key: cord-351868-w4d45fue authors: Zuwała, Kaja; Golda, Anna; Kabala, Wojciech; Burmistrz, Michał; Zdzalik, Michal; Nowak, Paulina; Kedracka-Krok, Sylwia; Zarebski, Mirosław; Dobrucki, Jerzy; Florek, Dominik; Zeglen, Sławomir; Wojarski, Jacek; Potempa, Jan; Dubin, Grzegorz; Pyrc, Krzysztof title: The Nucleocapsid Protein of Human Coronavirus NL63 date: 2015-02-20 journal: PLoS One DOI: 10.1371/journal.pone.0117833 sha: doc_id: 351868 cord_uid: w4d45fue Human coronavirus (HCoV) NL63 was first described in 2004 and is associated with respiratory tract disease of varying severity. At the genetic and structural level, HCoV-NL63 is similar to other members of the Coronavirinae subfamily, especially human coronavirus 229E (HCoV-229E). Detailed analysis, however, reveals several unique features of the pathogen. The coronaviral nucleocapsid protein is abundantly present in infected cells. It is a multi-domain, multi-functional protein important for viral replication and a number of cellular processes. The aim of the present study was to characterize the HCoV-NL63 nucleocapsid protein. Biochemical analyses revealed that the protein shares characteristics with homologous proteins encoded in other coronaviral genomes, with the N-terminal domain responsible for nucleic acid binding and the C-terminal domain involved in protein oligomerization. Surprisingly, analysis of the subcellular localization of the N protein of HCoV-NL63 revealed that, differently than homologous proteins from other coronaviral species except for SARS-CoV, it is not present in the nucleus of infected or transfected cells. Furthermore, no significant alteration in cell cycle progression in cells expressing the protein was observed. This is in stark contrast with results obtained for other coronaviruses, except for the SARS-CoV. Coronaviruses cause a variety of diseases in animals, whereas human infections are almost exclusively associated with respiratory tract infections (RTI). Contemporary taxonomy divides the Coronavirinae subfamily into four genera (alpha, beta, gamma, and delta). Only the alpha Cell culture LLC-MK2 cells (ATCC: CCL-7; Macaca mulatta kidney epithelial cell line) were maintained in minimal essential medium (MEM), containing 2 parts of Hank's MEM and 1 part of Earle's MEM (PAA Laboratories, Austria) supplemented with 3% heat-inactivated fetal bovine serum (FBS) (PAA Laboratories, Austria), penicillin (100 U/ml), and streptomycin (100 μg/ml). Cells were cultured on T75 flasks (TPP, Switzerland) at 37°C with 5% CO 2 . 293T cells (ECACC: 12022001; Human embryonic kidney SV40 transformed, genetically modified) were maintained in Dulbecco-modified Eagle's medium (DMEM; PAA Laboratories, Austria) supplemented with 3% heat-inactivated fetal bovine serum (FBS; PAA Laboratories, Austria), penicillin (100 U/ml), and streptomycin (100 μg/ml). Cells were cultured on T75 flasks (TPP, Switzerland) at 37°C with 5% CO 2 . 293T cells (ATCC CRL-3216) were transfected with the pLKO.1.-TRC-ACE2 plasmid using polyethylenimine (PEI; Sigma-Aldrich, Poland). The plasmid was based on the Addgene plasmid 10878 [42] . At 24 h post-transfection, the cells were washed with sterile 1 × PBS and cultured at 37°C for 48 h in media supplemented with puromycin (2 μg ml -1 ) at 37°C with 5% CO 2 . Following selection, cells were passaged and the surviving clones were collected and analyzed. ACE2-expressing (ACE2 + ) cells were maintained in Dulbecco's MEM (PAA Laboratories, Austria) supplemented with 10% FBS, penicillin (100 U ml -1 ), streptomycin (100 μg ml -1 ), ciprofloxacin (5 μg ml -1 ) and puromycin (1 μg ml -1 ). 293T_ACE2 + cells were maintained as wild type cells. Human tracheobronchial epithelial cells were obtained from airway specimens resected from patients undergoing surgery under Silesian Center for Heart Diseases approved protocols. This study was approved by the bioethical committee of the Medical University of Silesia in Katowice, Poland (approval no: KNW/0022/KB1/17/10 dated on 16.02.2010) . A written informed consent was obtained from all patients (2 adult patients). Primary cells detached from human bronchi and trachea with pronase E were expanded on collagen-coated (collagen type IV, Sigma-Aldrich) plastic in bronchial epithelial growth media (BEGM) to generate passage 1 cells and plated at density of 3×10 5 cells per well on permeable Transwell supports (6.5-mmdiameter; Corning Transwell-Clear) in BEGM. Cells were cultured at 37°C in presence of 5% CO 2 until confluence. Human airway epithelium (HAE) cultures were generated by changing the media to Air Liquid Interface media (ALI) and provision of an air-liquid interface for 6 to 8 weeks to form well-differentiated, polarized cultures that resemble the in vivo structure of pseudostratified mucociliary epithelium. All procedures were performed as previously described [43] . All cell cultures were routinely screened for Mycoplasma spp. contamination using Hoechst 33258 staining. HCoV-NL63 (Amsterdam I strain) stock was generated by infecting LLC-MK2 cells. Infected cells were lysed 6 days post-infection by two freeze-thaw cycles. The virus-containing fluid was cleared by centrifugation, aliquoted and stored at -80°C. A control from mock infected cells was prepared in the same manner as the virus stocks. Virus yield was assessed by titration on fully confluent LLC-MK2 cells, according to Reed and Muench formula [44] . Cells on 96-well plates were incubated at 32°C for 6 days and the cytopathic effect was scored using an inverted microscope. All experimental procedures were conducted as previously described [45] . Virus identity was confirmed by cDNA sequencing. RNA from viral and control cultures was extracted using GeneJet RNA purification kit (Thermo Scientific, Lithuania), according to manufacturer's protocol. Isolated RNA was stored at -80°C. DNA fragments were synthesized by a third party (Genomed, Poland). Isolated viral RNA was reverse transcribed using High Capacity cDNA kit (Life Technologies, Poland) and used as a template for subsequent amplification. In order to obtain eukaryotic expression vector, NL63-N gene was amplified using primers N_NL63_5HindIII (5 0 -GTA CAA GCT TGC CAC CAT GGC TAG TGT AAA TTG GGC C-3 0 ) and N_NL63_3BamHI (5 0 -GAC TGG ATC CGC ATG CAA AAC CTC GTT GAC AAT-3 0 ) with Marathon DNA polymerase (A&A Biotechnology, Poland). Resulting PCR product was subsequently gel purified using GeneJET Gel Extraction kit (Thermo Scientific, Lithuania) and digested with HindIII and BamHI restriction enzymes. Resulting fragment was cloned into the pmaxFP-Green-N plasmid (Lonza, Switzerland) using corresponding restriction sites. Plasmids (pmaxFP-Green-N/NL63-N) were recovered in DH5α Escherichia coli (Life Technologies, Poland) and their identity and sequence were confirmed by DNA sequencing (Genomed, Poland). Expression plasmids for prokaryotic expression of the NL63-N protein were prepared using pmaxFP-Green-N/NL63-N as a template. Briefly, the fragments of N gene were amplified using primers given in parentheses: procNL63-N (5 0 -ATG CCC ATG GGC CAT CAC CAT CAT CAC CAC TCT GGC GAC GAC GAC GAC AAG GCT AGT GTA AAT TGG GCC GAT G-3 0 and 5 0 -ATG CCT CGA GTT AAT GCA AAA CCT CGT TGA CAA T-3 0 ), procNL63-20/144-N (5 0 -CAT AGG ATC CAG AAA ACC TGT ATT TTC AGG GAT CAT TTT ACA TGC CTC TTT TG-3 0 and 5 0 -CAG CAA GCT TTT AAG AGC GAT CCT CAA ACT CAA C-3 0 ) and procNL63-221/340-N (5 0 -CAT AGG ATC CAG AAA ACC TGT ATT TTC AGG GAT CTC AAC CCA GGG CTG ATA AG-3 0 and 5 0 -CAG CAA GCT TTT ATG ACT GCA TTT CTT TGA TAG-3 0 ) and cloned into the pET Duet 1 plasmid (Clonetech, USA). The element encoding 6 × His tag was introduced at the N-terminus of the gene. Three plasmids for prokaryotic expression were generated: procNL63-N, procNL63-20/144-N (for expression of the N terminal-domain), and procNL63-221/340-N for expression of the Cterminal domain. Plasmid pmaxFP-Green-N/NL63-N or control plasmids were transfected to 293T cells using cationic carrier (polyethylenimine, PEI; Sigma-Aldrich, Poland). Briefly, 2 × 10 5 cells were seeded onto collagen-coated (Purecol; Advanced Biomatrix, USA) glass coverslips in a 6-well plate. Next day media was removed, cells were washed with 1 × PBS and overlaid with 2 ml of DMEM supplemented with 4 μg of PEI and 4 μg of plasmid. 24 h post-transfection coverslips were harvested for analysis. In order to test subcellular localization of the N protein in LLC-MK2 cells, the maxFP-Green-N/NL63-N encoding RNA was prepared based on the original plasmid. Briefly, the plasmid was used as a template with primers SP6_NEGFPmRNA (5 0 -ACT GAC TGA TTT AGG TGA CAC TAT AGA AGN GAA GCT TGC CAC CAT GGC TAG TG -3 0 ) and EGFPmR-NA_R (5 0 -TTT TTT TTT TTT TTT TTT TTT CAT TAA TGC AAA ACC TCG TTG AC -3 0 ), where the 5 0 primer carries the SP6 promoter. In vitro transcription was carried out using the mMessage mMachine SP6 kit (Life Technologies, Poland). Further, RNA was polyadenylated using Poly(A) Tailing Kit (Life Technologies, Poland) and transfected into cells using TransIT-mRNA Transfection Kit (Mirus, USA), as advised by the manufacturer. RNA encoding maxFP-Green protein (control) was prepared and transfected in the same manner using primers SP6_GFPmRNA (atg cAT TTA GGT GAC ACT ATA GAT GGA GAG CGA CGA GAG CGG CCT GC) and GFPmRNA_R (TTT TTT TTT TTT TTT TTT TTT TTT TTT TTT TTT TTT TTT TTC ATT ATT CTT CAC CGG CAT CTG CAT C). In order to assess the influence of the N protein on cell cycle, the N-encoding RNA was prepared based on the original plasmid and transfected in the same manner as described above. Primers SP6_NmRNA (5 0 -TCG GCC TCG TAG GCC ATT TAG GTG ACA CTA TAG AAG NCT GAG AGA ACC CAC TGC TTA C -3 0 ) and NmRNA_R (5 0 -TTT TTT TTT TTT TTT TTT TTT CAT TAA TGC AAA ACC TCG TTG AC -3 0 ) were used, where the 5 0 primer carries the SP6 promoter. Cells transfected with mRNA encoding the N protein or control cells were harvested 48 h posttransfection by trypsinization and pelleted in sterile 1 × PBS. After fixation in 70% EtOH for 2 h on ice, cells were incubated in staining solution (50 μg/ml propidium iodide and 10 μg/ml RNase A in sterile 1 × PBS; Sigma-Aldrich, Poland) for 30 min at 37°C. The N protein was visualized with monoclonal antibody specific to NL63-N (Ingenasa, Spain) and secondary Alexa Fluor 488 goat anti-mouse antibody (Life Technologies, Poland). Cells expressing the N-NL63 protein were analyzed by flow cytometry (FACSCalibur, Becton Dickinson) as previously described [46] . Cells treated with nocodazole (Sigma-Aldrich, Poland), a mitotic spindle poison, were sampled after 24 h and evaluated as positive control for cell cycle arrest. Obtained data were analyzed using ModFit LT software (Verity Software House, USA). All experiments were conducted independently at least three times. Cells were