设万维读者为首页 万维读者网 -- 全球华人的精神家园 广告服务 联系我们 关于万维
 
首  页 新  闻 视  频 博  客 论  坛 分类广告 购  物
搜索>> 发表日志 控制面板 个人相册 给我留言
帮助 退出
 
Pascal的博客  
“朝廷不是让我隐蔽吗?”“你也不看看,这是什么时候了?!”  
网络日志正文
微信群獐头鼠目微民疯传正丽造病毒留言 2020-02-05 12:29:51

image.png


Inline image

Inline image


Inline image


image.png


image.png


image.png


bats are an important source of highly lethal zoonotic viruses, such as Hendra, Nipah, Ebola and Marburg viruses16.

Here we report on a series of fatal swine disease outbreaks in Guangdong province, China, approximately 100 km from the location of the purported index case of SARS. Most strikingly, we found that the causative agent of this swine acute diarrhoea syndrome (SADS) is a novel HKU2-related coronavirus that is 98.48% identical in genome sequence to a bat coronavirus, which we detected in 2016 in bats in a cave in the vicinity of the index pig farm. This new virus (SADS-CoV) originated from the same genus of horseshoe bats (Rhinolophus) as SARS-CoV.

From 28 October 2016 onwards, a fatal swine disease outbreak was observed in a pig farm in Qingyuan, Guangdong province, China, very close to the location of the first known index case of SARS in 2002, who lived in Foshan (Extended Data Fig. 1a). Porcine epidemic diarrhoea virus (PEDV, a coronavirus) had caused prior outbreaks at this farm, and was detected in the intestines of deceased piglets at the start of the outbreak. However, PEDV could no longer be detected in deceased piglets after 12 January 2017, despite accelerating mortality (Fig. 1a), and extensive testing for other common swine viruses yielded no results (Extended Data Table 1). These findings suggested that this was an outbreak of a novel disease. Clinical signs are similar to those caused by other known swine enteric coronaviruses1718 and include severe and acute diarrhoea and acute vomiting, leading to death due to rapid weight loss in newborn piglets that are less than five days of age. Infected piglets died 2–6 days after disease onset, whereas infected sows suffered only mild diarrhoea and most sows recovered within two days. The disease caused no signs of febrile illness in piglets or sows. The mortality rate was as high as 90% in piglets that were five days or younger, whereas in piglets that were older than eight days, the mortality dropped to 5%. Subsequently, SADS-related outbreaks were found in three additional pig farms within 20–150 km of the index farm (Extended Data Fig. 1a) and, by 2 May 2017, the disease had caused the death of 24,693 piglets at these four farms (Fig. 1a). In farm A alone, 64% (4,659 out of 7,268) of all piglets that were born in February died. The outbreak has abated, and measures that were taken to control SADS included separation of sick sows and piglets from the rest of the herd. A qPCR test described below was used as the main diagnostic tool to confirm SADS-CoV infection.

Fig. 1: Detection of SADS-CoV infection in pigs in Guangdong, China.

figure1

a, Records of daily death toll on the four farms from 28 October 2016 to 2 May 2017. b, Detection of SADS-CoV by qPCR. The y axis shows the log(copy number per 106 copies of 18S rRNA). n = 12 sick piglets, 5 sick sows, 16 recovered sows and 10 healthy piglets. c, Tissue distribution of SADS-CoV in diseased pigs. n = 3. Data are mean ± s.d.; dots represent individual values. d, Detection of SADS-CoV antibodies. n = 46 sows from whom serum was first taken in the first three weeks of the outbreak (First bleed), n = 8 sows from whom serum was taken again (Second bleed) at more than one month after the onset of the outbreak, n = 8 sera from healthy pig controls, n = 35 human sera from pig farmers.

Source data

Full size image

A sample collected from the small intestine of a diseased piglet was analysed by metagenomics analysis using next-generation sequencing (NGS) to identify potential aetiological agents. Of the 15,256,565 total reads obtained, 4,225 matched sequences of the bat CoV HKU2, which was first detected in Chinese horseshoe bats in Hong Kong and Guangdong province, China19. By de novo assembly and targeted PCR, we obtained a 27,173-bp CoV genome that shared 95% sequence identity to HKU2-CoV (GenBank accession number NC_009988). Thirty-three full genome sequences of SADS-CoV were subsequently obtained (8 from farm A, 5 from farm B, 11 from farm C and 9 from farm D) that were 99.9% identical to each other (Supplementary Table 1).

Using qPCR targeting the nucleocapsid gene (Supplementary Table 2), we detected SADS-CoV in acutely sick piglets and sows, but not in recovered or healthy pigs on the four farms, nor in nearby farms that showed no evidence of SADS. The virus replicated to higher titres in piglets than in sows (Fig. 1b). SADS-CoV displayed tissue tropism of the small intestine (Fig. 1c), as observed for other swine enteric coronaviruses20. Retrospective PCR analysis revealed that SADS-CoV was present on farm A during the PEDV epidemic, where the first strongly positive SADS-CoV sample was detected on 6 December 2016. From mid-January onwards, SADS-CoV was the dominant viral agent detected in diseased animals (Extended Data Fig. 1b). It is possible that the presence of PEDV early in the SADS-CoV outbreak may have somehow facilitated or enhanced spillover and amplification of SADS. However the fact that the vast majority of piglet mortality occurred after PEDV infection had become undetectable suggests that SADS-CoV itself causes a lethal infection in pigs that was responsible for these large-scale outbreaks, and that PEDV does not directly contribute to its severity in individual pigs. This was supported by the absence of PEDV and other known swine diarrhoea viruses during the peak and later phases of the SADS outbreaks in the four farms (Extended Data Table 1).

We rapidly developed an antibody assay based on the S1 domain of the spike (S) protein using a luciferase immunoprecipitation system21. Because SADS occurs acutely and has a rapid onset in piglets, serological investigation was conducted only in sows. Among 46 recovered sows tested, 12 were seropositive for SADS-CoV within three weeks of infection (Fig. 1d). To investigate possible zoonotic transmission, serum samples from 35 farm workers who had close contact with sick pigs were also analysed using the same luciferase immunoprecipitation system approach and none were positive for SADS-CoV.

Although the overall genome identity of SADS-CoV and HKU2-CoV is 95%, the S gene sequence identity is only 86%, suggesting that the previously reported HKU2-CoV is not the direct progenitor of SADS-CoV, but that they may have originated from a common ancestor. To test this hypothesis, we developed a SADS-CoV-specific qPCR assay based on its RNA-dependent RNA polymerase (RdRp) gene (Supplementary Table 2) and screened 591 bat anal swabs collected between 2013 and 2016 from seven different locations in Guangdong province (Extended Data Fig. 1a). A total of 58 samples (9.8%) tested positive (Extended Data Table 2), all were from Rhinolophus spp. bats that are also the natural reservoir hosts of SARS-related coronaviruses3,4,5,6,7,8,9,10. Four complete genome sequences with the highest RdRp PCR-fragment sequence identity to that of SADS-CoV were determined by NGS. They are very similar in size (27.2 kb) compared to SADS-CoV (Fig. 2a) and we tentatively call them SADS-related coronaviruses (SADSr-CoV). Overall sequence identity between SADSr-CoV and SADS-CoV ranges from 96 to 98%. Most importantly, the S protein of SADS-CoV shared more than 98% sequence identity with sequences of two of the SADSr-CoVs (samples 162149 and 141388), compared to 86% with HKU2-CoV. The major sequence differences among the four SADSr-CoV genomes were found in the predicted coding regions of the S and NS7a and NS7b genes (Fig. 2a). In addition, the coding region of the S protein N-terminal (S1) domain was determined from 19 bat SADSr-CoVs to enable more detailed phylogenetic analysis.

Fig. 2: Genome and phylogenetic analysis of SADS-CoV and SADSr-CoV.

figure2

a, Genome organization and comparison. Colour-coding for different genomic regions as follows. Green, non-structural polyproteins ORF1a and ORF1b; yellow, structural proteins SEM and N; blue, accessory proteins NS3aNS7a and NS7b; Orange, untranslated regions. The level of sequence identity of SADSr-CoV to SADS-CoV is illustrated by different patterns of boxes: Solid colour, highly similar; Dotted fill, moderately similar; Dashed fill, least similar. b, Phylogenetic analysis of 57 S1 sequences (33 from SADS-CoV and 24 from SADSr-CoV). Different colours represent different host species as shown on the left. Scale bar, nucleotide substitutions per site.

Full size image

The phylogeny of S1 and the full-length genome revealed a high genetic diversity of alphacoronaviruses among bats and strong coevolutionary relationships with their hosts (Fig. 2b and Extended Data Fig. 2), and showed that SADS-CoVs were more closely related to SADSr-CoVs from Rhinolophus affinis than from Rhinolophus sinicus, in which HKU2-CoV was found. Both phylogenetic and haplotype network analyses demonstrated that the viruses from the four farms probably originated from their reservoir hosts independently (Extended Data Fig. 3), and that a few viruses might have undergone further genetic recombination (Extended Data Fig. 4). However, molecular clock analysis of the 33 SADS-CoV genome sequences failed to establish a positive association between sequence divergence and sampling date. Therefore, we speculate that either the virus was introduced into pigs from bats multiple times, or that the virus was introduced into pigs once, but subsequent genetic recombination disturbed the molecular clock.

For viral isolation, we tried to culture the virus in a variety of cell lines (see Methods for details) using intestinal tissue homogenates as starting material. Cytopathogenic effects were observed in Vero cells only after five passages (Extended Data Fig. 5a, b). The identity of SADS-CoV was verified in Vero cells by immunofluorescence microscopy (Extended Data Fig. 5c, d) and by whole-genome sequencing (GenBank accession number MG557844). Similar results were obtained by other groups2223.

Known coronavirus host cell receptors include angiotensin-converting enzyme 2 (ACE2) for SARS-related CoV, aminopeptidase N (APN) for certain alphacoronaviruses, such as human (H)CoV-229E, and dipeptidyl peptidase 4 (DPP4) for Middle East respiratory syndrome (MERS)-CoV24,25,26. To investigate the receptor usage of SADS-CoV, we tested live or pseudotyped SADS-CoV infection on HeLa cells that expressed each of the three molecules. Whereas the positive control worked for SARS-related CoV and MERS-CoV pseudoviruses, we found no evidence of enhanced infection or entry for SADS-CoV, suggesting that none of these receptors functions as a receptor for virus entry for SADS-CoV (Extended Data Table 3).

To fulfill Koch’s postulates for SADS-CoV, two different types of animal challenge experiments were conducted (see Methods for details). The first challenge experiment was conducted with specific pathogen-free piglets that were infected with a tissue homogenate of SADS-CoV-positive intestines. Two days after infection, 3 out of 7 animals died in the challenge group whereas 4 out of 5 survived in the control group. Incidentally, the one piglet that died in the control group was the only individual that did not receive colostrum due to a shortage in the supply. It is thus highly likely that lack of nursing and inability to access colostrum was responsible for the death (Extended Data Table 4). For the second challenge, healthy piglets were acquired from a farm in Guangdong that had been free of diarrheal disease for a number of weeks before the experiment, and were infected with the cultured isolate of SADS-CoV or tissue-culture medium as control. Of those inoculated with SADS-CoV, 50% (3 out of 6) died between 2 and 4 days after infection, whereas all control animals survived (Extended Data Table 5). All animals in the infected group suffered watery diarrhoea, rapid weight loss and intestinal lesions (determined after euthanasia upon experiment termination, Extended Data Tables 45). Histopathological examination revealed marked villus atrophy in SADS-CoV inoculated farm piglets four days after inoculation but not in control piglets (Fig. 3a, b) and viral N protein-specific staining was observed mainly in small intestine epithelial cells of the inoculated piglets (Fig. 3c, d).

