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

RPS8—a New Informative DNA Marker for Phylogeny of Babesia and Theileria Parasites in China

  • Zhan-Cheng Tian ,

    liuguangyuan2002@sina.com (GYL); tianzhancheng@caas.cn (ZCT)

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Guang-Yuan Liu ,

    liuguangyuan2002@sina.com (GYL); tianzhancheng@caas.cn (ZCT)

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Hong Yin,

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Jian-Xun Luo,

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Gui-Quan Guan,

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Jin Luo,

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Jun-Ren Xie,

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Hui Shen,

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Mei-Yuan Tian,

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Jin-feng Zheng,

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Xiao-song Yuan,

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

  • Fang-fang Wang

    Affiliation State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences, Lanzhou, China

Abstract

Piroplasmosis is a serious debilitating and sometimes fatal disease. Phylogenetic relationships within piroplasmida are complex and remain unclear. We compared the intron–exon structure and DNA sequences of the RPS8 gene from Babesia and Theileria spp. isolates in China. Similar to 18S rDNA, the 40S ribosomal protein S8 gene, RPS8, including both coding and non-coding regions is a useful and novel genetic marker for defining species boundaries and for inferring phylogenies because it tends to have little intra-specific variation but considerable inter-specific difference. However, more samples are needed to verify the usefulness of the RPS8 (coding and non-coding regions) gene as a marker for the phylogenetic position and detection of most Babesia and Theileria species, particularly for some closely related species.

Introduction

The piroplasms, comprising mainly the genera Babesia and Theileria, are tick-transmitted protozoa that are highly pathogenic to ruminants, horses, pigs, dogs, cats and cattle, and in some cases, even to humans. In the vertebrate hosts, the infection usually causes fever, anemia and haemoglubinuria, and in severe cases, death [1]. Animals that recover from acute or primary infections naturally remain persistently infected, and act as reservoirs for infecting ticks.

There are some controversial species placements such as T. equi and B. microti [2-6]. In the most thorough phylogenetic examination to date, incongruencies in the phylogenetic evolution with a taxonomically different dataset were displayed [6-9]. However, in the complex phylogenetic relationships between Babesia and Theileria spp., previous studies only relied on 18S rDNA gene [4]. The rDNA genes possess both conserved stems and variable loop regions which provide signals for different levels of phylogenetic inference [10], the 18S rDNA sequences support many piroplasm clades, but, being a slow-evolving marker it may fail to provide enough phylogenetic signal to resolve relationships at the species level such as some closely related Babesia species in China [11]. Internal transcribed spacer 2 (ITS2) may be a more ideal DNA barcode based on the current database for piroplasma [12], however, complex and unpredictable evolutionary behavior of ITS reduces its utility for phylogenetic analysis [11,13]. Information on mitochondrial DNA (mtDNA) from Babesia and Theileria spp., is limited, which precludes its use despite its advantage for use as a molecular marker for lower-level phylogeny [11,14,15]. Genome-wide analysis may aid in determining the taxonomy of species such as B. microti and T. equi but its usefulness as an everyday tool to classify Apicomplexan parasites is limited by availability of complete genome sequences and requirement for a large number of permutations [4-6]. It is therefore highly desirable to test other genes besides the 18S rRNA - to further improve phylogenetic analysis of Babesia and Theileria species.

Structure and sequence signatures in ribosomal RNA and proteins are defining characteristics of the three domains of life and instrumental in constructing the modern phylogeny [16]. Based on the Ribosomal protein S8 and L4, amino acid sequence alignments of orthologous ribosomal proteins found in Bacteria, Archaea, and Eukaryota display an unusual segment or block structure with major evolutionary implications [17]. A set of 50 informative genes that could be analyzed in a broader sampling of Piroplasmida taxa to gain a greater understanding of the evolutionary relationships of the piroplasms were revealed in the comparative genomic analysis of T. equi [5]. Among them, we selected the 40S ribosomal protein S8 (RPS8) gene locus (including coding and non-coding regions) as phylogenetic marker to evaluate the merits and shortcomings of the phylogeny based on the comparative analysis of the RPS8 and 18S rDNA of Babesia and Theileria isolates from China for phylogenetic analysis in this study.

Results and Discussion

RPS8 gene sequence information

The complete sequences of the RPS8 (coding and non-coding regions) gene were obtained from twenty-three isolates representing seven Babesia and six Theileria spp. in China (Table 1). The primers were designed to amplify 561 bp of coding sequence for Babesia species and 573 bp of coding sequence for Theileria species, which are known to be interrupted by an intron in Babesia species and two introns in Theileria species. The size of the introns varied from 146 bp in T. sinensis to 282 bp in Babesia sp. Kashi2 and were specifically located at the conserved position in all the species (Figure 1). The identity of nucleotide sequences of RPS8 between Babesia and Theileria species varied from 63.5% to 82.9% (interspecific variability), whereas within Babesia and Theileria species (intraspecific variability) the identity was approximately 89% to 99.7% respectively (data not shown). The nucleotide sequences of the RPS8 (coding and non-coding regions) genes of Babesia and Theileria species in this study have been deposited in the GenBank database under accession Nos. JN400408 to JN4004028 and Nos. JX402859 to JX402860. The nucleotide sequences of the 18S rDNA genes of several Theileria species in this study have been deposited in the GenBank database under accession Nos. KF559355 to KF559357. The nucleotide sequences of the RPS8 (coding and non-coding regions) genes of B. bovis USA isolate (NW_001820855), B. microti USA isolate (FO082874), T. annulata Ankara strain (NC_011099), T. parva Muguga strain (NC_007345), T. orientalis Shintoku strain (AP011947), T. equi Florida strain (ACOU00000000) were drawn from the GenBank database.

