Skip to main content
Advertisement
  • Loading metrics

Pathogenicity, tissue tropism and potential vertical transmission of SARSr-CoV-2 in Malayan pangolins

  • Xianghui Liang ,

    Contributed equally to this work with: Xianghui Liang, Xiaoyuan Chen, Junqiong Zhai, Xiaobing Li, Xu Zhang

    Roles Investigation, Validation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Xiaoyuan Chen ,

    Contributed equally to this work with: Xianghui Liang, Xiaoyuan Chen, Junqiong Zhai, Xiaobing Li, Xu Zhang

    Roles Investigation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Junqiong Zhai ,

    Contributed equally to this work with: Xianghui Liang, Xiaoyuan Chen, Junqiong Zhai, Xiaobing Li, Xu Zhang

    Roles Investigation, Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Xiaobing Li ,

    Contributed equally to this work with: Xianghui Liang, Xiaoyuan Chen, Junqiong Zhai, Xiaobing Li, Xu Zhang

    Roles Investigation

    Affiliations State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China, College of Life Sciences, Longyan University, Longyan, China

  • Xu Zhang ,

    Contributed equally to this work with: Xianghui Liang, Xiaoyuan Chen, Junqiong Zhai, Xiaobing Li, Xu Zhang

    Roles Investigation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Zhipeng Zhang,

    Roles Investigation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Ping Zhang,

    Roles Investigation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Xiao Wang,

    Roles Investigation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Xinyuan Cui,

    Roles Investigation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Hai Wang,

    Roles Investigation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Niu Zhou,

    Roles Investigation, Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Zu-Jin Chen,

    Roles Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Renwei Su,

    Roles Investigation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Fuqing Zhou,

    Roles Investigation

    Affiliation Department of Radiology, the First Affiliated Hospital, Nanchang University, Nanchang, China

  • Edward C. Holmes,

    Roles Investigation, Writing – review & editing

    Affiliation Sydney Institute for Infectious Diseases, School of Medical Sciences, University of Sydney, Sydney, New South Wales, Australia

  • David M. Irwin,

    Roles Writing – review & editing

    Affiliations Department of Laboratory Medicine and Pathobiology, University of Toronto, Toronto, Canada, Banting and Best Diabetes Centre, University of Toronto, Toronto, Canada

  • Rui-Ai Chen,

    Roles Writing – review & editing

    Affiliations State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China, Zhaoqing Branch Center of Guangdong Laboratory for Lingnan Modern Agricultural Science and Technology, Zhaoqing, China

  • Qian He,

    Roles Investigation

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Ya-Jiang Wu,

    Roles Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Chen Wang,

    Roles Software

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Xue-Qing Du,

    Roles Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Shi-Ming Peng,

    Roles Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Wei-Jun Xie,

    Roles Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Fen Shan,

    Roles Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Wan-Ping Li,

    Roles Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Jun-Wei Dai,

    Roles Resources

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  • Xuejuan Shen,

    Roles Investigation, Project administration

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Yaoyu Feng,

    Roles Writing – review & editing

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Lihua Xiao,

    Roles Writing – review & editing

    Affiliation State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China

  • Wu Chen ,

    Roles Investigation, Resources, Supervision, Writing – review & editing

    guangzhouchenwu@sina.com (WC); shenyy@scau.edu.cn (YS)

    Affiliation Guangzhou Zoo & Guangzhou Wildlife Research Center, Guangzhou, China

  •  [ ... ],
  • Yongyi Shen

    Roles Data curation, Funding acquisition, Project administration, Resources, Supervision, Writing – original draft

    guangzhouchenwu@sina.com (WC); shenyy@scau.edu.cn (YS)

    Affiliations State Key Laboratory for Animal Disease Control and Prevention, Guangdong Laboratory for Lingnan Modern Agriculture, Center for Emerging and Zoonotic Diseases, College of Veterinary Medicine, South China Agricultural University, Guangzhou, China, Guangdong Provincial Key Laboratory of Zoonosis Prevention and Control, Guangzhou, China

  • [ view all ]
  • [ view less ]

Abstract

Malayan pangolin SARS-CoV-2-related coronavirus (SARSr-CoV-2) is closely related to SARS-CoV-2. However, little is known about its pathogenicity in pangolins. Using CT scans we show that SARSr-CoV-2 positive Malayan pangolins are characterized by bilateral ground-glass opacities in lungs in a similar manner to COVID-19 patients. Histological examination and blood gas tests are indicative of dyspnea. SARSr-CoV-2 infected multiple organs in pangolins, with the lungs the major target, and histological expression data revealed that ACE2 and TMPRSS2 were co-expressed with viral RNA. Transcriptome analysis indicated that virus-positive pangolins were likely to have inadequate interferon responses, with relative greater cytokine and chemokine activity in the lung and spleen. Notably, both viral RNA and viral proteins were detected in three pangolin fetuses, providing initial evidence for vertical virus transmission. In sum, our study outlines the biological framework of SARSr-CoV-2 in pangolins, revealing striking similarities to COVID-19 in humans.

Author summary

SARS-CoV-2 has caused a severe and ongoing global pandemic. To date, bats and pangolins are the only groups of mammals known carry SARS-CoV-2-related coronaviruses (SARSr-CoV-2). Although the SARSr-CoV-2 detected in confiscated Malayan pangolins from Guangdong, China (PCoV-GD) has lower sequence similarity to SARS-CoV-2 than some bat-derived viruses, its receptor-binding domain within the S protein that binds to host ACE2 receptors for entry into cells is almost identical to that of SARS-CoV-2. Here, we determined the pathogenicity, tissue tropism and transcriptional response of PCoV-GD in deceased adult Malayan pangolins and their fetuses. Notably, virus-positive pangolins showed ground-glass opacities in their lungs similar to COVID-19 pneumonia and blood gas tests that were indicative of dyspnea. We detected viral RNA in multiple pangolin tissues, including liver, intestine, heart, kidney, spleen and skeletal muscle. This is similar to autopsy reports in COVID-19 patients and is associated with multiple organ dysfunction syndrome. The virus-positive pangolins showed an inadequate interferon response. In addition, PCoV-GD was detected in pangolin fetuses, suggesting the potential vertical transmission of the virus in utero. In sum, Malayan pangolins infected with a PCoV-GD display a pathology that is strikingly similar to human COVID-19.