Fig. 3: Immunohistopathology of SADS-CoV infected tissues.

figure3

ad, Sections of jejunum tissue from control (ac) and infected (bd) farm piglets four days after inoculation were stained with haematoxylin and eosin (ab) or rabbit anti-SADSr-CoV N serum (red), DAPI (blue) and mouse antibodies against epithelial cell markers cytokeratin 8, 18 and 19 (green) in (cd). SADS-CoV N protein is evident in epithelial cells and deeper in the tissue of infected piglets, which exhibit villus shortening. Scale bars, 200 μm (ab) and 50 μm (cd). The experiment was conducted three times independently with similar results.

Full size image

The current study highlights the value of proactive viral discovery in wildlife, and targeted surveillance in response to an emerging infectious disease event, as well as the disproportionate importance of bats as reservoirs of viruses that threaten veterinary and public health1. It also demonstrates that by using modern technological platforms, such as NGS, luciferase immunoprecipitation system serology and phylogenetic analysis, key experiments that traditionally rely on the isolation of live virus can be performed rapidly before virus isolation.


Methods

Sample collection

Bats were captured and sampled in their natural habitat in Guangdong province (Extended Data Fig. 1) as described previously4. Faecal swab samples were collected in viral transport medium (VTM) composed of Hank’s balanced salt solution at pH 7.4 containing BSA (1%), amphotericin (15 μg ml−1), penicillin G (100 units ml−1) and streptomycin (50 μg ml−1). Stool samples from sick pigs were collected in VTM. When appropriate and feasible, intestinal samples were also taken from deceased animals. Samples were aliquoted and stored at –80 °C until use. Blood samples were collected from recovered sows and workers on the farms who had close contact with sick pigs. Serum was separated by centrifugation at 3,000g for 15 min within 24 h of collection and preserved at 4 °C. Human serum collection was approved by the Medical Ethics Committee of the Wuhan School of Public Health, Wuhan University and Hummingbird IRB. Human, pigs and bats were sampled without gender or age preference unless indicated (for example, piglets or sows). No statistical methods were used to predetermine sample size.

Virus isolation

The following cells were used for virus isolation in this study: Vero (cultured in DMEM and 10% FBS); Rhinolophus sinicus primary or immortalized cells generated in our laboratory (all cultured in DMEM/F12 and 15% FBS): kidney primary cells (RsKi9409), lung primary cells (RsLu4323), lung immortalized cells (RsLuT), brain immortalized cells (RsBrT) and heart immortalized cells (RsHeT); and swine cell lines: two intestinal porcine enterocytes cell lines, IPEC (RPMI1640 and 10% FBS) and SIEC (DMEM and 10% FBS), three kidney cell lines PK15, LLC-PK1 (DMEM and 10% FBS for both) and IBRS (MEM and 10% FBS), and one pig testes cell line, ST (DMEM and 10% FBS). All cell lines were tested free of mycoplasma contamination, species were confirmed and authenticated by microscopic morphologic evaluation. None of the cell lines was on the list of commonly misidentified cell lines (by the ICLAC).

Cultured cell monolayers were maintained in their respective medium. PCR-positive pig faecal samples or the supernatant from homogenized pig intestine (in 200 μl VTM) were spun at 8,000g for 15 min, filtered and diluted 1:2 with DMEM supplemented with 16 μg ml−1 trypsin before addition to the cells. After incubation at 37 °C for 1 h, the inoculum was removed and replaced with fresh culture medium containing antibiotics (below) and 16 μg ml−1 trypsin. The cells were incubated at 37 °C and observed daily for cytopathic effect (CPE). Four blind passages (three-day interval between every passage) were performed for each sample. After each passage, both the culture supernatant and cell pellet were examined for the presence of virus by RT–PCR using the SADS-CoV primers listed in Supplementary Table 2. Penicillin (100 units ml−1) and streptomycin (15 μg ml−1) were included in all tissue culture media.

RNA extraction, S1 gene amplification and qPCR

Whenever commercial kits were used, the manufacturer’s instructions were followed without modification. RNA was extracted from 200 μl of swab samples (bat), faeces or homogenized intestine (pig) with the High Pure Viral RNA Kit (Roche). RNA was eluted in 50 μl of elution buffer and used as the template for RT–PCR. Reverse transcription was performed using the SuperScript III kit (Thermo Fisher Scientific).

To amplify S1 genes from bat samples, nested PCR was performed with primers designed based on HKU2-CoV (GenBank accession number NC_009988.1)19 (Supplementary Table 2). The 25-μl first-round PCR mixture contained 2.5 μl 10× PCR reaction buffer, 5 pmol of each primer, 50 mM MgCl2, 0.5 mM dNTP, 0.1 μl Platinum Taq Enzyme (Thermo Fisher Scientific) and 1 μl cDNA. The 50-μl second-round PCR mixture was identical to the first-round PCR mixture except for the primers. Amplification of both rounds was performed as follows: 94 °C for 5 min followed by 60 cycles at 94 °C for 30 s, 50 °C for 40 s, 72 °C for 2.5 min, and a final extension at 72 °C for 10 min. PCR products were gel-purified and sequenced.

For qPCR analysis, primers based on SADS-CoV RdRp and N genes were used (Supplementary Table 2). RNA extracted from above was reverse-transcribed using PrimeScript RT Master Mix (Takara). The 10 μl qPCR reaction mix contained 5 μl 2× SYBR premix Ex TaqII (Takara), 0.4 μM of each primer and 1 μl cDNA. Amplification was performed as follows: 95 °C for 30 s followed by 40 cycles at 95 °C for 5 s, 60 °C for 30 s, and a melting curve step.

Luciferase immunoprecipitation system assay

The SADS-CoV S1 gene was codon-optimized for eukaryotic expression, synthesized (GenScript) and cloned in frame with the Renilla luciferase gene (Rluc) and a Flag tag in the pREN2 vector21. pREN2-S1 plasmids were transfected into Cos-1 cells using Lipofectamine 2000 (Thermo Fisher Scientific). At 48 h post-transfection, cells were collected, lysed and a luciferase assay was performed to determine Rluc expression for both the empty vector (pREN2) and the pREN2-S1 construct. For testing of unknown pig or human serum samples, 1 μl of serum was incubated with 10 million units of Rluc alone (vector) or Rluc-S1, respectively, together with 3.5 μl of a 30% protein A/G UltraLink resin suspension (Pierce, Thermo Fisher Scientific). After extensive washing to remove unbounded luciferase-tagged antigens, the captured luciferase amount was determined using the commercial luciferase substrate kit (Promega). The ratio of Rluc-S1:Rluc (vector) was used to determine the specific S1 reactivity of pig and human sera. Commercial Flag antibody (Thermo Fisher Scientific) was used as the positive control, and various pig sera (from uninfected animals in China or Singapore; or pigs infected with PEDV, TGEV or Nipah virus) were used as a negative control.

Protein expression and antibody production

The N gene from SADSr-CoV 3755 (GenBank accession number MF094702), which shares a 98% amino acid sequence identity to the SADS-CoV N protein, was inserted into pET-28a+ (Novagen) for prokaryotic expression. Transformed Escherichia coli were grown at 37 °C for 12–18 h in medium containing 1 mM IPTG. Bacteria were collected by centrifugation and resuspended in 30 ml of 5 mM imidazole and lysed by sonication. The lysate, from which N protein expression was confirmed with an anti-His-tag antibody, was applied to Ni2+ resin (Thermo Fisher Scientific). The purified N protein, at a concentration of 400 μg ml−1, was used to immunize rabbits for antibody production following published methods27. After immunization and two boosts, rabbits were euthanized and sera were collected. Rabbit anti-N protein serum was used 1:10,000 for subsequent western blots.

Amplification, cloning and expression of human and swine genes

Construction of expression clones for human ACE2 in pcDNA3.1 has been described previously528. Human DPP4 was amplified from human cell lines. Human APN (also known as ANPEP) was commercially synthesized. Swine APN (also known as ANPEP), DPP4 and ACE2 were amplified from piglet intestine. Full-length gene fragments were amplified using specific primers (provided upon request). Human ACE2 was cloned into pCDNA3.1 fused with a His tag. Human APN and DPP4, swine APNDPP4 and ACE2 were cloned into pCAGGS fused with an S tag. Purified plasmids were transfected into HeLa cells. After 24 h, expression human or swine genes in HeLa cells was confirmed by immunofluorescence assay using mouse anti-His tag or mouse anti-S tag monoclonal antibodies (produced in house) followed by Cy3-labelled goat anti-mouse/rabbit IgG (Proteintech Group).

Pseudovirus preparation

The codon-humanized S genes of SADS-CoV or MERS-CoV cloned into pcDNA3.1 were used for pseudovirus construction as described previously528. In brief, 15 μg of each pHIV-Luc plasmid (pNL4.3.Luc.R-E-Luc) and the S-protein-expressing plasmid (or empty vector control) were co-transfected into 4 × 106 HEK293T cells using Lipofectamine 3000 (Thermo Fisher Scientific). After 4 h, the medium was replaced with fresh medium. Supernatants were collected 48 h after transfection and clarified by centrifugation at 3,000g, then passed through a 0.45-μm filter (Millipore). The filtered supernatants were stored at −80 °C in aliquots until use. To evaluate the incorporation of S proteins into the core of HIV virions, pseudoviruses in supernatant (20 ml) were concentrated by ultracentrifugation through a 20% sucrose cushion (5 ml) at 80,000g for 90 min using a SW41 rotor (Beckman). Pelleted pseudoviruses were dissolved in 50 μl phosphate-buffered saline (PBS) and examined by electron microscopy.

Pseudovirus infection

HeLa cells transiently expressing APN, ACE2 or DPP4 were prepared using Lipofectamine 2000 (Thermo Fisher Scientific). Pseudoviruses prepared above were added to HeLa cells overexpressing APN, ACE2 or DPP4 24 h after transfection. The unabsorbed viruses were removed and replaced with fresh medium at 3 h after infection. The infection was monitored by measuring the luciferase activity conferred by the reporter gene carried by the pseudovirus, using the Luciferase Assay System (Promega) as follows: cells were lysed 48 h after infection, and 20 μl of the lysates was taken for determining luciferase activity after the addition of 50 μl of luciferase substrate.

Examination of known CoV receptors for SADS-CoV entry/infection

HeLa cells transiently expressing APN, ACE2 or DPP4 were prepared using Lipofectamine 2000 (Thermo Fisher Scientific) in a 96-well plate, with mock-transfected cells as controls. SADS-CoV grown in Vero cells was used to infect HeLa cells transiently expressing APN, ACE2 or DPP4. The inoculum was removed after 1 h of absorption and washed twice with PBS and supplemented with medium. SARS-related-CoV WIV167 and MERS-CoV HIV-pseudovirus were used as positive control for human/swine ACE2 or human/swine DPP4, respectively. After 24 h of infection, cells were washed with PBS and fixed with 4% formaldehyde in PBS (pH 7.4) for 20 min at room temperature. SARS-related-CoV WIV16 replication was detected using rabbit antibody against the SARS-related-CoV Rp3 N protein (made in house, 1:100) followed by Cy3-conjugated goat anti-rabbit IgG (1:50, Proteintech)7. SADS-CoV replication was monitored using rabbit antibody against the SADSr-CoV 3755 N protein (made in house, 1:50) followed by FITC-conjugated goat anti-rabbit IgG (1:50, Proteintech). Nuclei were stained with DAPI (Beyotime). Staining patterns were examined using confocal microscopy on a FV1200 microscope (Olympus). Infection of MERS-CoV HIV-pseudovirus was monitored by luciferase 48 h after infection.

