key: cord-0909534-n6egr68s authors: Li, Mengxia; Ye, Gang; Si, Yu; Shen, Zhou; Liu, Zhu; Shi, Yuejun; Xiao, Shaobo; Fu, Zhen F.; Peng, Guiqing title: Structure of the multiple functional domains from coronavirus nonstructural protein 3 date: 2021-01-17 journal: Emerging microbes & infections DOI: 10.1080/22221751.2020.1865840 sha: 43d25a73b82dda267162a06ddf6a73c272083231 doc_id: 909534 cord_uid: n6egr68s Coronaviruses (CoVs) are potential pandemic pathogens that can infect a variety of hosts and cause respiratory, enteric, hepatic and neurological diseases. Nonstructural protein 3 (nsp3), an essential component of the replication/transcription complex, is one of the most important antiviral targets. Here, we report the first crystal structure of multiple functional domains from porcine delta-coronavirus (PDCoV) nsp3, including the macro domain (Macro), ubiquitin-like domain 2 (Ubl2) and papain-like protease (PLpro) catalytic domain. In the asymmetric unit, two of the subunits form the head-to-tail homodimer with an interaction interface between Macro and PLpro. However, PDCoV Macro-Ubl2-PLpro mainly exists as a monomer in solution. Then, we conducted fluorescent resonance energy transfer-based protease assays and found that PDCoV PLpro can cleave a peptide by mimicking the cognate nsp2/nsp3 cleavage site in peptide substrates and exhibits deubiquitinating and de-interferon stimulated gene(deISGylating) activities by hydrolysing ubiquitin-7-amino-4-methylcoumarin (Ub-AMC) and ISG15-AMC substrates. Moreover, the deletion of Macro or Macro-Ubl2 decreased the enzyme activity of PLpro, indicating that Macro and Ubl2 play important roles in maintaining the stability of the PLpro domain. Two active sites of PLpro, Cys260 and His398, were determined; unexpectedly, the conserved site Asp412 was not the third active site. Furthermore, the motif “NGYDT” (amino acids 409–413) was important for stabilizing the enzyme activity of PLpro, and the N409A mutant significantly decreased the enzyme activity of PLpro. These results provide novel insights into the replication mechanism of CoV and new clues for future drug design. Coronaviruses (CoVs) are enveloped, positive-sense single-stranded RNA (+ssRNA) viruses belonging to the family Coronaviridae of the order Nidovirales and have the largest genomes (26-32 kb) among known RNA viruses [1] . CoVs are divided into four genera: alpha-, beta-, gammaand delta-coronavirus (α-CoV, β-CoV, γ-CoV, and δ-CoV, respectively). CoVs can infect many species [2, 3] ; CoVs that infect humans are mainly from α-CoV and β-CoV. Human CoV 229E (HCoV-229E) and human CoV NL63 (HCoV-NL63) belong to α-CoV, severe acute respiratory syndrome CoV (SARS-CoV), Middle East respiratory syndrome (MERS-CoV) and emerging severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) belong to β-CoV [4] . CoVs infecting animals are found in four genera, including feline infectious peritonitis virus (FIPV), transmissible gastroenteritis virus (TGEV), porcine epidemic diarrhea coronavirus (PEDV), murine hepatitis virus (MHV), infectious bronchitis virus (IBV) and porcine delta-coronavirus (PDCoV) [2, 5] . Although CoVs cause severe threats to the health of humans and animals, effective commercially available vaccines or drugs are not available to treat CoVs except for the commercial vaccine against FIPV [6] . The 5'-terminal two-thirds of the CoV genome include two open reading frames (ORF1a and ORF1b) that encode two viral replicase polyproteins (ppla and pplab) [7, 8] . The two polyproteins are hydrolysed into 10-16 nonstructural proteins (nsps) by papain-like protease (PLpro of nsp3) and 3C-like protease (3CLpro or nsp5) and form mature functional nsps through posttranslational modification [9] . Then, the mature nsps together form the replication/transcription complex (RTC), which participates in the formation of double-membrane vesicles (DMVs) on the endoplasmic reticulum (ER) [10] . Nsp3 is essential for the formation of the RTC, and it functions as a scaffold protein by interacting with itself and other proteins (including viral nsps and host proteins) [5, 10] . As the largest multidomain protein in CoVs, although some domains are absent or repeated among different CoV genera, nsp3always contains the following conserved domains: ubiquitin-like domain 1 (Ubl1), hypervariable region (HVR), macrodomain (Macro, also named the ADRP or X domain), ubiquitin-like domain 2 (Ubl2), papain-like protease 2 (PLP2), ectodomain, Y1 and CoV-Y domains and two transmembrane regions (TMs). PDCoV contains 15 nsps, but the structures of most PDCoV nsps have not been reported so far except for the structure of nsp9 [11] . Macrodomains are ancient, conserved domains that occur widely across