Schematic organization of the HCoV-NL63 and HCoV-HKU1 genomes. All ORFs and the 3′ and 5′ untranslated regions (UTR) are marked. Below the scheme, the corresponding subgenomic (sg) mRNAs generated during the discontinuous transcription process are shown (putative for HCoV-HKU1). RFS represents the putative ribosomal frameshifting element between the 1a and 1b genes.

A common transcription regulatory sequence (TRS) is located upstream of the 5′ end of each ORF. This TRS sequence is essential for the formation of subgenomic mRNAs. The subgenomic mRNAs are generated by discontinuous minus-strand synthesis and are copied into plus-strand mRNAs, which all have a common leader sequence of about 70 nucleotides at their 5′ ends that is identical to the sequence at the 5′ end of the genomic RNA. Each subgenomic mRNA and the viral genomic RNA, which also serves as an mRNA, is translated to yield, in most cases, only a single protein encoded by the ORF located immediately downstream of the TRS sequence (27, 42, 67). Some subgenomic mRNAs are bicistronic, and in these cases, translation of a second ORF depends on an internal ribosomal entry site (68).

The replicase proteins are synthesized from the partially overlapping ORF1a and ORF1b, and ribosomal access to ORF1b is regulated by a hairpin-pseudoknot structure that acts as a ribosomal frameshifting element (RFS) (52). The coronavirus ORF1ab is translated as a single polyprotein that is cotranslationally cleaved by two proteases encoded in the 5′ region of the 1ab protein gene. A papain-like protease is encoded by the nonstructural protein 3 (nsp3) gene located near the 5′ terminus of the genome. This protease consists of two domains: PL1pro and PL2pro. The nsp5 gene encodes the second protease of coronaviruses. It has a predicted serine-like protease activity and is designated the “main” protease (Mpro) to stress its dominant function. This protease processes the majority of cleavage sites between the protein domains of the 1ab polyprotein (86).

GENOME ORGANIZATION OF HCoV-NL63 AND HCoV-HKU1

The RNA genome of HCoV-NL63 is 27,553 nucleotides, with a poly(A) tail (Fig. 1). With a GC content of 34%, HCoV-NL63 has one of the lowest GC contents of the coronaviruses, for which GC content ranges from 32 to 42% (74). Untranslated regions of 286 and 287 nucleotides are present at the 5′ and 3′ termini, respectively.

Genes predicted to encode the S, E, M, and N proteins are found in the 3′ part of the HCoV-NL63 genome. The HE gene, which is present in some group II coronaviruses, is absent, and there is only a single, monocistronic accessory protein ORF (ORF3) located between the S and E genes. Subgenomic mRNAs are generated for all ORFs (S, ORF3, E, M, and N), and the core sequence of the TRS of HCoV-NL63 is defined as AACUAAA (56). This sequence is situated upstream of every ORF except for the E ORF, which contains the suboptimal core sequence AACUAUA. Interestingly, a 13-nucleotide sequence with perfect homology to the leader sequence is situated upstream of the suboptimal E TRS. Annealing of this 13-nucleotide sequence to the leader sequence may act as a compensatory mechanism for the disturbed leader-TRS/body-TRS interaction (56).

The 1a/1b ORF of HCoV-NL63 contains a potential elaborated pseudoknot structure that triggers a −1 ribosomal frameshift at position 12439 to translate the complete 1ab polyprotein (54). The nsp3 and nsp5 genes of HCoV-NL63 encode the two putative viral proteases, papain-like protease and Mpro, respectively. The cleavage sites recognized by these enzymes are predicted to be essentially identical to those of the other coronaviruses, with one exception. At the nsp13/nsp14 junction, the cleavage site contains a histidine at position P1 instead of the regular glutamine (54). This altered cleavage site is present not only in laboratory strains of HCoV-NL63 (strain Amsterdam 1 and strain NL) but also in two clinical isolates (Amsterdam 57 and Amsterdam 496). It has not been formally demonstrated that this noncanonical cleavage site is indeed functional.

The genome of HCoV-HKU1 is a 29,926-nucleotide, polyadenylated RNA (81) (Fig. 1). The GC content is 32%, the lowest among all known coronaviruses. The genome organization is the same as that of other group II coronaviruses, with the characteristic gene order 1a, 1b, HE, S, E, M, and N. Furthermore, accessory protein genes are present between the S and E genes (ORF4) and at the position of the N gene (ORF8). The TRS is presumably located within the AAUCUAAAC sequence, which precedes each ORF except E. As in sialodacryoadenitis virus and mouse hepatitis virus (MHV), translation of the E protein possibly occurs via an internal ribosomal entry site. The 3′ untranslated region contains a predicted stem-loop structure immediately downstream of the N ORF (nucleotide position 29647 to 29711) (81). Further downstream, a pseudoknot structure is present at nucleotide position 29708 to 29760. Both RNA structures are conserved in group II coronaviruses and are critical for virus replication (26, 78).

The large 1ab polyprotein of HCoV-HKU1 is synthesized via a −1 ribosomal frameshift at a putative pseudoknot structure (position 13594). In the N terminus of nsp3 (containing PL1pro and PL2pro), tandem copies of a 30-base repeat encoding NDDEDVVTGD are present, followed by 30-base regions that encode NNDEEIVTGD and NDDQIVVTGD located within the acidic domain upstream of PL1pro. These acidic repeats are not observed in other coronaviruses, and their function is still unknown. The numbers of repeats vary among HCoV-HKU1 isolates (71, 83).

The characteristic amino acid sequences required for proteolytic cleavage of the HCoV-HKU1 1ab polyprotein by the papain-like protease and Mpro are present. However, it has been suggested that HCoV-HKU1 shares the unusual cleavage site at nsp13/nsp14 with HCoV-NL63 (80). When the nsp13/nsp14 cleavage site of HCoV-HKU1 is aligned with those of other group II coronaviruses, the P1 position is occupied by a histidine instead of the regular glutamine. This is the same alignment as described for HCoV-NL63 (54). However, HCoV-HKU1 also has a standard cleavage site with the regular glutamine at the P1 position for nsp13/nsp14. This standard site is recognized by Zcurve_CoV 2.0 software (24) and is located 24 residues amino terminally of the unusual histidine-containing cleavage site. Future experiments should resolve which site is cleaved by HKU1-Mpro.