High-throughput sequencing, pathogen screening and genome assembly

Tissue from the small intestine of deceased pigs was homogenized and filtered through 0.45-μm filters before nucleic acid extraction and ribosomal RNA was depleted using the NEBNext rRNA Depletion Kit (New England Biolabs). Metagenomics analysis of both RNA and DNA viruses was performed. For RNA virus screening, the sequencing library was constructed using Ion Total RNA-Seq Kit v2 (Thermo Fisher Scientific). For DNA virus screening, NEBNext Fast DNA Fragmentation & Library Prep Set for Ion Torrent (New England Biolabs) was used for library preparation. Both libraries were sequenced on an Ion S5 sequencer (Thermo Fisher Scientific). An analysis pipeline was applied to the sequencing data, which included the following analysis steps: (1) raw data quality filtering; (2) host genomic sequence filtering; (3) BLASTn search against the virus nucleotide database using BLAST; (4) BLASTx search against the virus protein database using DIAMOND v.0.9.0; (5) contig assembling and BLASTx search against the virus protein database. For whole viral genome sequencing, amplicon primers (provided upon request) were designed using the Thermo Fisher Scientific online tool with the HKU2-CoV and the SADS-CoV farm A genomes as references, and the sequencing libraries were constructed using NEBNext Ultra II DNA Library Prep Kit for Illumina and sequenced on an MiSeq sequencer. PCR and Sanger sequencing was performed to fill gaps in the genome. Genome sequences were assembled using CLC Genomic Workbench v.9.0. 5′-RACE was performed to determine the 5′-end of the genomes using SMARTer RACE 5′/3′ Kit (Takara). Genomes were annotated using Clone Manager Professional Suite 8 (Sci-Ed Software).

Phylogenetic analysis

SADS-CoV genome sequences and other representative coronavirus sequences (obtained from GenBank) were aligned using MAFFT v.7.221. Phylogenetic analyses with full-length genome, S gene and RdRp were performed using MrBayes v.3.2. Markov chain Monte Carlo was run for 20–50 million steps using the GTR+G+I model (general time reversible model of nucleotide substitution with a proportion of invariant sites and γ-distributed rates among sites). The first 10% was removed as burn-in. The association between phylogenies and phenotypes (for example, host species and farms) was assessed by BaTS beta-build2, with the trees obtained in the previous step used as input. For SADS-CoVs, a median-joining network analysis was performed using PopART v.1.7, with ɛ = 0. Phylogenetic analysis of the 33 full-length SADS-CoV genome sequences was performed using RAxML v.8.2.11, with GTRGAMMA as the nucleotide substitution model and 1,000 bootstrap replicates. The maximum likelihood tree was used to test the molecular clock using TempEst v.1.5. Potential genetic recombination events in our datasets were detected using RDP v.4.72.

Animal infection studies

Experiments were carried out strictly in accordance with the recommendations of the Guide for the Care and Use of Laboratory Animals of the National Institutes of Health. The use of animals in this study was approved by the South China Agricultural University Committee of Animal Experiments (approval number 201004152).

Two different animal challenge experiments were conducted. Pigs were used without gender preference. In the first experiment, which was conducted before the virus was isolated, we used three-day old specific pathogen-free (SPF) piglets of the same breeding line, cared for at a SPF facility, fed with colostrum (except one). These piglets were bred and reared to be free of PEDV, CSFV, SIV, PCV2 and PPV infections, and were routinely tested for viral infections using PCR. We also conducted NGS to further confirm that these were animals were free of infection of the above viruses before the animal experiment, and to demonstrate that the animals were free of SADS-CoV infection. The intestinal tissue samples from healthy and diseased animals (intestinal samples excised from euthanized piglets, then ground to make slurry for the inoculum and NGS was performed to confirm no other pig pathogens were found in the samples), were used to feed two groups of 5 (control) and 7 (infection) animals, respectively. For the second experiment, isolated SADS-CoV was used to infect healthy piglets from a farm in Guangdong, which had been free of diarrheal disease for a number of weeks. These piglets were from the same breed as those on SADS-affected farms, to eliminate potential host factor differences and to more accurately reproduce the conditions that occurred during the outbreak in the region. Both groups of piglets were cared for at a known pig disease-free facility. Again, qPCR and NGS were used to make sure that there was no other known swine diarrhoea virus present in the virus inoculum or any of the experimental animals. Two groups (6 for each group) of three-day old piglets were inoculated with SADS-CoV culture supernatant or normal cell culture medium as control. NGS and qPCR were used to confirm that there were no other known swine pathogens in the inoculum.

For both experiments, animals were recorded daily for signs of diseases, such as diarrhoea, weight loss and death. Faecal swabs were collected daily from all animals and screened for known swine diarrhoea viruses by qPCR. Weight loss was calculated as the percentage weight loss compared the original weight at day 0 with a threshold of >5%. It is important to point out that piglets when they are three days old tend to suffer from diarrhoea and weight loss when they are taken away from sows and the natural breast-feeding environment even without infection. At experimental endpoints, piglets were humanely euthanized and necropsies performed. Pictures were taken to record gross pathological changes to the intestines. Ileal, jejunal and duodenal tissues were taken from selected animals and stored at –80 °C for further analysis.

Haematoxylin and eosin and immunohistochemistry analysis

Frozen (–80 °C) small intestinal tissues including duodenum, jejunum and ileum taken from the experimentally infected pigs were pre-frozen at –20 °C for 10 min. Tissues were then embedded in optimal cutting temperature (OCT) compound and cut into 8-μm sections using the Cryotome FSE machine (Thermo Fisher Scientific). Mounted microscope slides were fixed with paraformaldehyde and stained with haematoxylin and eosin for histopathological examination.

For immunohistochemistry analysis, a rabbit antibody raised against the SADSr-CoV 3755 N protein was used for specific staining of SADS-CoV antigen. Slides were blocked by incubating with 10% goat serum (Beyotime) at 37 °C for 30 min, followed by overnight incubation at 4 °C with the rabbit anti-3755 N protein serum (1:1,000) and mouse anti-cytokeratin 8+18+19 monoclonal antibody (Abcam), diluted 1:100 in PBST buffer containing 5% goat serum. After washing, slides were then incubated for 50 min at room temperature with Cy3-conjugated goat-anti-rabbit IgG (Proteintech) and FITC-conjugated goat-anti-mouse IgG (Proteintech), diluted 1:100 in PBST buffer containing 5% goat serum. Slides were stained with DAPI (Beyotime) and observed under a fluorescence microscope (Nikon).

Reporting Summary

Further information on experimental design is available in the Nature Research Reporting Summary linked to this paper.

Data availability

Sequence data that support the findings of this study have been deposited in GenBank with accession codes MF094681MF094688MF769416MF769444MF094697MF094701MF769406MF769415 and MG557844. Raw sequencing data that support the findings of this study have been deposited in the Sequence Read Achieve (SRA) with accession codes SRR5991648SRR5991649SRR5991650SRR5991651SRR5991652SRR5991654SRR5991655SRR5991656SRR5991657SRR5991658 and SRR5995595.

Change history

  • 05 April 2018

    The Extended Data Figures and Tables section originally published with this article was missing Tables 1–5. This has now been corrected.


References

  1. 1.

    Olival, K. J. et al. Host and viral traits predict zoonotic spillover from mammals. Nature 546, 646–650 (2017).

  2. 2.

    Guan, Y. et al. Isolation and characterization of viruses related to the SARS coronavirus from animals in southern China. Science 302, 276–278 (2003).

  3. 3.

    Lau, S. K. et al. Severe acute respiratory syndrome coronavirus-like virus in Chinese horseshoe bats. Proc. Natl Acad. Sci. USA 102, 14040–14045 (2005).

  4. 4.

    Li, W. et al. Bats are natural reservoirs of SARS-like coronaviruses. Science 310, 676–679 (2005).

  5. 5.

    Ge, X. Y. et al. Isolation and characterization of a bat SARS-like coronavirus that uses the ACE2 receptor. Nature 503, 535–538 (2013).

  6. 6.

    He, B. et al. Identification of diverse alphacoronaviruses and genomic characterization of a novel severe acute respiratory syndrome-like coronavirus from bats in China. J. Virol. 88, 7070–7082 (2014).

  7. 7.

    Yang, X. L. et al. Isolation and characterization of a novel bat coronavirus closely related to the direct progenitor of severe acute respiratory syndrome coronavirus. J. Virol. 90, 3253–3256 (2016).

  8. 8.

    Wu, Z. et al. ORF8-related genetic evidence for Chinese horseshoe bats as the source of human severe acute respiratory syndrome coronavirus. J. Infect. Dis. 213, 579–583 (2016).

  9. 9.

    Wang, L. et al. Discovery and genetic analysis of novel coronaviruses in least horseshoe bats in southwestern China. Emerg. Microbes Infect. 6, e14 (2017).

  10. 10.

    Hu, B. et al. Discovery of a rich gene pool of bat SARS-related coronaviruses provides new insights into the origin of SARS coronavirus. PLoS Pathog. 13, e1006698 (2017).

  11. 11.

    Drosten, C. et al. Identification of a novel coronavirus in patients with severe acute respiratory syndrome. N. Engl. J. Med. 348, 1967–1976 (2003).

  12. 12.

    Ksiazek, T. G. et al. A novel coronavirus associated with severe acute respiratory syndrome. N. Engl. J. Med. 348, 1953–1966 (2003).

  13. 13.

    Marra, M. A. et al. The genome sequence of the SARS-associated coronavirus. Science 300, 1399–1404 (2003).

  14. 14.

    Peiris, J. S. et al. Coronavirus as a possible cause of severe acute respiratory syndrome. Lancet 361, 1319–1325 (2003).

  15. 15.

    Rota, P. A. et al. Characterization of a novel coronavirus associated with severe acute respiratory syndrome. Science 300, 1394–1399 (2003).

  16. 16.

    Wang, L.-F. & Cowled, C. (eds) Bats and Viruses: A New Frontier of Emerging Infectious Diseases 1st edn (John Wiley & Sons, Hoboken, 2015).

  17. 17.

    Dong, N. et al. Porcine deltacoronavirus in mainland China. Emerg. Infect. Dis. 21, 2254–2255 (2015).

  18. 18.

    Sun, D., Wang, X., Wei, S., Chen, J. & Feng, L. Epidemiology and vaccine of porcine epidemic diarrhea virus in China: a mini-review. J. Vet. Med. Sci. 78, 355–363 (2016).

  19. 19.

    Lau, S. K. et al. Complete genome sequence of bat coronavirus HKU2 from Chinese horseshoe bats revealed a much smaller spike gene with a different evolutionary lineage from the rest of the genome. Virology 367, 428–439 (2007).

  20. 20.

    Chen, J. et al. Molecular epidemiology of porcine epidemic diarrhea virus in China. Arch. Virol. 155, 1471–1476 (2010).

  21. 21.

    Burbelo, P. D. et al. Serological diagnosis of human herpes simplex virus type 1 and 2 infections by luciferase immunoprecipitation system assay. Clin. Vaccine Immunol. 16, 366–371 (2009).

  22. 22.

    Gong, L. et al. A new bat-HKU2-like coronavirus in swine, China, 2017. Emerg. Infect. Dis. 23, 1607–1609 (2017).

  23. 23.

    Pan, Y. et al. Discovery of a novel swine enteric alphacoronavirus (SeACoV) in southern China. Vet. Microbiol. 211, 15–21 (2017).

  24. 24.

    Li, W. et al. Angiotensin-converting enzyme 2 is a functional receptor for the SARS coronavirus. Nature 426, 450–454 (2003).

  25. 25.

    Masters, P. S. & Perlman, S. in Fields Virology Vol. 2 (eds Knipe, D. M. & Howley, P. M.) 825–858 (Lippincott Williams & Wilkins, Philadelphia, 2013).

  26. 26.

    Raj, V. S. et al. Dipeptidyl peptidase 4 is a functional receptor for the emerging human coronavirus-EMC. Nature 495, 251–254 (2013).

  27. 27.

    Harlow, E. & Lane, D. Antibodies: A Laboratory Manual (Cold Spring Harbor Laboratory Press, New York, 1988).

  28. 28.

    Ren, W. et al. Difference in receptor usage between severe acute respiratory syndrome (SARS) coronavirus and SARS-like coronavirus of bat origin. J. Virol. 82, 1899–1907 (2008).