parasiteHostLocationTick vectorIntron locationGenbank Accession No.
RPS818S rDNA
Babesia bovisCattleShaanxianRhipicephalus microplus212-458JN400408AY603398*
B. bovisCattleLushiR. microplus212-458JN400409JX495403*
B. bovisCattleUSAR. microplus212-415NW_001820855*-
B. bigeminaCattleKunmingR. microplus212-487JN400410AY603402*
B. bigeminaCattleLushiR. microplus212-487JN400411JX495402*
B. majorCattleYiliHaemaphysalis punctata212-485JN400412AY603399*
B. ovataCattleLushiH. longicornis212-486JN400413AY603401*
B. ovataCattleWenchuanH. longicornis212-484JN400414AY603403*
B. ovataCattleZhangjiachuanH. longicornis212-487JN400415AY603400*
Babesia sp. Kashi2CattleKashiHyalomma spp.212-493JN400416AY726557*
B. motasisheepLintan H. qinghaiensis 212-493JN400417AY260181*
B. motasisheepNingxianH. longicornis212-491JX402860AY260180*
B. motasisheepTianzhuH. qinghaiensis212-493JX402859DQ159072*
Babesia sp. Xinjiang-2005sheepKashiHyalomma anatolicum212-403JN400418DQ159073*
B. microtihumanUSAIxodes scapularis212-362FO082874* (2706040.2706769)AF231348*
Theileria annulataCattleShanmenxiaH. detritum212-328, 482-512JN400419KF559356
T. annulataCattleXinjiangH. scupense212-328, 482-512JN400420EU073963*
T. annulataCattleNingxiaH. detritum212-328, 482-512JN400428EU083800*
T. annulataCattleAnkaraH. detritum212-328, 482-512NC_011099*-
T. parvaCattleMugugaRhipicephalus appendiculatus212-328, 482-512NC_007345*HQ895968*
T. sergentiCattleLushiH. longicornis212-331, 485-516JN400421AF081137*
T. orientalisCattleShintoku (Japan)H. longicornis212-331, 485-516AP011947*-
T. sinensisYakLintanH. qinghaiensis212-328, 482-510JN400422EU274472*
T. sinensisYakWeiyuanH. qinghaiensis212-328, 482-510JN400423EU277003*
T. sinensisCattleLintaoH. qinghaiensis212-328, 482-510JN400427KF559355
T. luwenshunisheepNingxianH. qinghaiensis212-333, 487-515JN400424JF719833*
T. uilenbergisheepLongdeH. qinghaiensis212-329, 483-511JN400425JF719835*
T. equihorseZhaoyuanDemacntor spp.212-339, 493-528JN400426KF559357
T. equihorseFlorida (USA)D. variabilis ; Amblyomma cajennense ; R. microplus212-339, 493-528ACOU00000000*-

Table 1. The host, location, vector and RPS8 and 18S rDNA gene sequences for Babesia and Theileria species used in this study, * indicated that the sequence were drawn from the database.

CSV
Download CSV
thumbnail
Figure 1. The alignment of RPS8 (coding and non-coding regions) sequences from B. bovis and Theileria species.

B. bovis (GenBank accession no. JN400409); B. microti (GenBank accession no. FO082874); T. annulata (GenBank accession no. NC_011099); B. bovis (GenBank accession no. NW_001820855) were aligned using the ClustalW2 program. The non-coding region is marked with gray shading.

https://doi.org/10.1371/journal.pone.0079860.g001

Interspecific Genetic Distance

The ribosome, with its conserved central role in protein synthesis, has long constituted a prime subject for phylogenetic analysis [17]. The coding sequence of the RPS8 gene displays moderate conservation as a constitutive component of the ribosome. Compared with the moderate sequence identity seen in RPS8 coding regions from Babesia and Theileria species, the RPS8 non-coding regions are highly divergent. Using the Tamura–Nei model of sequence evolution, d values (Genetic Distance) were calculated independently across coding and non-coding region sequences (Data not shown). Across Babesia and Theileria species, d ranges were 0.07903 - 0.47956 (coding) and 0.12511- 1.14484 (non-coding regions). The d-values for non-coding regions are on average 2-fold greater than those for coding regions. Furthermore, the degree of variation in the RPS8 (coding and non-coding regions) dataset was considerable when observing pairwise differences between sequences. Among the eight Babesia and seven Theielria species sequenced, the average p-distances were 0.45, while for the 18S rDNA the average p-distance is 0.083, indicating the RPS8 (coding and non-coding regions) being more variable than the 18S rDNA (Data not shown).