fixed using 4% formaldehyde solution in sterile 1 × PBS for 15 minutes. Subsequently, cells were washed three times with 1 × PBS and incubated with 0.1% Triton solution in 1 × PBS to remove lipid fraction. Further, cells were incubated in blocking buffer (10% of BSA; Bio-Shop, Canada, 0.1% Tween 20; BioShop, Canada in 1 × PBS) for 60 minutes. For detection of HCoV-NL63 N protein mouse monoclonal anti-HCoV-NL63-N antibody (diluted 4000 ×, Ingenasa, Spain) was incubated with the sample for 1 hour at 4°C, followed by incubation with an anti-mouse Alexa Fluor 488 (dilution 400 ×, Thermo Fisher Scientific, Poland) for 1 hour at 4°C. For visualization of nucleic acids, DAPI dye (1 μg/ml; Sigma-Aldrich, Poland) was used. Fluorescent images were acquired with Leica TCS SP5 II confocal microscope (Leica Microsystems GmbH, Germany). Images were pre-processed using Leica Application Suite Advanced Fluorescence LAS AF v. 2.2.1 (Leica Microsystems GmbH) and further deconvolved with Huygens Essential package ver. 4.4 (Scientific Volume Imaging B.V., the Netherlands). All experiments were conducted independently at least three times. The NL63-N NTD (amino acids 2-144) and CTD (amino acids 221-340) expression constructs were designed in silico by analysis of sequence alignments, comparative modeling and literature data. The sequences and structures of the homologous coronavirus nucleocapsid polypeptides used are listed in Table 1 . The strategy is described in the Supporting Information section. Sequence sets were prepared using BLAST and SPDBV. The comparative modeling was performed with SPDBV, Coot and PyMOL [47] [48] [49] . Procedure of gene amplification and plasmid preparation is described above. In order to express the N protein, NTD and CTD in E. coli, respective plasmids were transformed to BL21 cells and further cultured in LB media supplemented with ampicillin (100 μg/ml) at 37°C, until the optical density (λ = 600nm) reached 0.5-0.6. Expression was induced by addition of IPTG (1 mM) and continued overnight at 20°C. Subsequently cells were pelleted by centrifugation and suspended in lysis buffer (50 mM Tris, 500 mM NaCl, 20 mM Imidazol pH 8.0). For the full length protein, the buffer was supplemented with 5 mM β-mercaptoethanol. Bacterial cells were lyzed by sonication, and cellular debris was removed by centrifugation. Proteins of interest were recovered by affinity chromatography (Ni Sepharose 6 Fast Flow, GE Healthcare, Poland), ion-exchange chromatography (Resource Q, GE Healthcare, Poland) and size exclusion chromatography (Superdex S75, GE Healthcare, Poland). Protein was detected using Western-blotting technique and anti-6 × His antibodies (Life Technologies, Poland). Samples for mass spectrometry were prepared by dialysis into 50 mM NH 4 HCO 3 , pH 7.8. Measurements were performed using the MicroTOF-QII mass spectrometer (Bruker, Germany) in positive ionization mode, using Appollo Source ESI sprayer. Prior to measurements the device was calibrated with TuneMix solution. The obtained MS spectra were analyzed using Data Analysis 4.0 software (Bruker, Germany). Molecular weight of proteins was confirmed using Maximum Entropy algorithm for MS spectra deconvolution (Bruker, Germany). Protein preparations were overlaid on the poly-L-lysine coated glass slides (diameter of 16 mm). Samples were fixed with 2.5% glutaraldehyde in 0.1M sodium cacodylate buffer pH = 7.4 for 20 minutes. Subsequently, samples were washed with the abovementioned buffer and gently dehydrated by using solutions of ethanol in a graded series of concentrations. Preparations were dried in a critical point dryer (Quorum Technologies, United Kingdom). Slides were mounted on holders using self-adhesive carbon discs (TAAB laboratories, United Kingdom) and sputter coated with gold (ion sputter JFC-1100E; JEOL, Japan). Electron micrographs were prepared using scanning electron microscope JSM-5410 (JEOL, Japan). is heat capacity at the T m measured with respect to the chemical baseline. The ratio C exc;max p ∕ DHðT m Þis sensitive to the shape (width) of the transition. If two-state model holds true, the van't Hoff and calorimetric enthalpies are equal within the experimental uncertainty, and so the ration = ΔH(T m )/ΔH vH should be equal to unity. Protein electrophoresis in denaturing conditions was carried out in Schagger & von Jagow system. Electrophoretic separation was carried out at 75 V (stacking) / 135 V (separation). Proteins were visualized using Coomassie Brilliant Blue G-250 (Serva, Germany). Page Ruler Plus (Thermo Scientific, Lithuania) was used as a prestained protein size marker. For EMSA assay 10 μg of RNA or DNA corresponding in sequence to the N-NL63 gene (prepared in the same manner as for the transfection of eukaryotic cells) was incubated in buffered solution (5 mM Tris, 50 mM NaCl, pH8.0) with 10 μg of the NTD or CTD for 30 minutes at room temperature. Subsequently, samples were separated on agarose gels and signal from nucleic acids was visualized with ethidium bromide staining. All experiments were conducted independently at least three times. In order to assess whether the NTD or CTD are able to form oligomers, 50 μg of the protein was mixed with glutaraldehyde (0.007%; Serva). Following 15 minute incubation at room temperature 0.5 μl of 1M Tris solution was added to the mixture, samples were mixed with protein sample buffer, denatured