Introduction

After SARS-CoV and MERS-CoV, SARS-CoV-2 is the third coronavirus to cause severe respiratory illness in humans identified in the past two decades. Although characterized by a lower infection fatality ratio [1], SARS-CoV-2 has caused an ongoing and severe global pandemic. SARS-CoV-2 belongs to the genus Betacoronavirus of the family Coronaviridae, sharing 79.5% overall genome nucleotide sequence identity with SARS-CoV. A better understanding of the pathobiology of SARS-CoV-2 is critical for its control and the development of therapeutics. Several animal models have been developed to help determine the pathogenicity of SARS-CoV-2, including transduced or transgenic mouse models that express human ACE2 [2,3], and macaques that can be infected and recapitulate moderate disease [4,5].

The origin and zoonotic source of SARS-CoV-2 remains uncertain. Bat-derived SARS-CoV-2-related coronavirus (SARSr-CoV-2), such as RaTG13 (from China) and the BANAL isolates (from Laos), are currently the most closely related to the SARS-CoV-2 at the whole genome level, with ~96–97% nucleotide sequence identify [6,7]. Strikingly, Malayan pangolins (Manis javanica) were also found to be infected by MERS-CoV-like viruses [8] and at least two SARSr-CoV-2 lineages [911]. Although these pangolin SARSr-CoV-2 have lower overall sequence similarity to SARS-CoV-2 than some bat-derived SARSr-CoV-2, the receptor-binding domain of the S protein of the pangolin-CoV-GD (PCoV-GD) is almost identical to that of SARS-CoV-2 [10].

Understanding the biology of SARSr-CoV-2 in their hosts may shed new light on key aspects of coronavirus disease. Previously, the functional comparison of SARS-CoV-2 with closely related pangolin and bat coronaviruses was performed using pseudotyped viruses [12] and spike glycoprotein structures [13,14]. However, the biology of SARSr-CoV-2 in pangolins remains unknown. Herein, we define the pathogenicity, tissue tropism and transcriptional response of PCoV-GD in Malayan pangolins.

Results

SARSr-CoV-2 infection in pangolins

We studied 28 adult Malayan pangolins that were confiscated in Guangdong Province between March 2019 to April 2020 and initially sent to Guangzhou wildlife rescue center. Notably, six of these animals were pregnant such that six pangolin fetuses were also available for study. We used the viral S (spike) gene to identify infection with PCoV-GD. This revealed that 14 animals, including six pregnant females, were naturally infected by the virus. All the PCoV-GD positive samples were from the batch confiscated in March 2019. Notably, the S genes from these animals exhibited no differences in the consensus genome sequence. RNA sequencing data enabled the de novo assembly of three near complete PCoV-GD genomes that exhibited 99.6–99.9% genetic identity with the PCoV-GD reference genome (EPI_ISL_410721), and 89.8%-90.4% with the SARS-CoV-2 (WHCV). As pangolins are wildlife species, it is possible that they might be infected by viruses other than PCoV-GD which may have in turn impacted pathogenicity. Accordingly, based on the transcriptome data, we found that these pangolins were mainly infected by Sendai virus, pestivirus and parvovirus (S1 Table).

Pathogenicity of PCoV-GD in Malayan pangolins

Blood gas tests revealed elevated levels of PCO2 (58.2, and 60.25 mmHg), HCO3 (29.9, and 34.6 mmol/L), and TCO2 (31.5 mmol/L) in the infected animals (S2 Table), indicative of dyspnea. CT scans in one PCoV-GD positive pangolin revealed a diffused bilateral distribution of ground-glass opacities in the lungs (Fig 1A). The other PCoV-GD positive pangolin exhibited air bronchogram, fibrotic streaks and a subpleural transparent line (Fig 1B).

thumbnail
Fig 1. Chest CTs of two PCoV-GD-positive and one virus negative pangolin.

(A) Multiple ground-glass opacities in bilateral lungs in one pangolin that was positive to both PCoV-GD and Sendai virus. Arrowheads 1: ground-glass opacities. (B) Bilateral focal consolidation, lobar consolidation, and patchy consolidation were clearly observed in the second PCoV-GD positive pangolin, which was negative for other viruses. Arrowheads 2: air bronchogram; Arrowheads 3: fibrotic streaks; Arrowheads 4: subpleural transparent line. (C) Chest CT of a virus-negative pangolin. This pangolin was negative to PCoV-GD, Sendai virus, pestivirus and parvovirus.

https://doi.org/10.1371/journal.ppat.1011384.g001

Our previous study revealed that four PCoV-GD positive Malayan pangolins had diffuse alveolar damage of varying severity in the lung [10]. To further examine the pathogenicity of the PCoV-GD, lungs from pangolin p60 which was positive to PCoV-GD but negative to other viruses, were stained with hematoxylin and eosin (H&E) for histological examination. Compared with virus-negative pangolins (Fig 2B), the most significant symptom in the lungs was severe interstitial pneumonia, in which the interstitium and the walls of the alveoli were thickened and displayed bronchiectasis (Fig 2A). To ensure these lesions were associated with PCoV-GD, we used immunohistochemistry to detect and localize the nucleocapsid (N protein) of the virus. Several pneumocytes were positive (Fig 2C). qRT-PCR and western blot analysis further supported the presence of PCoV-GD infection in the lungs (Fig 3A and 3B). Necropsy reports of this pangolin revealed hepatomegaly and mucosal injury of intestine (S1 Fig).

thumbnail
Fig 2. Pathological changes and virus antigen expression in the lungs of one PCoV-GD positive and one negative pangolin.

