Skip to main content

Comparative genomic analysis and phylogenetic position of Theileria equi

Abstract

Background

Transmission of arthropod-borne apicomplexan parasites that cause disease and result in death or persistent infection represents a major challenge to global human and animal health. First described in 1901 as Piroplasma equi, this re-emergent apicomplexan parasite was renamed Babesia equi and subsequently Theileria equi, reflecting an uncertain taxonomy. Understanding mechanisms by which apicomplexan parasites evade immune or chemotherapeutic elimination is required for development of effective vaccines or chemotherapeutics. The continued risk of transmission of T. equi from clinically silent, persistently infected equids impedes the goal of returning the U. S. to non-endemic status. Therefore comparative genomic analysis of T. equi was undertaken to: 1) identify genes contributing to immune evasion and persistence in equid hosts, 2) identify genes involved in PBMC infection biology and 3) define the phylogenetic position of T. equi relative to sequenced apicomplexan parasites.

Results

The known immunodominant proteins, EMA1, 2 and 3 were discovered to belong to a ten member gene family with a mean amino acid identity, in pairwise comparisons, of 39%. Importantly, the amino acid diversity of EMAs is distributed throughout the length of the proteins. Eight of the EMA genes were simultaneously transcribed. As the agents that cause bovine theileriosis infect and transform host cell PBMCs, we confirmed that T. equi infects equine PBMCs, however, there is no evidence of host cell transformation. Indeed, a number of genes identified as potential manipulators of the host cell phenotype are absent from the T. equi genome. Comparative genomic analysis of T. equi revealed the phylogenetic positioning relative to seven apicomplexan parasites using deduced amino acid sequences from 150 genes placed it as a sister taxon to Theileria spp.

Conclusions

The EMA family does not fit the paradigm for classical antigenic variation, and we propose a novel model describing the role of the EMA family in persistence. T. equi has lost the putative genes for host cell transformation, or the genes were acquired by T. parva and T. annulata after divergence from T. equi. Our analysis identified 50 genes that will be useful for definitive phylogenetic classification of T. equi and closely related organisms.

Background

Equine piroplasmosis of horses, mules, donkeys and zebras is caused by the tick-borne apicomplexan protozoan parasites Babesia caballi and Theileria equi, transmitted by ixodid ticks such as Dermacentor nitens (B. caballi) and Rhipicephalus microplus (T. equi)[1, 2]. Although endemic in most countries[3], the U. S., until recently, has been considered free of infection. Equine infections in Florida with B. caballi and T. equi were diagnosed between 1961 and 1969 leading to an eradication campaign which lasted twenty-five years and cost twelve million dollars[4]. The re-emergence of T. equi in Florida[4] and Texas[5] raised concern of its further spread within the U. S., and indeed, infected horses have been identified in 12 states[6, 7]. The cause of the 2008 Florida outbreak was due to iatrogenic transmission, but two tick species, Amblyomma cajennense and D. variabilis, were identified as novel vectors in the 2009 Texas outbreak[5]. The re-emergence of this pathogen in the U.S. impacts global movement and health of horses and affects the multi-billion dollar equine industry.

Additional members of the phylum Apicomplexa, important to global human and animal health include the organisms in the genus Plasmodium as well as T. parva and T. annulata, and Babesia bovis causes of malaria, bovine theileriosis and babesiosis, respectively. The phylogenetic position of T. equi has been controversial, and the organism has been renamed several times[8]. Molecular phylogenetic analyses indicate an intermediate position for T. equi between B. bovis and Theileria spp.[9, 10] and is supported by the genomic data presented here which provides the deepest phylogenetic analysis to date. Collective data supports the concept that a new genus placement sister to Theileria may be appropriate for T. equi.

Similar to bovine theileriosis caused by T. annulata, transmission of T. equi to equids eventually results in lysis of erythrocytes and prolonged anemia. Anemia associated with T. parva occurs later during infection and is comparatively and clinically mild[11]. Infection of B- and T-lymphocytes by T. parva and mononuclear phagocytes and B-lymphocytes by T. annulata lead to reversible cell transformation[12]. Infection of peripheral blood mononuclear cells (PBMCs) by T. equi has been reported[8, 13]. However the role of PBMC infection in the pathogenesis of T. equi, unlike T. parva and T. annulata remains unresolved, and PBMC proliferation and/or transformation have not been associated with clinical equine piroplasmosis.

The primary clinical outcome of acute T. equi infection is anemia and the associated erythrolysis is independent of parasite-specific immune responses[14]. Resolution of acute disease is followed by apparent life-long parasite persistence within equids[15]. Persistence is characterized by the continuous presence of 103 to 106 infected peripheral erythrocytes per ml/blood resulting in efficient acquisition and transmission by ticks[16]. A hallmark of pathogens that establish persistent infection and avoid immune elimination is the presence of an immunodominant, variable multigene family responsible for immune evasion, such as VESA1 (Variant Erythrocyte Surface Antigen 1) in B. bovis[17], PfEMP1 (Erythrocyte Membrane Protein 1) in P. falciparum[18] and VSG in T. brucei[19]. An analogous family was not detected in T. equi. A candidate multigene gene family in T. equi encodes Equi Merozoite Antigens (EMAs), which are immunodominant for antibody, however, this family contains just 10 members, and lacks an apparent structural basis for variation. The reemergence and persistence of T. equi in the U. S. prompted this genomic research due to the lack of a vaccine to block infection or clinical disease and the need for additional chemotherapeutics aimed at eliminating persistent infection and tick-borne transmission risk.