Download references

Acknowledgements

We thank S.-B. Xiao for providing pig cell lines, P. Burbelo for providing the luciferase immunoprecipitation system vector and L. Zhu for enabling the rapid synthesis of the S gene; the WIV animal facilities; J. Min for help with the preparation of the immunohistochemistry samples; and G.-J. Zhu and A. A. Chmura for assistance with bat sampling. This work was jointly supported by the Strategic Priority Research Program of the Chinese Academy of Sciences (XDPB0301) to Z.-L.S., China Natural Science Foundation (81290341 and 31621061 to Z.-L.S., 81661148058 to P.Z., 31672564 and 31472217 to J.-Y.M., 81572045, 81672001 and 81621005 to Y.-G.T.), National Key Research and Development Program of China (2015AA020108, 2016YFC1202705, AWS16J020 and AWS15J006) to Y.-G.T.; National Science and Technology Spark Program (2012GA780026) and Guangdong Province Agricultural Industry Technology System Project (2016LM1112) to J.-Y.M., State Key Laboratory of Pathogen and Biosecurity (SKLPBS1518) to Y.-G.T., Taishan Scholars program of Shandong province (ts201511056 to W.-F.S.), NRF grants NRF2012NRF-CRP001–056, NRF2016NRF-NSFC002-013 and NMRC grant CDPHRG/0006/2014 to L.-F.W., Funds for Environment Construction & Capacity Building of GDAS’ Research Platform (2016GDASPT-0215) to LBZ, United States Agency for International Development Emerging Pandemic Threats PREDICT project (AID-OAA-A-14-00102), National Institute of Allergy and Infectious Diseases of the National Institutes of Health (Award Number R01AI110964) to P.D. and Z.-L.S.

Reviewer information

Nature thanks C. Drosten, G. Palacios and L. Saif for their contribution to the peer review of this work.

Author information

L.-F.W., Z.-L.S., P.Z., Y.-G.T., and J.-Y.M. conceived the study. P.Z., W.Z., Y.Z., S.M., X.-S.Z., B.L., X.-L.Y., H.G., D.E.A., Y.L., X.L.L. and J.C. performed qPCR, serology and histology experiments and cultured the virus. H.F., Y.-W.Z., J.-M.L., G.-Q.P., X.-P.A., Z.-Q.M., T.-T.H., Y.H., Q.S., Y.-Y.W., S.-Z.X., X.-L.-L.Z., W.-F.S. and J.L. performed genome sequencing and annotations. T.L., Q.-M.X., J.-W.C., L.Z., K.-J.M., Z.-X.W., Y.-S.C., D.L., Y.S., F.C., P.-J.G. and R.H. prepared the samples and carried out animal challenge experiments. Z.-L.S., P.D., L.-B.Z., S.-Y.L. coordinated collection of bat samples. P.Z., L.-F.W., Z.-L.S. and P.D. had a major role in the preparation of the manuscript.

Correspondence to Peter Daszak or Lin-Fa Wang or Zheng-Li Shi or Yi-Gang Tong or Jing-Yun Ma.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Map of outbreak locations and sampling sites in Guangdong province, China and the co-circulation of PEDV and SADS-CoV during the initial outbreak on farm A.

a, SADS-affected farms are labelled (farms A–D) with blue swine silhouettes following the temporal sequence of the outbreaks. Bat sampling sites are indicated with black bat silhouettes. The bat SADSr-CoV that is most closely related to SADS-CoV (sample 162140) originated in Conghua. The red flag marks Foshan city, the site of the SARS index case. b, Pooled intestinal samples (n = 5 or more biological independent samples) were collected at dates given on the x axis from deceased piglets and analysed by qPCR. The viral load for each piglet is shown as copy number per milligram of intestine tissue (y axis).

Extended Data Fig. 2 Bayesian phylogenetic tree of the full-length genome and the ORF1a and ORF1b sequences of SADS-CoV and related coronaviruses.

a, Bayesian phylogenetic tree of the full-length genome. b, Bayesian phylogenetic tree of the ORF1a and ORF1b sequences. Trees were constructed using MrBayes with the average standard deviation of split frequencies under 0.01. The host of each sequence is represented as a silhouette. Newly sequenced SADS-CoVs are highlighted in red, bat SADSr-CoVs are shown in blue and previously published sequences are shown in black. Scale bars, nucleotide substitutions per site.

Extended Data Fig. 3 Phylogeny and haplotype network analyses of the 33 SADS-CoV strains from the four farms.

a, Phylogenetic tree constructed using MrBayes. The GTR+GAMMA model was applied and 20 million steps were run, with the first 10% removed as burn in. Viruses from different farms are labelled with different colours. Scale bar, nucleotide substitutions per site. b, Median-joining haplotype network constructed using ProART. In this analysis, ɛ = 0 was used. The size of the circles represents the number of samples. The larger the circle, the more samples it includes.

Extended Data Fig. 4 Recombination analysis for SADS-CoV and related CoVs.

The potential genetic recombination events were detected using RDP. For each virus strain, different colours represent different sources of the genomes. Source data

Extended Data Fig. 5 Isolation and antigenic characterization of SADS-CoV.

ab, Vero cells are shown 20 h after infection with mock (a) or SADS-CoV (b). cd, Mock or SADS-CoV-infected samples stained with rabbit serum raised against the recombinant SADSr-CoV N protein (red) and DAPI (blue). The experiment was conducted independently three times with similar results. Scale bars, 100 μm. Source data

Extended Data Table 1 List of all known swine viruses tested by PCR at the beginning of the of SADS outbreak investigation on the four farms

Full size table

Extended Data Table 2 List of SADSr-CoVs detected in bats in Guangdong, China

Full size table

Extended Data Table 3 Test of SADS-CoV entry and infection in Hela cells expressing known coronavirus receptors

Full size table

Extended Data Table 4 Experimental infection of SPF piglets using intestine tissue homogenate

Full size table

Extended Data Table 5 Experimental animal infection of farm piglets using cultured SADS-CoV

Full size table

Supplementary information

Supplementary Information

This file contains two supplementary tables. Supplementary table 1 contains a list of nucleotide variants among the 33 SADS-CoV genomes obtained from the four farms. Supplementary table 2 contains a list of PCR primers used in this study

Reporting Summary

Source data

Source Data Figure 1

Source Data Extended Data 4

Source Data Extended Data 5

Rights and permissions

Reprints and Permissions

About this article

Cite this article

Zhou, P., Fan, H., Lan, T. et al. Fatal swine acute diarrhoea syndrome caused by an HKU2-related coronavirus of bat origin. Nature 556, 255–258 (2018). https://doi.org/10.1038/s41586-018-0010-9

Download citation

Share this article

Anyone you share the following link with will be able to read this content:

Get shareable link

Subjects

Further reading



https://www.nature.com/ar

ticles/s41586-018-0010-9


image.png

image.png


image.png


image.png


image.png


image.png


image.png

image.png


image.png



image.png


image.png



Abstract

The discovery of SARS-like coronavirus in bats suggests that bats could be the natural reservoir of SARS-CoV. However, previous studies indicated the angiotensin-converting enzyme 2 (ACE2) protein, a known SARS-CoV receptor, from a horseshoe bat was unable to act as a functional receptor for SARS-CoV. Here, we extended our previous study to ACE2 molecules from seven additional bat species and tested their interactions with human SARS-CoV spike protein using both HIV-based pseudotype and live SARS-CoV infection assays. The results show that ACE2s of Myotis daubentoni and Rhinolophus sinicus support viral entry mediated by the SARS-CoV S protein, albeit with different efficiency in comparison to that of the human ACE2. Further, the alteration of several key residues either decreased or enhanced bat ACE2 receptor efficiency, as predicted from a structural modeling study of the different bat ACE2 molecules. These data suggest that M. daubentoni and R. sinicus are likely to be susceptible to SARS-CoV and may be candidates as the natural host of the SARS-CoV progenitor viruses. Furthermore, our current study also demonstrates that the genetic diversity of ACE2 among bats is greater than that observed among known SARS-CoV susceptible mammals, highlighting the possibility that there are many more uncharacterized bat species that can act as a reservoir of SARS-CoV or its progenitor viruses. This calls for continuation and expansion of field surveillance studies among different bat populations to eventually identify the true natural reservoir of SARS-CoV.

Introduction

Severe acute respiratory syndrome coronavirus (SARS-CoV) is the aetiological agent responsible for the SARS outbreaks during 2002–2003, which had a huge global impact on public health, travel and the world economy [411]. The host range of SARS-CoV is largely determined by the specific and high-affinity interactions between a defined receptor-binding domain (RBD) on the SARS-CoV spike protein and its host receptor, angiontensin-converting enzyme 2 (ACE2) [679]. It has been hypothesized that SARS-CoV was harbored in its natural reservoir, bats, and was transmitted directly or indirectly from bats to palm civets and then to humans [10]. However, although the genetically related SARS-like coronavirus (SL-CoV) has been identified in horseshoe bats of the genus Rhinolophus [581218], its spike protein was not able to use the human ACE2 (hACE2) protein as a receptor [13]. Close examination of the crystal structure of human SARS-CoV RBD complexed with hACE2 suggests that truncations in the receptor-binding motif (RBM) region of SL-CoV spike protein abolish its hACE2-binding ability [710], and hence the SL-CoV found recently in horseshoe bats is unlikely to be the direct ancestor of human SARS-CoV. Also, it has been shown that the human SARS-CoV spike protein and its closely related civet SARS-CoV spike protein were not able to use a horseshoe bat (R. pearsoni) ACE2 as a receptor [13], highlighting a critical missing link in the bat-to-civet/human transmission chain of SARS-CoV.

There are at least three plausible scenarios to explain the origin of SARS-CoV. First, some unknown intermediate hosts were responsible for the adaptation and transmission of SARS-CoV from bats to civets or humans. This is the most popular theory of SARS-CoV transmission at the present time [10]. Second, there is an SL-CoV with a very close relationship to the outbreak SARS-CoV strains in a non-bat animal host that is capable of direct transmission from reservoir host to human or civet. Third, ACE2 from yet to be identified bat species may function as an efficient receptor, and these bats could be the direct reservoir of human or civet SARS-CoV. Unraveling which scenario is most likely to have occurred during the 2002–2003 SARS epidemic is critical for our understanding of the dynamics of the outbreak and will play a key role in helping us to prevent future outbreaks. To this end, we have extended our studies to include ACE2 molecules from different bat species and examined their interaction with the human SARS-CoV spike protein. Our results show that there is great genetic diversity among bat ACE2 molecules, especially at the key residues known to be important for interacting with the viral spike protein, and that ACE2s of Myotis daubentoni and Rhinolophus sinicus from Hubei province can support viral entry.

Materials and methods

Cell lines and antibodies

HeLa cells were grown in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% fetal bovine serum (Gibco, USA). Rabbit polyclonal antibodies against ACE2 of R. pearsoni (RpACE2) were generated using R. pearsoni ACE2 protein expressed in Escherichia coli at the Wuhan Institute of Virology following standard procedures.

Bat sample collection and identification

Bats were sampled from their natural habitats in Hubei, Guangxi, Guizhou, Henan and Yunnan provinces in China as described previously [8]. Bat identification was initially determined in the field by morphology and later confirmed in the laboratory by sequencing the mitochondrial cytochrome b gene from samples of blood cells or rectal tissue as described previously [1].

Bat ACE2 amplification and cloning

Total RNA was extracted from bat rectal tissue using TRIzol Reagent (Invitrogen, USA) and treating with RNase-free DNase I at 37°C for 30 min. First-strand cDNA was synthesized from total RNA by reverse transcription with random hexamers. Full-length bat ACE2 fragments were amplified using the forward primer bAF2 (5′-CTTGGTACCATGTCAGGCTCTTYCTGG-3′) and the reverse primer bAR2 (5′-CCGCTCGAGCTAAAAB[G/T/C]GAV[G/A/C]GTCTGAACATCATC-3′). The PCR mixture (25 μL) contained 0.5 μL cDNA, 1.5 mM MgCl2 and 0.2 μM of each primer, and the fragments were amplified using the following parameters: 95°C for 5 min, 35 cycles of 94°C for 30 s, 55°C for 45 s and 68°C for 3 min, with a final elongation step at 68°C for 10 min. All bat ACE2s were cloned into pCDNA3.1 with KpnI and XhoI, and this was verified by sequencing.