(A) lung sections of a PCoV-GD positive compared to a negative individual (B) were stained by hematoxylin and eosin (H&E). Arrowheads 1, thickened walls of the alveoli; Arrowheads 2, proliferation of alveolar epithelial cell; and red blood cells (Arrowheads 3) are seen. (C-D) Immunofluorescence for virus antigen expression in pneumocytes. (C) nucleocapsid of PCoV-GD (light red, arrowheads 4); (D) negative control. Scale bars of (A) and (B) are 100um, while those in (C) and (D) are 50um. The PCoV-GD positive pangolin (p60) was negative to other viruses. Pangolin tissues were fixed since their necropsy in 2019, hence they were not sufficiently fresh for high quality immunofluorescence, such that the immunofluorescence result was less clear than using fresh tissues.

https://doi.org/10.1371/journal.ppat.1011384.g002

thumbnail
Fig 3. Expression patterns of PCoV-GD and the ACE2 and TMPRSS2 receptor genes in different tissues of adult pangolins, pregnant pangolins and their fetuses.

(A) Quantitative real-time PCR (qRT-PCR) of the S gene of PCoV-GD. (B) Western blot of the N protein from the virus-positive tissues. Tissue lysates were subjected to SDS-PAGE and immunoblotting with antisera against SARS-CoV-2 N protein. In this study, we aim to demonstrate the presence of pangolin-CoV N protein in multiple tissues, and it is very difficult to choose a reference protein that expressed in all these tissues. Therefore three reference proteins were used. As the expression of reference proteins inherently varies greatly among tissues, loading control based on the grayscale of the reference proteins is not appropriate for our study. We therefore measured the total protein concentration for each sample, and controlled the consistency of protein loading by adding the same amount of total protein for each sample (30ug). These pangolins died in 2019, prior to the emergence of COVID-19. The importance of these samples was not realized at the time, so that sample collection was incomplete and the numbers of different tissues inconsistent. Fold change of mean values of relative expression of the ACE2 gene (C) and TMPRSS2 gene (D).

https://doi.org/10.1371/journal.ppat.1011384.g003

Tissue tropism and potential vertical transmission of PCoV-GD

Necropsy of virus-positive pangolins also revealed multiple organ damage (S1 Fig). We therefore used qRT-PCR to examine the tissue tropism of PCoV-GD. Viral RNA was mainly detected in the lungs, although other organs, including the liver, intestine, heart, kidney, spleen and skeletal muscle of some individuals, also had detectable levels of virus RNA. The six pregnant pangolins had relative lower ΔCt values, and thus higher levels of viral RNA in the lung, liver, kidney and muscle than in adult pangolins. Although pregnant pangolins had relative lower ΔCt values in spleen and muscle compared to fetuses, the different was not statistically significant (Fig 3A and S3 Table). Western blotting further supported the presence of viral infection in these tissues (Fig 3B). PCoV-GD RNA and the virus N protein were detected in the intestine, spleen, and skeletal muscle of three of the six fetuses (Fig 3A and 3B and S3 Table).

Expression of ACE2 and TMPRSS2 was detected in multiple tissues, and the lung, liver, and spleen had the highest levels of ACE2 and TMPRSS expression compared to the other tissues examined (Fig 3C and 3D). PCoV-GD RNA was also detected in the tissues expressing the highest levels of ACE2 and TMPRSS. However, PCoV-GD was also detected in the intestine and kidney, tissues that do not have high expression of these receptor genes in pangolins.

Host transcriptional response to PCoV-GD

As shown in S1 Table, most PCoV-GD positive pangolins were co-infected by Sendai virus, pestivirus and parvovirus. To reduce the influence of these viruses, individual animals infected by any of these viruses were excluded. After exclusion, two PCoV-GD positive and two negative adult pangolins remained, as well as two PCoV-GD positive and two negative fetuses.

Up-regulated different expressed genes (DEGs) were associated with the host immune response, such as ISG15, IL17D, CXCL9, CXCL10, LTB and BMP6 in the lung, ISG15, IFIT2, IFIT3, IL27, CCL24 and CXCL10 in spleen, and ISG15, IL12A and CXCL10 in muscle (Fig 4). These DEGs were enriched in the immune response and inflammatory response et al. in the lungs, while the type I interferon signaling pathway, defense response to virus, and innate immune response were up-regulated in the spleens of adult animals (Fig 4D–4F and S4 Table). In the case of the fetuses there were no up-regulated ISGs and very few up-regulated DEGs associated with immune response in lungs. In contrast, ISG15, CXCL8, CXCL13 et al. were up-regulated in the spleen, and IL1RN, TNFRSF11B, TNFSF18 et al. up-regulated in muscle. Functional enrichment analysis of DEGs showed that the fetuses were likely characterized by an inadequate immune response (Fig 4D–4F and S4 Table).

thumbnail
Fig 4. Host transcriptional response to PCoV-GD in the lung, spleen and muscle in adult pangolins and their fetuses.