Comparison of the RPS8 (coding and non-coding regions) gene-based phylogenetic trees and the 18S rDNA gene-based phylogenetic trees

Phylogenetic trees based on the 18S rDNA and RPS8 (coding and non-coding regions) genes were constructed by the best-fit model of Bayesian and Maximum likelihood (ML) analysis (Figure 2 and 3). The Bayesian trees were congruent with those obtained under the ML criterion. Nodes receiving ≥85% bootstrap (BP) support in the Maximum likelihood analysis and/or ≥0.98 posterior probability (PP) in the Bayesian analysis were considered strongly statistically supported, and nodes receiving <60 BP and /or <0.90 PP were considered poorly supported, revealing significant levels of genetic diversity. In the RPS8 tree, the Bayesian and ML analyses returned nearly identical topologies, however, for Theileria species, there was less resolution in the ML analysis given that the RPS8 gene is highly variable. The datasets with significant heterogeneity result in the long-branch attraction and poor resolution at relatively deeper nodes when using the ML approach [18].

thumbnail
Figure 2. Inferred phylogenetic relationship among representative Babesia and Theileria species based on 18S rDNA sequences.

The 18S rDNA sequences were analyzed utilizing Bayesian analysis (Bayes) and maximum likelihood (ML), using Toxoplasma gondii as outgroup. The numbers along branches indicate posterior probability (PP) and bootstrap probability (BP) values resulting from different analyses in the order: Bayes/ ML. Highly statistically supported nodes were BP≥85; PP≥0.98; while poorly statistically supported nodes were BP<60; PP<0.90.

https://doi.org/10.1371/journal.pone.0079860.g002

thumbnail
Figure 3. Inferred phylogenetic relationship among representative Babesia and Theileria species based on RPS8 (coding and non-coding regions) gene sequences.

The RPS8 (coding and non-coding regions) gene sequences were analyzed utilizing Bayesian analysis (Bayes) and maximum likelihood (ML), using Toxoplasma gondii as outgroup. The numbers along branches indicate posterior probability (PP) and bootstrap probability (BP) values resulting from different analyses in the order: Bayes/ ML. The accession numbers of the isolates used in the phylogenetic tree were listed in Table 1. High statistically supported nodes had BP≥85; PP≥0.98; while poorly statistically supported nodes had BP<60; PP<0.90.

https://doi.org/10.1371/journal.pone.0079860.g003

Both Babesia and Theileria were polyphyletic with four major clades being defined (Figure 3). Theileria species were clustered within clade I expect T. equi, this clade was strongly supported by both ML and Bayesian analyses (PP > 0.95, BP > 85). Clade II consisted of only three representatives of a single species, T. equi/B. equi (PP = 1.00, BP = 98-100), for which the phylogenetic status is controversial [2,3]. In our analysis, this clade was strongly supported as a separated taxon to other Theileria spp. in the 18S rDNA tree (PP = 1.00), furthermore, there are poorly supported as a sister taxon to other Theileria spp. with ML (BP = 16) analysis of RPS8. However, there was relatively moderate statistical support for T. equi as a sister taxon to other Theileria spp. in the Bayesian analysis (PP = 0.91). Its position relative to Clade I and Clade III comprising the majority of Babesia sp. was unresolved, indicating the unclear evolutionary position of T. equi [4,14]. In both the trees, B. microti was placed at the root of piroplasms, thereby, separating it from the Babesia and Theileria clades (Figure 2, 3), which is consistent with Genome-wide phylogenetic analyses suggesting a new genus for the B. microti group of strains [6].

Overall topology of the two trees showed similar major branching orders, beginning with the outgroup Toxoplasma gondii followed by polyphyletic Theileria and Babesia species. However, phylogenetic reconstructions show that, although it is the least variable, the 18S rDNA tree is more resolved than the RPS8 tree in the nodes of some clades, For example, B. motasi isolates were clearly separated from the branch composed of B. ovata and B. bigemina with statistical support (PP=0.97, BP=64) in the 18S rDNA tree as well as there was moderate statistical support (PP=0.92, BP=78) in the RPS8 tree. These data indicate that 18S rDNA and RPS8 reliably distinguishes the deeper branches among some Babesia species. Within B. ovata isolates, the phylogenetic relationship between the B. ovata Zhangjiachuan isolate and the Wenchuan isolate is reliable in the RPS8 tree (PP=0.99, BP=84) as showed in the 18S rDNA tree (PP=1.00, BP=83). The RPS8 gene seems to be equal to the 18S rDNA in recognizing close lineages among some Babesia and Theileria species.

In the 18S rDNA tree, the phylogenetic relationship between B. bovis and other two Babesia species (Babesia sp. Xinjiang-2005 and Babesia sp. Kashi2) were poorly statistically supported (PP=0.87, BP=37) but the latter two Babesia species are closely related with high statistical support (PP=1.00, BP=95), which was not reconciled with morphological and biological data in previous studies (Figure 2) [19-24]. On the contrary, the phylogeny generated using the RPS8 gene did provide extremely strong support for the close sister-taxon relationship between B. bovis and the other two Babesia species (Babesia sp. Xinjiang-2005 and Babesia sp. Kashi2) (PP=1.00, BP=92) and unreliable sister-taxon relationship between Babesia sp. Xinjiang-2005 and Babesia sp. Kashi2 (PP=0.69, BP=50) (Figure 3). This is probably due to the use of a small dataset and two markers with different modes of evolution. Thus, for more accurate definition of certain Babesia isolates, it may be necessary to incorporate more markers that have differing evolutionary rates.