Results and discussion

Chromosome arrangement and content

The genomic complement of T. equi contains six molecules, including four chromosomes (Figure1) of ~11.6 Mbp, an apicoplast genome of 47.8 kbp and a mitochondrial genome of 9 kbp. Chromosomes 1 (3,677,484 bp) and 3 (2,338,319 bp) were completely assembled, while chromosome 2 (2,060,349 bp) contains one assembly gap. Chromosome 4 (3,480,987 bp) is fragmented into six pieces with four physical gaps and one assembly gap. The chromosome assemblies agree with the sizes of the chromosomes seen on pulsed field gel electrophoresis, suggesting that there are no large gaps in the genome sequence.

Figure 1
figure 1

Chromosomal map and depiction of synteny between piroplasms. T. equi (Te), T. annulata (Ta), T. parva (Tp) and B. bovis (Bb) chromosomes are shown with the T. equi chromosome represented by the thin black line, and corresponding syntenic regions from other species’ chromosomes shown as color blocks. S hades of red represent chromosomes 1, shades of yellow represent chromosomes 2, shades of green represent chromosomes 3, and shades of blue represent chromosomes 4. To the left of the Te chromosome are indicated genes discussed in the manuscript: purple bars represent locations of ema family members, black bars represent the location of genes used in phylogenetic analysis and brown bars represent the two rRNA operons.

The larger genome size of T. equi as compared to other piroplasms (Theileria and Babesia) (Table1) is also reflected in the number of predicted protein coding genes (5330), ~ 25% greater than found for T. parva, T. annulata and B. bovis. While T. equi contains homologs of genes found only in the two Theileria spp. (366), or B. bovis (137), it has far more unique genes (1985), which account for the increased size of the genome. The four species have 2,482 genes in common. T. equi has two rRNA operons, and 46 tRNA genes.

Table 1 Genome characteristics of Apicomplexans

The apicoplast

The apicoplast is a plastid-like organelle thought to be derived from a secondary endosymbiotic event with green algae[20, 21]. Like plastids, many of the genes for metabolic processes in the apicoplast have migrated to the nuclear genome leaving a remnant genome[22]. In T. equi, the A+T rich (71%) 47.8 kb apicoplast genome is larger than those of other piroplasms, due primarily to expansion of the repertoires of three hypothetical genes (Additional file1). There are 43 unidirectionally encoded CDSs in the T. equi molecule, which includes 11 ribosomal protein coding sequences. Additionally, each of the 20 tRNA and two rRNA genes are present.

T. equi contains 509 nuclear-encoded proteins potentially targeted to the apicoplast as predicted by PlasmoAP[23], ApicoAP[24] and/or by homology to genes for pathways predicted to occur in the apicoplast (Additional file2). Similar to other apicomplexans, T. equi has a complete set of nuclear-encoded enzymes for isoprenoid precursor biosynthesis via the methylerythritol phosphate pathway and these activities are predicted to occur in the apicoplast.

Mitochondrial genome

The mitochondrial (mt) genome is 9001 bp in length, longer than in other piroplasms and shows evidence of gene duplication and rearrangement. The T. equi mt sequence was recently reported and the authors suggest that a duplicated CDS (BEWA_044660 and BEWA_044650) is actually cytochrome c oxidase subunit III (cox3)[25]. Although there is no sequence similarity to known cox3 sequences, T. equi could not conduct respiration without cox3, and these genes are the only candidate for this function. Our findings corroborate the linear structure of the mt genome and the long inverted terminal repeat structure[25].

Metabolism

Energy Production

The predicted metabolic profile of T. equi is similar to other piroplasms: most elements of the core pathways for energy production are present including glycolysis, pentose phosphate pathway and the tricarboxylic acid (TCA) cycle. Glycolysis is fully intact with putative enzymes identified for each stage of the conversion of sugars to pyruvate for the production of energy, making carbon the primary energy source. However, since pyruvate dehydrogenase (EC 1.2.4.1) is missing from the T. equi genome (similar to B. bovis and other Theileria spp.), glycolysis does not seem to be coupled to the TCA cycle. Although the TCA cycle is intact, the lack of pyruvate dehydrogenase suggests that the primary function may be to generate precursors rather than produce energy. For example, succinyl CoA is a critical biosynthetic precursor for the synthesis of protoheme, used in cytochromes and many enzymes. Electron transport pathways in T. equi resemble those reported for P. falciparum[26]. Lack of the full complement of NADH dehydrogenase subunits and ATP synthase peptides makes it questionable as to whether the electron transport pathway is efficient for the generation of ATP from the products of glycolysis.

Small molecule synthesis

Metabolic pathway similarities with other sequenced hemoparasites include the ability to synthesize pyrimidines, limited amino acid biosynthesis, and the lack of a urea cycle. De novo purine biosynthesis is absent, however, unlike other piroplasms, hypoxanthine/guanine phosphoribosyl transferase (BEWA_017710) and adenine phosphoribosyltransferase (BEWA_017730) are present, indicating that purine salvage can occur, similar to P. falciparum. Fatty acid synthesis, heme biosynthesis and the shikimic acid pathway do not occur in the piroplasms, although these pathways are present in P. falciparum[2629].