CORONAVIRUS FUSION PROTEIN

The coronavirus spike protein is a membrane-anchored glycoprotein that is found on the virion surface in a trimeric form. The N-terminal half of the protein (S1) contains the receptor-binding domain, and the C-terminal half (S2) is the membrane-anchored domain with fusogenic activity (Fig. 2). Similar to those of other class I fusion proteins, the S2 domain contains two heptad repeat (HR) regions, of which one (HR2) is located close to the transmembrane anchor. Class I fusion proteins carry a hydrophobic fusion peptide at or close to the N terminus of the membrane fusion subunit. For SARS-CoV, a fusion peptide was recognized at the N terminus of the S2 subunit (59). Thus, the spike protein displays strong similarities to class I virus fusion proteins, but there is one unusual characteristic. It does not require cleavage into S1 and S2 for coronavirus infection, unlike other class I fusion proteins (14).

FIG. 2.

FIG. 2.

Schematic representation of the spike protein of HCoV-NL63 and HCoV-HKU1. The signal peptide was identified by using the SignalIP 3.0 server (4, 53). The border between the S1 and S2 domains corresponds to the predicted furin-like enzyme cleavage site identified with the ProP 1.0 server (17). The yellow dots represent N-glycosylation sites predicted with the NetNGlyc1.0 server (R. Gupta, E. Jung, and S. Brunak, unpublished data). Heptad repeat regions (HR1 and HR2) were identified by using protein alignments. The transmembrane domain was identified with the TMHMM 2.0 server (39, 66).

Coronavirus entry into the target cell is facilitated either by direct fusion with the plasma membrane or by endocytosis and subsequent fusion to the endosomal membrane. Viruses that use pH-dependent endocytosis are sensitive to treatment with lysosomotropic agents that lower the endosomal pH. SARS-CoV and HCoV-229E are examples of coronaviruses that use endosomal entry (5, 30, 63). Treatment with lysosomotropic agents like bafilomycin A, chloroquine, and NH4Cl inhibits entry of these viruses.

SPIKE PROTEIN OF HCoV-NL63 AND HCoV-HKU1

The N-terminal part of the S protein of HCoV-NL63 contains a unique 179-amino-acid domain that is not present in other coronaviruses, including its closest relative, HCoV-229E. This domain of the HCoV-NL63 S protein contains several potential glycosylation sites (Fig. 2) and, in fact, represents the most-variable region of the complete HCoV-NL63 genome, suggesting a role in immune evasion (73). On the other hand, strains of MHV that differ in virulence and tissue tropism have deletions of different sizes in the N-terminal domain of S (58). Furthermore, the difference between transmissible gastroenteritis virus (TGEV; an enterotropic porcine virus) and porcine respiratory coronavirus (PRCoV; a respiratory porcine coronavirus) is also pinpointed to a 200-amino-acid deletion in the N-terminal region of S (77). Thus, the N-terminal region of NL63-S might also be a determinant of cell tropism and the pathogenicity of the virus.

Some group II and group III coronaviruses possess a furin-specific cleavage site between the S1 and S2 domains. A furin cleavage site is lacking in HCoV-NL63. This is not unusual, as none of the group I coronaviruses harbors a furin cleavage site between the S1 and S2 domains. The S1 domain contains the receptor-binding region (see Cell Tropism and Receptor Usage, below), and the S2 domain is involved in fusion of viral and cellular membranes. The S2 domain contains two heptad repeat regions: HR1 (residues 949 to 1064) and HR2 (residues 1234 to 1293). As in the other group I coronaviruses, the NL63-HRs are longer than those of group II or III coronaviruses (Fig. 2) (6). Both HRs of HCoV-NL63 contain a 14-amino-acid insert (two extra helical turns). After binding of a virus to its target cell, the HR regions change their conformation, which is required for membrane fusion. The HR1 and HR2 peptides eventually assemble into the hexameric form, the very stable six-helix bundle (55). An NL63-HR2 peptide can efficiently block HCoV-NL63 infection, which confirms that HR1 and HR2 interactions are vital during HCoV-NL63 replication (55).

It is not very clear whether HCoV-NL63 entry occurs via fusion with the plasma membrane or via endocytosis and subsequent fusion with the endosomal membrane. HCoV-NL63 entry is not very sensitive to lysosomotropic agents (32, 34). Incubation with bafilomycin A only mildly affects HCoV-NL63 S-mediated entry, indicating that internalization of HCoV-NL63 into its host cell is less dependent on low pH than are SARS-CoV and HCoV-229E entry. It is possible that HCoV-NL63 uses the endosomes but is not strictly dependent on pH-dependent cleavage or that fusion can occur via the plasma membrane.

The S protein of HCoV-HKU1 does contain a potential furin cleavage site between the S1 and S2 domains: RRKRR, with cleavage predicted to occur between residues 760 and 761 (Fig. 2) (81). The S2 domain contains two heptad repeats, located at residues 982 to 1083 (HR1) and 1250 to 1297 (HR2), and the S1 domain presumably contains the receptor-binding site. However, which cellular surface molecule is used as a receptor by HCoV-HKU1 is currently unknown.