Chimeric ACE2 construction

For samples in which full-length ACE2 amplification was unsuccessful, the N-terminal region (1–1,170 bp) was amplified using the forward primer bAF2 and the reverse primer RMR (5′-TTAGCTCCATTTCTTAGCAGGTAGG-3′). Chimeric ACE2 was constructed by combining the N-terminal region of bat ACE2 with the C-terminal portion of human ACE2 at the unique BamHI site (1,070–1,075 bp). The chimera was subsequently cloned into pCDNA3.1 with KpnI and XhoI and sequenced as above.

Construction of bat ACE2 mutants

ACE2 from M. daubentoni was chosen to generate a series of ACE2 mutants using a QuikChange II Site-Directed Mutagenesis Kit (Stratagene, USA). The altered amino acid codon for each mutant is indicated as follows: I27T, N31K, K35E, and H41Y. Mutants were confirmed by sequencing.

Sequence analysis

All bat ACE2s were submitted to GenBank (EF569964, GQ999931–GQ999938). Sequence alignment was performed using ClustalX version 1.83 [15] and corrected manually. A phylogenetic tree based on amino acid (aa) sequences was constructed using the neighbor-joining (NJ) method in MEGA version 4.1. [14].

Analysis of ACE2 expression by western blotting

Lysates of HeLa cells expressing human ACE2 or bat ACE2 were separated on a 4–10% SDS-PAGE gel, followed by transfer to a polyvinylidene difluoride (PVDF) membrane using a semi-dry protein transfer apparatus (Bio-Rad, USA). The membrane was probed with a rabbit polyclonal antibody against the bat ACE2 protein (1:200) at room temperature for 1 h, followed by incubation with alkaline-phosphatase-conjugated goat anti-rabbit IgG (1:1,000) (Chemicon, Australia). The probed proteins were visualized using NBT and BCIP color development (Promega, USA).

Pseudotype virus infection assays

An HIV-1-luciferase pseudotype virus carrying the SARS-CoV BJ01 S protein, HIV/BJ01-S, was prepared as described previously [13]. HeLa cells were seeded onto 96-well plates for 18 h and then transfected with 0.2 μg recombinant plasmid containing bat or human ACE2 using 0.5 μL Lipofectamine 2000 (Invitrogen, USA) according to the manufacturer’s protocol. At 24 h post-transfection, 30 μL medium containing HIV/BJ01-S was added to each well. At 2–3 h postinfection, unadsorbed viruses were removed, and fresh medium was added. The infection was monitored by measuring luciferase activity, expressed from the reporter gene carried by the pseudovirus, using a luciferase assay kit (Promega, USA). Cells were lysed at 48 h postinfection by adding 20 μL lysis buffer provided with the kit, and 10 μL of the resulting lysates were tested for luciferase activity by the addition of 20 μL luciferase substrate in a Turner Designs TD-20/20 luminometer. Each infection experiment was conducted in triplicate, and all experiments were repeated three times.

Live virus infection assays

Live SARS-CoV infection was carried out under BSL4 conditions at the Australian Animal Health Laboratory (AAHL) as described previously [1617]. Briefly, 48 h after transfection, the time at which expression of the ACE2 receptor on the HeLa cell surface is optimal, 2 × 106 TCID50 of virus was added to the cells for infection. The cells were fixed 24 h later by treatment with 100% methanol for 10 min and washed five times with PBST. The primary antibody, chicken anti-SARS-CoV S (produced against the recombinant S protein expressed in E. coli at AAHL), at a 1:500 dilution in 1% BSA/PBS, was added and incubated with the cells for 1 h at room temperature. An FITC anti-chicken conjugate (Chemicon, Australia) at 1:1,000 in 1% BSA/PBS was added after washing the cells five times and incubated with the cells for 1 h. Infection was monitored by immunofluorescent microscopic analysis.

Results and discussion

Cloning and expression of ACE2 genes from different bat species

ACE2 genes from seven bat species were amplified and cloned (Fig. 1, sFig. 1). Full-length genes were obtained from Rhinolophus ferrumequinum from Hubei province (Rf-HB), R. macrotis from Hubei province (Rm-HB), R. pearsoni from Guangxi (Rp-GX), R. pusillus from Hubei province (Rpu-HB), R. sinicus from Guangxi province (Rs-GX) and R. sinicus from Hubei province (Rs-HB). For the following bat species, amplification of the full-length coding region was not successful, and instead,the N-terminal region was cloned in frame with the C-terminal region of the human ACE2 gene to form a chimeric full-length ACE molecule: R. pearsoni from Guizhou province (Rp-GZ), Myotis daubentonii bat from Yunnan province (Md-YN) and Hipposideros pratti bat from Henan province (Hp-HN). The full-length sequences of bat ACE2 are identical in size to that of hACE2 (805 aa in total). Sequence comparison showed that bat ACE2s are closely related to ACE2s of other mammals and have an aa sequence identity of 80–82% to human and civet ACE2. The aa identity of ACE2 from different bat families ranges from 78 to 84%, and within the genus Rhinolophus, the sequence identity increases to 89–98%. The major sequence variation among bat ACE2s is located in the N-terminal region, which has been identified in structural studies as the SARS-CoV-binding region [67]. A phylogenetic tree was constructed based on the sequences of bat ACE2 (sFig. 2) using the MEGA package [14].

Fig. 1figure1

Sequence alignment of SARS-CoV binding regions of ACE2s from 9 bats, civet and human. The GenBank accession numbers of bat, civet and human ACE2 are as follows: human (NM021804), civet (AY881174), Rf-HB (GQ999931), Rm-HB (GQ999932), Rs-GX (GQ999933), Rp-GX (EF569964), Hp-HN (GQ999934), Rp-GZ (GQ999935), Rs-HB (GQ999936), Md-YN(GQ999937) and Rpu-HB (GQ999938). The alignment was generated using ClustalX v1.83. In black are single, fully conserved residues. In gray are strongly conserved residues. In light gray are weakly conserved residues. Asterisks indicate residues that interact directly with the receptor-binding domain of the SARS-CoV S protein

Full size image

All ACE2 genes were cloned into a eukaryotic expression vector and used to transfect HeLa cells. Western blot analysis showed that all ACE2s were expressed efficiently and at very similar levels and were recognized by a rabbit anti-bat ACE2 antibody with an apparent molecular weight of approximately 100–130 kDa (Fig. 2c).

Fig. 2figure2

Testing of the ability of bat ACE2 proteins to mediate pseudovirus HIV/BJ01-S and live SARS- CoV infection. a HeLa cells transfected with plasmids encoding bat and human ACE2s were infected with pseudovirus HIV/BJ01-S. Infectivity was determined by measuring the activity of reporter luciferase as described in “Materials and methods”. HeLa cells transfected with pcDNA3.1 and human ACE2 were used as the negative and positive controls, respectively. All tests were performed in triplicate, and the experiments were repeated three times. The error bar represents the calculated standard deviation. I27T, N31K, K35E, and H41Y are mutants of MdACE2 that were made using a QuikChange II Site-Directed Mutagenesis Kit. b SARS-CoV live virus infection using the ACE2s from bats as described in “Materials and methods”. HeLa cells transfected with pcDNA3.1 and human ACE2 were used as the negative and positive controls, respectively. c Expression of bat or human ACE2. Lysates from HeLa cells transfected with plasmid expressing human or bat ACE2 were analyzed by western blot. Rabbit anti-bat ACE2 polyclonal antibody (upper panel) or β-actin monoclonal antibody (lower panel) was used as the primary antibody. Lane 1 vector pcDNA3.1 control; lanes 2–10 bat ACE2 from samples Rf-HB, Rm-HB, Rpu-HB, Hp-HN, Rp-HB, Rp-GZ, Rs-GX, Rs-HB and Md-YN; lanes 11–14 Md-YN ACE2 mutant I27T, N31K, K35E and H41Y; lane 15 human ACE2. The abbreviations of bat species are given in the main text

Full size image

Functionality of bat ACE2 as an SARS-CoV entry receptor

To examine the susceptibility of different bat ACE2 molecules to SARs-CoV entry, the HIV/BJ01-S pseudovirus system was used to infect HeLa cells transiently expressing bat ACE2 or human ACE2 genes. Among the bat ACE2s, only MdACE2 (MdACE2) and Rs-HB ACE2 demonstrated significant pseudovirus infection, as deduced from the significantly higher level of luciferase activity in comparison to background activity in the negative control (Fig. 2a). Although such assays are not to be viewed as an absolute quantification of receptor activity, it is nevertheless worth noting that MdACE2-mediated infection seemed to be more efficient than with Rs-HB ACE2. In the same context, it is clear that the bat ACE2s were less efficient overall than the human ACE2 in this particular assay system. The biological significance of this observation remains to be determined. The functionality of MdACE2 and Rs-HB ACE2 as SARS-CoV entry receptors was further confirmed by infection with live virus. As shown in Fig. 2b, both bat ACE2 proteins could clearly support SARs-CoV infection. No attempt was made to quantify infection efficiency in this study due to difficulties encountered in conducting experiments under BSL4 conditions.

Structural modeling of bat ACE2 molecules

Homologous structural modeling of human SARS-CoV RBD complexed with MdACE2 supports MdACE2 as a receptor for human SARS-CoV S protein. The crystal structure of human SARS-CoV RBD complexed with hACE2 shows that two salt bridges at the SARS-CoV-hACE2 interface, between hACE2 Lys31 and Glu35 and between hACE2 Lys353 and hACE2 Glu38, are both buried in a hydrophobic environment and contribute critically to the SARS-CoV-hACE2 interactions (Fig. 3a, c) [7]. Disturbance of the formation of either of these salt bridges weakens SARS-CoV-hACE2 binding. The Lys31-Glu35 salt bridge at the SARS-CoV-hACE2 interface becomes an Asn31-Lys35 hydrogen bond at the SARS-CoV-Md-YNACE2 interface (Fig. 3b), which possibly weakens virus-receptor binding but still is largely compatible with the virus-receptor interface. Thr27 on hACE2 supports the Lys31-Gu35 salt bridge through hydrophobic interactions with Tyr475 (Fig. 3a); Ile27 on MdACE2 supports the Asn31-Lys35 hydrogen bond more efficiently than Thr27 through tighter hydrophobic interactions with Tyr475 (Fig. 3b). Moreover, Tyr41 on hACE2 supports the Lys353-Glu38 salt bridge (Fig. 3c); His41 on MdACE2 supports the same salt bridge less efficiently than Tyr41 (Fig. 3d). Overall, MdACE2 is an efficient receptor for SARS-CoV, despite the fact that its receptor activity is lower than that of hACE2.

Fig. 3figure3

Homologous structural modeling of SARS-CoV and Md-YN ACE2 (MdACE2) interactions. a Critical salt bridge between hACE2 Lys31 and Glu35 and the hydrophobic residues surrounding it, based on the experimentally determined crystal structure of SARS-CoV RBD complexed with hACE2 (PDB 2AJF). b Homologous structural modeling of the hydrogen bond between MdACE2 Asn31 and Lys35. The modeling was done in the program O [3]. c Critical salt bridge between hACE2 Lys353 and Glu38 and the hydrophobic residues surrounding it, based on the structure of SARS-CoV RBD complexed with hACE2. d Homologous structural modeling of the salt bridge between MdaACE2 Lys353 and SARS-CoV Glu38 and the hydrophobic residues surrounding it. Structural illustrations were prepared using the program Povscript [2]

Full size image

Compared with MdACE2, Rs-HB ACE2 contains Glu31 and Glu35, which are not compatible with each other due to their same negative charges, which disfavor virus-receptor binding. However, Rs-HB ACE2 also contains Thr27 and Tyr41, both of which support SARS-CoV entry by contributing favorably to the hydrophobic interactions at the virus-receptor interface. Thus, Rs-HB is a low-efficiency receptor for SARS-CoV. All of the other bat ACE2 molecules contain combinations of the aforementioned key residues that are completely incompatible with virus–receptor interactions. More specifically, they either contain same-charged residues at the 31 and 35 positions, which repel each other, or contain His41 and Lys27, which disfavor SARS-CoV binding (Fig. 1). In particular, Lys27 on some of these bat ACE2 molecules is incompatible with certain hydrophobic residues, such as Leu443 and Phe460, on SARS-CoV RBD (Fig. 3a, b). Therefore, these bat ACE2 molecules are not receptors for SARS-CoV.