at 95°C and loaded onto the polyacrylamide gel. All experiments were conducted independently at least three times. The sequences of DNA, RNA and proteins used within the study correspond to those of HCoV-NL63 isolate Amsterdam 1 (GenBank accession number: NC_005831). Accession numbers for N proteins of different coronaviruses are provided in S1 File. In silico analysis of NL63-N NL63-N is a basic protein (predicted pI, 9.78) comprising 377 amino acids (aa). The predicted molecular weight is 42,252.47 Da. Literature data indicate that the full length coronaviral nucleocapsid protein consists of two folded domains linked by an unstructured region. In more details the N protein includes following elements: N-tail, N-terminal domain (NTD), R.S.A.G. rich linker, C-terminal domain (CTD) and C-tail [50] . The constructs of NTD and CTD used in this study were designed based on literature data, HCoV-NL63 N protein amino acid sequence alignment with known homologs and on the comparative analysis of currently available crystal structures of these homologs. According to Saikatendu et al. the NTD of HCoV-NL63 N encompasses residues 17-141 [51] . Our sequence alignment and structural analysis suggests that NL63-N 2-144 fragment better reflects the full N-terminal domain. NL63-N fragment encompassing its CTD was chosen exclusively on the basis of sequence alignment and structural analysis which suggests that fragment 221-340 contains full, structurally stable CTD. The sequences and structures of N proteins used in above analysis are listed in Table 1 . The analysis strategy is summarized in S1 File and the amino acid sequences of the final constructs of CTD and NTD are presented in S2 File. In silico analysis conducted using PSORT II revealed that two nuclear localization signals (NLS) are buried within NL63-N: pat4 (aa 232-KKPR-235) and pat7 (aa 234-PRWKRVP-240). No bipartite NLS were detected. The complete NL63-N protein and its CTD and NTD were expressed in E. coli BL21 cells. NTD and CTD were purified to homogeneity whereas the full length N-protein was purified to about 80% homogeneity as demonstrated by SDS-PAGE (Fig. 1) . The identity of purified proteins was confirmed with mass spectrometry (data not shown). We next used differential scanning calorimetry (DSC) to examine thermal stability of the proteins. The DSC curves for the first heating scans obtained for NL63-N, the CTD, and the NTD are shown in Fig. 2 . NL63-N underwent irreversible denaturation and showed a broad transition curve. The denaturation temperature was estimated at 45.7°C, with an enthalpy change of approximately 80 kcal/mol ( Table 2 ). However, due to low signal to noise ratio the resulting baseline was variable and these values can only be treated as rough estimates. Nevertheless, the low enthalpy value suggests that NL63-N protein is relatively unstable. Unfolding of the NTD was irreversible and accompanied by protein aggregation (indicated by the exotherm present in the high temperature region of the DSC curve). The transition temperature was 45°C and the ΔH cal was 104.4 kcal/mol. Surprisingly, thermal transition of the CTD was fully reversible, showing T t of 55.7°C, a ΔH cal of 143.6 kcal/mol, and a ΔS cal of 0.44 kcal/ Kmol. Thermal transition of the CTD was cooperative, with a van't Hoff enthalpy/calorimetric enthalpy ratio (ΔH van'Hoff / ΔH cal ) of 1.03. The coronaviral nucleocapsid forms a protective scaffold around the viral RNA [52] [53] [54] . The N protein forms oligomers via specific interactions between different regions within the protein [55] [56] [57] . To confirm this assumption for HCoV-NL63, we performed mass spectrometry analyses. The results confirmed the presence of complete N protein dimers. Furthermore, similar results were obtained for the CTD but not NTD, suggesting that CTD harbors the sites responsible for N protein dimerization ( Table 3) . The mass spectrometry results were confirmed by protein crosslinking studies. Incubating the CTD in the presence of glutaraldehyde followed by SDS-PAGE analysis revealed the presence of protein dimers and higher molecular weight oligomers (Fig. 3A) . Similar results were obtained using size exclusion chromatography, showing that~40% of the protein is present as dimers (Fig. 3B) . Dimerization was not observed for NTD. We next performed an electrophoretic mobility shift assay to determine whether the N protein interacts with nucleic acids. Briefly, samples containing nucleic acids were separated in agarose gel under native conditions in the presence/absence of the CTD or the NTD. Nucleic acids were detected by ethidium bromide staining. As shown in Fig. 4 , the NTD binds nucleic acids (both DNA and RNA), as demonstrated by retarded RNA and DNA migration. The CTD did not bind nucleic acids. We also conducted similar analysis for the complete N protein (data not shown). Obtained results suggested that the complete N protein has lower nucleic acid binding ability or is more specific compared to the NTD. However, due to rapid degradation of the complete N protein into separate domains and resulting presence of the free NTD in the solution these results were inconclusive. Coronaviral N proteins localize to the cytoplasm, where they are involved in virus replication and assembly. However, the N proteins of almost all coronaviruses (except for SARS-CoV) also localize to the nucleus or to micronuclei [58] [59] [60] [61] . To examine the subcellular localization of the NL63-N, cultures of 293T_ACE2 + , LLC-MK2, and HAE cultures were infected with the HCoV-NL63 virus. Subsequently, the cells were fixed and stained with antibodies specific for the N protein. In all cell types tested the protein localized exclusively in the cytoplasm (Fig. 5, S3, S4, S5 Files) . To test whether the observed lack of nuclear localization of NL63-N does not result from insufficient nuclear staining, 293T cells were also transfected with pmaxFP-Green-N/NL63-N plasmid. LLC-MK2 cells were transfected with maxFP-Green/NL63-N encoding mRNA due to poor transfection efficiency using conventional DNA delivery methods. MaxFP-Green protein was used as a control. We then examined the subcellular localization of NL63-N using confocal microscopy ( Fig. 6A and 6B ). maxFP-Green-labeled NL63-N localized exclusively to the cytoplasm in all tested cell types and no staining was observed in the nucleus. Overexpression of the N protein resulted in formation of large deposits of the N protein in the cytoplasm, which may be attributed to vast overexpression of the protein, as infection of these cells did not result in formation of such structures. Similar patterns were previously seen for other coronaviruses [62] . Previous reports show that expression of the coronaviral N protein results in delayed cellular growth. These observations were supported by biochemical studies showing that the N protein may actively participate in cell cycle regulation in infected cells by localizing to the nucleus and interacting with cyclins [36] [37] [38] 61] . To examine whether overexpression of the N protein affects cell cycle progression, we transfected LLC-MK2 and 293T cells with RNA encoding NL63-N. In this particular case, we used RNA transfection because it was more efficient than transfecting cells with DNA vectors (unpublished observations). Only cells expressing NL63-N (as identified by staining with specific antibodies) were used for subsequent analyses. There was no significant difference in cell cycle progression in NL63-N-expressing and non-expressing cells (Fig. 7) . Since its discovery, HCoV-NL63 was considered to be the closest relative of HCoV-229E. However, later identification of novel bat coronavirus species revealed several alphacoronaviruses closely clustering with the two human pathogens. Such an observation raised speculations on zoonotic origin of HCoV-NL63 [63, 64] . Considering that HCoV-NL63 is rather an atypical alphacoronavirus (e.g., in terms of receptor usage and the predicted protease active site [26, 29] , it has been suggested that, despite the high similarity between HCoV-229E and HCoV-NL63 [65] , these viruses may represent two distinct species that evolved from a common ancestor in bats, and were then introduced into human population via two independent zoonotic transmission events [63, 64] . The NL63-N is a basic protein comprising of 377 aa. Both the NTD and the complete N protein are unstable, as shown by the broad, irreversible DSC curves and obtained thermal parameters values: T m below 50°C and low ΔH value for complete N protein. In general, ΔH represents the energy amount of non-covalent bonds occurring within native protein. This is HCoV-NL63 N Protein consistent with our finding that the N protein rapidly degrades in aqueous solutions (data not shown). Obtained results are consistent with previous reports [55, 66, 67] . However, the CTD was surprisingly stable; no protein aggregation was observed upon heating and thermal denaturation was fully reversible, moreover judging from ΔH(T m )/ ΔH vH ratio it undergoes cooperative transition. Thermodynamic protein stability is defined as the Gibbs energy difference between the denatured and native states but, for some practical purposes, the denaturation temperature, T m (where ΔG = 0 for two-state reversible transition) may be more useful to measure the protein stability. Irreversibility is a common feature of DSC measurements when the large, multidomain proteins are considered. Irreversibility limits the use of a standard equilibrium thermodynamic analysis. For such a process, the denaturation temperature measured in DSC experiment is significantly lower than that corresponding to ΔG = 0. In these cases, the T m (so called operational thermal stability) is kinetically controlled. The simplest model of irreversible unfolding can be described in the terms of kinetic analysis (rate equations) according to general Lumry-Erying scheme: where K is the equilibrium unfolding constant characterizing reversible unfolding step, k is the first order rate constant describing following irreversible step; N, U and D correspond to native, reversibly unfolded and irreversibly denatured monomer of the protein, respectively. Therefore, D represents a modified denatured state existing in a quasi-two-state equilibrium with the native state [68, 69] . In fact, for irreversible denaturation the enthalpy and T m become only apparent values since both parameters are dependent on the scan rate. The existence of the exotherm on the downhill part of endotherm DSC peak is the obvious evidence for the presence of association-aggregation processes-the most common reason of the protein unfolding irreversibility. Lower stability of NTD and the complete N protein in comparison to CTD could be attributed to the distinct