T. equi encodes dihydrofolate synthase (EC 6.3.2.17, BEWA_029790) and a bifunctional dihydrofolate reductase-thymidylate synthase (EC 1.5.1.3, DHFR-TS, BEWA_008170) and is predicted to carry out a limited folate biosynthetic pathway similar to Theileria spp[27, 28]. Notably, a large number of folate-biopterin transporters were predicted, suggesting that T. equi imports folate which is then modified into other compounds, chiefly nucleotide precursors. Folate biosynthesis predicts sensitivity to the drug pyrimethamine cycloguaryl, which has been observed in previous studies[30], even though T. equi DHFR-TS encodes the S125F mutation that reportedly confers pyrimethamine resistance in B. bovis and certain Plasmodium sp DHFR-TS[31].

Phospholipid metabolism

Phospholipid metabolism in apicomplexans is well-documented, and highlighted by studies in B. bovis showing a markedly greater phospholipid composition in infected erythrocytes compared to uninfected erythrocytes[32]. T. equi has an increased number of choline/ethanolamine kinase genes, even relative to B. bovis, which has a demonstrated increase in phosphatidylcholine relative to uninfected bovine erythrocytes[33]. Although only a partial set of enzymes for synthesis of glycosylphosphatidylinositol (GPI) anchors were detected, T. equi has been demonstrated to incorporate GPI anchors on membrane proteins[34]. Approximately 2% of the proteome (132 proteins; Additional file3) were predicted to contain both required signatures for GPI anchors; however, this prediction should be used with caution as metabolic labeling studies indicate that there are relatively few GPI anchored protein species within infected erythrocytes, with members of the EMA family being the predominantly labeled proteins[34].

Transporter families

TransportDB predicts T. equi to have the most transporters of any hemoparasite genome sequenced to date[2629], with 142 in total (Additional file4). The most profound increase is within the ATP-binding cassette (ABC) superfamily of transporters, with T. equi having 45 members of this family, compared to just 17 in T. parva, and 9 in B. bovis. Comparatively, Plasmodium falciparum has 16 members of the ABC family, most notably ABCB1 (MDR1), a known mediator of chloroquine and mefloquine resistance[35]. Resistance to chemotherapeutics also occurs in T. equi, however specific mechanisms of resistance are unknown. Given the high number of ABC transporter family members in T. equi, including those of the MDR1 type and orthologues of the other known drug transport members ABCC1 and ABCG2, it is reasonable to hypothesize that ABC-mediated transport contributes to chemotherapeutic resistance in T. equi. Another expanded family is the Type II general secretory pathway, which contains 11 transporters, ~three times the number in the other hemoparasite genomes (Additional file4). This pathway moves signal peptide containing proteins across the cell membrane. The abundance of signal peptide-containing proteins without predicted transmembrane domains suggests that T. equi has a large secretome that utilizes this pathway.

As noted earlier, the folate/biopterin family of transporters is increased (at 5), and thus, T. equi may import additional folate to contribute to the one carbon pool. Finally, T. equi has 23 transporter genes in the Major Facilitator Superfamily (MFS) about twice as many as other hemoparasites. The MFS transporters include drug efflux systems, organophosphate:phosphate exchangers and oligosaccharide:H+ symport permeases and are single-polypeptide secondary carriers capable only of transporting small solutes in response to chemiosmotic ion gradients[36].

Chromosomal synteny

Figure1 shows blocks of synteny shared between T. equi and B. bovis, T. parva, or T. annulata. There are relatively few large regions that do not have syntenic matches in the other piroplasm genomes. Analysis of chromosome 1 demonstrated that these regions contain unique proteins (184), without Pfam hits or other functional assignment. The T. parva and T. annulata chromosomes are highly syntenic, and a much more fragmented pattern of synteny is seen between Theileria spp. and B. bovis[29, 37]. Large regions of synteny are observed between the Theileria spp. and T. equi, suggesting more recent shared ancestry than for T. equi and B. bovis, where the blocks of synteny are more fragmented. T. equi chromosome 1 appears to be evolutionarily related to Thelieria spp. chromosomes 1 and 2. The ancestral Theileria chromosome 2 has split to provide elements of T. equi chromosomes 1 and 2. Approximately 2/3 of each of T. equi chromosomes 2 and 3 share synteny with Theileria spp. chromosomes 1 and 4, respectively. The remaining 1/3 of each of these chromosomes contains more of the B. bovis specific lineage sequences or unique sequences. With a few exceptions, T. equi chromosome 4 and is syntenic to both T. parva and T. annulata chromosome 3. Notably, the orthologous gene matches extend to the ends of each of the T. equi chromosomes demonstrating the lack of telomerically located species-specific gene families and repeats seen in other hemoparasites[26, 28].

Paralogous families

TribeMCL placed 2614 proteins in 334 families ranging in size from 2 to 356 members (Additional file5). Many of the families, including the two largest, have little significant sequence identity, for any two proteins, in pairwise comparisons; however, this result did not change when we used more stringent parameters for clustering. Therefore, we analyzed the families for common functional annotation and show that most members of a given family share a common functional domain or feature (Additional file5). Exceptions are family 1 and family 17. A few families have high levels of sequence conservation and readily identifiable functional attributes found through Pfam hits. For example, family 11 contains 42 members with hits to Pfam PF00005, a family of ABC transporters. A survey of T. equi protein families that have at least 20 members revealed that most of these families are comprised of functional attributes similar to those represented in similar scans of B. bovis and Theileria spp. None of the T. equi families appear to contain a family of known immunodominant antigenically variable genes, nor are any of the families telomerically associated. Notable families are presented below.