CELL TROPISM AND RECEPTOR USAGE

HCoV-NL63 can be cultured in monkey epithelial cell lines (LLC-MK2, Vero E6, or Vero B4) (23, 60, 74). This preference for monkey epithelial cells is shared with SARS-CoV, but the closest relative of HCoV-NL63, HCoV-229E, cannot replicate on these cells. It was generally thought that all group I coronaviruses use CD13, also known as aminopeptidase N, as a receptor. CD13 usage has been described for HCoV-229E, porcine coronaviruses TGEV and PRCoV, and the feline and canine coronaviruses (15, 69, 85). Surprisingly, HCoV-NL63 is not able to use CD13 as a receptor for cell entry (31). A detailed analysis of cell tropism revealed that certain human cells, namely Huh-7 cells (a human hepatocellular carcinoma cell line), can also be infected by HCoV-NL63 (31). This underscores the similarity of HCoV-NL63 and SARS-CoV, which can readily infect Huh-7 cells. This shared cell tropism is suggestive of shared receptor usage. SARS-CoV is unique among the coronaviruses in that it uses angiotensin-converting enzyme 2 (ACE2) to bind and enter target cells (45). ACE2 is a surface molecule that was not previously known as a receptor for coronaviruses. Hofmann et al. demonstrated that HCoV-NL63 is the only other coronavirus that uses ACE2 as an entry receptor (31, 32). There is no similarity between the HCoV-NL63 and SARS-CoV spike proteins that could explain this shared receptor usage (45). The interaction between HCoV-NL63 and ACE2 is probably different from the interaction between SARS-CoV and ACE2. Mutations in ACE2 known to affect binding by SARS-CoV do not affect binding by HCoV-NL63 (32). Furthermore, the receptor binding domain of HCoV-NL63, which is located between residues 232 and 694 in the S protein, is not linear, whereas the SARS-CoV receptor binding domain is linear (79). Correct folding of the S protein is probably needed for HCoV-NL63-S to bind the ACE2 molecule (32).

The ACE2 molecule is a surface molecule that is localized to arterial and venous endothelial cells, arterial smooth muscle cells, and epithelia of the small intestine, as well as to ciliated tracheobronchial airway epithelia, thus supporting virus infection in the airways (28, 64). ACE2 is a homologue of the ACE protein, and both are key enzymes of the renin-angiotensin system. ACE activates the renin-angiotensin system by cleaving angiotensin I into angiotensin II, whereas ACE2 negatively regulates this system by inactivating angiotensin II. The renin-angiotensin system has a role in severe acute lung injury (35), with ACE2 playing a protective role in lung failure and its counterpart ACE promoting lung edema and impaired lung function during acute lung injury.

Interestingly, infection with SARS-CoV suppresses the expression of ACE2 protein, and it is hypothesized that the reduced level of ACE2 during SARS-CoV infection is the main cause of the severe pneumonia and acute, often lethal, lung failure associated with this virus (41). It is important to compare the pathogenicity of SARS-CoV and HCoV-NL63 with respect to ACE2 binding and downregulation. The reduced pathogenicity of HCoV-NL63 suggests that ACE2 receptor usage by the virus is not the only factor that determines the severity of viral pathogenicity. Further research on this subject, especially on the modulation of ACE2 expression levels during HCoV-NL63 infection, is warranted in order to understand the difference in lung pathogenicity of SARS-CoV and HCoV-NL63.

It has not been possible to culture HCoV-HKU1, despite attempts to do so in the following cell lines: RD (human rhabdomyosarcoma), I13.35 (murine macrophage), L929 (murine fibroblast), HRT-18 (colorectal adenocarcinoma), B95a (marmoset B-lymphoblastoid), mixed neuron-glia culture, MDCK (canine kidney), LLC-MK2 (rhesus monkey kidney), HEp-2 (human epithelial carcinoma), MRC-5 (human lung fibroblast), FRhK-4 (rhesus monkey kidney), A-549 (lung epithelial adenocarcinoma), BSC-1 (African green monkey kidney), CaCO2 (human colorectal adenocarcinoma), Huh-7 (human hepatoma), and Vero E6 (African green monkey kidney) (81).

DISEASES ASSOCIATED WITH HCoV-NL63 AND HCoV-HKU1 INFECTIONS

After the discovery of HCoV-NL63 and HCoV-HKU1, several groups reported infections by these viruses in different countries, illustrating that these viruses have spread worldwide (1, 2, 12, 13, 18, 21, 25, 36, 51, 65, 71, 72). The viruses can be detected in 1 to 10% of patients with acute respiratory tract infections, and double infections with other respiratory viruses are common (73). The first described cases of HCoV-NL63 infections were in young children with severe lower respiratory tract infections (LRTIs) in hospital settings (23, 74). Another study reported one HCoV-NL63-infected elderly Canadian patient who died 5 days after the onset of disease (2), showing that the severity of the respiratory disease can be substantial. On the other hand, several of the HCoV-NL63-infected patients showed relatively mild symptoms like fever, cough, sore throat, and rhinitis (2).

A clear link between HCoV-NL63 and respiratory diseases was established in the German prospective population-based study on LRTI in children less than 3 years of age (22, 38). Of the children with HCoV-NL63 infections, 45% had laryngotracheitis (croup) compared to only 6% in the control group. Multivariate analysis demonstrated that the chance of croup is 6.6 times higher in HCoV-NL63-positive children than in HCoV-NL63-negative children (75). Croup is a common manifestation of LRTIs in children. The cause was generally assumed to be a respiratory virus, e.g., one of the parainfluenza viruses, but now it is apparent that HCoV-NL63 also plays a major role in this disease.

HCoV-NL63 has also been associated with Kawasaki disease (20). Kawasaki disease is one of the most common forms of childhood vasculitis (10). It presents with prolonged fever and a polymorphic exanthem, oropharyngeal erythema, and bilateral conjunctivitis. A number of epidemiological and clinical observations previously suggested that an infectious agent might be the cause of Kawasaki disease (reviewed in reference 9). That there may be a link between HCoV-NL63 and Kawasaki disease is fascinating, but multiple groups recently questioned this association. These groups screened for HCoV-NL63 in respiratory material from Kawasaki disease patients, but none were able to confirm the link between HCoV-NL63 and Kawasaki disease (3, 11, 19, 61).

The first described HCoV-HKU1 cases were among elderly patients with major underlying diseases, in particular of the respiratory and cardiovascular systems. Of 10 patients with HCoV-HKU1 infections, 2 patients (aged 66 and 74), both with another serious underlying disease, died (82). Another Hong Kong study described 11 infected children, all with upper or lower respiratory tract infections, of which the majority had an underlying illness (8 of 11). A third Hong Kong study identified 13 HKU1-infected children with acute respiratory tract infections (44). Of these, eight had an underlying illness. In the United States, HCoV-HKU1 was identified in nine children with upper or lower respiratory tract infections; five of them had an underlying disease (21). In France, three of six HCoV-HKU1 infections were observed in patients with an underlying disease (71). These studies suggest that HCoV-HKU1 infection may aggravate the conditions of persons with an underlying disease, thus requiring hospitalization.