Heatmap of DEGs belonging to ISGs and cytokine-related genes of lung (A), spleen (B) and muscle (C). The graphs show the log2(fold change) of DEGs of positive compared with negative animals. The genes included have a log2(fold change) of more than 1. Functional enrichment analysis of DEGs of lung (D), spleen (E) and muscle (F).

https://doi.org/10.1371/journal.ppat.1011384.g004

Discussion

COVID-19 is characterized by a range of symptoms, in most cases including fever, cough, dyspnea, and myalgia [15]. Malayan pangolins that were positive for PCoV-GD exhibited respiratory symptoms such as cough and shortness of breath. Bilateral opacities on x-ray or patchy shadows and ground glass opacities visible in CT scans are the most common features of severe COVID-19 cases [16,17]. Pangolin CT scans showed a diffused bilateral distribution of ground-glass opacities in the lungs of one virus positive pangolin (Fig 1A), similar to early-phase COVID-19 pneumonia [16,17]. The other virus-positive pangolin exhibited air bronchogram, fibrotic streaks and a subpleural transparent line (Fig 1B), similar to advanced COVID-19 disease [18]. Histological examination further revealed the interstitium and the walls of the alveoli were thickened and displayed bronchiectasis, supporting the severe interstitial pneumonia (Fig 2). These features are indicative of dyspnea and consistent with the blood gas tests that revealed elevated levels of PCO2, HCO3, and TCO2 in the infected animals (S2 Table). Hence, pangolins naturally infected with PCoV-GD show similar symptoms and CT features as COVID-19 pneumonia.

The spike protein of SARS-CoV-2 binds to ACE2 receptors for entry into cells, while TMPRSS2 mediates spike protein activation and facilitates viral entry [1921]. In humans, ACE2 is abundant in the epithelia of the lung, kidney, and small intestine [22]. Although respiratory symptoms dominate the clinical presentation of COVID-19, SARS-CoV-2 is also associated with multiple organ dysfunction syndrome [15], and SARS-CoV-2 RNA is detected in multiple organs [23]. In pangolins, ACE2 and TMPRSS are also expressed in multiple tissues (Fig 3C and 3D), and similar to COVID-19 patients, viral RNA was detected in these tissues (Fig 3A and S3 Table). Hence, these data are suggestive of viral damage to multiple organs (S1 Fig). However, whether the expression levels of ACE2 and TMPRSS2 drive virus tropism needs further study.

The ACE2 receptor is widely expressed in the placenta during pregnancy [24], presenting a theoretical risk of vertical transmission of SARS-CoV and SARS-CoV-2. The risk of SARS-CoV-2 infection in pregnant women and the potential risk of vertical transmission of SARS-CoV-2 are of great concern [25,26]. Although neonatal cases of SARS-CoV-2 infection and antibodies in newborns have been identified [2729], testing was only performed in newborns such that infection after birth, or the transfer of antibodies from the placenta and breast milk, cannot be excluded. We detected PCoV-GD RNA at a relative higher level in the six virus-positive pregnant pangolins (Fig 3A). This may mean that pregnant animals are more susceptible to the virus, although this will need to be confirmed with a larger sample size. Further, viral RNA and the N protein were detected in three pangolin fetuses (Fig 3A and 3B and S3 Table). Hence, these results are compatible with the vertical transmission of PCoV-GD in utero, although this again needs to be confirmed with a larger sample size.

The S genes from these virus-positive pangolins exhibited no changes in the consensus sequence. That the virus genomes across all the animals sampled exhibited 99.6–99.9% genetic identity is indicative of a localized outbreak in these animals due to a single source of infection. As such, it is appropriate to compare their transcriptional response to the PCoV-GD. Interferon induction of ISGs is an essential part of the antiviral response after cellular detection of viral entry into a host cell [30,31]. In a similar manner to SARS-CoV and SARS-CoV-2 [32,33], transcriptome analysis revealed only small significant increases of ISG expression in virus-positive pangolins in all tissues, implying an inadequate interferon response (Fig 4). It is therefore possible that PCoV-GD may inhibit IFN1 by regulating IFN-β synthesis and signaling to avoid the innate immune response in a similar manner to other coronaviruses [3436]. CXCL10 is associated with severe disease in SARS-CoV [15,33,37]. Similarly, we observed the up-regulation of CXCL10 in the lung, spleen and muscle of adult pangolins. While the fetuses showed inadequate immune responses, especially interferon induction of ISGs in all tissues, this may due to their underdeveloped immune system.

Our study has some limitations. As pangolins are endangered and protected, all their tissues were necessarily collected after their death. In addition, this is not a controlled challenge study as pangolins studied were naturally infected by PCoV-GD. As the pangolins were wild caught, they were infected with additional pathogens, and after co-infected individuals were excluded only a small number of animals were available for study. These caveats might impact analyses of their immune responses.

In sum, although PCoV-GD is genetically distinct from SARS-CoV-2, we show that infected Malayan pangolins exhibit a strikingly similar pathogenicity to human COVID-19. In addition, we show that PCoV-GD was present in pangolin fetuses and was present at high levels in pregnant pangolins, highlighting the potential for virus transmission in utero.