Family 3

Containing 109 members, this family has ~30% sequence conservation with a repeat gene family in T. annulata (Tar). The Tar genes were reported to be analogous to the Tpr genes of T. parva, however Family 3 does not share significant sequence identity with Tpr[27]. Most members of Family 3 encode a protein of ~50 KDa, with no signal peptide. Although the function of these proteins in Theileria is unknown, it is thought that these genes in T. parva are involved in the generation of diversity[28], with the Tpr repertoires appearing to be isolate specific[38]. Like T. annulata, the T. equi Family 3 genes are distributed throughout all four chromosomes. Expressed Sequence Tags (ESTs) were found for only seven of the Family 3 genes suggesting that only a small number of these genes are expressed at a given time.

Equi merozoite antigen (EMA) family

Family 29 contains 10 members related to the best characterized protein in T. equi, Equi Merozoite Antigen 1 (EMA1). EMA1, a 34 kDa immunodominant protein, is used as the basis of a cELISA diagnostic test for T. equi, as this highly conserved antigen is recognized by sera from infected animals worldwide[39]. EMA1 has a GPI anchor, a putative erythrocyte-binding domain shared with certain hemotoxins, and surface-exposed epitopes[34]. Immunoprecipitation of T. equi proteins by serum from a horse challenged twice at a 2-month interval with T. equi revealed that in addition to EMA1, EMA2 has a GPI anchor and detectable antibody responses were limited to EMA1-3 proteins[34, 40]. Although ema1 was originally described as a single copy gene, genomic analysis has revealed a total of 10 genes in this family. The amino acid identity ranged from 17 to 55% in pairwise comparisons of family members. EMA2 was previously characterized as 50% identical to EMA1, and is also an immunodominant GPI-anchored protein. EMA1 is more highly expressed in blood stages of the parasite relative to EMA2 while EMA2 is more highly expressed in the tick salivary gland during transmission feeding than EMA1[2]. These two proteins, along with a third member of this family (EMA3) have been shown to interact with the erythrocyte cytoskeleton[40, 41]. A single ortholog of these proteins have been identified in T. annulata (TAMS1) and T. parva (mMPSA) but not in other Apicomplexa[42, 43]. At least one EST was found for each ema family member, with the exception of BEWA_028210 and BEWA_047350, demonstrating that transcription in blood stage parasites takes place for most family members.

A core set of 8 ema genes are predicted to encode proteins of ~30 kDa, with the remaining two encoding markedly differently sized proteins, one being truncated (BEWA_034050) and the other (BEWA_047350) being much longer and only having similarity in the C-terminal domain (Additional file6). Interestingly, the truncated gene produces a transcript, based on EST analysis, as long as the other family members, but which encodes multiple stops in the 5’ region preventing formation of a full length protein. Perhaps this family member is in the process of becoming a pseudogene by disruption of the full reading frame, and an indication that the family may be reduced in size over time.

The ema genes are distributed across the genome, with no evidence of clustering or telomeric association. There is no evidence that the ema genes undergo recombination or dynamic sequence variation, as might be expected if these genes are involved in immune evasion[16, 44]. Therefore, despite their immunodominance, their limited number and lack of a structural basis for variation do not fit with the paradigms seen for P. falciparum var genes (encoding PfEMP1) or B. bovis ves1 genes (encoding VESA1) where evasion of the immune response through emergence of variable surface antigens is a well characterized mechanism of parasite persistence[18, 45].

FAINT domains

While not a protein family per se, the FAINT domain was originally detected in T. annulata and T. parva as a stretch of 70 amino acids and was named “Frequently Associated IN Theileria”; subsequently, Pfam04385, the FAINT domain – a domain of unknown function (DUF529) was established and was found to be over represented in proteins predicted to be secreted[27, 28]. T. equi proteins may have multiple FAINT domains and are typically classified into several protein families. T. equi contains 271 proteins with 560 hits to Pfam04385. FAINT domains are not reported for B. bovis or P. falciparum. The FAINT domain containing genes are distributed throughout the T. equi genome.

T. equi does not contain homologs of putative Theileria host cell transforming genes

Although host cell transformation has not been reported for T. equi, it has been reported to invade the lymphocyte and develop to a macroshizont stage[13]. In T. parva and T. annulata, the macroschizont life cycle can induce transformation of the infected cell[46]. To assure that the T. equi isolate used in this study infected equine lymphocytes/PBMCs, equine PBMCs were infected with sporozoites obtained from infected adult male Rhipicephalus microplus ticks. To verify infection of equine PBMC cultures, EMA-1 and EMA-2 monoclonal antibodies were tested by immunofluorescence and bound corresponding T. equi antigens within equine PBMCs (Figure2). The lack of observed transformation in this close relative of pathogens (T. parva and T. annulata) that do transform their host led us to search the T. equi genome for putative host cell transformation genes to see if these genes were retained in the genome. A set of genes[46] identified as potential manipulators of host cell phenotype in T. annulata and T. parva was used to explore the T. equi genome for orthologous genes, along with the EST expression data. T. parva and T. annulata express prohibitin in macro- and microschizonts, while T. equi expresses the ortholog (BEWA_014000) in erythrocyte-stage parasite ESTs. Prohibitin is reported to be a tumor suppressor that is involved in pathways that lead to immortalization, however, this gene was also identified in B. bovis, which does not undergo transformation[46]. Similarly, T. equi has several homologs of cyclophilins, which have been linked to transformation. This gene is found in B. bovis, T. parva and T. annulata, again, limiting the implication of this gene product as a host cell transforming factor. T. equi does not have a homolog of the Theileria schizont AT hook (TashAT) family of proteins. The AT hooks function as nuclear localization signals and DNA binding domains, and may function in cell transformation. Likewise, TashHN (host nucleus) and SuAT1 do not have T. equi orthologs. TashHN is expressed by Theileria macroschizonts that have lost the ability to form merozoites, while SuAT1 is a schizont protein with characteristics of the TashAT family, that can modulate the phenotype of cultured bovine macrophages by altering the expression profile of cytoskeletal proteins in a manner similar to that seen during infection.