The respiratory symptoms accompanying a HCoV-HKU1 infection are usually rhinorrhea, fever, coughing, and wheezing, and disease manifestations include bronchiolitis and pneumonia (44, 65), but Vabret et al. suggested that HCoV-HKU1 might also be involved in gastrointestinal disease (71). They described six HCoV-HKU1 infections in France, of which the majority included upper respiratory tract illness, but three patients were admitted to the hospital because of acute enteric disease. The fact that HCoV-HKU1 was detected in stool samples from two of the patients supports the idea that this virus may also play a role in gastrointestinal disease. A large study of children with acute respiratory tract infections demonstrated an interesting association between HCoV-HKU1 and febrile seizures (44). These symptoms were more frequently observed in HCoV-HKU1-infected children than in HCoV-OC43-infected children.

HCoV-HKU1 and HCoV-NL63 infections are generally not life threatening, certainly not in otherwise healthy persons. This suggests that HCoV-NL63 and HCoV-HKU1, like HCoV-229E and HCoV-OC43, are common cold viruses (48) that can cause more-severe clinical symptoms in young children, elderly persons, and those that are immunocompromised (49, 76).

VARIABILITY

Clinical HCoV-NL63 isolates display sequence heterogeneity, and phylogenetic analyses revealed that there are two genotypic subgroups (1, 12, 51, 72, 74). This was first demonstrated by analyzing a part of the 1a ORF, and S ORF sequences have confirmed the presence of two genotypes. Surprisingly, typing based on the 1a ORF does not match S-based typing for some isolates (51), suggesting that recombinant forms circulate. Sequencing of the full-length genomes of two additional clinical isolates demonstrated that the HCoV-NL63 genome is a mosaic structure comprised of fragments of two donor sequences (56a).

The genetic variability among HCoV-HKU1 isolates suggests that this virus was introduced into the human population some time ago. The sequences of HCoV-HKU1 isolates display marked genetic variability. In one study in Hong Kong, the S and N ORF sequences from nine patients were analyzed. The majority of isolates (n = 7) belonged to one genotype (genotype A), and the other two isolates formed a separate subgroup (genotype B) (82). HCoV-HKU1 isolates from Australia belong to one genotype (n = 8) (65). Woo et al. investigated whether recombinant forms of genotype A and genotype B circulate (83). Full-genome sequencing of 22 HCoV-HKU1 isolates illustrated that, besides genotype A and genotype B, there is another strain (genotype C) that is the resultant of recombination between genotypes A and B (83).

CONCLUDING REMARKS

The renewed interest in coronaviruses since the outbreak of SARS-CoV has revealed that at least four human coronaviruses circulate worldwide, generally causing relatively mild respiratory symptoms. However, in young children, immunocompromised patients, or otherwise-weakened persons, these viruses can cause more-serious respiratory tract disease that requires hospitalization.

The ACE2 molecule, which is used as a receptor by HCoV-NL63 and SARS-CoV, has recently been associated with respiratory disease. The fact that ACE2 is also involved in protection against lung damage is intriguing and suggests opportunities to treat complications during respiratory tract infections. The fact that SARS-CoV and HCoV-NL63 use the same receptor, while their pathogenicities and disease courses are utterly different, calls for detailed comparisons of the immune responses, inflammation processes, and in vivo replication characteristics of these two viruses.

The identification and characterization of novel respiratory viruses is of obvious clinical importance. Diagnostics can be improved, and new therapies can be developed. An antiviral strategy that would target all human coronaviruses, e.g., broad-spectrum inhibitors of the Mpro enzymes (84), is promising. For HCoV-NL63, several compounds that efficiently inhibit distinct steps of the viral replication cycle have been described (55). A short interfering RNA (siRNA) targeting the S gene of HCoV-NL63 is one of these potent inhibitors. Inhalation of a cocktail of siRNAs targeting all the different coronaviruses or perhaps all respiratory viruses may be an effective and simple therapy to block viral replication in the lungs.

Acknowledgments

L.V.D.H. is supported by VIDI grant 016.066.318 from The Netherlands Organization for Scientic Research (NWO).

Footnotes

Published ahead of print on 1 November 2006.