Materials and methods

Pangolin samples

The Malayan pangolin (Manis javanica) is classified as critically endangered in accordance with the International Union for Conservation of Nature (IUCN) Red List of Threatened Species R, and included in Appendix I of the Convention on International Trade in Endangered Species of Wild Fauna and Flora (CITES I). The 28 adult Malayan pangolins, six of which were pregnant, used in this study were confiscated by Customs and the Department of Forestry of Guangdong Province between March 2019 to April 2020. Of these animals, 21 were part of our previous study [10]. These animals were initially sent to Guangzhou wildlife rescue center, during which time some presented with respiratory symptoms. During animal rescue, computed tomographic (CT) scans were performed on two sick animals. For comparison, the CT scan of another pangolin without respiratory symptoms was also performed. Unfortunately, despite exhaustive rescue efforts, all of the pangolins, including the virus-negative ones, died. To determine the cause of their death, veterinarians at the rescue center performed autopsies. Tissues, including lung, liver, spleen, skeletal muscle, kidney, intestine, and heart were collected and stored at -80°C. As most pangolins died in 2019, prior to the emergence of COVID-19, the importance of these samples was not appreciated at the time so that sample collection was incomplete. Available tissue samples are shown in S3 Table. Secondary tissue usage was approved by Guangzhou Wildlife Research Center (permit number: GZZOO20190827C).

qRT-PCR analysis

PCoV-GD viral RNA, as well as ACE2 and TMPRSS2 transcript levels, were measured by real-time PCR using SYBR Green chemistry. Total RNA was isolated using the QIAamp Viral RNA Mini Kit (Qiagen) according to the kit instructions. 5μg of total RNA was reverse transcribed into cDNA in a 20 μL reaction volume using by PrimeScript IV 1st strand cDNA Synthesis Mix (Takara, Code No. 6215A). For quantitative real-time PCR (qRT-PCR), cDNA was used with ChamQ Universal SYBR qPCR Master Mix (Vazyme, Code No. Q711-02) in a CFX Connect Real-Time System (BIO-RAD). Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) was selected as a host reference gene to normalize the relative viral loads, evaluated by calculating -ΔΔCt values. Each quantitative PCR was performed in triplicate. For PCoV-GD viral RNA detection, a PCoV-GD positive and a negative sample were used as control. The primers used are shown in S5 Table, and are specific such that they do not cross-react with other sequences in the pangolin genome.

Western blots

Protein was extracted with radio-immuno precipitation assay (RIPA) buffer (50 mM Tris, pH 8.0, 150 mM NaCl, 1.0% NP-40, 0.5% sodium deoxycholate, 0.1% SDS) and a protease inhibitor cocktail. 10mg of total protein was separated on SDS-PAGE and then transferred to PVDF membranes using the Bio-Rad Trans-Blot protein transfer system. The membrane was blocked at room temperature for 90 min with 5% skim milk in 1 × TBST (20 mM Tris–HCl, pH 7.4, 150 mM NaCl, and 0.1% Tween 20). After blocking, the membrane was incubated with SARS-CoV-2 nucleocapsid antibody (kindly provided by Dr. Zhengli Shi, Wuhan Institute of Virology), recombinant anti-GAPDH antibody (Abcam, USA), and recombinant anti-actin antibody (Abcam, USA) at 4°C overnight, washed three times with 1 × TBST, and incubated with corresponding peroxidase-conjugated (Thermo Fisher Scientific) or fluorescently labeled (IRDye 680RD or 800CW: Li-Cor Biosciences) secondary antibody. The membranes were washed three times with 1 × TBST, and then scanned on an Odyssey Infrared Imaging System (Li-Cor Biosciences).

Histopathology and immunofluorescence

Lungs from six Malayan pangolins, including four positive for PCoV-GD, were used for histopathology and immunofluorescence. Five of these animals (four virus positive and one virus negative) were used in our previous study [10]. As only ten hematoxylin and eosin (H&E) sections for each sample were available in our previous study [10], and without immunofluorescence results, for the current study their tissue blocks were re-sectioned, and histopathology and immunofluorescence were performed. Histopathology of kidneys and livers from two virus-positive pangolins were also performed.

Tissues were fixed in 10% buffered formalin since their necropsy in 2019 and hence were not of sufficient freshness for detailed analysis. They were washed to remove formalin, dehydrated in ascending grades of ethanol, cleared with chloroform, and embedded with molten paraffin wax in a template. The tissue blocks were sectioned with a microtome, and sections were transferred onto grease-free glass slides, deparaffinized, and rehydrated through descending grades of ethanol and distilled water.

Sections were stained with a hematoxylin and eosin staining kit (Baso Diagnostics Inc., Wuhan Servicebio Technology Co., Ltd.). Other sections were incubated in 3% H2O2 solution in methanol to block endogenous peroxidase activity after dewaxing in xylene and hydrating in different concentrations of ethanol (100%, 95%, 85%, 75% and 50%). Subsequently, sections were incubated in permeabilizing solution containing 1% Triton X-100 (Beijing Solarbio Science & Technology Co., Ltd) in PBS for one hour and then treated with microwaves for antigen retrieval and washed with PBS. Sections were then blocked using 5% BSA at room temperature for one hour and incubated overnight in a humidified chamber with primary antibody against SARS-CoV-2 nucleocapsid (1:500, SinBiological China) at 4°C. Following three washes in PBS, sections were incubated with goat anti-rabbit IgG H&L (HRP) (1,2000, Abcam, USA)in a humidified chamber for 30 min at 37°C and then incubated with Cyanine 3 Tyramide (Cy3 TSA Fluorescence System Kit, Runnerbio Co. Ltd, China), for 30 min at 37°C. After washing five times with PBS, sections were mounted with Antifade Mounting Medium with DAPI (Beijing ZSGB Biotechnology Co., Ltd). Tissue fluorescence was visualized using a fluorescence microscope (OLYMPUS, Tokyo, Japan).

RNA sequencing

TRIzol (Invitrogen) was used for the extraction of total RNA, with DNA digested with DNase I before library construction. Ribosome RNA (rRNA) was also removed before library preparation. Sequencing libraries were generated using NEBNext Ultra Directional RNA Library Prep Kit for Illumina (NEB, USA) following the manufacturer’s recommendations and index codes were added to attribute the sequences to each sample. Sequencing of the strand-specific libraries was performed on an Illumina NovaSeq 6000 platform, generating 150 bp paired-end reads. Approximately 12 Gb of raw data were generated for each sample (S1 Table). RNA sequencing progress was performed by Novogene (Beijing Novogene Technologies, Beijing, China).