Figure 2
figure 2

Analysis of schizont-infected PBMC in vitro. The capacity of T. equi (Florida) to infect PBMC in vitro was assessed by light microscopy (A, B) and IFA (C, D). Fresh PBMC from adult Arabian horses were cocultured with tick salivary gland homogenates containing T. equi sporozoites. Infected and uninfected control PBMC were sampled daily for cytospin preparation and Diff-Quick staining [representative photomicrographs of uninfected control cultures (A) and T. equi infected cultures (B) on day 9]. Panel B includes three schizont-infected leukocytes, with multiple, oval to round, 1–2 μm diameter, purple nuclei (developing merozoites). To confirm the intracytoplasmic organisms were T. equi, uninfected control and infected cultures were labeled with antibody specific for equine merozoite antigens 1 and 2 [mAb 36/133.97 (anti-EMA 1/2)]. In the infected culture wells (D), intracytoplasmic schizonts and developing merozoites were specifically labeled with anti-EMA 1/2 (secondary goat anti-mouse IgG1 conjugated with FITC-green; Nuclear stain = DAPI). Cells from the uninfected control cultures were not labeled with anti-EMA 1/2 (representative data in panel C). Scale bar = 10 μm.

Vaccine candidates

The RAP-1 Family

Members of the rhoptry associated protein-1 (rap 1) gene family are candidates for vaccine development in Babesia spp. Immunization of cattle with purified native RAP-1 of B. bigemina resulted in partially protective responses, and is one of three antigens in a recombinant vaccine developed in Australia[47]. However, recombinant B. bovis RAP-1 failed to elicit protective immune responses in vaccine trials. The RAP-1 proteins of Babesia spp. are characterized by a signal peptide followed by a cysteine-rich region and several short conserved sequence motifs within the first 300 amino acids. Thus, the N-terminal region of all known Babesia RAP-1 proteins contains the “rap-1 domain”. The C-terminal regions are less conserved among Babesia species. B. bovis contains two identical rap 1 genes in tandem and a shorter rap 1 related protein gene (RRP) located ~30 kbp from the rap 1 locus. Three rap 1 genes were identified in T. parva and four in T. annulata. The general structure of rap 1 in Theileria spp. is different from that in B. bovis, with the Theileria proteins containing more than one rap-1 domain. The T. equi rap 1 locus contains two tandemly arranged genes: BEWA_037610 contains two rap-1 domains, and BEWA_037600 contains three rap-1 domains. Thus, the arrangement of rap-1 domains in T. equi resembles that found in other Theileria spp.

Apical Membrane Antigen 1

T. equi Apical Membrane Antigen-1 (AMA-1; BEWA_036830) is a microneme protein with surface exposed epitopes that has orthologs in hemoprotozoan Apicomplexa, as well as Toxoplasma gondii[48]. AMA-1 antibody has inhibitory effects on parasite replication in vitro, and has been protective in animal models[4951]. T. equi and B. bovis AMA-1 lacks the N-terminal extension found in other Theileria spp. AMA-1 has a set of conserved cysteine residues which may contribute to a conserved architecture through disulfide bonds. Plasmodium AMA-1 is polymorphic[52], and the degree of polymorphism must be assessed in T. equi before AMA-1 can be considered as a vaccine candidate for this species.

Thrombospondin-related anonymous protein (TRAP)

Conserved hypothetical proteins BEWA_005690 and BEWA_005710 fall into a Cluster of Orthologous Groups (COG) with apicomplexan genes annotated as TRAP, sporozoite surface protein or microneme protein 2[53]. Although these molecules have been effective immunogens in Plasmodium[54], the T. equi proteins have low sequence similarity and it is unclear whether these would be effective vaccine candidates.

Other prophylaxis opportunities suggested by comparative genomics

HAP2 is a protein essential for membrane fusion during zygote formation in Plasmodium that has a homolog in numerous species, including Leishmania and T. parva[55]. A positional homolog i.e., a gene with very low sequence identity couched within a highly conserved locus, exists in B. bovis (BBOV_III006770) and T. equi (BEWA_7380). This target could provide a novel method of transmission blocking through gene knock out, thus preventing gamete fusion and tick transmission of a modified live vaccine. Membrane Occupation and Recognition Nexus protein (MORN1) seems to be related to cell division in all Apicomplexa studied, and functions in both asexual and sexual reproduction[56]. A common and effective prophylaxis for members of the phylum could come from targeting this molecule. BEWA_033160 is annotated as a MORN repeat domain-containing protein, and falls in to a COG with proteins from Plasmodium, Theileria, and B. bovis, as well as Toxoplasma gondii and Cryptosporidium parvum. Heme Detoxification Protein (HDP) has a role in detoxifying heme that results from parasite metabolism of hemoglobin by polymerizing the heme into hemozoin[57]. BEWA_011450, a putative HDP, is in a COG conserved across apicomplexan organisms. This hypothetical target for prophylaxis would allow toxic levels of heme to accumulate through inactivation of HDP.