REFERENCES

  • 1.Arden, K. E., M. D. Nissen, T. P. Sloots, and I. M. Mackay. 2005. New human coronavirus, HCoV-NL63, associated with severe lower respiratory tract disease in Australia. J. Med. Virol. 75:455-462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Bastien, N., K. Anderson, L. Hart, P. Van Caeseele, K. Brandt, D. Milley, T. Hatchette, E. C. Weiss, and Y. Li. 2005. Human coronavirus NL63 infection in Canada. J. Infect. Dis. 191:503-506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Belay, E. D., D. D. Erdman, L. J. Anderson, T. C. Peret, S. J. Schrag, B. S. Fields, J. C. Burns, and L. B. Schonberger. 2005. Kawasaki disease and human coronavirus. J. Infect. Dis. 192:352-353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Bendtsen, J. D., H. Nielsen, G. von Heijne, and S. Brunak. 2004. Improved prediction of signal peptides: SignalP 3.0. J. Mol. Biol. 340:783-795. [DOI] [PubMed] [Google Scholar]
  • 5.Blau, D. M., and K. V. Holmes. 2001. Human coronavirus HCoV-229E enters susceptible cells via the endocytic pathway. Adv. Exp. Med. Biol. 494:193-198. [DOI] [PubMed] [Google Scholar]
  • 6.Bosch, B. J., B. E. Martina, R. van der Zee, J. Lepault, B. J. Haijema, C. Versluis, A. J. Heck, R. de Groot, A. D. Osterhaus, and P. J. Rottier. 2004. Severe acute respiratory syndrome coronavirus (SARS-CoV) infection inhibition using spike protein heptad repeat-derived peptides. Proc. Natl. Acad. Sci. USA 101:8455-8460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Bradburne, A. F., M. L. Bynoe, and D. A. Tyrrell. 1967. Effects of a “new” human respiratory virus in volunteers. Br. Med. J. 3:767-769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Bradburne, A. F., and B. A. Somerset. 1972. Coronative antibody tires in sera of healthy adults and experimentally infected volunteers. J. Hyg. (London) 70:235-244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Burgner, D., and A. Harnden. 2005. Kawasaki disease: what is the epidemiology telling us about the etiology? Int. J. Infect. Dis. 9:185-194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Burns, J. C., and M. P. Glode. 2004. Kawasaki syndrome. Lancet 364:533-544. [DOI] [PubMed] [Google Scholar]
  • 11.Chang, L. Y., B. L. Chiang, C. L. Kao, M. H. Wu, P. J. Chen, B. Berkhout, H. C. Yang, and L. M. Huang. 2006. Lack of association between infection with a novel human coronavirus (HCoV), HCoV-NH, and Kawasaki disease in Taiwan. J. Infect. Dis. 193:283-286. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Chiu, S. S., K. H. Chan, K. W. Chu, S. W. Kwan, Y. Guan, L. L. Poon, and J. S. Peiris. 2005. Human coronavirus NL63 infection and other coronavirus infections in children hospitalized with acute respiratory disease in Hong Kong, China. Clin. Infect. Dis. 40:1721-1729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Choi, E. H., H. J. Lee, S. J. Kim, B. W. Eun, N. H. Kim, J. A. Lee, J. H. Lee, E. K. Song, S. H. Kim, J. Y. Park, and J. Y. Sung. 2006. The association of newly identified respiratory viruses with lower respiratory tract infections in Korean children, 2000-2005. Clin. Infect. Dis. 43:585-592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.de Haan, C. A., K. Stadler, G. J. Godeke, B. J. Bosch, and P. J. Rottier. 2004. Cleavage inhibition of the murine coronavirus spike protein by a furin-like enzyme affects cell-cell but not virus-cell fusion. J. Virol. 78:6048-6054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Delmas, B., J. Gelfi, R. L'Haridon, L. K. Vogel, H. Sjostrom, O. Noren, and H. Laude. 1992. Aminopeptidase N is a major receptor for the entero-pathogenic coronavirus TGEV. Nature 357:417-420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Drosten, C., S. Gunther, W. Preiser, S. van der Werf, H. R. Brodt, S. Becker, H. Rabenau, M. Panning, L. Kolesnikova, R. A. Fouchier, A. Berger, A. M. Burguiere, J. Cinatl, M. Eickmann, N. Escriou, K. Grywna, S. Kramme, J. C. Manuguerra, S. Muller, V. Rickerts, M. Sturmer, S. Vieth, H. D. Klenk, A. D. Osterhaus, H. Schmitz, and H. W. Doerr. 2003. Identification of a novel coronavirus in patients with severe acute respiratory syndrome. N. Engl. J. Med. 348:1967-1976. [DOI] [PubMed] [Google Scholar]
  • 17.Duckert, P., S. Brunak, and N. Blom. 2004. Prediction of proprotein convertase cleavage sites. Protein Eng. Des. Sel. 17:107-112. [DOI] [PubMed] [Google Scholar]
  • 18.Ebihara, T., R. Endo, X. Ma, N. Ishiguro, and H. Kikuta. 2005. Detection of human coronavirus NL63 in young children with bronchiolitis. J. Med. Virol. 75:463-465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Ebihara, T., R. Endo, X. Ma, N. Ishiguro, and H. Kikuta. 2005. Lack of association between New Haven coronavirus and Kawasaki disease. J. Infect. Dis. 192:351-352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Esper, F., E. D. Shapiro, C. Weibel, D. Ferguson, M. L. Landry, and J. S. Kahn. 2005. Association between a novel human coronavirus and Kawasaki disease. J. Infect. Dis. 191:499-502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Esper, F., C. Weibel, D. Ferguson, M. L. Landry, and J. S. Kahn. 2006. Coronavirus HKU1 infection in the United States. Emerg. Infect. Dis. 12:775-779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Forster, J., G. Ihorst, C. H. Rieger, V. Stephan, H. D. Frank, H. Gurth, R. Berner, A. Rohwedder, H. Werchau, M. Schumacher, T. Tsai, and G. Petersen. 2004. Prospective population-based study of viral lower respiratory tract infections in children under 3 years of age (the PRI.DE study). Eur. J. Pediatr. 163:709-716. [DOI] [PubMed] [Google Scholar]