Site-directed mutagenesis analysis

To confirm the above homologous structural analysis, we carried out site-directed mutagenesis on MdACE2. Our results show that mutations E31K, K35E, and I27T all dramatically decrease the receptor activity of MdACE2, whereas mutation H41Y greatly increases its receptor activity (Fig. 2a). Therefore, our mutagenesis data further confirmed that key residues in ACE2 determine the receptor activity of MdACE2.

Our finding that M. daubentoni and R. sinicus could support SARS-CoV infection has important implications in relation to the origin of SARS-CoV. Since all lines of investigation have indicated that ACE2-binding affinity is among the important determinants for SARS-CoV host range, our data would suggest that M. Daubentonii and R. sinicus have the potential to serve as the direct reservoirs for human SARS-CoV or its highly related civet SARS-CoV. To further investigate the potential of M. Daubentonii and R. sinicus as reservoirs for SARS-CoV, more efforts will have to be directed toward widening the surveillance of bats in these families and in different geographical locations.

Another important finding of our current study is the great genetic diversity of bat ACE2 proteins, which is in contrast to the genetically homogenous hACE2 [10]. Sequence variations of bat ACE2, especially in positions that are critical to SARS-CoV binding, such as residues 27, 31, 35, and 41, suggest that, in addition to the Md-YN and Rs-HB ACE2s, there may be many other bats with an ACE2 protein that makes them susceptible to SARS-CoV entry. This again highlights the need for more field surveillance and molecular characterization of different bat ACE2 proteins until the true reservoir host of SARS-CoV is identified and its spillover mechanisms and transmission pathways are fully characterized.

References

  1. 1.

    Cui J, Han N, Streicker D, Li G, Tang X, Shi Z, Hu Z, Zhao G, Fontanet A, Guan Y, Wang L, Jones G, Field HE, Daszak P, Zhang S (2007) Evolutionary relationships between bat coronaviruses and their hosts. Emerg Infect Dis 13:1526–1532

  2. 2.

    Fenn TD, Ringe D, Petsko GA (2003) POVScript+: a program for model and data visualization using persistence of vision ray-tracing. J Appl Crystallogr 36:944–947

  3. 3.

    Jones TA, Zou JY, Cowan SW, Kjeldgaard M (1991) Improved methods for building protein models in electron density maps and the location of errors in these models. Acta Crystallogr A 47:110–119

  4. 4.

    Ksiazek TG, Erdman D, Goldsmith CS, Zaki SR, Peret T, Emery S, Tong S, Urbani C, Comer JA, Lim W, Rollin PE, Dowell SF, Ling AE, Humphrey CD, Shieh WJ, Guarner J, Paddock CD, Rota P, Fields B, DeRisi J, Yang JY, Cox N, Hughes JM, LeDuc JW, Bellini WJ, Anderson LJ (2003) A novel coronavirus associated with severe acute respiratory syndrome. N Engl J Med 348:1953–1966

  5. 5.

    Lau SK, Woo PC, Li KS, Huang Y, Tsoi HW, Wong BH, Wong SS, Leung SY, Chan KH, Yuen KY (2005) Severe acute respiratory syndrome coronavirus-like virus in Chinese horseshoe bats. Proc Natl Acad Sci USA 102:14040–14045

  6. 6.

    Li F, Li W, Farzan M, Harrison SC (2005) Structure of SARS coronavirus spike receptor-binding domain complexed with receptor. Science 309:1864–1868

  7. 7.

    Li F (2008) Structural analysis of major species barriers between humans and palm civets for severe acute respiratory syndrome coronavirus infections. J Virol 82:6984–6991

  8. 8.

    Li W, Shi Z, Yu M, Ren W, Smith C, Epstein JH, Wang H, Crameri G, Hu Z, Zhang H, Zhang J, McEachern J, Field H, Daszak P, Eaton BT, Zhang S, Wang LF (2005) Bats are natural reservoirs of SARS-like coronaviruses. Science 310:676–679

  9. 9.

    Li W, Zhang C, Sui J, Kuhn JH, Moore MJ, Luo S, Wong SK, Huang IC, Xu K, Vasilieva N, Murakami A, He Y, Marasco WA, Guan Y, Choe H, Farzan M (2005) Receptor and viral determinants of SARS-coronavirus adaptation to human ACE2. EMBO J 24:1634–1643

  10. 10.

    Li W, Wong SK, Li F, Kuhn JH, Huang IC, Choe H, Farzan M (2006) Animal origins of the severe acute respiratory syndrome coronavirus: insight from ACE2-S-protein interactions. J Virol 80:4211–4219

  11. 11.

    Peiris JS, Lai ST, Poon LL, Guan Y, Yam LY (2003) Coronavirus as a possible cause of severe acute respiratory syndrome. Lancet 361:1319–1325

  12. 12.

    Ren W, Li W, Yu M, Hao P, Zhang Y, Zhou P, Zhang S, Zhao G, Zhong Y, Wang S, Wang LF, Shi Z (2006) Full-length genome sequences of two SARS-like coronaviruses in horseshoe bats and genetic variation analysis. J Gen Virol 87:3355–3359

  13. 13.

    Ren W, Qu X, Li W, Han Z, Yu M, Zhou P, Zhang SY, Wang LF, Deng H, Shi Z (2008) Difference in receptor usage between severe acute respiratory syndrome (SARS) coronavirus and SARS-like coronavirus of bat origin. J Virol 82:1899–1907

  14. 14.

    Tamura K, Dudley J, Nei M, Kumar S (2007) MEGA4: molecular evolutionary genetics analysis (MEGA) software version 4.0. Mol Biol Evol 24:1596–1599

  15. 15.

    Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG (1997) The CLUSTAL_X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res 25:4876–4882

  16. 16.

    Tu C, Crameri G, Kong X, Chen J, Sun Y, Yu M, Xiang H, Xia X, Liu S, Ren T, Yu Y, Eaton BT, Xuan H, Wang LF (2004) Antibodies to SARS coronavirus in civets. Emerg Infect Dis 10:2244–2248

  17. 17.

    Yu M, Stevens V, Berry JD, Crameri G, McEachern J, Tu C, Shi Z, Liang G, Weingartl H, Cardosa J, Eaton BT, Wang LF (2008) Determination and application of immunodominant regions of SARS coronavirus spike and nucleocapsid proteins recognized by sera from different animal species. J Immunol Methods 331:1–12

  18. 18.

    Yuan J, Hon CC, Li Y, Wang D, Xu G, Zhang H, Zhou P, Poon LL, Lam TT, Leung FC, Shi Z (2010) Intraspecies diversity of SARS-like coronaviruses in Rhinolophus sinicus and its implications for the origin of SARS coronaviruses in humans. J Gen Virol 91:1058–1062

Download references

Acknowledgments

This work was jointly funded by the State Key Program for Basic Research Grants (2005CB523004, 2010CB530100) from the Chinese Ministry of Science, Technology and the Knowledge Innovation Program Key Project administered by the Chinese Academy of Sciences (KSCX1-YW-R-07) to Z.S. and the CSIRO CEO Science Leader Award to L.-F.W. We thank Gary Crameri and Jennifer Barr for help with live virus infection studies.

Author information

    • Yuxuan Hou

    • , Cheng Peng

    • , Yan Li

    • , Zhenggang Han

    •  & Zhengli Shi

    • Meng Yu

    •  & Lin-Fa Wang

    • Fang Li

Correspondence to Lin-Fa Wang or Zhengli Shi.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (DOC 150 kb)

Rights and permissions

Reprints and Permissions

About this article

Cite this article

Hou, Y., Peng, C., Yu, M. et al. Angiotensin-converting enzyme 2 (ACE2) proteins of different bat species confer variable susceptibility to SARS-CoV entry. Arch Virol 155, 1563–1569 (2010). https://doi.org/10.1007/s00705-010-0729-6

Download citation

Share this article

Anyone you share the following link with will be able to read this content:

Get shareable link

Keywords

  • Salt Bridge

  • Severe Acute Respiratory Syndrome

  • ACE2 Gene

  • Pseudotype Virus

  • Severe Acute Respiratory Syndrome Coronavirus


浏览(1271) (4) 评论(16)
发表评论
文章评论
作者:Pascal 回复 一冰 留言时间:2020-02-07 14:28:58

Image result for 王岐山胡舒立

回复 | 0
作者:一冰 留言时间:2020-02-07 07:36:17

为何财新参合进来,啥意思?

回复 | 0
作者:Pascal 留言时间:2020-02-06 00:22:17

Image

Image

Image

回复 | 0
作者:Pascal 回复 体育老师 留言时间:2020-02-05 18:25:31

老师所言极是。

回复 | 0
作者:Pascal 留言时间:2020-02-05 17:56:14

Image result for 财新网 胡舒立

Image result for 财新网

Image

Image

Image

Image

国内外专家学者一致判断新冠病毒不可能是人工制造的、基因工程的产物。石正丽表示希望国家专业部门来调查,以还团队一个清白多位专家学者都认为,新冠病毒中有一些其他冠状病毒所没有的插入物,类似于HIV中发现的序列。但关键在于,这些遗传密码片段也存在于其他的病毒中,没有理由相信它们来自HIV病毒。图/视觉中国相关报道【财新周刊】封面报道|37位记者四万字全景调查:新冠病毒何以至此?【财新周刊】封面报道之一|现场篇:武汉围城【财新周刊】封面报道之二|疑似病人难题:谁来关心“移动的传染源”?【财新周刊】封面报道之三|解毒篇:溯源新冠病毒【财新周刊】封面报道之四|国际篇:全球共济2018年的“新型冠状病毒”,只会感染猪,不会传染人

  【财新网】(记者 杨睿 冯禹丁 赵今朝)“阴谋论者不相信科学。我希望国家专业部门来调查,给我们一个清白。”中国科学院武汉病毒研究所研究员石正丽2月4日回复财新记者称,“我自己的话没有说服力,我不能控制别人的思想和言论。”

  石正丽,中科院新发和烈性病原与生物安全重点实验室主任、武汉病毒所新发传染病研究中心主任、湖北省科技厅“2019新型肺炎应急科技攻关研究项目”应急攻关专家组组长,却在新型冠状病毒肺炎疫情异常严峻的关头,陷入到一场“质疑”她所在实验室可能是新冠病毒源头的舆论风波之中。

  “新型冠状病毒是人造的生化武器”“新冠病毒是从武汉病毒所实验室泄漏出来的”……自新冠疫情暴发以来,不断有人在社交媒体上宣称,新冠病毒与中科院位于武汉的病毒研究所有关。公开资料显示,中科院位于武汉的病毒研究所拥有中国唯一一个P4级别的生物安全实验室,石正丽是该实验室副主任和生物安全3级实验室主任。P4是生物安全实验室的最高防护级别,专用于研究高度危险、至今无已知疫苗或治疗方式的病原体。

  流言契合了人们心中的一些疑惑,不免引发联想。比如,为什么偏偏在武汉出现新型病毒;为什么病毒传人的源头——中间宿主始终找不到;病毒的自然宿主是蝙蝠,而石正丽实验室正是蝙蝠病毒研究的学术权威。

  石正丽团队曾于2017年确定SARS病毒是经过几个蝙蝠SARS样冠状病毒重组而来。新冠疫情暴发后,石正丽团队1月23日在生物学论文预印平台bioRxiv上即发表文章“一种新型冠状病毒的发现及其可能的蝙蝠起源”。该研究表明,新冠病毒与2003年非典病毒(SARS-CoV)的基因序列一致性达79.5%,与云南菊头蝠中存在的RaTG13冠状病毒一致性高达96%,表明其自然宿主很有可能是蝙蝠。这一研究最早在实验层面证明了新冠病毒与SARS病毒的关联,以及其可能的自然宿主为蝙蝠(参见财新周刊封面报道“新冠病毒何以至此”之三:溯源新冠病毒)。2月3日,在经过同行评审后,这篇论文已发表在国际顶尖学术期刊《自然》上。