p67 locus synteny

T. parva p67 is an abundant sporozoite surface antigen, and along with its homolog from T. annulata (SPAG-1) has been shown to induce neutralizing antibody, and immunity to theileriosis and has been a target of vaccine development[58]. A positional homolog was identified in B. bovis (BBOV_IV007750)[29, 59] and T. equi contains a gene at the same location (BEWA_015160). EST data show that BEWA_015160 is transcribed. The prospect that the genes occupying this position in syntenic loci have similar functions during parasite infection but tailored to different hosts makes them interesting candidates for vaccine studies.

The T. parva six

A set of six T. parva antigens that are the targets of CD8+ T lymphocyte responses in immune cattle have been designated Tp1,Tp2, Tp4,Tp5, Tp7 and Tp8[60]. Tp2 is a surface protein whose homolog is also expressed in T. annulata. All of these vaccine candidates except hypothetical protein Tp1 have orthologs in T. equi and have been demonstrated to be transcribed in erythrocyte stages via EST analyses. While the significance of piroplasm stage expression is unclear, these T. equi proteins may have utility as vaccine components based on observations in T. parva.

Phylogenetic analysis of piroplasms

Mehlhorn and Schein renamed Babesia equi as T. equi in 1998 due to the description of a lymphocytic stage for this parasite[8]; however, while this classification has been well adopted, it is not borne out by phylogenetic studies. The phylogenetic position of T. equi was explored using 150 deduced amino acid sequences (Additional file7) from eight fully sequenced apicomplexan genomes. Results of the Bayesian analysis (Figure3) support two possible scenarios: the four piroplasm species, T. parva, T. annulata, B. bovis and T. equi, represent a single genus or alternatively, these species represent three separate genera. Maximum Parsimony (MP) analyses of individual sequences resulted in many different tree topologies; however, the four most commonly recovered topologies (Additional file8) represent 78% of all trees. Consistent with Bayesian inference, the three most common topologies (Additional file8, topologies a-c) suggest that T. equi shares a more recent common ancestor with the two Theileria species than with B. bovis. MP analysis of the concatenated datasets unambiguously recovered this same relationship of the four piroplasm species, and completely corroborated the tree estimated using Bayesian inference. Finally, maximum likelihood analyses of the concatenated datasets using a variety of substitution models (JTT, MTREV, WAG, RTREV, CPREV, VT) consistently yielded the same estimates, which completely corroborate the trees generated by Bayesian inference and MP analyses of the concatenated datasets (Additional file8). Collectively, robust phylogeny estimation, which utilized three separate optimality criteria and various models of protein evolution, consistently placed T. equi as a sister taxon to the Theileria clade, with all three of these piroplasm species subtended by B. bovis.

Figure 3
figure 3

Phylogenetic tree of sequenced apicomplexans. Posterior probability distribution tree representing best likelihood score (probability of 1.0) following Bayesian analysis of 150 concatenated polypeptides across eight taxa. Taxa codes are Pf (Plasmodium falciparum), Pv (Plasmodium vivax), Tg (Toxoplasma gondii), Cp (Cryptosporidium parvum), Bb (Babesia bovis), Te (Theileria equi), Tp (Theileria parva), Ta (Theileria annulata).

Previous studies employing 18S rRNA sequences across many piroplasmid taxa also revealed uncertainty in the phylogenetic position of T. equi and several other piroplasms[61, 62]. Our robust phylogenetic analyses, which strongly imply the placement of T. equi as sister to the Theileria clade provide stability to piroplasm systematics. Our goal was to determine if a selection of orthologous polypeptides from completed Babesia and Theileria genomes could be used to discriminate the appropriate phylogenetic position of T. equi. However, species sampling was restricted by available complete genome sequences and is too limited to completely resolve this issue. A more appropriate solution may be to place T. equi in a genus distinct from both Theileria and Babesia, as has been previously suggested[9]. This option would likely result in the renaming of some related taxa including Cytauxzoon felis, which groups as the closest relative of T. equi in the 18S rRNA study[61]. Importantly, our analysis revealed 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 (Boxed sequences in Additional file7).

Conclusions

Characteristics of genes involved in immune evasion leading to persistence include immunodominance, highly variable domains, and membership in a multigene family. Within the phylum Apicomplexa, B. bovis contains the 119 member ves1 gene family which encodes the VESA1, known to vary by gene conversion and participate in immune evasion; and P. falciparum contains the 59 member var gene family encoding PfEMP1, which utilizes several mechanisms to generate antigenic variation resulting in immune evasion[26]. An analogous gene family wasn’t detected in T. equi, instead, it contains a small family of immunodominant proteins including EMA1, EMA2, EMA3 and seven additional members which have marked variation throughout each gene. While the precise mechanism(s) of antigenic variation provided by the ema family isn’t yet known, we hypothesize that EMA family members provide antigenic variation and immune escape through a protein reassortment strategy that generates immune escape variants by creating novel heteromers of EMA on the cell surface. The genome sequence also revealed a 109 member gene family (Family 3); however these proteins have never been detected by immune serum, suggesting that they are not contributing to immune evasion. Genome sequencing allowed an exploration of the phylogenetic position of T. equi at a depth not previously feasible; demonstrating that this species clusters consistently with, and as a sister lineage to other Theileria spp. Genome sequences for closely related piroplasms may allow “fine tuning” of the T. equi phylogenetic position and appropriate naming of the lineage to which this species belongs. Finally, the genome sequence will provide an invaluable tool for researchers developing methods to control equine piroplasmosis.