various organisms, such as bacteria, archaea and eukaryotes [12] , and even in hepatitis E virus (HEV) and CoVs [13, 14] . Macro is named after the nonhistone domain of the histone macroH2A. Affinity for adenosine diphosphate ribose (ADPr) or poly-ADP-ribose is a well-known characteristic of Macro, and in some cases, it also shows ADP-ribose 1 ′′ -monophosphatase (ADRP) activity [15] ; however, the biological significance of this enzyme activity remains unclear [16] [17] [18] [19] . According to previous reports, CoVs Macros may mediate resistance to the antiviral interferon response, and the SARS-CoV Macro exhibits the ability to inhibit the expression of innate immunity-related genes [3, 18] . These possible biological functions are consistent with the functions of SARS-CoV PLpro and HCoV-229E PLP2, revealing that synergy may exist between the two domains. Excitingly, a recent report indicates that the interplay between Macro and PLpro may affect viral replication and pathogenesis [20] . Several unliganded or liganded structures of CoV Macro have been reported to date, including SARS-CoV, SARS-CoV-2, MERS-CoV, HCoV-229E, IBV and FCoV Macros [16, [21] [22] [23] . Based on these studies, Macro can be a drug target, and the different domains of nps3 may exhibit cooperative interactions. PLpro is responsible for the hydrolytic release of nsp1/2 -nsp4; in addition to hydrolytic activity, PLpro also possesses deubiquitinating and deISGylating activities [24] [25] [26] [27] . Lys48-, K63-and linear ubiquitin are three types of polyubiquitination that are associated with immune signalling pathways, and PLPs remove the K48-Ub and K63-Ub from target proteins by recognizing the LRGG motif [28] [29] [30] . Interferon-stimulated gene 15 protein is a ubiquitin-like molecule that conjugates to the target protein, such as RIG-I, JAK1, STAT1, PKR and MxA. PLPs have the ability to remove conjugated ISG15 cellular proteins [30] [31] [32] . Thus, the deubiquitinating and deISGylating activities of PLPS are considered two pathways by which CoVs escape the immune response. PLPs block the IFN regulatory factor 3 (IRF3) induced IFN-β and TNFα-mediated NF-κB activation [33] . As previous reports, PLPs contain the classic Cys-His-Asp triad. Interestingly, HCoV-229E PLP1 exhibits a Cys-His dyad [34] , while the homologue PLpro of equine arteritis virus (EAV) comprises a Cys-His-Asn triad [35] . Ubl2 exists in CoVs and among host ubiquitin-specific proteases (USPs), which regulate the catalytic activity of proteases [36, 37] . At present, the structure and function of PDCoV PLpro have not been reported. Here, we first report the structure of PDCoV Macro-Ubl2-PLpro from δ-CoV with 2.5 Å resolution. The structure contains several important structural features, including the arrangement of domains from Macro to PLpro in nsp3, the interaction between Macro and PLpro in the crystal homodimer, and the catalytic core of PDCoV PLpro. We performed in vitro assays using substrates representing viral and cellular targets to characterize the enzyme activities of the purified protease. Our results show that PDCoV PLpro has hydrolytic activity, deubiquitinating and deISGylating activities, and Ubl2 plays important roles in stabilizing the PLpro enzyme activities. The sequence encoding the PDCoV Macro-Ubl2-PLpro domain (residues 939-1384 of the polyprotein ppla from PDCoV strain CHN-HB-2014, GenBank accession number KP757891.1) was amplified by PCR from the parental plasmid pCAGGS-HA-nsp3, which was kindly provided by Professor Xiao and cloned into the pET-42b vector. Macro-Ubl2-PLpro mutants (C260A, H398A, D412A, NGYDT409-413AAAAA, N409A, G410A, Y411A, and T413A) and truncations (Ubl2-PLpro, PLpro) were also cloned into pET-42b with C-terminal His 6 tags. All of the recombinant plasmids were sequenced. The recombinant plasmids were transformed into E. coli BL21 (DE3) cells for expression. Cultures were grown in LB medium at 37°C until the optical density at 600 nm (OD600) reached 0.6-0.8, induced with 1 mM isopropyl-β-D-thiogalactopyranoside (IPTG), and incubated with shaking overnight at 27°C . To solve the phase problem, selenomethionine (Se-Met)-labelled PDCoV Macro-Ubl2-PLpro was expressed in E. coli BL21 (DE3) cells according the instructions of the M9 Se-Met High-Yield Growth Media Kit (Medicion, China), and cultured with 50 µg/mL kanamycin, M9 salt medium, 15 mineral supplements, vitamins (thiamine, vitamin B12), 0.5% glycerol at 37°C until reaching the OD600 of 1.2. Then, the amino acid mixture was added (lysine, phenylalanine, threonine, isoleucine, leucine, valine, and selenomethionine), 15 min later, IPTG was added at a final concentration of 1 mM and cells grown at 37°C for 5 h. For protein