  面对外界的质疑和指责,2月2日,石正丽在微信朋友圈愤怒回应道:“2019新型冠状病毒是大自然给人类不文明生活习惯的惩罚,我石正丽用我的生命担保,与实验室没有关系。奉劝那些相信并传播不良媒体的谣传的人、相信印度学者不靠谱的所谓学术分析的人,闭上你们的臭嘴。”

  事后,石正丽向财新记者解释,专业问题她不想与非专业人士讨论,“说不清”“没有用的,浪费我时间”,她说,“我能告诉你的是,我们是合法合规地开展实验活动。”

社交媒体上的质疑声

  最早把石正丽实验室与新型冠状病毒联系起来的流言之一,是在1月下旬,有人发现2018年央视报道过中科院武汉病毒研究所牵头的科研团队发现了一种源自蝙蝠的新型冠状病毒(SADS-CoV),便揣测新冠病毒(2019-nCoV)是否与其有关。实际上,2018年被发现的是一种“猪急性腹泻综合征冠状病毒”,与新冠病毒在分类学上不是一个种,两者的基因组相似度仅仅略高于50%。此乃天壤之别。(详见财新网报道“2018年的新型冠状病毒,只会感染猪,不会传染人”)

  随后,又有人发现石正丽曾参与2015年11月9日发表在《自然医学》上的一篇研究。该研究是关于一种在中国马蹄蝠身上发现的类SARS冠状病毒(SHC014-CoV)可能会引发的疾病,研究人员利用SARS反向遗传学系统,生成并鉴别出一种嵌合病毒。简单来说,这种嵌合病毒由SHC014的表面蛋白和SARS病毒的骨架构成。嵌合病毒可感染人类的呼吸道细胞,证明了SHC014的表面蛋白具有与细胞上的关键受体结合并感染细胞的必要结构。嵌合体能引发小鼠疾病,但无法致死。研究表明,目前蝙蝠种群中流行的病毒可能会再次引发SARS-CoV(非典病毒)疫情的潜在风险。

  需要指出的是,这篇论文共有15位作者,分别来自美国北卡罗来纳大学、阿肯色州杰斐逊市食品和药物管理局国家毒理学研究中心等单位,其中包括中科院武汉病毒研究所的两名科研人员葛行义和石正丽。在作者贡献中,葛行义负责假型实验,而石正丽则提供了SHC014棘突蛋白序列以及质粒。实验的设计和实施都是在北卡罗来纳大学教堂山分校的实验室进行的。

  质疑者据此宣称,新冠病毒可能是4年前这一实验中经改造的病毒从实验室泄漏而来。他们还援引《自然》资深记者德兰·巴特勒(Declan Butler)2015年撰写的文章称,巴黎巴斯德(Pasteur)研究所的病毒学家韦恩·霍布森(Simon Wain-Hobson)曾指出,“如果病毒逃脱,那么谁也无法预测它的传播轨迹。”

  此后,这一掐头去尾的素材在社交媒体上广为流传,甚至有人据此“实名举报”石正丽及其团队,或要求“当面对质”。值得一提的是,这些质疑者的身份背景均不是病毒相关领域的专业人士。

  财新记者核查了上述《自然》记者巴特勒撰写的报道。实际上,巴特勒的这篇报道立场中立,他援引正反两方的观点,呈现了“功能获得性研究”所引发的争议。病毒“功能获得性研究”(Gain of Function Research),是指在实验室中增加病原体的毒力、易传播性或宿主范围,以研究病毒的特性及评估新兴传染病。

  巴特勒的报道写道,有的专家反对进行此类研究,比如韦恩·霍布森反对这一实验的理由是,它没有什么益处,并未揭示出蝙蝠体内的野生SHC014病毒对人类构成什么风险。美国罗格斯大学的分子生物学家、瓦克斯曼微生物研究所实验室主任理查德·埃布赖特(Richard Ebright)认为,“这项工作的唯一影响是在实验室中创造了一种新的非自然风险。”韦恩·霍布森和理查德·埃布赖特都是“功能获得性研究”的长期批评者。

  报道称,美国国立卫生研究院(NIH)2013年10月起暂停了对所有此类研究的资助,但曾允许这一实验在该机构审查期间继续进行,因为NIH得出的结论是,这项工作的风险并没有达到令其暂停的程度。

  但也有学者认为,这类研究确实有好处。比如研究组织“生态健康联盟”总裁彼得·达斯扎克(Peter Daszak)认为,此类实验可以帮助识别应该优先考虑的病原体,以引起进一步的关注。他举例说,若没有这项实验,SHC014病毒仍将不被视作威胁——此前科学家基于分子水平建模和其他研究认为它不会感染人类细胞,而新的实验表明该病毒已经能够锁存在人类受体上。彼得·达斯扎克曾与石正丽团队有过科研合作。

  一位国内生物化学专家向财新记者解释,在保证生物安全的前提下,这类病原体功能获得性研究有助于人们更加深刻地认识病毒的作用传播机制、特异性,从而更好地预防未知病毒。“人类对细菌的研究相对透彻,对动物病毒的理解相对较少。如果要研究病毒,的确需要一些获得性的功能研究,但是要防止各类泄露,”该专家说,“科学都是双刃剑。”

印度学者论文风波

  截止此时,对石正丽及其实验室提出质疑者,都是非专业人士。而之后印度学者发表在bioRxiv上的一篇论文(目前已撤稿),则又引发了新一轮的舆论风波。

  1月31日,印度德里大学和印度理工学院的研究人员在bioRxiv上发表一篇题为“2019新冠病毒棘突蛋白中含有独特的插入序列,并与艾滋病的HIV-1 dp120和Gag蛋白有着奇特相似性”的文章。简单来说,印度学者比对了新冠病毒和SARS的棘突蛋白序列,发现与SARS病毒相比,新冠病毒的棘突蛋白上有4段新的插入序列。之后,他们将这4段插入序列与数据库中的序列进行比对,发现这4段都能够在艾滋病的蛋白序列中找到。研究称,这种不寻常的同一性/相似性在自然界中不太可能是偶然现象;这4个插入序列均为新冠病毒独有,在其他冠状病毒中不存在。

  这一新闻再次引发了一些网民的联想和演绎——新冠病毒可能是SARS病毒与艾滋病病毒人工合成的结果。

  但随后,印度学者撤下了这篇论文。目前,这篇论文的网址只剩下论文标题,摘要部分被替换为:“该论文已被作者撤稿。他们打算根据研究同行对该研究技术方法及结论阐释的反馈进行修正”。论文的一位作者还在BioRxiv平台上留言称:“这是初始研究。我们无意于为阴谋论提供原材料。尽管我们尊重科研同行在BioRxiv及其他地方作出的批评与评论,但这个故事已经在社交平台和新闻媒体上以不同的方式阐释和分享了。为避免在世界范围内造成进一步的误解和混乱,我们决定撤回当前预印本,并在重新分析后提交修正版”。

  bioRxiv预印平台也在官网上添加了一条黄色的“警告”横幅:“bioRxiv收到许多与新冠病毒有关的新论文。有必要提醒您:这些是未经同行评审的原始报告,不应被认为是具有结论性的,不能指导临床实践/与健康相关的行为,或作为既成事实信息在新闻媒体上报道。”

  事实上,印度学者论文的结论在撤稿前已遭到多位国际专家的质疑。美国斯坦福大学生物化学助理教授Silvana Konermann发推特称,她检查了印度学者的结果,却发现所谓的相似性是假的。“他们比对新冠病毒和SARS,发现4个新插入的序列。其中有两个已经在蝙蝠的冠状病毒中发现了。剩下两个插入序列中,其中有一个与HIV病毒序列最为相似,但是非常短,以至于都不比偶然性高;另一个插入序列与HIV之外的另13种病毒序列都更为相似,且这些相似性都不比偶然性高(鉴于插入片段的大小/病毒蛋白质数据库的大小)。

  美国华盛顿大学医学系和基因组科学系副教授、Fred Hutchinson癌症研究中心的生物信息学专家特雷弗·贝德福德(Trevor Bedford)则在推特上公开了自己重复比对的结果,他确认这些短插入序列确实存在于新冠病毒中,但是这些插入序列与多种生物(的序列)都能匹配,没有理由得出是HIV病毒序列的结论。

  美国俄亥俄州立大学病毒与新发病原学中心主任刘善虑在接受《知识分子》采访时也指出,从科学和病毒进化的角度讲,冠状病毒和属于逆转录病毒的艾滋病毒差别很大,发生基因间重组的可能性不大,因为同源性太差。再者,根据经验,我们会时常看到所要比较的序列中有几十个核苷酸或几个氨基酸与某些毫不相关的东西完全或非常相似,但不具有任何生物学意义。

  多位专家学者都认为,新冠病毒中有一些其他冠状病毒所没有的插入物,类似于HIV中发现的序列。但关键在于,这些遗传密码片段也存在于其他的病毒中,没有理由相信它们来自HIV病毒。

专家:新冠不可能人工制造

  2020年1月21日,《中国科学:生命科学》英文版在线发表文章,揭示了武汉冠状病毒感染人类的机制。它是通过S-蛋白与人ACE2相互作用的分子机制,来感染人的呼吸道上皮细胞。

  研究作者之一中国科学院上海巴斯德研究所研究员郝沛告诉财新记者,从人体作用机制一致角度来看,武汉冠状病毒感染能力与SARS类似,但是影响病毒传播因素感染能力只是一方面,还有病毒复制、病毒传播途径等其他因素影响。

  1月22日,北京大学、广西中医药大学、宁波大学及武汉生物工程学院学者联合攻关,在Journal of Medical Virology 在线发表的一篇研究论文称,研究结果表明,新型冠状病毒2019-nCoV似乎是蝙蝠冠状病毒与起源未知的冠状病毒之间的重组病毒。

  那么,到底新冠病毒有没有可能是人工制造的、基因工程的产物?财新记者采访了多位专家学者,他们一致的判断是,不可能。

  特雷弗·贝德福德告诉财新记者,没有证据表明新冠病毒源自基因工程。他解释说,相较于蝙蝠身上的病毒(RaTG13冠状病毒),新冠病毒中存在的基因差异与自然进化是一致的。“如果是基因编辑的病毒,那么需要替换掉大量的遗传物质,目前没有观察到这种痕迹。相反,只看到一些符合自然进化的零星、分散的改变。”

  特雷弗·贝德福德等人建立了一个开源网站nextstrain.org,能够对各种病原体的基因序列进行分析和可视化,其中包括此前已知冠状病毒家族(蝙蝠、果子狸、SARS)的基因序列系谱,以及此次共享在全球共享流感病毒数据库GISAID中的53例新冠肺炎患者的病毒基因组全序列。贝德福德结合病毒之间核苷酸的差异和其他冠状病毒的假定突变率估算,RaTG13和新冠病毒这两种病毒“在25-65年前有一个共同的祖先”。因此,RaTG13病毒突变为新冠病毒可能需要几十年的时间。

  根据贝德福德的分析,RaTG13冠状病毒与新冠病毒之间有近1100个核苷酸的差异。可兹对比的是,果子狸冠状病毒与人类SARS病毒之间的差异仅为10个核苷酸。

  中科院一位生物信息学领域的研究员告诉财新记者,新冠病毒“肯定是天然的,它不可能是人工的”。其依据在于,新冠病毒基因组全序列分析表明,其基因序列与云南菊头蝠中存在的RaTG13冠状病毒最为接近,一致性高达96%。但这4%的基因差异也是极大的,人和老鼠的基因组相似度也高达90%。如此差异绝非人工所能填补,因为新冠病毒有近3万个核苷酸,4%就是近1200个点位的基因突变,“只有大自然才有这种本事,经过很多年的进化造出这样一个病毒,”该人士说。此外,病毒丢失基因片段很正常,它在精简自己,病毒每个周期要合成基因组,对它来说功能越精简效率越高,但病毒获得基因片段插入其实是很难的,“(阴谋论者)把人类想得太伟大了。就算人类能做到(插入这些基因片段),为什么要弄1200个基因突变?稍微改一下不就行了?为什么搞那么复杂,最后自己也控制不住?”