Materials and methods

Ethics statement

Animal experiments were approved by the Institutional Animal Care and Use Committee at University of Idaho, USA, in accordance with institutional guidelines based on the U.S. National Institutes of Health (NIH) Guide for the Care and Use of Laboratory Animals.

Parasite culture

For sequencing

Theileria equi USDA (Florida) stabilate (PPE: 0.64%) generated from infection of a spleen intact horse (H2) was inoculated intravenously into a splenectomized horse (H047). When the parasitemia reached 16% as determined by Giemsa stained blood smear examination, blood was collected into a flask containing glass beads for defibrination. The blood was washed three times with HBSS/10mM EDTA to remove the white blood cells (WBC). Erythrocytes were purified using histopaque 1077 following the manufacturer’s protocol. Microscopic examination determined WBC contamination to be 5 WBC per 1 ul of blood. T. equi genomic DNA was extracted from the purified erythrocytes using components of the Gentra Puregene DNA Purification System (Kit D-5000).

T. equi sporozoite infection of PBMC in vitro:

To obtain the sporozoites need for in vitro PBMC infection, one thousand adult male Rhipicephalus microplus ticks were acquisition fed on a T. equi (Florida strain) merozoite stabilate-infected adult horse during ascending parasitemia for eight days (acquisition feeding). Following acquisition feeding, fed ticks were incubated for two days at 15°C with 94% relative humidity and a 12 hr photo period, and subsequently transmission fed on a naïve horse to induce final maturation of sporozoites into their infectious stage. On day seven of the transmission feeding, 240 live adult males were recovered and salivary gland pairs were dissected aseptically, and sporozoite extracts prepared using established methods. Briefly, collected salivary glands were washed in 0.5 ml of complete media (RPMI 1640 supplemented with 10% heat inactivated fetal bovine serum, 200 IU/ml benzyl penicillin, 200 ug/ml streptomycin sulphate, 50 μg/ml gentamycin and 5 × 10-5 M 2-mercaptoethanol) and crushed in a glass homogenizer to release sporozoites. The homogenate was then centrifuged at 300 g for 5 min and the supernatant containing sporozoites was collected and adjusted to 10 tick equivalents/ml. One ml of the sporozoite extract was inoculated into each of ten wells of a 24-well plate containing 2 × 106 PBMC from an uninfected horse. Starting on day nine post-inoculation, cell culture aliquots were cytospun and examined microscopically for infection following Diff-Quick staining.

Sequencing

DNA from T. equi USDA (Florida) strain was used to construct a small and large insert plasmid library as well as a bacterial artificial chromosome library. A total of 100,835 high quality sequence reads (822 bp average read length) were generated and assembled using the Celera Assembler (http://sourceforge.net/projects/wgs-assembler/), resulting in 119 scaffolds consisting of 495 contigs. A BAC library was end sequenced to generate an additional 3023 reads which were used to confirm the assembly and for targeted sequencing in the closure phase. Gaps were closed by a combination of primer walking and transposon based or shotgun sequencing of medium insert clones, BAC clones or PCR products. The genome sequence has been deposited in GenBank under accession number ACOU00000000. A cDNA library was constructed using cultured infected erythrocytes (Florida strain) using the Creator SMART cDNA library construction kit (Clontech). cDNA in the size range of 0.3 -2kb was normalized using the Trimmer normalization kit (Evrogen). 7629 sequences (GenBank Accession #s HS032712 - HS040340) were generated and assembled into 2355 contigs using PASA[63].

Genome annotation

Genes encoding tRNA’s were identified using tRNA scan-SE[64]. Gene models were predicted using the ab initio gene finding programs GeneZilla, GlimmerHMM[65], Phat[66] and Snap[67] that used 395 partial and full length high confidence genes in the training set. The training data was manually constructed and inspected for its alignment against highly conserved protein sequences using the AAT package[68] and PASA[63] to align the ESTs to the genomic sequence with a stringent criteria of 95% identity over 90% length using gmap[69]. Data were combined and consensus gene models were derived using Evidence Modeler EVM[70]. The consensus gene models were manually checked for obvious errors. Such models were corrected using a Java based tool called Neomorphic Annotation Station[71].

Functional annotation was as for B. bovis[29]. In addition, EC numbers were assigned in an automated fashion using PRIAM[72] and metabolic pathways constructed using SRI’s pathway tools[73]. TransportDB is a relational database that was used to determine transporter complement[74]. TribeMCL[75] was used with default parameters to construct a database of genes that are part of paralogous families of proteins represented in the genome. The method is a sequence similarity matrix-based Markov clustering method.

Comparative genome analysis

Sybil (http://sybil.sourceforge.net) was used to create an all-versus-all BLASTP search using the proteomes of T. equi, B. bovis, T. parva, T.annulata, P. falciparum, P. vivax, C. parvum and T. gondii. These outputs were subjected to Jaccard clustering, placing proteins into distinct clusters for each proteome. Clusters from different proteomes were linked based on best bidirectional BLASTP hits between them to provide Jf-COGs. A minimum block size of five with one gap was allowed in the analyses.