structural flexibility of these proteins. In turn, higher flexibility allows NTD to bind the nucleic acids which has obvious biological relevance for NTD and complete N protein function. One of the major functions of the N protein is to bind viral RNA to form a nucleoprotein, which is then packed into new virions. Therefore, we next examined the ability of NL63-N to bind nucleic acids in an electrophoretic mobility shift assay. Our results confirmed that the NTD efficiently bound RNA and DNA. However, previous studies show that coronaviral N protein has a higher affinity for the viral RNA than non-viral nucleic acids and that only nucleoproteins with coronaviral genomic RNA are efficiently incorporated into new virions [51, [70] [71] [72] . Observed limited binding of the nucleic acids by the complete N protein (data not shown) hypothetically resulted from interaction between the NTD and CTD. However, rapid N-protein degradation at the conditions of the experiment did not allow us to obtain conclusive answer. This phenomena requires further studies, as we were not able to distinguish at this stage HCoV-NL63 N Protein the mechanism of NTD-CTD interaction that limits nucleic acid binding and most likely increases the binding specificity. Previous studies demonstrated that coronaviral N proteins form dimers, and to lesser extent higher order oligomers [55] . The crystal structures of the SARS-CoV, infectious bronchitis virus (IBV), and mouse hepatitis virus (MHV)-N CTDs revealed that all those domains are characterized by a similar polypeptide fold and are dimeric, strongly suggesting that the dimeric N protein is the unit that functions in vivo [56, 57] . Furthermore, Tang et al. compared the N proteins from SARS-CoV and HCoV-229E [55] demonstrating that they formed oligomers and bound to nucleic acids. Moreover, Lo et al. showed that oligomerization of the 229E-N protein is most likely also mediated by the CTD [57] . Here we show that the complete NL63-N protein and the CTD can self-associate to form dimers and higher order oligomers. The role of aforementioned N-N interaction is, however, debatable, as NL63-N strongly binds nucleic acids during formation of the ribonucleocapsid. It is therefore possible that this interaction may be important for stabilizing and shaping the ribonucleocapsid; however, it may also be important for other N-mediated processes [73] . N protein of most coronaviruses is abundantly present in the nucleus of the infected cell [58] [59] [60] [61] . Analysis of the NL63-N protein using PSORT II revealed the presence of two NLS (pat4 and pat7), both of which are homologous to those observed in other coronaviral N proteins. For example, the IBV virus carries two very similar putative NLS signals: pat4 RPKK [aa 359-362] and pat7 PKKEKKL [aa 360-366]. One may, therefore, expect that the NL63-N protein would also localize to the nucleus. Surprisingly, no sign of NL63-N localization to the nuclei of infected or transfected cells was detected, suggesting that these NLS motifs are buried within the NL63-N structure, as previously proposed for SARS-CoV [58, 74] . To validate that the solely cytoplasmic localization of NL63-N is found not only in established cell lines, we tested a fully differentiated human airway epithelium cultures, mimicking the natural environment of the human airway epithelium; identical results were obtained. It is, however, possible that the lack of nuclear staining observed after natural infection or transfection with the NL63-N-encoding plasmid may be due to poor antibody staining within the nucleus [75] . To address this problem, we used a vector encoding maxFP-Green-N/NL63-N fusion protein, but found no difference in the subcellular localization of the fusion protein and the native protein [60, 76, 77] . Also, we found that overexpression of maxFP-Green-N/NL63-N in the cells yielded a staining pattern identical to that shown by the native N protein. Presented data show that the NL63-N protein does not localize to the nucleus (or does so in a very limited fashion), similarly to the SARS-CoV N protein and differently than other coronaviral N proteins [60, 76] . Nuclear localization of the N protein may be important for several processes, including direct interference with the cell cycle. Dysregulation of the cell cycle is a common strategy used by many DNA and RNA viruses, which enables them to hijack and exploit the host cell machinery for their own benefit. Indeed, the N proteins of many coronaviruses (including SARS-CoV; although the N protein does not localize into the nucleus in this case) inhibit cell cycle progression [36] [37] [38] 61] . However, we found no difference in the proportion of cells in each phase of cell cycle for NL63-N protein-expressing and non-expressing cells. In conclusion, we demonstrated here that although NL63-N is largely similar to other coronaviral N proteins, it possesses some unique characteristics: it does not localize to the nucleus of the infected cell and its expression does not appear to affect the cell cycle progression. Supporting Information S1 File. Design of the N-and C-terminal domains of HCoV-NL63 N protein expression constructs was based on the sequence alignment with homologous N proteins together with the structural comparative analysis. Upper panel: In order to predict polypeptide sequence encompassing the N-terminal domain of N protein