  • 23.Fouchier, R. A., N. G. Hartwig, T. M. Bestebroer, B. Niemeyer, J. C. de Jong, J. H. Simon, and A. D. Osterhaus. 2004. A previously undescribed coronavirus associated with respiratory disease in humans. Proc. Natl. Acad. Sci. USA 101:6212-6216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Gao, F., H. Y. Ou, L. L. Chen, W. X. Zheng, and C. T. Zhang. 2003. Prediction of proteinase cleavage sites in polyproteins of coronaviruses and its applications in analyzing SARS-CoV genomes. FEBS Lett. 553:451-456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Gerna, G., G. Campanini, F. Rovida, E. Percivalle, A. Sarasini, A. Marchi, and F. Baldanti. 2006. Genetic variability of human coronavirus OC43-, 229E-, and NL63-like strains and their association with lower respiratory tract infections of hospitalized infants and immunocompromised patients. J. Med. Virol. 78:938-949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Goebel, S. J., B. Hsue, T. F. Dombrowski, and P. S. Masters. 2004. Characterization of the RNA components of a putative molecular switch in the 3′ untranslated region of the murine coronavirus genome. J. Virol. 78:669-682. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Gorbalenya, A. E., L. Enjuanes, J. Ziebuhr, and E. J. Snijder. 2006. Nidovirales: evolving the largest RNA virus genome. Virus Res. 117:17-37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Hamming, I., W. Timens, M. L. Bulthuis, A. T. Lely, G. J. Navis, and H. van Goor. 2004. Tissue distribution of ACE2 protein, the functional receptor for SARS coronavirus. A first step in understanding SARS pathogenesis. J. Pathol. 203:631-637. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Hamre, D., and J. J. Procknow. 1966. A new virus isolated from the human respiratory tract. Proc. Soc. Exp. Biol. Med. 121:190-193. [DOI] [PubMed] [Google Scholar]
  • 30.Hofmann, H., K. Hattermann, A. Marzi, T. Gramberg, M. Geier, M. Krumbiegel, S. Kuate, K. Uberla, M. Niedrig, and S. Pohlmann. 2004. S protein of severe acute respiratory syndrome-associated coronavirus mediates entry into hepatoma cell lines and is targeted by neutralizing antibodies in infected patients. J. Virol. 78:6134-6142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Hofmann, H., K. Pyrc, L. van der Hoek, M. Geier, B. Berkhout, and S. Pohlmann. 2005. Human coronavirus NL63 employs the severe acute respiratory syndrome coronavirus receptor for cellular entry. Proc. Natl. Acad. Sci. USA 102:7988-7993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Hofmann, H., G. Simmons, A. J. Rennekamp, C. Chaipan, T. Gramberg, E. Heck, M. Geier, A. Wegele, A. Marzi, P. Bates, and S. Pohlmann. 2006. Highly conserved regions within the spike proteins of human coronaviruses 229E and NL63 determine recognition of their respective cellular receptors. J. Virol. 80:8639-8652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Holmes, K. V., and M. M. C. Lai. 1996. Coronaviridae: the viruses and their replication, p. 1075-1093. In B. N. Fields, D. M. Knipe, P. M. Howley, et al. (ed.), Fields virology. Lippincott-Raven Publishers, Philadelphia, PA.
  • 34.Huang, I. C., B. J. Bosch, F. Li, W. Li, K. H. Lee, S. Ghiran, N. Vasilieva, T. S. Dermody, S. C. Harrison, P. R. Dormitzer, M. Farzan, P. J. Rottier, and H. Choe. 2006. SARS coronavirus, but not human coronavirus NL63, utilizes cathepsin L to infect ACE2-expressing cells. J. Biol. Chem. 281:3198-3203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Imai, Y., K. Kuba, S. Rao, Y. Huan, F. Guo, B. Guan, P. Yang, R. Sarao, T. Wada, H. Leong-Poi, M. A. Crackower, A. Fukamizu, C. C. Hui, L. Hein, S. Uhlig, A. S. Slutsky, C. Jiang, and J. M. Penninger. 2005. Angiotensin-converting enzyme 2 protects from severe acute lung failure. Nature 436:112-116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Kaiser, L., N. Regamey, H. Roiha, C. Deffernez, and U. Frey. 2005. Human coronavirus NL63 associated with lower respiratory tract symptoms in early life. Pediatr. Infect. Dis. J. 24:1015-1017. [DOI] [PubMed] [Google Scholar]
  • 37.Kapikian, A. Z., H. D. James, Jr., S. J. Kelly, J. H. Dees, H. C. Turner, K. McIntosh, H. W. Kim, R. H. Parrott, M. M. Vincent, and R. M. Chanock. 1969. Isolation from man of “avian infectious bronchitis virus-like” viruses (coronaviruses) similar to 229E virus, with some epidemiological observations. J. Infect. Dis. 119:282-290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.König, B., W. König, R. Arnold, H. Werchau, G. Ihorst, and J. Forster. 2004. Prospective study of human metapneumovirus infection in children less than 3 years of age. J. Clin. Microbiol. 42:4632-4635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Krogh, A., B. Larsson, G. von Heijne, and E. L. Sonnhammer. 2001. Predicting transmembrane protein topology with a hidden Markov model: application to complete genomes. J. Mol. Biol. 305:567-580. [DOI] [PubMed] [Google Scholar]
  • 40.Ksiazek, T. G., D. Erdman, C. S. Goldsmith, S. R. Zaki, T. Peret, S. Emery, S. Tong, C. Urbani, J. A. Comer, W. Lim, P. E. Rollin, S. F. Dowell, A. E. Ling, C. D. Humphrey, W. J. Shieh, J. Guarner, C. D. Paddock, P. Rota, B. Fields, J. DeRisi, J. Y. Yang, N. Cox, J. M. Hughes, J. W. LeDuc, W. J. Bellini, and L. J. Anderson. 2003. A novel coronavirus associated with severe acute respiratory syndrome. N. Engl. J. Med. 348:1953-1966. [DOI] [PubMed] [Google Scholar]