  美国宾夕法尼亚大学医学院副研究员李懿泽告诉财新记者,新冠病毒不可能是实验室制造出来的,“实验室制造病毒需要反向遗传系统,而这个系统的核心是序列拼接,序列拼接必须要人工引入内切酶位点,会留下人工改造的痕迹。只要在序列中发现了人工引入的位点,就可以基本上认定这个病毒是人工制造的,不可能做到天衣无缝。 而新型冠状病毒没有人工引入内切酶位点的痕迹,因此不可能是实验室制造的。”

  前述“功能获得性研究”的批评者理查德·埃布赖特告诉财新记者,基于目前病毒的基因序列分析,没有实质性证据证明,病毒经过基因编辑(was engineered)。

  但他也提醒道,把病毒是否被基因编辑(这种可能性已被排除)与病毒是因为实验室事故进入人群(目前还不能排除)区分开,是重要的。

  “当下抗击疫情是最重要的事情。当疫情结束后,需要发起法医调查(forensic investigation)以确定此次疫情暴发的源头。”他说。

  美国俄亥俄州立大学预防兽医系终身教授王秋红近日接受《中国科学报》采访时,也呼吁官方介入。她指出,自从疫情发生以来出现了很多谣言,一种是说病毒是中科院武汉病毒所制造出来的,一种是说是北卡罗来纳大学教堂山分校Ralph Baric实验室泄露的,Baric实验室曾用SARS病毒感染小鼠进行实验。“我特别希望国家成立一个专家团来辟谣,”她说,“现在基因序列已经发布,从这个序列分析,没有人工制造的位点,不可能是从实验室泄露的,完全是自然界的病毒。”

  1月31日,《科学》杂志撰文称,大多数研究人员认为,病毒基因序列驳斥了新冠病原体来自武汉病毒研究所的观点。曾与石正丽合作研究的Peter Daszak对《科学》称:“每当一种新的疾病或病毒出现时,都会有同样的故事:这是某个机构释放或泄露的,又或者这是基因编辑的病毒。这是一种耻辱(It’s just a shame)。”

  的确,每逢重大疫情总有类似阴谋论产生。比如2003年非典疫情期间,“SARS是人工武器”的流言也曾盛传一时。2014年,埃博拉在西非几内亚等国大暴发后,也有言论称是美国人制造了埃博拉病毒。麻省理工学院政治学系教授Vipin Narang近日在推特上表示,没有证据显示这场疫情中有生物武器的存在,刻意传播这种流言“完全不负责任”。“事实上,(如果是生物武器的话)这是一个很糟糕的生物武器,因为存在回爆的风险。好的生物武器应该被设计成高致命性,但低传染性。”他说。

  (记者 徐路易 张子竹 实习记者 黄晏浩 对本文亦有贡献)

*******************************************************************

Image

Image

回复 | 0
作者:体育老师 回复 Pascal 留言时间:2020-02-05 17:40:47

这群不知天高地厚,只惦记自己出人头地的,能称为科学家吗?

科学家首先要知道“what to do?” 他们知道吗?以为”撸起子“什么都能做!

回复 | 3
作者:Pascal 留言时间:2020-02-05 16:39:40

Image

Image

******************************************************************

回复 | 0
作者:gmuoruo 回复 老度 留言时间:2020-02-05 15:54:26

军方的意思是暗示美国在武汉投毒,想用这个法子来忽悠国内的脑残们,想把美国当成这次瘟疫的背锅侠。

-----

哈,哈,哈,老度兄,不要排除美帝生物战的可能哦。

说不定美帝发现中国俄杂的基因不同,比如万维的几只显然不像普通中国人的。美帝生物战不用耽心伤及无辜。

回复 | 0
作者:老度 回复 一冰 留言时间:2020-02-05 14:20:27

【不过他举例西路网,不太能说服我,因为有时文章仅代表作者观点。】

西陆网是军方网站,文章说的是武汉新型肺炎病毒是人工制造,是一种生物武器,所列举的例子也很有说服力。

军方的意思是暗示美国在武汉投毒,想用这个法子来忽悠国内的脑残们,想把美国当成这次瘟疫的背锅侠。

目前土共的宣传出现了不同调的地方,按照王沪宁的意思,先是把武汉肺炎病毒推给野生动物,如果不行,就让美国人来背锅,军方也是在配哈中宣部的舆论导向,不过上得也太块了一些,第一步还没走完,就上第二步了,出现了不协调的地方。

回复 | 0
作者:Pascal 留言时间:2020-02-05 13:41:46

http://news.creaders.net/china/2020/02/05/2186975.html

http://tech.creaders.net/2020/02/05/2186945.html

回复 | 0
作者:Pascal 留言时间:2020-02-05 13:22:46

Image

回复 | 0
作者:Pascal 留言时间:2020-02-05 13:12:22

这篇文章是那位波董事长公开举报时列出的证据吗

不是,源自这里:

https://twitter.com/SylarSu2018/status/1225004471664762881

回复 | 0
作者:一冰 留言时间:2020-02-05 13:03:28

这篇文章是那位波董事长公开举报时列出的证据吗,因为都是术语我没有打开,估计看不懂。以前我也追踪过一个类似的案子,需要有化学家和药学家·病毒学家共同协助工作,郭文贵和他的团队也很能干,大胆假设,小心求证,不过他举例西路网,不太能说服我,因为有时文章仅代表作者观点。

回复 | 0
作者:Pascal 留言时间:2020-02-05 13:01:33

https://twitter.com/MgOqkzLBRPLCHyN/status/1224498503876935681

Image result for 人民日报 logo"

Image

Image

Image

Image

Image

Image result for 人民日报 logo"

image.png

image.png

image.png

image.png

image.png

Image

Image

Image

*******************************************************************

回复 | 0
作者:一冰 留言时间:2020-02-05 13:00:26

这个五毒所真是罪该万死!

作为科学家,如果没有人性,没有良知,没有方向,就成了科学恶魔,石正丽和她的团队与纳粹门格尔医生和日本731部队有一大区别:日本和德国是用外族人做实验,可是石正丽他们竟然用本国同胞实地大规模试验起来,这真是打开了潘多拉的盒子,王毅真是乌鸦嘴。

回复 | 0
作者:Pascal 留言时间:2020-02-05 12:49:19

******************************************************************

Image

Image

image.png

image.png

Image

image.png

image.png

image.png

image.png

image.png

Image

Image

Image

Image

Image

回复 | 0
我的名片
Pascal
注册日期: 2014-10-22
访问总量: 8,564,085 次
点击查看我的个人资料
Calendar
最新发布
· 认认八万年前没有走出非洲东亚人
· 或是世界第1位在职医生公开向接
· 中共中央笑开了怀.说好的百年未
· 中央军委排兵布阵平台灭美.4600
· 你小子不是死不打苗躲过一劫吗.
· 盖茨吐真言.终于找到党卫军首脑
· 传福音!种疫苗与苗后心脏猝死没
分类目录
【他山之石】
· 认认八万年前没有走出非洲东亚人
· 或是世界第1位在职医生公开向接
· 中共中央笑开了怀.说好的百年未
· 中央军委排兵布阵平台灭美.4600
· 你小子不是死不打苗躲过一劫吗.
· 盖茨吐真言.终于找到党卫军首脑
· 传福音!种疫苗与苗后心脏猝死没
· 18年前先知卡扎菲.穆斯林将不费
· 美国政府拟公布中共最高层在美资
· 美参议员披露.15名联邦高官2018
存档目录
2024-04-01 - 2024-04-22
2024-03-01 - 2024-03-31
2024-02-01 - 2024-02-29
2024-01-01 - 2024-01-31
2023-12-01 - 2023-12-31
2023-11-01 - 2023-11-30
2023-10-01 - 2023-10-31
2023-09-01 - 2023-09-30
2023-08-01 - 2023-08-31
2023-07-01 - 2023-07-31
2023-06-01 - 2023-06-30
2023-05-01 - 2023-05-31
2023-04-01 - 2023-04-30
2023-03-01 - 2023-03-31
2023-02-01 - 2023-02-28
2023-01-01 - 2023-01-31
2022-12-01 - 2022-12-31
2022-11-01 - 2022-11-30
2022-10-01 - 2022-10-31
2022-09-01 - 2022-09-29
2022-08-01 - 2022-08-31
2022-07-01 - 2022-07-31
2022-06-01 - 2022-06-30
2022-05-01 - 2022-05-31
2022-04-02 - 2022-04-29
2022-03-01 - 2022-03-31
2022-02-01 - 2022-02-28
2022-01-01 - 2022-01-31
2021-12-01 - 2021-12-31
2021-11-01 - 2021-11-30
2021-10-01 - 2021-10-31
2021-09-01 - 2021-09-30
2021-08-01 - 2021-08-31
2021-07-01 - 2021-07-31
2021-06-01 - 2021-06-30
2021-05-01 - 2021-05-31
2021-04-01 - 2021-04-30
2021-03-01 - 2021-03-31
2021-02-01 - 2021-02-28
2021-01-01 - 2021-01-31
2020-12-01 - 2020-12-31
2020-11-01 - 2020-11-30
2020-10-01 - 2020-10-31
2020-09-01 - 2020-09-30
2020-08-01 - 2020-08-31
2020-07-01 - 2020-07-31
2020-06-01 - 2020-06-30
2020-05-01 - 2020-05-31
2020-04-01 - 2020-04-30
2020-03-02 - 2020-03-31
2020-02-01 - 2020-02-29
2020-01-01 - 2020-01-31
2019-12-01 - 2019-12-31
2019-11-01 - 2019-11-30
2019-10-01 - 2019-10-31
2019-09-01 - 2019-09-30
2019-08-01 - 2019-08-31
2019-07-01 - 2019-07-31
2019-06-01 - 2019-06-30
2019-05-01 - 2019-05-30
2019-04-01 - 2019-04-30
2019-03-01 - 2019-03-31
2019-02-01 - 2019-02-28
2019-01-02 - 2019-01-31
2018-12-01 - 2018-12-31
2018-11-01 - 2018-11-30
2018-10-01 - 2018-10-31
2018-09-02 - 2018-09-24
2018-08-01 - 2018-08-31
2018-07-04 - 2018-07-31
2018-06-01 - 2018-06-30
2018-05-01 - 2018-05-31
2018-04-01 - 2018-04-30
2018-03-02 - 2018-03-31
2018-02-01 - 2018-02-28
2018-01-10 - 2018-01-30
2017-11-01 - 2017-11-30
2017-10-01 - 2017-10-30
2017-09-22 - 2017-09-29
2017-08-02 - 2017-08-30
2017-07-01 - 2017-07-31
2017-06-02 - 2017-06-30
2017-05-02 - 2017-05-30
2017-04-01 - 2017-04-29
2017-03-01 - 2017-03-31
2017-02-02 - 2017-02-28
2017-01-02 - 2017-01-31
2016-12-03 - 2016-12-30
2016-11-05 - 2016-11-28
2016-10-01 - 2016-10-29
2016-09-01 - 2016-09-29
2016-08-01 - 2016-08-30
2016-07-01 - 2016-07-31
2016-06-02 - 2016-06-30
2016-05-01 - 2016-05-27
2016-04-01 - 2016-04-30
2016-03-01 - 2016-03-31
2016-02-04 - 2016-02-28
2016-01-01 - 2016-01-28
2015-12-03 - 2015-12-31
2015-11-03 - 2015-11-29
2015-10-02 - 2015-10-30
2015-09-10 - 2015-09-28
2015-08-02 - 2015-08-31
2015-07-01 - 2015-07-28
2015-06-02 - 2015-06-30
2015-05-01 - 2015-05-31
2015-04-02 - 2015-04-29
2015-03-02 - 2015-03-31
2015-02-02 - 2015-02-27
2015-01-03 - 2015-01-31
2014-12-01 - 2014-12-31
2014-11-01 - 2014-11-30
2014-10-26 - 2014-10-31
 
关于本站 | 广告服务 | 联系我们 | 招聘信息 | 网站导航 | 隐私保护
Copyright (C) 1998-2024. CyberMedia Network /Creaders.NET. All Rights Reserved.