Bioinformatic analyses

Adaptor and low-quality sequences were trimmed using fastp (v0.19.7) [38]. Viral sequences were counted by mapping the clean reads to the PCoV-GD reference genome (GIASID: EPI_ISL_410721) through BWA-MEM (v0.7.17) [39]. Viral genomes were de novo assembled using Megahit (v1.0.3) [40]. For RNA-seq analysis, the Malayan pangolin genome (YNU_ManJav_2.0) was used as the reference host genome [41]. Sequence alignments were performed using the program Hisat2 [42]. Gene counts were summarized using the featureCounts program [43] as part of the Subread package release 2.0.0 (http://subread.sourceforge.net/). Reads mapping and gene counts were performed using the strand-specific pattern for NEB library. Samples with a pangolin genome mapping rate of >70% were used for downstream analysis. Differential expression analysis of two groups was performed using the edgeR package in R [44]. DEGs were characterized for each pair group (|log2 foldchange| > 1, P value < 0.01).

To identify pangolin orthologs and make a GO enrichment analysis of different expressed genes (DEGs), the unigenes were conducted by sequence similarity comparisons against the Swiss-Prot database (https://web.expasy.org/groups/swissprot/) with BLASTp (v2.7.1+). A GO enrichment analysis of different expressed genes (DEGs) was performed using clusterProfiler [45] in the R package. Heatmaps of gene expression levels were constructing using pheatmap in R. Dolt plots were constructed using ggplot2 [46]. The heatmap of Type-I IFN responses was constructed on genes with log2 foldchange > 1 belonging to the following GO annotations: GO:0035457, GO:0035458, GO:0035455, GO:0035456, GO:0034340. The heatmap of cytokine activity and chemokine activity was constructed on genes with log2 foldchange > 1 belonging GO annotations for GO: 0005125 and GO: 0008009.

Supporting information

S1 Fig. Photos taken at necropsy from a PCoV-GD positive Malayan.

(A) hepatomegaly; (B) mucosal injury of intestine. This pangolin (p60) was positive to PCoV-GD but negative to other viruses.

https://doi.org/10.1371/journal.ppat.1011384.s001

(TIF)

S1 Table. Summary of the pangolin transcriptome data.

https://doi.org/10.1371/journal.ppat.1011384.s002

(DOCX)

S2 Table. Blood gas analysis and routine blood tests of pangolins.

https://doi.org/10.1371/journal.ppat.1011384.s003

(DOCX)

S3 Table. ΔCt of qRT-PCR of S gene of PCoV-GD, ACE2 and TMPRSS2 genes.

https://doi.org/10.1371/journal.ppat.1011384.s004

(DOCX)

S4 Table. Functional annotation clustering for DEGs.

https://doi.org/10.1371/journal.ppat.1011384.s005

(XLSX)

Acknowledgments

We thank Dr. Zhengli Shi, Wuhan Institute of Virology, for providing the SARS-CoV-2 nucleocapsid antibody.

References

  1. 1. Verity R, Okell LC, Dorigatti I, Winskill P, Whittaker C, Imai N, et al. Estimates of the severity of coronavirus disease 2019: a model-based analysis. Lancet Infect Dis. 2020;20(6):669–77. Epub 2020/04/03. pmid:32240634; PubMed Central PMCID: PMC7158570.
  2. 2. Muñoz-Fontela C, Widerspick L, Albrecht RA, Beer M, Carroll MW, de Wit E, et al. Advances and gaps in SARS-CoV-2 infection models. PLoS Pathog. 2022;18(1):e1010161. Epub 2022/01/14. pmid:35025969; PubMed Central PMCID: PMC8757994.
  3. 3. Sun SH, Chen Q, Gu HJ, Yang G, Wang YX, Huang XY, et al. A mouse model of SARS-CoV-2 infection and pathogenesis. Cell Host Microbe. 2020;28(1):124–33.e4. Epub 2020/06/03. pmid:32485164; PubMed Central PMCID: PMC7250783.
  4. 4. Munster VJ, Feldmann F, Williamson BN, van Doremalen N, Pérez-Pérez L, Schulz J, et al. Respiratory disease in rhesus macaques inoculated with SARS-CoV-2. Nature. 2020;585(7824):268–72. Epub 2020/05/13. pmid:32396922.
  5. 5. Rockx B, Kuiken T, Herfst S, Bestebroer T, Lamers MM, Oude Munnink BB, et al. Comparative pathogenesis of COVID-19, MERS, and SARS in a nonhuman primate model. Science. 2020;368(6494):1012–5. Epub 2020/04/19. pmid:32303590; PubMed Central PMCID: PMC7164679.
  6. 6. Zhou P, Yang XL, Wang XG, Hu B, Zhang L, Zhang W, et al. A pneumonia outbreak associated with a new coronavirus of probable bat origin. Nature. 2020;579(7798):270–3. Epub 2020/02/06. pmid:32015507; PubMed Central PMCID: PMC7095418.
  7. 7. Temmam S, Vongphayloth K, Baquero E, Munier S, Bonomi M, Regnault B, et al. Bat coronaviruses related to SARS-CoV-2 and infectious for human cells. Nature. 2022;604(7905):330–6. Epub 2022/02/17. pmid:35172323.
  8. 8. Chen J, Yang X, Si H, Gong Q, Que T, Li J, et al. A bat MERS-like coronavirus circulates in pangolins and utilizes human DPP4 and host proteases for cell entry. Cell. 2023;186(4):850–63.e16. Epub 2023/02/22. pmid:36803605.
  9. 9. Wacharapluesadee S, Tan CW, Maneeorn P, Duengkae P, Zhu F, Joyjinda Y, et al. Evidence for SARS-CoV-2 related coronaviruses circulating in bats and pangolins in Southeast Asia. Nat Commun. 2021;12(1):972. Epub 2021/02/11. pmid:33563978.
  10. 10. Xiao K, Zhai J, Feng Y, Zhou N, Zhang X, Zou JJ, et al. Isolation of SARS-CoV-2-related coronavirus from Malayan pangolins. Nature. 2020;583(7815):286–9. Epub 2020/05/08. pmid:32380510.
  11. 11. Lam TT, Jia N, Zhang YW, Shum MH, Jiang JF, Zhu HC, et al. Identifying SARS-CoV-2-related coronaviruses in Malayan pangolins. Nature. 2020;583(7815):282–5. Epub 2020/03/29. pmid:32218527.
  12. 12. Nie J, Li Q, Zhang L, Cao Y, Zhang Y, Li T, et al. Functional comparison of SARS-CoV-2 with closely related pangolin and bat coronaviruses. Cell Dis. 2021;7(1):21. Epub 2021/04/08. pmid:33824288; PubMed Central PMCID: PMC8022302.
  13. 13. Zhang S, Qiao S, Yu J, Zeng J, Shan S, Tian L, et al. Bat and pangolin coronavirus spike glycoprotein structures provide insights into SARS-CoV-2 evolution. Nat Commun. 2021;12(1):1607. Epub 2021/03/13. pmid:33707453; PubMed Central PMCID: PMC7952905.
  14. 14. Wrobel AG, Benton DJ, Xu P, Calder LJ, Borg A, Roustan C, et al. Structure and binding properties of Pangolin-CoV spike glycoprotein inform the evolution of SARS-CoV-2. Nat Commun. 2021;12(1):837. Epub 2021/02/07. pmid:33547281; PubMed Central PMCID: PMC7864994.
  15. 15. Huang C, Wang Y, Li X, Ren L, Zhao J, Hu Y, et al. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. Lancet. 2020;395(10223):497–506. Epub 2020/01/28. pmid:31986264; PubMed Central PMCID: PMC7159299.
  16. 16. Zhu N, Zhang D, Wang W, Li X, Yang B, Song J, et al. A novel coronavirus from patients with pneumonia in China, 2019. N Engl J Med. 2020;382(8):727–33. Epub 2020/01/25. pmid:31978945.
  17. 17. Shi H, Han X, Jiang N, Cao Y, Alwalid O, Gu J, et al. Radiological findings from 81 patients with COVID-19 pneumonia in Wuhan, China: a descriptive study. Lancet Infect Dis. 2020;20(4):425–34. Epub 2020/02/28. pmid:32105637; PubMed Central PMCID: PMC7159053.
  18. 18. Zhou S, Wang Y, Zhu T, Xia L. CT features of Coronavirus Disease 2019 (COVID-19) pneumonia in 62 patients in Wuhan, China. AJR Am J Roentgenol. 2020;214(6):1287–94. Epub 2020/03/07. pmid:32134681.
  19. 19. Hoffmann M, Kleine-Weber H, Schroeder S, Krüger N, Herrler T, Erichsen S, et al. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and Is blocked by a clinically proven protease inhibitor. Cell. 2020;181(2):271–80.e8. Epub 2020/03/07. pmid:32142651; PubMed Central PMCID: PMC7102627.
  20. 20. Murgolo N, Therien AG, Howell B, Klein D, Koeplinger K, Lieberman LA, et al. SARS-CoV-2 tropism, entry, replication, and propagation: Considerations for drug discovery and development. PLoS Pathog. 2021;17(2):e1009225. Epub 2021/02/18. pmid:33596266; PubMed Central PMCID: PMC7888651.
  21. 21. Mittal A, Manjunath K, Ranjan RK, Kaushik S, Kumar S, Verma V. COVID-19 pandemic: Insights into structure, function, and hACE2 receptor recognition by SARS-CoV-2. PLoS Pathog. 2020;16(8):e1008762. Epub 2020/08/22. pmid:32822426; PubMed Central PMCID: PMC7444525.
  22. 22. Hamming I, Timens W, Bulthuis ML, Lely AT, Navis G, van Goor H. Tissue distribution of ACE2 protein, the functional receptor for SARS coronavirus. A first step in understanding SARS pathogenesis. J Pathol. 2004;203(2):631–7. Epub 2004/05/14. pmid:15141377; PubMed Central PMCID: PMC7167720.
  23. 23. Stein SR, Ramelli SC, Grazioli A, Chung JY, Singh M, Yinda CK, et al. SARS-CoV-2 infection and persistence in the human body and brain at autopsy. Nature. 2022;612(7941):758–63. Epub 2022/12/15. pmid:36517603; PubMed Central PMCID: PMC9749650.
  24. 24. Levy A, Yagil Y, Bursztyn M, Barkalifa R, Scharf S, Yagil C. ACE2 expression and activity are enhanced during pregnancy. Am J Physiol Regul Integr Comp Physiol. 2008;295(6):R1953–61. Epub 2008/10/24. pmid:18945956.
  25. 25. Schmid MB, Fontijn J, Ochsenbein-Kölble N, Berger C, Bassler D. COVID-19 in pregnant women. Lancet Infect Dis. 2020;20(6):653. Epub 2020/03/21. pmid:32197098; PubMed Central PMCID: PMC7270341.
  26. 26. Chen H, Guo J, Wang C, Luo F, Yu X, Zhang W, et al. Clinical characteristics and intrauterine vertical transmission potential of COVID-19 infection in nine pregnant women: a retrospective review of medical records. Lancet. 2020;395(10226):809–15. Epub 2020/03/11. pmid:32151335; PubMed Central PMCID: PMC7159281.
  27. 27. Dong L, Tian J, He S, Zhu C, Wang J, Liu C, et al. Possible vertical transmission of SARS-CoV-2 from an infected mother to her newborn. JAMA. 2020;323(18):1846–8. Epub 2020/03/28. pmid:32215581; PubMed Central PMCID: PMC7099527.
  28. 28. Zeng H, Xu C, Fan J, Tang Y, Deng Q, Zhang W, et al. Antibodies in infants born to mothers with COVID-19 pneumonia. JAMA. 2020;323(18):1848–9. Epub 2020/03/28. pmid:32215589; PubMed Central PMCID: PMC7099444.
  29. 29. Zeng L, Xia S, Yuan W, Yan K, Xiao F, Shao J, et al. Neonatal early-onset infection with SARS-CoV-2 in 33 neonates born to mothers with COVID-19 in Wuhan, China. JAMA Pediatr. 2020;174(7):722–5. Epub 2020/03/28. pmid:32215598; PubMed Central PMCID: PMC7099530.
  30. 30. Bailey CC, Zhong G, Huang IC, Farzan M. IFITM-family proteins: the cell’s first line of antiviral defense. Annu Rev Virol. 2014;1:261–83. Epub 2015/01/20. pmid:25599080; PubMed Central PMCID: PMC4295558.
  31. 31. Schneider DS, Ayres JS. Two ways to survive infection: what resistance and tolerance can teach us about treating infectious diseases. Nat Rev Immunol. 2008;8(11):889–95. Epub 2008/10/18. pmid:18927577; PubMed Central PMCID: PMC4368196.
  32. 32. Reghunathan R, Jayapal M, Hsu LY, Chng HH, Tai D, Leung BP, et al. Expression profile of immune response genes in patients with Severe Acute Respiratory Syndrome. BMC Immunol. 2005;6:2. Epub 2005/01/19. pmid:15655079; PubMed Central PMCID: PMC546205.
  33. 33. Blanco-Melo D, Nilsson-Payant BE, Liu WC, Uhl S, Hoagland D, Møller R, et al. Imbalanced host response to SARS-CoV-2 drives development of COVID-19. Cell. 2020;181(5):1036–45. Epub 2020/05/18. pmid:32416070
  34. 34. Kopecky-Bromberg SA, Martínez-Sobrido L, Frieman M, Baric RA, Palese P. Severe acute respiratory syndrome coronavirus open reading frame (ORF) 3b, ORF 6, and nucleocapsid proteins function as interferon antagonists. J Virol. 2007;81(2):548–57. Epub 2006/11/17. pmid:17108024; PubMed Central PMCID: PMC1797484.
  35. 35. Perlman S, Netland J. Coronaviruses post-SARS: update on replication and pathogenesis. Nat Rev Microbiol. 2009;7(6):439–50. Epub 2009/05/12. pmid:19430490; PubMed Central PMCID: PMC2830095.
  36. 36. Yuen CK, Lam JY, Wong WM, Mak LF, Wang X, Chu H, et al. SARS-CoV-2 nsp13, nsp14, nsp15 and orf6 function as potent interferon antagonists. Emerg Microbes Infect. 2020;9(1):1418–28. Epub 2020/06/13. pmid:32529952; PubMed Central PMCID: PMC7473193.
  37. 37. McKechnie JL, Blish CA. The innate immune system: Fighting on the front lines or fanning the flames of COVID-19? Cell Host Microbe. 2020;27(6):863–9. pmid:32464098
  38. 38. Chen S, Zhou Y, Chen Y, Gu J. fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics (Oxford, England). 2018;34(17):i884–i90. Epub 2018/11/14. pmid:30423086; PubMed Central PMCID: PMC6129281.
  39. 39. Li H, Durbin R. Fast and accurate long-read alignment with Burrows-Wheeler transform. Bioinformatics. 2010;26(5):589–95. Epub 2010/01/19. pmid:20080505; PubMed Central PMCID: PMC2828108.
  40. 40. Li D, Liu CM, Luo R, Sadakane K, Lam TW. MEGAHIT: an ultra-fast single-node solution for large and complex metagenomics assembly via succinct de Bruijn graph. Bioinformatics. 2015;31(10):1674–6. Epub 2015/01/23. pmid:25609793.
  41. 41. Hu JY, Hao ZQ, Frantz L, Wu SF, Chen W, Jiang YF, et al. Genomic consequences of population decline in critically endangered pangolins and their demographic histories. Natl Sci Rev. 2020;7(4):798–814. Epub 2020/04/01. pmid:34692098; PubMed Central PMCID: PMC8288997.
  42. 42. Kim D, Paggi JM, Park C, Bennett C, Salzberg SL. Graph-based genome alignment and genotyping with HISAT2 and HISAT-genotype. Nature biotechnology. 2019;37(8):907–15. Epub 2019/08/04. pmid:31375807; PubMed Central PMCID: PMC7605509.
  43. 43. Liao Y, Smyth GK, Shi W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics. 2014;30(7):923–30. Epub 2013/11/15. pmid:24227677.
  44. 44. Robinson MD, McCarthy DJ, Smyth GK. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26(1):139–40. Epub 2009/11/17. pmid:19910308; PubMed Central PMCID: PMC2796818.
  45. 45. Yu G, Wang LG, Han Y, He QY. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS. 2012;16(5):284–7. Epub 2012/03/30. pmid:22455463; PubMed Central PMCID: PMC3339379.
  46. 46. Wickham H. ggplot2: elegant graphics for data analysis: springer; 2016.