from HCoV-NL63 the following sequences of N proteins were aligned: Q6Q1R8 (Uniprot ID), HCoV-NL63; P69596, avian infectious bronchitis virus (IBV, strain Beaudette), sequence referring to the 2bxx crystal structure (Protein Data Bank ID); P69598, IBV (strain Beaudette US), sequence referring to the 2c86 crystal structure; P32923, IBV (strain Gray), sequence referring to the 2gec crystal structure; P03416, murine hepatitis virus (strain A59), sequence referring to the 3hd4 crystal structure; P59595, SARS CoV, sequence referring to the 2ofz crystal structure, P59595, SARS CoV, sequence referring to the 2og3 crystal structure. Sequences referring to residues 1-210 of N protein from HCoV-NL63 are presented as it is enough to reflect the full N-terminal domains of each sequence aligned. Sequences reflecting residues defined by the electron density maps of the crystal structures (2bxx, 2c86, 2gec, 3hd4, 2ofz, 2og3) are colored green. The sequence of HCoV-NL63 N protein predicted to constitute the structurally stable N-terminal domain is colored blue. To predict polypeptide sequence encompassing the C-terminal domain of N protein from HCoV-NL63 the following sequences of N proteins were aligned: Q6Q1R8, HCoV HCoV-NL63 N Protein NL63; P69596, avian infectious bronchitis virus (IBV, strain Beaudette), sequence referring to the 2ca1 crystal structure; P32923, IBV (strain Gray), sequence referring to the 2ge7 crystal structure; P32923, IBV (strain Gray), sequence referring to the 2ge8 crystal structure; P59595, human SARS coronavirus, sequence referring to the 2cjr crystal structure; P59595, SARS CoV, sequence referring to the 2jw8 solution structure. Sequences referring to residues 194-377 of N protein from HCoV-NL63 are presented as it is enough to reflect the full C-terminal domains of each sequence aligned. Sequences reflecting residues defined by the electron density maps of the N protein CTD crystal structures (2ca1, 2ge7, 2ge8, 2cjr) and the residues of 2jw8 solution structure are colored violet. The sequence of HCoV-NL63 N protein predicted to constitute the structurally stable C-terminal domain is colored red. Lower panel: Superposition of N protein NTD and CTD structures used in the comparative analysis. Multiple molecules composing the asymmetric unit of given structure are colored the same. Multialignment was performed in MUSCLE [78] . Comparative structural analysis was done in SPDBV, Coot and PyMol [47, 79] . S5 File. Subcellular localization of NL63-N protein in fully differentiated human airway epithelium cultures cells infected with HCoV-NL63. Virions were labelled with antibodies specific to the N-NL63 protein (green); DNA was stained with DAPI (blue). Analysis was carried on with confocal microscopy using Leica TCS SP5 II confocal microscope. Voxel size: 49.9 × 49.9 × 125.9 nm, step size: 130 nm, scale bar: 10 μm. Taxonomy of Viruses., King AMQ (2012) Virus taxonomy: classification and nomenclature of viruses: ninth report of the International Committee on Taxonomy of Viruses A new virus isolated from the human respiratory tract Growth in suckling-mouse brain of "IBV-like" viruses from patients with upper respiratory tract disease Cultivation of a Novel Type of Common-Cold Virus in Organ Cultures The morphology of three previously uncharacterized human respiratory viruses that grow in organ culture Identification of a novel coronavirus in patients with severe acute respiratory syndrome A novel coronavirus associated with severe acute respiratory syndrome Coronavirus as a possible cause of severe acute respiratory syndrome Identification of a new human coronavirus Identification of new human coronaviruses Characterization and complete genome sequence of a novel coronavirus, coronavirus HKU1, from patients with pneumonia Isolation of a novel coronavirus from a man with pneumonia in Saudi Arabia Detection of new viruses by VIDISCA. Virus discovery based on cDNA-amplified fragment length polymorphism Croup is associated with the novel coronavirus NL63 Epidemiological and clinical features of human coronavirus infections among different subsets of patients. Influenza Other Respir Viruses Coronavirus infections in hospitalized pediatric patients with acute respiratory tract disease Human coronavirus NL63 in children: epidemiology, disease spectrum, and genetic diversity The dominance of human coronavirus OC43 and NL63 infections in infants Human coronavirus NL-63 infection in a Brazilian patient suspected of H1N1 2009 influenza infection: description of a fatal case Fatal lower respiratory tract disease with human corona virus NL63 in an adult haematopoietic cell transplant recipient Clinical manifestations of human coronavirus NL63 infection in children in Taiwan Role of human coronavirus NL63 in hospitalized children with croup Human aminopeptidase N is a receptor for human coronavirus 229E Feline aminopeptidase N serves as a receptor for feline, canine, porcine, and human coronaviruses in serogroup I Identification of a putative cellular receptor 150 kDa polypeptide for porcine epidemic diarrhea virus in porcine enterocytes Human coronavirus NL63 employs the severe acute respiratory syndrome coronavirus receptor for cellular entry Replication-dependent downregulation of cellular angiotensin-converting enzyme 2 protein expression by human coronavirus NL63 Genome structure and transcriptional regulation of human coronavirus