  • 41.Kuba, K., Y. Imai, S. Rao, H. Gao, F. Guo, B. Guan, Y. Huan, P. Yang, Y. Zhang, W. Deng, L. Bao, B. Zhang, G. Liu, Z. Wang, M. Chappell, Y. Liu, D. Zheng, A. Leibbrandt, T. Wada, A. S. Slutsky, D. Liu, C. Qin, C. Jiang, and J. M. Penninger. 2005. A crucial role of angiotensin converting enzyme 2 (ACE2) in SARS coronavirus-induced lung injury. Nat. Med. 11:875-879. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Lai, M. M., C. D. Patton, R. S. Baric, and S. A. Stohlman. 1983. Presence of leader sequences in the mRNA of mouse hepatitis virus. J. Virol. 46:1027-1033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Lau, S. K., P. C. Woo, K. S. Li, Y. Huang, H. W. Tsoi, B. H. Wong, S. S. Wong, S. Y. Leung, K. H. Chan, and K. Y. Yuen. 2005. Severe acute respiratory syndrome coronavirus-like virus in Chinese horseshoe bats. Proc. Natl. Acad. Sci. USA 102:14040-14045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Lau, S. K., P. C. Woo, C. C. Yip, H. Tse, H. W. Tsoi, V. C. Cheng, P. Lee, B. S. Tang, C. H. Cheung, R. A. Lee, L. Y. So, Y. L. Lau, K. H. Chan, and K. Y. Yuen. 2006. Coronavirus HKU1 and other coronavirus infections in Hong Kong. J. Clin. Microbiol. 44:2063-2071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Li, W., M. J. Moore, N. Vasilieva, J. Sui, S. K. Wong, M. A. Berne, M. Somasundaran, J. L. Sullivan, K. Luzuriaga, T. C. Greenough, H. Choe, and M. Farzan. 2003. Angiotensin-converting enzyme 2 is a functional receptor for the SARS coronavirus. Nature 426:450-454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Li, W., Z. Shi, M. Yu, W. Ren, C. Smith, J. H. Epstein, H. Wang, G. Crameri, Z. Hu, H. Zhang, J. Zhang, J. McEachern, H. Field, P. Daszak, B. T. Eaton, S. Zhang, and L. F. Wang. 2005. Bats are natural reservoirs of SARS-like coronaviruses. Science 310:676-679. [DOI] [PubMed] [Google Scholar]
  • 47.Masters, P. S. 2006. The molecular biology of coronaviruses. Adv. Virus Res. 66:193-292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.McIntosh, K. 1996. Coronaviruses, p. 1095. In B. N. Fields, D. M. Knipe, P. M. Howley, et al. (ed.), Fields virology. Lippincott-Raven Publishers, Philadelphia, PA.
  • 49.McIntosh, K., R. K. Chao, H. E. Krause, R. Wasil, H. E. Mocega, and M. A. Mufson. 1974. Coronavirus infection in acute lower respiratory tract disease of infants. J. Infect. Dis. 130:502-507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.McIntosh, K., J. H. Dees, W. B. Becker, A. Z. Kapikian, and R. M. Chanock. 1967. Recovery in tracheal organ cultures of novel viruses from patients with respiratory disease. Proc. Natl. Acad. Sci. USA 57:933-940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Moës, E., L. Vijgen, E. Keyaerts, K. Zlateva, S. Li, P. Maes, K. Pyrc, B. Berkhout, L. van der Hoek, and M. Van Ranst. 2005. A novel pancoronavirus RT-PCR assay: frequent detection of human coronavirus NL63 in children hospitalized with respiratory tract infections in Belgium. BMC Infect. Dis. 5:6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Namy, O., S. J. Moran, D. I. Stuart, R. J. Gilbert, and I. Brierley. 2006. A mechanical explanation of RNA pseudoknot function in programmed ribosomal frameshifting. Nature 441:244-247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Nielsen, H., J. Engelbrecht, S. Brunak, and G. von Heijne. 1997. Identification of prokaryotic and eukaryotic signal peptides and prediction of their cleavage sites. Protein Eng. 10:1-6. [DOI] [PubMed] [Google Scholar]
  • 54.Pyrc, K., B. Berkhout, and L. van der Hoek. 2005. Molecular characterization of human coronavirus NL63. Recent research developments in infection and immunity. Transworld Research Network, Kerala, India.
  • 55.Pyrc, K., B. J. Bosch, B. Berkhout, M. F. Jebbink, R. Dijkman, P. Rottier, and L. van der Hoek. 2006. Inhibition of HCoV-NL63 infection at early stages of the replication cycle. Antimicrob. Agents Chemother. 50:2000-2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Pyrc, K., M. F. Jebbink, B. Berkhout, and L. van der Hoek. 2004. Genome structure and transcriptional regulation of human coronavirus NL63. Virol. J. 1:7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56a.Pyrc, K., R. Dijkman, L. Deng, M. F. Jebbink, H. A. Ross, B. Berkhout, and L. van der Hoek. 2006. Mosaic structure of human coronavirus NL63, one thousand years of evolution. J. Mol. Biol. 364:964-973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Reed, S. E. 1984. The behaviour of recent isolates of human respiratory coronavirus in vitro and in volunteers: evidence of heterogeneity among 229E-related strains. J. Med. Virol. 13:179-192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Rowe, C. L., S. C. Baker, M. J. Nathan, and J. O. Fleming. 1997. Evolution of mouse hepatitis virus: detection and characterization of spike deletion variants during persistent infection. J. Virol. 71:2959-2969. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Sainz, B., Jr., J. M. Rausch, W. R. Gallaher, R. F. Garry, and W. C. Wimley. 2005. Identification and characterization of the putative fusion peptide of the severe acute respiratory syndrome-associated coronavirus spike protein. J. Virol. 79:7195-7206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Schildgen, O., M. F. Jebbink, M. de Vries, K. Pyrc, R. Dijkman, A. Simon, A. Muller, B. Kupfer, and L. van der Hoek. 7 September 2006. Identification of cell lines permissive for human coronavirus NL63. J. Virol. Methods 138:207-210. [Epub ahead of print.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Shimizu, C., H. Shike, S. C. Baker, F. Garcia, L. van der Hoek, T. W. Kuijpers, S. L. Reed, A. H. Rowley, S. T. Shulman, H. K. Talbot, J. V. Williams, and J. C. Burns. 2005. Human coronavirus NL63 is not detected in the respiratory tracts of children with acute Kawasaki disease. J. Infect. Dis. 192:1767-1771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Siddell, S. 1983. Coronavirus JHM: coding assignments of subgenomic mRNAs. J. Gen. Virol. 64:113-125. [DOI] [PubMed] [Google Scholar]
  • 63.Simmons, G., J. D. Reeves, A. J. Rennekamp, S. M. Amberg, A. J. Piefer, and P. Bates. 2004. Characterization of severe acute respiratory syndrome-associated coronavirus (SARS-CoV) spike glycoprotein-mediated viral entry. Proc. Natl. Acad. Sci. USA 101:4240-4245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Sims, A. C., R. S. Baric, B. Yount, S. E. Burkett, P. L. Collins, and R. J. Pickles. 2005. Severe acute respiratory syndrome coronavirus infection of human ciliated airway epithelia: role of ciliated cells in viral spread in the conducting airways of the lungs. J. Virol. 79:15511-15524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Sloots, T. P., P. McErlean, D. J. Speicher, K. E. Arden, M. D. Nissen, and I. M. Mackay. 2006. Evidence of human coronavirus HKU1 and human bocavirus in Australian children. J. Clin. Virol. 35:99-102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Sonnhammer, E. L., G. von Heijne, and A. Krogh. 1998. A hidden Markov model for predicting transmembrane helices in protein sequences. Proc. Int. Conf. Intel. Syst. Mol. Biol 6:175-182. [PubMed] [Google Scholar]