For Theileria species, clade I in the 18S rDNA tree was composed of two subclades, one subclade comprising T. annulata isolates and T. parva, with the other subclade consisting of T. uilenbergi, T. luwenshuni, T. sinensis isolates and T. sergenti. The phylogenetic status of two subclades resolved with strong statistical support (PP=1.00, BP=96). Furthermore, T. annulata and T. parva were closely related with a strong statistical support (PP=1.00, BP=100). T. luwenshuni formed a separate branch that is distantly related to T. sergenti and T. sinensis with statistical support (PP=0.76, BP=84) (Figure 3). T. uilenbergi formed a separate branch with statistical support (PP=0.89, BP=86). In the RPS8 tree, T. luwenshuni formed a separate branch with strong statistical support (PP=0.98, BP=76) [25,26].

In short, 18S rDNA as a slow-evolving marker more reliably distinguishes deeper branches among some Babesia species than the RPS8 gene. However, the RPS8 gene seems to be equal to the 18S rDNA in recognizing lineages among some closely related Babesia and Theileria species.

Combined analysis

A phylogenetic tree was reconstructed based on the combined dataset of the two markers. The resulting tree is presented in Figure 4. It is mostly reconciled with the phylogenetic trees constructed based on the RPS8 gene and 18S rDNA apart from the phylogenetic relationship of T. equi to other Theileria spp. The clade of T. equi was strongly supported as sister taxon to other Theileria spp. with ML (BP=97) analysis of the combined dataset of the two markers. However, the phylogenetic tree constructed based on the combined dataset of the two markers is not reliable with poor statistical support (Data not shown).

thumbnail
Figure 4. Inferred phylogenetic relationship among representative Babesia and Theileria species based on the combined data of 18S rDNA and RPS8 (coding and non-coding regions) gene sequences.

The combined data of 18S rDNA and RPS8 (coding and non-coding regions) gene sequences were analyzed utilizing maximum likelihood (ML), using Toxoplasma gondii as outgroup. The numbers along branches indicate bootstrap probability (BP) values. The accession numbers of the isolates used in the phylogenetic tree were listed in Table 1. High statistically supported nodes had BP≥85; while poorly statistically supported nodes had BP<60.

https://doi.org/10.1371/journal.pone.0079860.g004

Materials and Methods

Ethics statement

This study was approved by the Animal Ethics Committee of the Lanzhou Veterinary Research Institute, Chinese Academy of Agricultural Sciences. All sheep and calves, were handled in accordance with good animal practices required by the Animal Ethics Procedures and Guidelines of the People's Republic of China.

Animals

All sheep and calves were purchased from a Babesia and Theileria-free area and maintained in an isolated stable. All sheep and calves were splenectomized and treated with antibiotics to promote the healing of wounds one month before the study. During this time, blood films were taken weekly from the ears of the sheep and calves to be examined by Giemsa stain for the presence of haemoprotozoan parasites. The experimental animals were tested by PCR with the universal primers for Babesia and Theileria species based on the 18S rDNA sequences prior to use to ensure that they were free of hemoparasites. After collecting blood containing parasites, the experimental animals were treated with anti-piroplasmosis drug and they were rehabilitated.

Parasite species

The isolates used in this study are listed in Table 1. Babesia bovis (Shanxian and Lushi) [27], B. bigemina (Kunming and Lushi) [28], B. major (Yili) [29], B. ovata (Wenchuan and Lushi and Zhangjiachuan) [30,31,32], Babesia sp. Kashi2 (Kashi) [24], B. motasi (Ningxian and Tianzhu and Lintan) [33,34], Babesia sp Xingjiang-2005 (Kashi) [21], Theileria annulata (Xingjiang and Ningxia and Shanmenxia) [30,35], T. sergenti (Lushi) [36], T. sinensis (Weiyuan and Lintan and Lintao) [37,38], T. uilenbergi (Longde) [26], T. luwenshuni (Ningxian) [26], T. equi (Zhaoyuan) [39]. The reference parasite species are as follows: T. annulata (Ankara strain) [40], T. orientalis (Shintoku strain) [41], T. equi (USDA strain) [5], T. parva (Muguga strain) [42], B. microti (RI isolate) [6], Toxoplasma gondii (ME49 strain) [43], B. bovis (T2Bo strain) [44].

DNA extraction

Sheep and calves were infected intravenously with 15 ml of cryopreserved infected blood stock of these Babesia and Theileria isolates. Daily post-infection rectal temperatures were measured to monitor for disease and blood smears were examined to monitor for presence of piroplasms. When the parasitemia reached more than 5% of whole blood, blood was collected into heparinised tubes. Parasite DNA was isolated using a genomic DNA Purification Kit (Gentra, USA) according to the manufacturer’s instructions. The amount of DNA isolated was assessed spectro-photometrically. Control DNA was isolated from venous blood of uninfected sheep and calves.

PCR, cloning, and DNA sequencing

The partial RPS8 gene was amplified from the genomic DNA extracted from each sample except Babesia microti by conventional PCR using forward primer 5’- ATGGGTATT(A/C)TCA(G/T/)C(A)GT(C/G)GAC(T)AG-3’ and reverse primer 5’- GCGTTTCTTCTTA(G)TCCATACG -3’. The reaction mixture consisted of 10×PCR buffer, 1.5 mM MgCl2, 200 mM each deoxynucleotide triphosphate, 40 pmol each primer, 1.5 U of Taq polymerase (Takara) and approximately 10 ng of DNA, in a final volume of 50 ml. Each PCR consisted of 35 cycles of denaturation at 94°C for 45 s, annealing at 52°C for 60 s, and extension at 72°C for 60 s; an initial denaturation step consisting of incubation at 94°C for 5 min and a final extension step consisting of incubation at 72°C for 10 min was also included. After PCR amplification, the PCR fragment was cloned into the pGEM-T Easy vector (Promega) according to the manufacturer’s recommendations, At least two positive clones from each sample were sequenced by using ABI PRISMTM 377XL DNA sequencer (TaKaRa). All new data has been deposited in GenBank.

Bioinformatic study of genetic distances and phylogenetic analysis

Both ML and Bayesian approaches were used to evaluate each of the individual loci separately. The outgroup taxon Toxoplasma gondii was used for all loci. ML analyses were conducted with PhyML [45] as implemented within Geneious [46], using the best-fit models of nucleotide substitution detected by jModelTest. Support for nodes was estimated by analyzing 1000 bootstrap pseudoreplicates for each locus. Bayesian phylogenetic analysis was conducted on the total dataset using MrBayes v3.1.2 [47]. Akaike information criterion was used to identify the most appropriate model of nucleotide substitution for the Bayesian analysis in the program MrModeltest v2.2, and the best-fit model determined by jModelTest for 18S and RPS8 was GTR with both a proportion of invariable sites (I) and variation among sites (G). The Bayesian analysis was run for 15,000,000 generations with phylogenies sampled every 1,000 generations, and values for the substitution model parameters were not defined a priori, but were treated as unknown variables with uniform priors. Resulting burn-in values (the point at which the model parameters and tree scores reached stationarity) were determined empirically by evaluating likelihood scores. The nucleotide distance matrices were created under a ML correction in MEGA 4.0. The extent of sequence disparity between specimens was calculated by averaging pairwise comparisons of sequence differences across all specimens [48].

The sequence data were also examined using Maximum likelihood analyses in PhyML [45]. The combined data of 18S rDNA and RPS8 (coding and non-coding regions) gene sequences from Toxoplasma gondii was used as an outlier group. The phylogenetic trees were constructed based on the combined data of 18S rDNA and RPS8 (coding and non-coding regions) gene sequences from Theileria and Babesia species determined in our laboratory or obtained from the GenBank database.

The program Genescan for predicting the locations and exon-intron structures of genes in genomic sequences from a variety of organisms (http://genes.mit.edu/GENSCAN.html). The RPS8 deduced amino acid sequence was analyzed with the Expert Protein Analysis System (http://us.expasy.org/). Theoretical molecular weights and isoelectric point (pI) were determined using Peptide-Mass (http://us.expasy.org/tools/peptide-mass.html), and prediction of potential phosphorylation sites was carried out at the NetPhos website. Identification of other motifs was conducted using the Motif scan program (http://hits.isb-sib.ch/cgi-bin/PFSCAN).

Acknowledgments

We thank five anonymous reviewers and editor Kelly A. Brayton for their valuable comments and suggestions. The authors also thank Sueyakant D. Waghela (Texas A&M University) for improving the manuscript.

Author Contributions

Conceived and designed the experiments: ZCT GYL. Performed the experiments: ZCT. Analyzed the data: ZCT. Contributed reagents/materials/analysis tools: GQG JRX HY JXL JL MYT HS XSY FFW JFZ. Wrote the manuscript: ZCT.

References

  1. 1. Kuttler KL (1988) World-wide impact of Babesiosis, M. Ristic, Babesiosis of domestic animals and man. Boca Raton, FL: CRC Press. pp. 1–22. Available: . http://scholar.google.com.hk/scholar?
  2. 2. Mehlhorn H, Shein E (1984) The piroplasms: life cycle and sexual stages. Adv Parasitol 23: 37-103. Available: http://www.ncbi.nlm.nih.gov/pubmed/6442536. PubMed: 6442536.
  3. 3. Schein E, Rehbein G, Voigt WP, Zweygarth E (1981) Babesia equi (Laveran 1901) -Development in horses and in lymphocyte culture. Tropenmed Parasitol 32: 223-227. Available: . http://www.ncbi.nlm.nih.gov/pubmed/7345686. PubMed: 7345686.
  4. 4. Lack JB, Reichard MV, Van Den Bussche RA (2012) Phylogeny and evolution of the Piroplasmida as inferred from 18S rRNA sequences. Int J Parasitol 42: 353-363. doi:https://doi.org/10.1016/j.ijpara.2012.02.005. PubMed: 22429769; Ncbi website. Available: http://www.ncbi.nlm.nih.gov/pubmed/22429769. Accessed: 13 March 2012.
  5. 5. Kappmeyer LS, Thiagarajan M, Herndon DR, Ramsay JD, Caler E et al. (2012) Comparative genomic analysis and phylogenetic position of Theileria equi. BMC Genomics 13: 603. doi:https://doi.org/10.1186/1471-2164-13-603. PubMed: 23137308; Ncbi website. Available: http://www.ncbi.nlm.nih.gov/pubmed/23137308. Accessed: 9 November 2012.
  6. 6. Cornillot E, Hadj-Kaddour K, Dassouli A, Noel B, Ranwez V et al. (2012) Sequencing of the smallest Apicomplexan genome from the human pathogen Babesia microti. Nucleic Acids Res 40: 9102-9114. doi:https://doi.org/10.1093/nar/gks700. PubMed: 22833609; Ncbi website. Available: http://www.ncbi.nlm.nih.gov/pubmed/22833609. Accessed: 24 July 2012.
  7. 7. Criado-Fornelio A, Martinez-Marcos A, Buling-Saraña A, Barba-Carretero JC (2003) Molecular studies on Babesia, Theileria and Hepatozoon in southern Europe part II. Phylogenetic analysis and evolutionary history. Vet Parasitol 114: 173–194. Available: http://www.ncbi.nlm.nih.gov/pubmed/12788253 Accessed: 11 June 2003. doi:https://doi.org/10.1016/S0304-4017(03)00141-9. PubMed: 12788253.
  8. 8. Allsopp MTE, Allsopp BA (2006) Molecular sequence evidence for the reclassification of some Babesia species. Ann NY Acad Sci 1081: 509–517. Available: . http://www.ncbi.nlm.nih.gov/pubmed/17135560 Accessed: 2006 Oct.
  9. 9. Morrison DA (2009) Evolution of the Apicomplexa: where are we now? Trends Parasitol 25: 375–382. doi:https://doi.org/10.1016/j.pt.2009.05.010. PubMed: 17135560. Available: http://www.ncbi.nlm.nih.gov/pubmed/19635681 Accessed: 25 July 2009.
  10. 10. Voigt O, Erpenbeck D, Wörheide G (2008) Molecular evolution of rDNA in early diverging Metazoa: first comparative analysis and phylogenetic application of complete SSU rRNA secondary structures in Porifera. BMC Evol Biol 8: 69. doi:https://doi.org/10.1186/1471-2148-8-69. PubMed: 18304338; Ncbi website. Available: http://www.ncbi.nlm.nih.gov/pubmed/18304338. Accessed: 27 February 2008.
  11. 11. Tian Z, Luo J, Zheng J, Xie J, Shen H et al. (2013a) Phylogenetic analysis of Babesia species in China based on cytochrome b (COB) gene. Infect Genet Evol 13: 36-40. doi:https://doi.org/10.1016/j.meegid.2012.09.001. PubMed: 23041715; Available: http://www.ncbi.nlm.nih.gov/pubmed/23041715. Accessed: 4 October 2012.
  12. 12. Gou H, Guan G, Liu A, Ma M, Xu Z et al. (2012) A DNA barcode for Piroplasmea. Acta Trop 124: 92-97. doi:https://doi.org/10.1016/j.actatropica.2012.07.001. PubMed: 22796448; Ncbi website. Available: http://www.ncbi.nlm.nih.gov/pubmed/22796448. Accessed: 14 July 2012.
  13. 13. Alvarez I, Wendel JF (2003) Ribosomal ITS sequences and plant phylogenetic inference. Mol Phylogenet Evol 29: 417–434. Available: http://www.ncbi.nlm.nih.gov/pubmed/14615184 Accessed: December 2003.
  14. 14. Hikosaka K, Watanabe Y, Tsuji N, Kita K, Kishine H, et al. (2010) Divergence of the mitochondrial genome structure in the apicomplexan parasites, Babesia and Theileria. Mol Biol Evol 27: 1107-1116. doi:https://doi.org/10.1093/molbev/msp320. PubMed: 14615184. Available: . http://www.ncbi.nlm.nih.gov/pubmed/20034997 Accessed:. 24 December 2009.
  15. 15. Tian Z, Liu G, Yin H, Luo J, Guan G et al. (2013b) Cytochrome c oxidase subunit III (COX3) gene, an informative marker for phylogenetic analysis and differentiation of Babesia species in China. Infect Genet Evol 18: 13-17. doi:https://doi.org/10.1016/j.meegid.2013.04.002. PubMed: 23619098; Ncbi website. Available: http://www.ncbi.nlm.nih.gov/pubmed/23619098. Accessed: 22 April 2013.
  16. 16. Chen K, Eargle J, Sarkar K, Gruebele M, Luthey-Schulten Z (2010) Functional role of ribosomal signatures. Biophys J 99: 3930-3940. doi:https://doi.org/10.1016/j.bpj.2010.09.062. PubMed: 21156135; Available: http://www.ncbi.nlm.nih.gov/pubmed/21156135. Accessed: 15 December 2010.
  17. 17. Vishwanath P, Favaretto P, Hartman H, Mohr SC, Smith TF (2004) Ribosomal protein-sequence block structure suggests complex prokaryotic evolution with implications for the origin of eukaryotes. Mol Phylogenet Evol 33: 615-625. Available: . http://www.ncbi.nlm.nih.gov/pubmed/15522791 Accessed:. December 2004
  18. 18. Felsenstein J (2004) Inferring Phylogenies. Sunderland, MA, USA: Sinauer Associates, Inc.. Available: . http://scholar.google.com.hk/scholar. doi:https://doi.org/10.1016/j.ympev.2004.07.003. PubMed: 15522791?
  19. 19. Guan G, Yin H, Luo JX, Lu WS, Zhang QC et al. (2001) Isolation of a large ovine Babesia sp. in Xinjiang, China. Chin J Vet Sci Technol 31: 35–36. (in Chinese) Available: http://epub.cnki.net/kns/brief/result.aspx?dbPrefix=CJFQ Accessed: 22 October 2013.
  20. 20. Guan GQ, Yin H, Luo JX, Lu WS, Zhang QC et al. (2002) Transmission of Babesia sp. to sheep with field-collected Haemaphysalis qinghaiensis. Parasitol Res 88: S22–S24. Available: . http://www.ncbi.nlm.nih.gov/pubmed/12051601 Accessed:. May 2002
  21. 21. Guan G, Ma M, Moreau E, Liu J, Lu B, et al. (2009) A new ovine Babesia species transmitted by Hyalomma anatolicum anatolicum. Exp Parasitol 122: 261-267. doi:https://doi.org/10.1016/j.exppara.2009.05.001. PubMed: 12051601. Available: http://www.ncbi.nlm.nih.gov/pubmed/19460377 Accessed: 19 May 2009.
  22. 22. Luo J, Chen F, Lu W, Guan G, Ma M et al. (2003) Experimental transmission of an unnamed bovine Babesia by Hyalomma spp., Haemaphysalis longicornis and Boophilus microplus. Vet Parasitol 116: 115-124. Available: http://www.ncbi.nlm.nih.gov/pubmed/14519316 Accessed: 8 October 2003. doi:https://doi.org/10.1016/S0304-4017(03)00262-0. PubMed: 14519316.
  23. 23. Luo J, Yin H, Liu Z, Yang D, Guan G et al. (2005a) Molecular phylogenetic studies on an unnamed bovine Babesia sp. based on small subunit ribosomal RNA gene sequences. Vet Parasitol 133: 1-6. Available: http://www.ncbi.nlm.nih.gov/pubmed/15982821 Accessed: 10 October 2005, doi:https://doi.org/10.1016/j.vetpar.2005.02.014. PubMed: 15982821.
  24. 24. Luo J, Yin H, Guan G, Yang D, Liu A et al. (2005b) A comparison of small-subunit ribosomal RNA gene sequences of bovine Babesia species transmitted by Haemaphysalis spp. in China. Parasitol Res 95: 145-149. Available: http://www.ncbi.nlm.nih.gov/pubmed/15609061 Accessed: 18 December 2004, doi:https://doi.org/10.1007/s00436-004-1268-4.
  25. 25. Schnittger L, Yin H, Gubbels MJ, Beyer D, Niemann S et al. (2003) Phylogeny of sheep and goat Theileria and Babesia parasites. Parasitol Res 91: 398–406. Available: http://www.ncbi.nlm.nih.gov/pubmed/14505044 Accessed: 18 September 2003. doi:https://doi.org/10.1007/s00436-003-0979-2. PubMed: 14505044.
  26. 26. Yin H, Schnittger L, Luo J, Seitzer U, Ahmed JS (2007) Ovine theileriosis in China: a new look at an old story. Parasitol Res 101: S191-S195. Available: . http://www.ncbi.nlm.nih.gov/pubmed/17823827 Accessed:. September 2007.
  27. 27. Bai Q, Liu GY, Zhou JY (1991) Isolation of single species of Babesia bovis. Chin J Vet Sci Technol 1: 20-22. (in Chinese. ) Available: . http://epub.cnki.net/kns/brief/result.aspx?dbPrefix=CJFQ Accessed:. 22 October 2013.. doi:https://doi.org/10.1007/s00436-007-0689-2. PubMed: 17823827.
  28. 28. Bai Q, Yin SX, Chen ZH (1987) Isolation of single species of Babesia bigemina. Chin J Vet Sci Technol 9: 25-27. (in Chinese) Available: http://epub.cnki.net/kns/brief/result.aspx?dbPrefix=CJFQ Accessed: 22 October 2013.
  29. 29. Yin H, Lu W, Luo J, Zhang Q, Lu W et al. (1996) Experiments on the transmission of Babesia major and Babesia bigemina by Haemaphysalis punctata. Vet Parasitol 67: 89-98. Available: http://www.ncbi.nlm.nih.gov/pubmed/9011018. doi:https://doi.org/10.1016/S0304-4017(96)01022-9. PubMed: 9011018.
  30. 30. Luo JX, Yin H, Liu GY, Guan GQ, Liu ZJ et al. (2006) Collection and identification of Piroplasma infected to cattle and sheep in China. Chin J Parasitol Parasit Dis 24: S48–S53. (in Chinese) Available: http://epub.cnki.net/kns/brief/result.aspx?dbPrefix=CJFQ Accessed: 22 October.
  31. 31. Bai Q, Liu GY, Zhang L (1990) Discovey and isolation of Babesia ovata in China. Chin J Vet Sci 16: 2-4. (in Chinese) Available: http://epub.cnki.net/kns/brief/result.aspx?dbPrefix=CJFQ Accessed: 22 October 2013.
  32. 32. Bai Q, Liu GY, Han GF, Hui Y (1994) Isolation and complementary transmission of Babesia ovata in Zhangjiachuan of Gansu Province. Chin J Vet Sci Technol 24: 9–10. (in Chinese) Available: http://epub.cnki.net/kns/brief/result.aspx?dbPrefix=CJFQ Accessed: 22 October 2013.
  33. 33. Yin H, Luo JX, Lu WS, Lu WX, Zhang QC et al. (1997) Isolation and Morphology survey of the Babesia motasi and Babesia ovis in China. Chin J Vet Sci Technol 27: 7–9. (in Chinese) Available: http://epub.cnki.net/kns/brief/result.aspx?dbPrefix=CJFQ Accessed: 22 October 2013.
  34. 34. Yin H, Lu W, Luo JX (1997) Babesiosis in China. Trop Anim Health Prod 29: 11–15. Available: http://www.ncbi.nlm.nih.gov/pubmed/9512738. doi:https://doi.org/10.1007/BF02632908.
  35. 35. Luo JX, Lu WS (1997) Cattle theileriosis in China. Trop Anim Health Prod 29: 4–7. Available: http://www.ncbi.nlm.nih.gov/pubmed/9512736. doi:https://doi.org/10.1007/BF02632906.
  36. 36. Liu A, Guan G, Liu Z, Liu J, Leblanc N et al. (2010) Detecting and differentiating Theileria sergenti and Theileria sinensis in cattle and yaks by PCR based on major piroplasm surface protein (MPSP). Exp Parasitol 126: 476-481. doi:https://doi.org/10.1016/j.exppara.2010.05.024. PubMed: 20685208; Ncbi website. Available: http://www.ncbi.nlm.nih.gov/pubmed/20685208. Accessed: 1 June 2010.
  37. 37. Bai Q, Liu GY, Yin H, Zhao QZ, Liu DK et al. (2002) Theileria Sinensis sp nov: A New Species of Bovine Theileria - Classical Taxonomic Studies. Acta Veterinar ia et Zootechnica Sinica 33: 73- 77. (in Chinese) Available: http://epub.cnki.net/kns/brief/result.aspx?dbPrefix=CJFQ Accessed: 22 October 2013.
  38. 38. Yin H, Luo J, Guan G, Gao Y, Lu B et al. (2002) Transmission of an unidentified Theileria species to small ruminants by Haemaphysalis qinghaiensis ticks collected in the field. Parasitol Res 88: S25–S27. Available: http://www.ncbi.nlm.nih.gov/pubmed/12051602 Accessed: May 2002, PubMed: 12051602.
  39. 39. Luo J, Liu GY, Tian ZC, Xie JR, Zhang P et al. (2011) Taxonomic Status Assays of Theileria equi Based on Hsp70 Gene. Sci Agricultura Sinica 44: 3694-3700. (in Chinese) Available: http://epub.cnki.net/kns/brief/result.aspx?dbPrefix=CJFQ Accessed: 22 October 2013.
  40. 40. Hayashida K, Hara Y, Abe T, Yamasaki C, Toyoda A et al. (2012) Comparative genome analysis of three eukaryotic parasites with differing abilities to transform leukocytes reveals key mediators of Theileria-induced leukocyte transformation. mBio. 3: e00204-212. doi:https://doi.org/10.1128/mBio.00204-12. PubMed: 22951932; Available: http://www.ncbi.nlm.nih.gov/pubmed/22951932. Accessed: 4 September 2012.
  41. 41. Pain A, Renauld H, Berriman M, Murphy L, Yeats CA et al. (2005) Genome of the host-cell transforming parasite Theileria annulata compared with T. parva. Science 309: 131-133. Available: http://www.ncbi.nlm.nih.gov/pubmed/15994557 Accessed: 1 July 2005. doi:https://doi.org/10.1126/science.1110418. PubMed: 15994557.
  42. 42. Gardner MJ, Bishop R, Shah T, de Villiers EP, Carlton JM et al. (2005) Genome sequence of Theileria parva, a bovine pathogen that transforms lymphocytes. Science 309: 134-137. Available: http://www.ncbi.nlm.nih.gov/pubmed/15994558 Accessed: 1 July 2005, doi:https://doi.org/10.1126/science.1110439. PubMed: 15994558.
  43. 43. Bontell IL, Hall N, Ashelford KE, Dubey JP, Boyle JP et al. (2009) Whole genome sequencing of a natural recombinant Toxoplasma gondii strain reveals chromosome sorting and local allelic variants. Genome Biol 10: R53. doi:https://doi.org/10.1186/gb-2009-10-5-r53. PubMed: 19457243; Ncbi website. Available: http://www.ncbi.nlm.nih.gov/pubmed/19457243. Accessed: 20 May 2009.
  44. 44. Brayton KA, Lau AO, Herndon DR, Hannick L, Kappmeyer LS et al. (2007) Genome sequence of Babesia bovis and comparative analysis of apicomplexan hemoprotozoa. PLOS Pathog 3: 1401-1413. Available: http://www.ncbi.nlm.nih.gov/pubmed/17953480 Accessed: 19 October 2007, PubMed: 17953480.
  45. 45. Guindon S, Gascuel O (2003) A simple, fast and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst Biol 52: 696–704. Available: http://www.ncbi.nlm.nih.gov/pubmed/14530136 Accessed: October 2003, doi:https://doi.org/10.1080/10635150390235520. PubMed: 14530136.
  46. 46. Drummond AJ, Ashton B, Cheung M, Heled J, Kearse M et al. (2009) Geneious v4.7. Available: http://www.geneious.com.
  47. 47. Huelsenbeck JP, Ronquist F (2001) MrBayes: Bayesian inference of phylogeny. Bioinformatics 17: 754–755. Available: http://www.ncbi.nlm.nih.gov/pubmed/11524383 Accessed: August 2001, doi:https://doi.org/10.1093/bioinformatics/17.8.754. PubMed: 11524383.
  48. 48. Felsenstein J (1981) Evolutionary trees from DNA sequences: a maximum likelihood approach. J Mol Evol 17: 368–376. Available: http://www.ncbi.nlm.nih.gov/pubmed/7288891. doi:https://doi.org/10.1007/BF01734359. PubMed: 7288891.