NL63 Molecular characterization of human coronavirus NL63 Quantification of individual subgenomic mRNA species during replication of the coronavirus transmissible gastroenteritis virus Fields virology. Philadelphia: Wolters Kluwer/Lippincott Williams & Wilkins Health Gene N proximal and distal RNA motifs regulate coronavirus nucleocapsid mRNA transcription Coronavirus nucleocapsid protein facilitates template switching and is required for efficient transcription M and N proteins of SARS coronavirus induce apoptosis in HPF cells The SARS coronavirus nucleocapsid protein induces actin reorganization and apoptosis in COS-1 cells in the absence of growth factors The nucleocapsid protein of severe acute respiratory syndrome coronavirus inhibits cell cytokinesis and proliferation by interacting with translation elongation factor 1alpha Characterisation of cyclin D1 downregulation in coronavirus infected cells The nucleocapsid protein of severe acute respiratory syndrome-coronavirus inhibits the activity of cyclin-cyclin-dependent kinase complex and blocks S phase progression in mammalian cells SARS-CoV nucleocapsid protein antagonizes IFN-beta response by targeting initial step of IFN-beta induction pathway, and its C-terminal region is critical for the antagonism PSORT: a program for detecting sorting signals in proteins and predicting their subcellular localization A knowledge base for predicting protein localization sites in eukaryotic cells A lentiviral RNAi library for human and mouse genes applied to an arrayed viral high-content screen Well-differentiated human airway epithelial cell cultures A simple method of estimating fifty per cent endpoints Infection with human coronavirus NL63 enhances streptococcal adherence to epithelial cells Practical flow cytometry Coot: model-building tools for molecular graphics Automated protein modelling-the proteome in 3D Basic local alignment search tool The nucleocapsid protein of coronavirus infectious bronchitis virus: crystal structure of its N-terminal domain and multimerization properties Ribonucleocapsid formation of severe acute respiratory syndrome coronavirus through molecular action of the N-terminal domain of N protein Isolation and morphology of the internal component of human coronavirus, strain 229E Ribonucleoprotein-like structures from coronavirus particles Ribonucleoprotein of avian infectious bronchitis virus Biochemical and immunological studies of nucleocapsid proteins of severe acute respiratory syndrome and 229E human coronaviruses X-ray structures of the N-and C-terminal domains of a coronavirus nucleocapsid protein: implications for nucleocapsid formation Oligomerization of the carboxyl terminal domain of the human coronavirus 229E nucleocapsid protein Trafficking motifs in the SARS-coronavirus nucleocapsid protein Characterization of the nuclear export signal in the coronavirus infectious bronchitis virus nucleocapsid protein Subcellular localization of the severe acute respiratory syndrome coronavirus nucleocapsid protein Localization to the nucleolus is a common feature of coronavirus nucleoproteins, and the protein may disrupt host cell division Phosphorylation of the arginine/serine dipeptide-rich motif of the severe acute respiratory syndrome coronavirus nucleocapsid protein modulates its multimerization, translation inhibitory activity and cellular localization Evidence supporting a zoonotic origin of human coronavirus strain NL63 Distant relatives of severe acute respiratory syndrome coronavirus and close relatives of human coronavirus 229E in bats Human coronaviruses 229E and NL63: close yet still so far High affinity interaction between nucleocapsid protein and leader/intergenic sequence of mouse hepatitis virus RNA Low stability of nucleocapsid protein in SARS virus Differential scanning calorimetry of proteins A differential scanning calorimetry study of tetracycline repressor Specific interaction between coronavirus leader RNA and nucleocapsid protein Mass spectroscopic characterization of the coronavirus infectious bronchitis virus nucleoprotein and elucidation of the role of phosphorylation in RNA binding by using surface plasmon resonance Coronavirus nucleocapsid protein is an RNA chaperone Characterization of N protein self-association in coronavirus ribonucleoprotein complexes Nuclear/nucleolar localization properties of C-terminal nucleocapsid protein of SARS coronavirus A higher concentration of an antigen within the nucleolus may prevent its proper recognition by specific antibodies Intracellular localization of the severe acute respiratory syndrome coronavirus nucleocapsid protein: absence of nucleolar accumulation during infection and after expression as a recombinant protein in vero cells The coronavirus infectious bronchitis virus nucleoprotein localizes to the nucleolus MUSCLE: multiple sequence alignment with high accuracy and high throughput SWISS-MODEL and the Swiss-PdbViewer: an environment for comparative protein modeling Coronavirus N protein N-terminal domain (NTD) specifically binds the transcriptional regulatory sequence (TRS) and melts TRS-cTRS RNA duplexes Structure of the SARS coronavirus nucleocapsid protein RNA-binding dimerization domain suggests a mechanism for helical packaging of viral RNA Solution structure of the c-terminal dimerization domain of SARS coronavirus nucleocapsid protein solved by the SAIL-NMR method