  • 67.Spaan, W., H. Delius, M. Skinner, J. Armstrong, P. Rottier, S. Smeekens, B. A. van der Zeijst, and S. G. Siddell. 1983. Coronavirus mRNA synthesis involves fusion of non-contiguous sequences. EMBO J. 2:1839-1844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Thiel, V., and S. G. Siddell. 1994. Internal ribosome entry in the coding region of murine hepatitis virus mRNA 5. J. Gen. Virol. 75:3041-3046. [DOI] [PubMed] [Google Scholar]
  • 69.Tresnan, D. B., R. Levis, and K. V. Holmes. 1996. Feline aminopeptidase N serves as a receptor for feline, canine, porcine, and human coronaviruses in serogroup I. J. Virol. 70:8669-8674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Tyrrell, D. A. J., and M. L. Bynoe. 1965. Cultivation of novel type of common-cold virus in organ cultures. Br. Med. J. 1:1467-1470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Vabret, A., J. Dina, S. Gouarin, J. Petitjean, S. Corbet, and F. Freymuth. 2006. Detection of the new human coronavirus HKU1: a report of 6 cases. Clin. Infect. Dis. 42:634-639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Vabret, A., T. Mourez, J. Dina, L. van der Hoek, S. Gouarin, J. Petitjean, J. Brouard, and F. Freymuth. 2005. Human coronavirus NL63, France. Emerg. Infect. Dis. 11:1225-1229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.van der Hoek, L., K. Pyrc, and B. Berkhout. 2006. Human coronavirus NL63, a new respiratory virus. FEMS Microbiol. Rev. 30:760-773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.van der Hoek, L., K. Pyrc, M. F. Jebbink, W. Vermeulen-Oost, R. J. Berkhout, K. C. Wolthers, P. M. Wertheim-van Dillen, J. Kaandorp, J. Spaargaren, and B. Berkhout. 2004. Identification of a new human coronavirus. Nat. Med. 10:368-373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.van der Hoek, L., K. Sure, G. Ihorst, A. Stang, K. Pyrc, M. F. Jebbink, G. Petersen, J. Forster, B. Berkhout, and K. Uberla. 2005. Croup is associated with the novel coronavirus NL63. PLoS. Med. 2:e240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.van Elden, L. J., A. M. van Loon, F. van Alphen, K. A. Hendriksen, A. I. Hoepelman, M. G. van Kraaij, J. J. Oosterheert, P. Schipper, R. Schuurman, and M. Nijhuis. 2004. Frequent detection of human coronaviruses in clinical specimens from patients with respiratory tract infection by use of a novel real-time reverse-transcriptase polymerase chain reaction. J. Infect. Dis. 189:652-657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Vaughn, E. M., P. G. Halbur, and P. S. Paul. 1995. Sequence comparison of porcine respiratory coronavirus isolates reveals heterogeneity in the S, 3, and 3-1 genes. J. Virol. 69:3176-3184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Williams, G. D., R. Y. Chang, and D. A. Brian. 1999. A phylogenetically conserved hairpin-type 3′ untranslated region pseudoknot functions in coronavirus RNA replication. J. Virol. 73:8349-8355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Wong, S. K., W. Li, M. J. Moore, H. Choe, and M. Farzan. 2004. A 193-amino acid fragment of the SARS coronavirus S protein efficiently binds angiotensin-converting enzyme 2. J. Biol. Chem. 279:3197-3201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Woo, P. C., Y. Huang, S. K. Lau, H. W. Tsoi, and K. Y. Yuen. 2005. In silico analysis of ORF1ab in coronavirus HKU1 genome reveals a unique putative cleavage site of coronavirus HKU1 3C-like protease. Microbiol. Immunol. 49:899-908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Woo, P. C., S. K. Lau, C. M. Chu, K. H. Chan, H. W. Tsoi, Y. Huang, B. H. Wong, R. W. Poon, J. J. Cai, W. K. Luk, L. L. Poon, S. S. Wong, Y. Guan, J. S. Peiris, and K. Y. Yuen. 2005. Characterization and complete genome sequence of a novel coronavirus, coronavirus HKU1, from patients with pneumonia. J. Virol. 79:884-895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Woo, P. C., S. K. Lau, H. W. Tsoi, Y. Huang, R. W. Poon, C. M. Chu, R. A. Lee, W. K. Luk, G. K. Wong, B. H. Wong, V. C. Cheng, B. S. Tang, A. K. Wu, R. W. Yung, H. Chen, Y. Guan, K. H. Chan, and K. Y. Yuen. 2005. Clinical and molecular epidemiological features of coronavirus HKU1-associated community-acquired pneumonia. J. Infect. Dis. 192:1898-1907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Woo, P. C., S. K. Lau, C. C. Yip, Y. Huang, H. W. Tsoi, K. H. Chan, and K. Y. Yuen. 2006. Comparative analysis of 22 coronavirus HKU1 genomes reveals a novel genotype and evidence of natural recombination in coronavirus HKU1. J. Virol. 80:7136-7145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Yang, H., W. Xie, X. Xue, K. Yang, J. Ma, W. Liang, Q. Zhao, Z. Zhou, D. Pei, J. Ziebuhr, R. Hilgenfeld, K. Y. Yuen, L. Wong, G. Gao, S. Chen, Z. Chen, D. Ma, M. Bartlam, and Z. Rao. 6 September 2005. Design of wide-spectrum inhibitors targeting coronavirus main proteases. PLoS. Biol. 3:e324. [Epub ahead of print.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Yeager, C. L., R. A. Ashmun, R. K. Williams, C. B. Cardellichio, L. H. Shapiro, A. T. Look, and K. V. Holmes. 1992. Human aminopeptidase N is a receptor for human coronavirus 229E. Nature 357:420-422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Ziebuhr, J. 2005. The coronavirus replicase. Curr. Top. Microbiol. Immunol. 287:57-94. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Virology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES