Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Metagenomic Profile of the Viral Communities in Rhipicephalus spp. Ticks from Yunnan, China

  • Han Xia,

    Affiliations State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China, Department of Microbiology and Immunology, University of Texas Medical Branch, Galveston, TX, United States of America

  • Cheng Hu,

    Affiliation International School of Software, Wuhan University, Wuhan, Hubei, China

  • Dake Zhang,

    Affiliation Laboratory of Genome Variations and Precision Biomedicine, Beijing Institute of Genomics, Chinese Academy of Sciences. NO.1 Beichen West Road, Chaoyang District, Beijing, China

  • Shuang Tang,

    Current address: State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China

    Affiliation State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China

  • Zhong Zhang,

    Current address: State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China

    Affiliation State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China

  • Zheng Kou,

    Affiliation State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China

  • Zhaojun Fan,

    Current address: State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China

    Affiliation State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China

  • Dennis Bente,

    Affiliation Department of Microbiology and Immunology, University of Texas Medical Branch, Galveston, TX, United States of America

  • Changqing Zeng ,

    czeng@big.ac.cn (CZ); litx@wh.iov.cn (TL)

    Affiliation Laboratory of Genome Variations and Precision Biomedicine, Beijing Institute of Genomics, Chinese Academy of Sciences. NO.1 Beichen West Road, Chaoyang District, Beijing, China

  • Tianxian Li

    czeng@big.ac.cn (CZ); litx@wh.iov.cn (TL)

    Current address: State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China

    Affiliation State Key Laboratory of Virology, Wuhan Institute of Virology, Chinese Academy of Sciences, Wuhan, Hubei, China

Abstract

Besides mosquitoes, ticks are regarded as the primary source of vector-borne infectious diseases. Indeed, a wide variety of severe infectious human diseases, including those involving viruses, are transmitted by ticks in many parts of the world. To date, there are no published reports on the use of next-generation sequencing for studying viral diversity in ticks or discovering new viruses in these arthropods from China. Here, Ion-torrent sequencing was used to investigate the presence of viruses in three Rhipicephalus spp. tick pools (NY-11, NY-13, and MM-13) collected from the Menglian district of Yunnan, China. The sequencing run resulted in 3,641,088, 3,106,733, and 3,871,851 reads in each tick pool after trimming. Reads and assembled contiguous sequences (contigs) were subject to basic local alignment search tool analysis against the GenBank database. Large numbers of reads and contigs related to known viral sequences corresponding to a broad range of viral families were identified. Some of the sequences originated from viruses that have not been described previously in ticks. Our findings will facilitate better understanding of the tick virome, and add to our current knowledge of disease-causing viruses in ticks living under natural conditions.

Introduction

Ticks are second only to mosquitoes as important arthropod vectors for spreading viruses from wildlife to domestic animals and humans. They are also a source of unknown viruses. To date, at least 38 known viral species are transmitted by ticks, and some of them are a significant threat to human health [1]. Such viruses include tick-borne encephalitis virus [2], Crimean-Congo hemorrhagic fever virus (CCHFV) [3], Kyasanur forest disease virus [4], Alkhurma virus, severe fever with thrombocytopenia syndrome virus (SFTSV) [1, 5], and Heartland virus (HRTV)[6]. Additionally, several tick-borne viruses also threaten the health of livestock; these include Africa swine fever virus, Nairobi sheep disease virus (NSDV), and louping ill virus [7, 8]. Among ixodid ticks, viral disease-causing vectors are found mostly in the following genera: Ixodes, Haemaphysalis, Hyalomma, Amblyomma, Dermacentor, Rhipicephalus, and Boophilus [1].

In recent years, novel tick-borne viral diseases have emerged worldwide. From 2009 to 2011, an acute febrile illness of tick-borne origin was noted in several Chinese provinces and killed about 30% of the people infected. The novel virus, SFTSV, which belongs to the Phlebovirus genus of the family Bunyaviridae, was found to be the causative agent of severe fever with thrombocytopenia syndrome. Later on, Haemaphysalis longicornis was identified as the primary vector of SFTSV [5]. HRTV, another novel tick-borne phlebovirus associated with two cases of critical febrile illness in humans, was found in the United States in 2009. Amblyomma americanum ticks have been suggested as potential vectors of this disease [6, 9].

Next generation sequencing (NGS) technologies provide a powerful means of studying viral metagenomics, and can help us to gain better understanding of viral populations and discover unknown viruses in a number of environments [10, 11]. NGS is especially useful for assessment of uncultured samples, such as feces, blood, water, air and potential viral reservoirs [1217]. The study which used NGS to explore the viral community in mosquitoes presented that mosquito virome contained sequences related to a broad range of animal, plant, insect and bacterial viruses. And the majority of the sequences from viral community in mosquitoes were novel [16]. Recently a research about tick virome from Amblyomma americanum, Dermacentor variablilits, and Ixodes scapularis ticks in the United States was published. Their results reveal novel highly divergent viruses in ticks, which include nairoviruses, phleboviruses, monoegavirusand viruses with similarity to plant and insect viruses [17]. However, there are no published studies on the use of NGS for exploring the viral diversity present in ticks from China.

Recent reports indicate that many tick-borne diseases exist in the Yunnan Province of China, such as Kyasanur forest disease, Crimean-Congo hemorrhagic fever, Colorado tick fever, and severe fever with thrombocytopenia syndrome [1820]. The Menglian district, which belongs to subtropical lower mountainous areas, is located in southwest Yunnan close to the borders of China and Laos and China and Myanmar. The climate in Menglian is typically subtropical, warm and moist, and the abundant plants and wild animals (e.g., bearcat, deer, hare, mouse, and monkey) in this region create a suitable habitat for ticks. Additionally, many residents depend on raising livestock (e.g., goat, dog, cattle, buffalo, and horse), planting tobacco or tea for their economic livelihoods, and some even share their houses with livestock. These life habits may increase the risk of infection with tick-borne diseases from tick bites through close contact with livestock and plantations. An epidemiological study in the southeast region of Yunnan in 2008 revealed that partial sequences related to the CCHFV S segment were detected in some tick samples [21]. Rhipicephalus is known to associate with many viral pathogens, such as Thogoto virus (Orthomyxoviridae, genus Thogoto), Wad Medani virus (Reoviridae, genus Orbivirus), CCHFV and NSDV (Bunyaviridae, genus Nairovirus), Kismayo virus and Chim virus (Bunyaviridae, genus Phlebovirus), and Kandam virus (Flaviviridae, genus Flavivirus), all of which can cause disease in livestock and humans [1]. Rhipicephalus is the one of the most common tick genus in Menglian and generally throughout southwest China [22, 23], which preferably feeds on livestock and wild animals. The Rhipicephalus microplus, R. haemaphysaloides and R. sanguineus are very common tick species in this region.

In this study, Ion-torrent sequencing was used to investigate the presence of viruses in Rhipicephalus spp. ticks collected from the field in the Menglian district of Yunnan, China. The viral communities from three tick pools from two collection sites were analyzed and compared, and numerous virus-related sequences were identified.

Methods

Sample collection and taxonomy identification

A total of 387 ticks were collected in the Menglian district of Yunnan, China in 2011 and 2013 (Fig. 1). One-hundred and twenty-seven of them were collected from Nayun (Latitude: 22.3, Longitude: 99.5, Altitude: 1143 m) in May, 2011 (NY-11), 150 from Nayun in May, 2013 (NY-13), and 110 from Mengma (Latitude: 22.8, Longitude: 100.9, Altitude: 1301 m) in May, 2013 (MM-13). No specific permissions for tick collection were required in these locations.

thumbnail
Fig 1. Tick species and sample collection map.

(a) In the Menglian district, tick pools NY-11 and NY-13 were collected from Nayun, while MM-13 was collected from Mengma (b) A tick representing the Rhipicephalus genus from this region.Ticks were collected by the Menglian Center for Disease Control and Prevention (China) using a standard dragging method.

https://doi.org/10.1371/journal.pone.0121609.g001

Ticks were stored in tubes and transported to the laboratory immediately. Morphological and molecular identification were used for tick taxonomic classification. The primers ITS-1 and ITS-2 (S1 Table), which were used for molecular identification, target the internal transcribed spacers of ribosomal DNA [24]. From each group, we selected 20 adult unfed Rhipicephalus spp. ticks as a pool for Ion-torrent sequencing (Fig. 1 (B)). The remaining Rhipicephalus spp. ticks in each group were stored at −80°C. To prevent contamination, the ticks used for Ion-torrent sequencing were surface-sterilized with 3% hydrogen peroxide, followed by 95% ethanol and 1M sodium hypochlorite. Ticks treated in this manner were washed and preserved in PBS before viral nucleic acids were extracted.

Viral nucleic acid extraction

Each tick pool was homogenized through Tissuelyser II (Qiagen, Germany), 30 cycles/sec with 1 ml of phosphate-buffered saline and 5 mm stainless steel beads, and then then centrifuged at 5,000 × g for 10 min. The supernatant was transferred to a fresh tube and then centrifuged at 13,000 × g for 15 min. The supernatant was filtered through a 0.45 μM Millex filter (Millipore, USA) to remove eukaryotic and bacterial cell-sized particles. Total viral nucleic acids (DNA and RNA) were extracted from a 140-μl sample (from the above step) using a QIAamp Viral RNA Mini kit (Qiagen, Germany) according to the manufacturer’s instructions. Nucleic acids were eluted into 40 μl of AVE buffer [13].

cDNA library preparation for Ion-torrent sequencing

First-strand cDNA synthesis was performed with a SuperScript III First-Strand System Kit (Invitrogen, USA). An 8 μl volume of purified viral nucleic acids from each tick pool was mixed with 1 μl of 10 mM deoxynucleoside triphosphate and 1 μl (50 ng/μl) of Brs primer (a random primer with the tag sequence at the 5’ end) (S1 Table), after which the solution was denatured at 65°C for 5 min, and placed on ice for 1 min. A 10 μl aliquot of the cDNA synthesis mix containing 2 μl of 10 × RT buffer, 4 μl of 25 mM MgCl2, 2 μl of 0.1 M DTT, 1 μl of RNaseOUT(40 U/μl) and 1 μl of SuperScript III RT (200 U/μl) was added to each nucleic acid–primer mixture. The reaction mixture was incubated at 25°C for 10 min and 50°C for 50 min, followed by enzyme inactivation at 85°C for 5 min and chilling on ice for 1 min. For the second-strand cDNA synthesis, 0.5 μl (20 pmol) of Brs primer, 2.5 μl of 10 × Klenow fragment buffer, and 2 μl of Klenow fragment (3.5 U/ μl, Takara, Japan) were added. The reaction mixture was incubated at 25°C for 10 min and 37°C for 60 min, followed by enzyme inactivation at 75°C for 10 min [13].

PCR assays were performed using a primer identical to the tag sequence of the Bra primer and using Phusion High-Fidelity DNA polymerase (New England Biolabs, USA), and lower cycles (≤20 cycles) were used to reducing amplification bias. Next, the products were purified to acquire three cDNA libraries. Each cDNA library was loaded onto one single Ion-Torrent 318 chip, and Ion-Torrent sequencing (Life Technologies, USA) was performed by the Beijing Institute of Genomics at the Chinese Academy of Sciences.

Bioinformatics

Raw data were filtered into short (<100 bp) and long (>350 bp) reads, low quality scores (<20), exact duplicate reads and trimmed primer sequences by PRINSEQ 0.20.4 (http://edwards.sdsu.edu/cgi-bin/prinseq/prinseq.cgi) and Mothur v.1.33.0 (http://www.mothur.org/). The metagenomic reads were subjected to a basic local alignment search tool (BLASTx) comparison against the National Center for Biotechnology Information (NCBI, http://www.ncbi.nlm.nih.gov/) non-redundant database (nr) and its viral database by using an E-value cutoff of 1e-04. The taxonomy of the viral community was analyzed and compared using the MEGAN5 analysis toolset (http://ab.inf.uni-tuebingen.de/software/megan/). The reads that hit viral database sequences were assembled by a metagenomic de novo assembly for the Ion-Torrent model of SeqMan NGen 11.2.1 (DNASTAR, http://www.dnastar.com/), and the contiguous sequences (contigs) obtained were subject to BLAST against GenBank using BLASTn and BLASTx.

PCR screening

cDNA from tick pools NY-11, NY-13 and MM-13 were amplified and then subjected to PCR with specific primers based on the sequence-selected contigs from the tick viromes (S1 Table). The PCR products were sequenced to verify the accuracy of the assembled contigs.

Phylogenetic studies

The nucleotide sequences and translated amino acid sequences of contigs with high similarity to known viral nucleic acids and proteins in GenBank were used for phylogenetic analysis. Alignments were performed by ClustalW and phylogenetic trees were constructed by the Maximum Likelihood method with a bootstrap of 1,000 replicates through MEGA 6.06 (http://www.megasoftware.net/megamacBeta.php). Gaps were regarded as a pairwise deletion unless specifically noted.

Results

Taxonomic identification of ticks

Each tick collected was identified to genus level using a stereomicroscope. Through morphological identification, 80 ticks in NY-11, 135 in NY-13, and 97 in MM-13 groups belonged to the genus Rhipicephalus. Molecular identification indicated the presence of multiple species in each group; these included R. microplus, R. haemaphysaloides, and R. sanguineus.

Ion-torrent sequencing reads

The sequencing run resulted in 4,038,018, 3,835,222, and 4,820,504 reads in tick pools NY-11, NY-13, and MM-13, respectively. After filtering and trimming, there were 3,641,088 reads with a 197-bp average length for NY-11, 3,106,733 reads with a 179-bp average length for NY-13, and 3,871,851 reads with a 180-bp average length for MM-13.

Viral-related reads in Rhipicephalus spp. ticks

Based on the most significant BLASTx similarities (e-value<10−6), about 12.2% (445,818/3,641,088) of reads from NY-11, 3.0% (92,715/3,106,733) of reads from NY-13, and 10.3% (399,360/3,871,851) of reads from MM-13 were corresponded to known viral sequences.The reads identified from the three tick viromes were classified into bacteria, invertebrates, vertebrates, plants, algae, and protozoan viruses (Fig. 2, S2 Table). In NY-11 and NY-13, the majority of the viral reads classified were vertebrate-human-invertebrate viruses (24% and 43%, respectively), followed by bacterial viruses (11% and 12%, respectively). However, no reads corresponded to vertebrate-human-invertebrate viruses or vertebrate–human viruses in MM-13. In MM-13, most viral-related reads belonged to bacterial viruses (57%). The viral category distribution of NY-11 was similar to NY-13.

thumbnail
Fig 2. Taxonomic classification overview of metagenomic sequences from the three tick viromes.

(a) Classified by viral host. Breakdown of the viral sequences into seven categories: vertebrate-human-invertebrate, vertebrate-human, vertebrates, invertebrates, plants, algae, and protozoa viruses. (b) Classified by viral family. Viral metagenomic sequences were compared with the viral protein database in GenBank. Comparisons were performed with the BLASTx algorithm using an E-value cutoff 1e-04. The MEGAN5 program (http://ab.inf.uni-tuebingen.de/software/megan5/) was used for acceptance of the output of a BLAST enquiry based on the lowest-common-ancestor algorithm taxonomy analysis.

https://doi.org/10.1371/journal.pone.0121609.g002

From the taxonomic classification overview (Fig. 2, S2 Table), the virus-related reads exist in three viromes distributed across 24 virus families. Anelloviridae, Microviridae, Virgaviridae, Bidnaviridae were only found in NY-11, Ascoviridae, Baculovirdae, Polydnaviridae, Caulimoviridae, Retroviridae, Nanoviridae, Rhabdoviridae were only found in NY-13, and Iridoviridae was only found in MM-13. We found Bunyaviridae and Picornaviridae in NY-11 and NY-13 and Circoviridae and Parvoviridae in NY-11 and MM-13. Finally, Myoviridae, Podoviridae, Siphoviridae, Mimiviridae, and Phycodnaviridae were shared across all three viromes.

De novo consensus assembly of three tick virome reads

The following contigs were assembled from the reads: 1144 contigs from NY-11, 516 contigs from NY-13, and 629 contigs from MM-13. Also, 40, 14 and 64 large contigs (≥1000bp) in three viromes were identified (Table 1). Next, the contigs were aligned against the NCBI nucleotide (nt) and viral genome database using BLASTn; contigs lacking similarity to BLASTn were then aligned against the nr and viral protein database using BLASTx.

From the BLASTn results (e-value <10−6, S3 Table), there were 32 contigs from NY-11 with 79–98%, 27 contigs from NY-13 with 74–99%, and 83 contigs from MM-13 with 72–96% identities to known viral genomic sequences. In the BLASTx alignment results (contigs ≥1000bp, e-value<10−6; Table 2), there were 22 contigs from NY-11 with 36–66%, 8 contigs from NY-13 with 39–53%, and 13 contigs from MM-13 with 41–93% sequence identities to known viral protein sequences. Most contigs had low similarity values to known viral proteins, indicating that these sequences may represent novel viruses.

thumbnail
Table 2. Large contigs with significant BLASTx similarities to known viruses in three Rhipicephalus spp. tick samples.

https://doi.org/10.1371/journal.pone.0121609.t002

Animal and human viruses identified in ticks

Contigs from three tick viromes were similar to those from three families of animal viruses, namely, Bunyaviridae, Anelloviridae, and Rhabdoviridae.

Nairovirus (Bunyaviridae family).

The Nairovirus genus contains the following seven serogroups: Crimean-Congo hemorrhagic fever, Dera Ghazi Khan, Hughes, Nairobi sheep disease, Qalyub, Sakhalin, and Thiafora. All of them are tick-borne and contain some of the most important tick-borne viruses [25]. CCHFV, a representative member of the Nairovirus genus, has one of the widest geographical distributions of medically important arboviruses and causes severe hemorrhagic fever syndrome in humans with a mortality rate ranging from 30–50% [3]. NSDV, another important virus, is highly pathogenic to sheep and goats and can cause disease in humans [26].

Nairovirus has three separate negative-stranded RNA (L, M, and S) which encoding RNA polymerase, pre-glycoprotein and nucleoprotein, respectively [27]. Numerous reads and contigs from NY-11 and NY-13 viromes were found to have statistically significant (e-value < 10−6) relationships with Nairovirus (Table 2, S3 Table). Contig 326 (from NY-11, Genbank Accession Number: KP141755) and contig 3 (from NY-13, Genbank Accession Number: KP141756) covered 99.9% of the open reading frame (ORF) of the Nairovirus nucleoprotein (40.66% identity). The contigs corresponding to the Nairovirus L protein cover most of the L protein ORF. Additionally, the cross-blastn results revealed overlapping nucleic acid regions between NY-11 and NY-13 contigs (related to Nairovirus) that had almost the same sequences (99.8% identity). The specific primers (S1 Table) based on known contigs related to Nairovirus L protein were used to fill the gaps between the contigs, thereby generating a consensus sequence (11435 nt) with 90% coverage of the Nairovirus L protein.

The tick nairovirus-related nucleoprotein amino acid sequence was used to determine phylogenetic relationships. The result showed it clustered with the Nairovirus genus, but was distantly related to the known nairoviruses (Fig. 3), suggesting that the sequence may represent a novel virus belonging to the Nairovirus genus. This virus was named as Nayun tick nairovirus (NTNV).

thumbnail
Fig 3. Phylogenetic analysis of nairovirus-related sequences from ticks.

(a) Schematic representation of the S segment encoded nucleoprotein (∼480 amino acids) of Nairovirus. The orange and green bars represent the contig sequences from NY-11 and NY-13. (b) Phylogenetic analysis of the selected contigs related to Nairovirus is based on alignments with 480 amino acids of the nucleoprotein. All five Bunyaviridae genera are indicated on the right of the tree. NP denotes nucleoprotein.

https://doi.org/10.1371/journal.pone.0121609.g003

Thetatorquevirus (Anelloviridae family).

Thetatorquevirus, a recently discovered genus of the Anelloviridae family, is a small, non-enveloped DNA virus with a circular single-stranded DNA genome (∼2–3 kb) containing three or four overlapping ORFs (ORF1, ORF2 and ORF3) [28]. The known hosts for anelloviruses include humans, non-human primates and domestic animals. Viruses belonging to this family are usually identified in blood [29], and ticks may obtain viruses from viremic hosts during blood feeding. The top BLASTx hit for NY-11 contig240 (Genbank Accession Number: KP141758) had an amino acid identity of 43% and covered 85% of ORF1 of the Torque teno canis virus (TTcV) (Table 2). Phylogenetic analysis based on the amino acid sequence encoded by ORF1 and the topology of the tree indicated that the contig sequence related to Anelloviridae in sample NY-11(Nayun tick torquevirus, NTTV) is a novel virus cluster with TTcV (Fig. 4).

thumbnail
Fig 4. Phylogenetic analysis of tick thetatorquevirus-related sequences.

(a) Schematic representation of three ORFs for Thetatorquevirus. The orange bar represents the contig sequence from NY-11. (b) Phylogenetic analysis of the selected contig related to Thetatorquevirus is based on an alignment with the predicted 440 amino acid sequence of ORF1. Thetatorquevirus and three other genera from the Anelloviridae family are indicated on the right of the tree.

https://doi.org/10.1371/journal.pone.0121609.g004

Rhabdovirus (Rhabdoviridae family).

The family Rhabdoviridae family presents in vertebrates, invertebrates and plants. The viral genome is a negative-sense single-stranded RNA containing at least five ORFs that encode its structural protein (N, P, M, G and L). Currently, Rhabdoviridae consist of nine named genera [30], and many rhabdoviruses still await taxonomic classification. Rabies virus (genus Lyssavirus) is the best-known member of this family which can cause lethal disease of human. Many viruses from genus Vesiculovirus are typical arboviruses. There were many reports about rhabdovirus from ticks. Such as Isfahan virus which belongs to the genus Vesiculovirus has been isolated from Hyalomma asiaticm ticks in Turkmenia [31]. And the other rhabdoviruses isolated from ticks are currently unassigned. Such as Kolente virus which isolated from Amblyomma variegatum ticks in the Republic of Guinea [32], and Long island tick rhabdovirus from Amblyomma americanum ticks in New York State of American [30]. In sample NY-13, three large contigs (contig 2, 5 and 7) shared amino acid sequence similarity levels of 40–53% with the large polymerase protein (L) of rhabdovirus which belongs to the genus Vesiculovirus. A partial sequence from the large polymerase protein (L) region (Genbank Accession Number: KP141757) was used for determining phylogenetic relationships. However the phylogenetic tree revealed that tick rhabdovirus from NY-13 represents a distinct and divergent lineage which shows no clear relationship to Vesiculovirus and any other named genus (Fig. 5). This tick unclassified rhabdovirus (Nayun tick rhabdovirus, NTRV) represents a novel species in the family Rhabdoviridae.

thumbnail
Fig 5. Phylogenetic analysis of tick rhabdovirus-related sequences.

(a) Schematic representation of the large polymerase protein of rhabdovirus. The green bar represents the contig sequence from NY-13. (b) Phylogenetic analysis of the selected contig related to rhabdovirus is based on an alignment with an 889 amino acid sequence from the large polymerase protein region of rhabdovirus. All nine Rhabdoviridae genera are indicated on the right of the tree.

https://doi.org/10.1371/journal.pone.0121609.g005

Plant viruses

Tobamovirus belongs to the family Virgaviridae, which comprises 29 definitive and 6 unclassified species. Tobamovirus can cause disease in a wide range of hosts, including tobacco, tomato, pepper, orchid, cucumber, melon, bean, and cruciferous plants, resulting in serious economic losses in both field and greenhouse-grown crops. Tobamovirus contains positive-sense single-stranded RNA encoding at least four proteins. Tobacco Mosaic Virus (TMV) is the best characterized member of this family, with a genome size of about 6.4 kb in length [33]. We identified 26 contigs in NY-11 sharing high nucleotide sequence similarity scores with Tobamovirus. All of them relating to TMV had 89–98% sequence identity scores and comprised ∼52% coverage of the TMV genome.

Phages

The three tick viromes contained a large diversity of phages, including members from Myoviridae, Podoviridae and Siphoviridae as well as unclassified phages (Fig. 2). In the tick virome of MM-13 particularly, all of the contigs related to these phages (Table 2, S3 Table). Most of the phage sequences found in the ticks only shared amino acid identity with known phages. However, there were 31 contigs with 82–96%, 31 contigs with 74–91%, 13 contigs with 82–96%, and 5 contigs with 84–93% nucleotide sequence identity scores shared with the Staphylococcus phage known as Twort, Salmonella phage FSL SP-031, Vibrio phage pYD38-A, and Staphylococcus phage EW (S3 Table). These findings indicated that these closely related phages were present in the MM-13 tick virome.

The phages identified in the tick virome may have originated from the bacterial flora of the tick or that of the hosts they had fed upon. Staphylococcus (the host for the Staphylococcus phages Twort and EW) and Salmonella (the host for the phage FSL SP-031) bacteria have both been reported in ticks previously [34, 35].

Other viruses

Two large contigs (contigs 16 and 839) in NY-11 share amino acid sequence similarities with the major capsid protein of unclassified ssDNA viruses denoted as Dragonfly-associated microphage 1. This microphage is a single-stranded, circular DNA virus with a 4.5-kb genome encoding five ORFs (replication-associated protein, major capsid protein, two hypothetical proteins and a putative DNA pilot protein). Furthermore, seven contigs in NY-11 shared amino acid sequence similarities with unclassified ssDNA viruses including the fungal virus Sclerotinia sclerotiorum hypovirulence-associated DNA virus 1 (SSHSDV-1) [36]. SSHSDV-1, which infects the fungus Sclerotinia sclerotiorum, is a single-stranded DNA virus with a circular genome (∼ 2.2 kb) containing two major ORFs, a 312 amino acid coat protein and a replication-associated protein of 324 amino acids. This virus is the only known example of a DNA virus that infects a fungus. In seven contigs, five of them (contig10, 130, 454, 817, and 913) were related to the coat protein, and two of them (contig637 and 846) to the replication-associated protein of SSHSDV-1.

Discussion

Rhipicephalus spp. tick is well suited as a vector of zoonotic disease because it feeds on a wide assortment of animal hosts in the wild, as well as humans. Moreover, it is a vector for many pathogenic agents, such as CCHFV, NSDV, Chim virus, Thogoto virus, and Kandam virus [1, 37, 38]. Furthermore, the geographic distribution of Rhipicephalus spp. is extensive in southwest China. Therefore, it is important to use highly sensitivity methods to identify or monitor medically important viruses in Rhipicephalus spp. The present study is the first where NGS was used to explore the variety of viral communities that exist in Rhipicephalus spp. ticks in China. High-throughput sequencing was performed by the Ion-torrent technique (Invitrogen). It can result in higher speed, lower cost, and smaller instrument size. Compared to the Hiseq 2000 technique, the sequencing quality of Ion Torrent is more stable, has a higher map rate, and the GC depth distribution is better [39]. The Ion 318 chip we used can get > 1 GB data in 2 hours, and the maximum reads length is ∼ 400 bp.

Through BLASTx and BLASTn analyses, a large number of viral sequences were found in ticks, some of which had a close relationship with known viral sequences, while some additional sequences potentially corresponded to novel viruses.

By comparing the three tick viromes, the most abundant sequences were from animal viruses (especially those related to Nairovirus), followed by phages in NY-11 and NY-13. However, the most abundant sequences in MM-13 were phages, followed by invertebrate viruses, while there were only a few sequences corresponding to animal viruses (Fig. 2, S2 Table).

When we compared the virome of MM-13 with those of NY-11 and NY-13, only some bacteriophages were commonly shared. The distance between the tick sampling sites at Nayun and Mengma was about 15–20 km; however, the viral communities differed extensively at these two sites, possibly indicating that ticks in these two regions have separate natural ecological environments. Or this could be a result of individual differences in the pooled samples.

Almost all known nairoviruses are transmitted by ticks, and some of them, such as CCHFV and NSDV, have been isolated from or identified in Rhipicephalus spp.; these viruses can cause serious diseases in humans and livestock [1]. In the present study, a new tick nairovirus was identified in Nayun (NY-11 and NY-13) through deep sequencing, and then confirmed by specific PCR amplification. For the assembled contigs from NY-11 and NY-13, many contigs shared amino acid similarity (∼40%) with the S and L protein of Nairovirus. Additionally, cross-BLASTn showed very high similarity (> 98% nucleic acid identity) scores between the contigs relating to Nairovirus from NY-11 and NY-13. These findings indicated that the same nairovirus-related agent was present in the NY-11 and NY-13 tick samples. The phylogenetic analysis indicated that the NTNV clustered with Nairovirus, but had a distant relationship with known nairoviruses. Unfortunately, we did not obtain the large, assembled contigs corresponding to the Nairovirus M protein (glycoprotein). The lack of recognizable M segment is consistent with the reports about the tick virome in the U.S. [17]. They recovered > 90% of the L and S segments for one nairovirus and two phleboviruses, however the sequences with similarity to Bunyavirdae M segments were unable to identify. This may be due to the complicated secondary structure of M segments which inhibit efficient cDNA synthesis and interfere with amplification. Or because the glycoprotein has greater variation than nucleoprotein and L polymerase, we can’t get the related sequences through the alignment with the known database.

Our analysis uncovered the anellovirus and phages in ticks, which is very similar to what’s found and proposed in mosquito [16]. The anelloviurs related sequences in tick virome were novel, which suggested that the animal hosts the ticks feed on contain uncharacterized anellovirus. Like the phages identified in mosquitoes, the tick viromes contained a large diversity of phages sequences (Fig. 2 and Table 2, S2 Table). It is possible that the tick acquires phages during blood feeding, or it is possible phages originate from the tick.

Additionally all three tick viromes contained some sequences related to the following plant viruses: Caulimoviridae, Nanoviridae, Geminiviridae, Virgaviridae and Sobemovirus (Fig. 2, S2 Table); in the tick virome study in the U.S., they found plant related viruses which belong to genus Sobemovirus [17].

In conclusion, this study revealed the presence of highly novel and diverse viral communities in ticks. This information will provide better understanding of the virome via knowledge about the presence and transmission of disease-causing viruses in ticks under natural conditions.

Supporting Information

S1 Table. List of primers used in this study.

https://doi.org/10.1371/journal.pone.0121609.s001

(DOC)

S2 Table. Reads related viral family and genus.

Table shows the viral genus, family and host information of the reads from three tick viromes.

https://doi.org/10.1371/journal.pone.0121609.s002

(DOCX)

S3 Table. Analysis of the contigs with nucleotide identities (BLASTn, e-value < 10−6) to known viruses.

https://doi.org/10.1371/journal.pone.0121609.s003

(DOC)

Acknowledgments

We thank Dr. Aysen Gargili for helping with taxonomic identification of ticks and Dr. Sandra Junglen for her valuable suggestions and discussions for improving the quality of the manuscript. We also appreciate the excellent assistance from the Menglian Center for Disease Control and Prevention.

Author Contributions

Conceived and designed the experiments: HX TL CZ. Performed the experiments: HX ST ZZ. Analyzed the data: HX CH DZ ZF. Contributed reagents/materials/analysis tools: ST ZZ. Wrote the paper: HX DB. Contributed to the improvement of analysis tools: ZK.

References

  1. 1. Labuda M, Nuttall PA. Tick-borne viruses. Parasitology. 2004; 129: S221–S245. pmid:15938513
  2. 2. HavlÍKovÁ S, LiČKovÁ M, Klempa B. Non-viraemic transmission of tick-borne viruses. Acta Virol. 2013; 57: 123–129. pmid:23600870
  3. 3. Bente DA, Forrester NL, Watts DM, McAuley AJ, Whitehouse CA, Bray M. Crimean-Congo hemorrhagic fever: history, epidemiology, pathogenesis, clinical syndrome and genetic diversity. Antivir Res. 2013; 100: 159–189. pmid:23906741
  4. 4. Holbrook MR. Kyasanur forest disease. Antivir Res. 2012; 96: 353–362. pmid:23110991
  5. 5. Yu XJ, Liang MF, Zhang SY, Liu Y, Li JD, Sun YL, et al. Fever with Thrombocytopenia Associated with a Novel Bunyavirus in China. New Engl J Med. 2011; 364: 1523–1532. pmid:21410387
  6. 6. McMullan LK, Folk SM, Kelly AJ, MacNeil A, Goldsmith CS, Metcalfe MG, et al. A new phlebovirus associated with severe febrile illness in Missouri. New Engl J Med. 2012; 367: 834–841. pmid:22931317
  7. 7. Dantas-Torres F, Chomel BB, Otranto D. Ticks and tick-borne diseases: a One Health perspective. Trends Parasitol. 2012; 28: 437–446. pmid:22902521
  8. 8. de la Fuente J, Estrada-Pena A, Venzal JM, Kocan KM, Sonenshine DE. Overview Ticks as vectors of pathogens that cause disease in humans and animals. Front Biosci. 2008; 1: 9.
  9. 9. Savage HM, Godsey MS Jr, Lambert A, Panella NA, Burkhalter KL, Harmon JR, et al. First detection of heartland virus (Bunyaviridae: Phlebovirus) from field collected arthropods. Am J Trop Med Hyg. 2013; 89: 445–452. pmid:23878186
  10. 10. Bzhalava D, Dillner J. Bioinformatics for Viral Metagenomics. J Data Mining Genomics Proteomics. 2013; 4: 134.
  11. 11. Radford AD, Chapman D, Dixon L, Chantrey J, Darby AC, Hall N. Application of next-generation sequencing technologies in virology. J Gen Virol. 2012; 93: 1853–1868. pmid:22647373
  12. 12. Dacheux L, Cervantes-Gonzalez M, Guigon G, Thiberge JM, Vandenbogaert M, Maufrais C, et al. A Preliminary Study of Viral Metagenomics of French Bat Species in Contact with Humans: Identification of New Mammalian Viruses. PLoS ONE. 2014; 9: e87194. pmid:24489870
  13. 13. Ge X, Li Y, Yang X, Zhang H, Zhou P, Zhang Y, et al. Metagenomic analysis of viruses from bat fecal samples reveals many novel viruses in insectivorous bats in China. J Virol. 2012; 86: 4620–4630. pmid:22345464
  14. 14. Steffen MM, Li Z, Effler TC, Hauser LJ, Boyer GL, Wilhelm SW. Comparative Metagenomics of Toxic Freshwater Cyanobacteria Bloom Communities on Two Continents. PLoS ONE. 2012; 7: e44002. pmid:22952848
  15. 15. Whon TW, Kim MS, Roh SW, Shin NR, Lee HW, Bae JW. Metagenomic characterization of airborne viral DNA diversity in the near-surface atmosphere. J Virol. 2012; 86: 8221–8231. pmid:22623790
  16. 16. Ng TF, Willner DL, Lim YW, Schmieder R, Chau B, Nilsson C, et al. Broad Surveys of DNA Viral Diversity Obtained through Viral Metagenomics of Mosquitoes. PLoS ONE. 2011; 6: e20579. pmid:21674005
  17. 17. Tokarz R, Williams SH, Sameroff S, Sanchez Leon M, Jain K, Lipkin WI. Virome analysis of Amblyomma americanum, Dermacentor variabilis, and Ixodes scapularis ticks reveals novel highly divergent vertebrate and invertebrate viruses. J Virol. 2014; 88: 11480–11492. pmid:25056893
  18. 18. Wang J, Zhang H, Fu S, Wang H, Ni D, Nasci R, et al. Isolation of Kyasanur Forest Disease Virus from Febrile Patient, Yunnan, China. Emerg Infect Dis. 2009; 15: 326–328. pmid:19193286
  19. 19. Wu XB, Na RH, Wei SS, Zhu JS, Peng HJ. Distribution of tick-borne diseases in China. Parasite Vector. 2013, 6: 119. pmid:23617899
  20. 20. Gao X, Nasci R, Liang G. The Neglected Arboviral Infections in Mainland China. PLoS Negl Trop Dis. 2010; 4: e624. pmid:20436960
  21. 21. Xia H, Li P, Yang J, Pan L, Zhao J, Wang Z, et al. Epidemiological survey of Crimean-Congo hemorrhagic fever virus in Yunnan, China, 2008. Int J Infect Dis. 2011; 15: e459–463. pmid:21546303
  22. 22. Gong ZD, Zi DY, Feng XG. Composition and distribute of ticks in the Hengduan Mountain region of western Yunnan, China [In Chinese]. China J Pest Co. 2001; 2: 3.
  23. 23. Chen Z, Yang XJ, Yang XH, Liu JZ. Geographical Distribution and Fauna of Chinese Ticks [in Chinese]. Sichuan J Zool. 2008; 27: 4.
  24. 24. Chitimia L, Lin RQ, Cosoroaba I, Braila P, Song HQ, Zhu XQ. Molecular characterization of hard and soft ticks from Romania by sequences of the internal transcribed spacers of ribosomal DNA. Parasitol Res. 2009, 105: 907–911. pmid:19462182
  25. 25. Honig JE, Osborne JC, Nichol ST. The high genetic variation of viruses of the genus Nairovirus reflects the diversity of their predominant tick hosts. Virology. 2004; 318: 10–16. pmid:14972529
  26. 26. Marczinke B. Nairobi Sheep Disease Virus, an Important Tick-Borne Pathogen of Sheep and Goats in Africa, Is Also Present in Asia. Virology. 2002; 303: 146–151. pmid:12482666
  27. 27. Lasecka L, Baron M. The molecular biology of nairoviruses, an emerging group of tick-borne arboviruses. Arch Virol. 2014; 159: 1249–1265. pmid:24327094
  28. 28. de Villiers EM, Borkosky SS, Kimmel R, Gunst K, Fei JW. The diversity of torque teno viruses: in vitro replication leads to the formation of additional replication-competent subviral molecules. J Virol. 2011; 85: 7284–7295. pmid:21593173
  29. 29. Biagini P, Pierre G, Cantaloube JF, Attoui H, de Micco P, de Lamballerie X. Distribution and genetic analysis of TTV and TTMV major phylogenetic groups in French blood donors. J Med Virol. 2006; 78: 298–304. pmid:16372292
  30. 30. Tokarz R, Sameroff S, Leon MS, Jain K, Lipkin WI. Genome characterization of Long Island tick rhabdovirus, a new virus identified in Amblyomma americanum ticks. Virol J. 2014; 11: 26. pmid:24517260
  31. 31. Karabatsos N. International Catalogue of Arboviruses Including Certain Other Viruses of Vertebrates, Third edn. 1985. San Antonio, Texas, AmericanSociety of Tropical Medicine and Hygiene. pmid:646031
  32. 32. Ghedin E, Rogers MB, Widen SG, Guzman H, Travassos da Rosa AP, Wood TG, et al. Kolente virus, a rhabdovirus species isolated from ticks and bats in the Republic of Guinea. J Gen Virol. 2013; 12: 2609–2615. pmid:24062532
  33. 33. Gibbs A. Tobamovirus Classification. In: Van Regenmortel M.H.V., Fraenkel-Conrat Heinz, editors. The Plant Viruses. Springer US. 1986. pp. 167–180.
  34. 34. Nersesov VA, Beridze LP, Giorgadze TS, Manvelian D. Spontaneous infection of ixodes ticks with salmonella. Med Parazitol. 2003; 3: 30–31. pmid:14564841
  35. 35. Andreotti R, Pérez de León A, Dowd S, Guerrero F, Bendele K, Scoles GA. Assessment of bacterial diversity in the cattle tick Rhipicephalus (Boophilus) microplus through tag-encoded pyrosequencing. BMC Microbiol. 2011; 11: 6. pmid:21211038
  36. 36. Jiang D, Fu Y, Guoqing L, Ghabrial SA. Viruses of the Plant Pathogenic Fungus Sclerotinia sclerotiorum. In: Said A. G., editor. Advances in Virus Research. Academic Press; 2013. pp. 215–248. https://doi.org/10.1016/B978-0-12-394315-6.00008-8 pmid:23498908
  37. 37. Reuben Kaufman W, Nuttall PA. Rhipicephalus appendiculatus (Acari: Ixodidae): dynamics of Thogoto virus infection in female ticks during feeding on guinea pigs. Exp Parasitol. 2003; 104: 20–25. pmid:12932755
  38. 38. Lwande OW, Lutomiah J, Obanda V, Gakuya F, Mutisya J, Mulwa F, et al. Isolation of tick and mosquito-borne arboviruses from ticks sampled from livestock and wild animal hosts in Ijara District, Kenya. Vector Borne Zoonotic Dis. 2013; 13: 637–642. pmid:23805790
  39. 39. Liu L, Li Y, Li S, Hu N, He Y, Pong R, et al. Comparison of next-generation sequencing systems. J Biomed Biotechnol. 2012; 50: 2881–2887.