purification, cells were pelleted by centrifugation at 8,000 rpm (5 min at 4°C) from 1 L of culture, resuspended in 50 mL of phosphate-buffered saline (PBS) and lysed using an ultrahigh-pressure cell disrupter (ATS Engineering Inc.). After centrifugation at 8,000 rpm for 30 min, the supernatant was filtered through a 0.45-µm-pore-size filter and loaded onto a His Trap HP column (GE Healthcare). The target protein was eluted with a linear gradient between the binding buffer (20 mM Tris-HCl [pH 7.4] and 500 mM NaCl) and elution buffer (20 mM Tris-HCl [pH 7.4], 500 mM NaCl and 500 mM imidazole). The protein was further purified using a Superdex200 gel filtration column (GE Healthcare) equilibrated with buffer (20 mM Tris-HCl [pH 7.4] and 200 mM NaCl). For crystallization, the purified protein was concentrated to 12 mg/mL using a 30-kDa-molecularmass-cutoff centrifuge concentrator, and the concentration was determined by measuring the absorbance at 280 nm with a Nano Drop spectrophotometer (Thermo Scientific). The protein sample was flash-frozen with liquid nitrogen and stored at −80°C. Notably, to avoid unexpected degradation, all of the purification procedures should be performed at 4°C. Crystallization screening for native Macro-Ubl2-PLpro at a concentration of 12 mg/mL was performed at 20°C using hanging drop vapour diffusion with 96well plates. Crystals were obtained from a solution containing 0.1 M imidazole and 12% polyethylene glycol (PEG) 20,000. Further optimization of the crystallization conditions was performed with 24-well plates through vapour diffusion in sitting drops consisting of a 1 µL drop of 12 mg/mL protein, 0.1 M imidazole, and 10-16% PEG 20,000. The crystals grew overnight and were cryoprotected by the addition of 20% ethylene glycol and flash-cooled in liquid nitrogen. The selenomethionine derivative Macro-Ubl2-PLpro was crystallized under similar conditions. Single-wavelength X-ray diffraction data were collected from single crystals at the BL17U1 beam line (wave-length=0.97910 Å, temperature=100 K) of the Shanghai Synchrotron Radiation Facility (SSRF). All data were processed with HKL-3000 software [38] , and the resulting statistics are listed in Table 1 . The initial structure was solved using the single-wavelength anomalous dispersion (SAD) method from the Se-Met derivative and molecular replacement with PHA-SER [39] . All five potential selenium atoms in the Macro-Ubl2-PLpro monomer were located, and the initial phases were calculated using the program Auto-Sol from the PHENIX software suite [40] . Manual model rebuilding was performed using Coot [41] . Refinement was carried out using the program PHE-NIX. Structural figures were drawn using the program PyMOL [42] . The amino acid sequences of CoV Macro and Ubl2-PLpro were aligned using ClustalW2 [43] and visualized with the ESPript 3 server (http:// espript.ibcp.fr) [44] . Small angle X-ray scattering analysis SAXS data were collected at the BL19U2 beamline of the Shanghai Synchrotron Radiation Facility (SSRF) at room temperature. For the SAXS measurement, 25 µM PDCoV Macro-Ubl2-PLpro was prepared in buffer (20 mM Tris-HCl [pH 7.4] and 200 mM NaCl). For each measurement, 20 consecutive frames with 1-sec exposure were recorded and averaged, with difference between the first and the last frames. The background scattering was recorded for the matching buffer and was subtracted from the protein scattering data. The data were visualized and analysed using the software package ATSAS [45] . The theoretical SAXS where F o and F c are the observed and calculated structure factors, respectively. R free is equivalent to R work , but 5% of the measured reflections have been excluded from the refinement and set aside for cross-validation. profiles of monomer and dimer crystal structures were calculated with the software CRYSOL. Assays to determine the peptide cleavage of a fluorescent resonance energy transfer (FRET) substrate, phenyl]azo benzoic acid; EDANS, 5-[(2-aminoethyl) amino] naphthalene-1-sulfonic acid) (GenScript, Nanjing, China), were performed using different substrate concentrations (10-50 µM) and 2 µM purified PDCoV Macro-Ubl2-PLpro. Assays were performed in 20 mM Tris-HCl, pH 7.5, 200 mM NaCl at 30°C with a 96-well microplate using the multimode reader platform (Tecan). The rate of substrate hydrolysis was determined by monitoring the fluorescence as a function of time (excitation λ, 336 nm; emission λ 490 nm) and calculated from the linear part of the curves. Since no saturation was observed in the plot of initial velocities versus substrate concentrations, data points were fit to the equation v/[E] Total =k app [S] to determine the pseudo-first-order rate constant k app assuming that [S]<