Immunofluorescence assay

To verify PBMC cultures were infected with T. equi, cytospin preparations were made for immunofluorescence antibody microscopy on day nine after cell culture inoculation. After drying, cytospin preparations were fixed in 1% formalin for 2 min and primarily labeled with T. equi EMA-1 and EMA-2-specific monoclonal antibody 36/133.97. Bound antibody was detected using fluorescein-conjugated goat anti-mouse IgG1 and subsequently visualized by fluorescence microscopy.

Phylogenetic analysis

The phylogenetic position of T. equi relative to other piroplasm species with published genomes, namely T. parva, T. annulata, and B. bovis was determined based on 150 polypeptide sequences, and rooted with four other apicomplexan taxa: P. vivax, P. falciparum, T. gondii, and C. parvum. Polypeptides were selected from single copy genes identified in all eight species. COGs (clusters of orthologous genes) were defined using an in-house comparative pipeline, which starts with BLASTP among protein sequences within and across species. We used the BLOSUM62 matrix with expected value 10-5. Jaccard clustering was then performed twice, once to form within-species clusters of paralogous genes and a second time to derive a set of multi-species COGs. In the first case, we used an 80% identity cutoff over a minimum of 70% of the length of the smallest protein and a link score of 0.6. In the second, we set the identity cutoff at 50% over >70% length, and Jaccard coefficient cutoff of 0 for edge pruning. In order to minimize ambiguous amino acid homology assignments that can occur when indels are present, peptides were chosen that have the smallest protein size variation across all species (Additional file7 for list). Protein sequences within a COG were aligned with ClustalW using default parameters[76].

Three methods were used to reconstruct the phylogenetic relationships among the sequences: i) maximum parsimony (MP) with exhaustive tree search, and otherwise default parameters as implemented in PAUP v4.0b10[77], ii) Bayesian inference (BI) implemented in MrBayes v3.1.2[78], and iii). maximum likelihood (ML) estimation as implemented in RAxML v.7.2.8[79]. Initially, the individual protein datasets were analyzed with maximum parsimony, with branch support assessed with 1000 bootstrap replicates. The 150 datasets were subsequently concatenated and again analyzed using MP with exhaustive tree search and similar bootstrapping procedure. For BI of the concatenated datasets, the amino acid transition matrix was set to a mixture of models with fixed rate matrices (Poisson, Jones, Dayhoff, Mtrev, Mtmam, Wag, Rtrev, Cprev, Vt and Blosum) of equal prior probabilities, and otherwise default parameters. Four runs of MrBayes were conducted, each with 4 chains. Two of those runs ran for 567,000 generations and the other two ran for 2,365,000 generations. Convergence was achieved (Potential Scale Reduction Factor, PSRF=1.00) for all model parameters estimated, including tree length (mean-1.79), the amino acid model (Wag, with posterior probability=1.00), and the tree topology. ML analyses of the concatenated datasets all implemented a gamma model of rate heterogeneity with estimation of the proportion of invariable sites. Six separate amino acid substitution models (JTT, MTREV, WAG, RTREV, CPREV, VT) were utilized. Branch support was assessed with 1000 bootstrap pseudoreplications.

Note: Two additional genome sequences of interest were published while this article was in review. T. orientalis does not transform it’s host cell and has been compared to T. parva and T. annulata to identify putative mediators of leukocyte transformation[80]. B. microti phylogeny was analyzed using 316 single copy genes and found to clade separately to T. parva, T. annulata and B. bovis[81].

Abbreviations

ABC:

ATP-binding cassette

AMA:

Apical Membrane Antigen

COG:

Cluster of Orthologous Groups

DHFR-TS:

Dihydrofolate reductase-thymidylate synthase

EMA:

Equi Merozoite Antigens

EST:

Expressed sequence tag

FAINT:

Frequently Associated IN Theileria

GPI:

Glycosylphosphatidylinositol

MFS:

Major Facilitator Superfamily

MP:

Maximum Parsimony

mt:

Mitochondrial

MORN:

Membrane Occupation and Recognition Nexus

PBMC:

Peripheral blood mononuclear cells

rap :

Rhoptry associated protein

TCA:

Tricarboxylic acid cycle

VESA:

Variant erythrocyte surface antigen

WBC:

White blood cells.

References

  1. Schwint ON, Knowles DP, Ueti MW, Kappmeyer LS, Scoles GA: Transmission of Babesia caballi by Dermacentor nitens (Acari: Ixodidae) is restricted to one generation in the absence of alimentary reinfection on a susceptible equine host. J Med Entomol. 2008, 45 (6): 1152-1155. 10.1603/0022-2585(2008)45[1152:TOBCBD]2.0.CO;2.

    Article  CAS  PubMed  Google Scholar 

  2. Ueti MW, Palmer GH, Scoles GA, Kappmeyer LS, Knowles DP: Persistently infected horses are reservoirs for intrastadial tick-borne transmission of the apicomplexan parasite Babesia equi. Infect Immun. 2008, 76 (8): 3525-3529. 10.1128/IAI.00251-08.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  3. Maurer FD: Equine piroplasmosis–another emerging disease. J Am Vet Med Assoc. 1962, 141: 699-702.

    CAS  PubMed  Google Scholar 

  4. Short MA, Clark CK, Harvey JW, Wenzlow N, Hawkins IK, Allred DR, Knowles DP, Corn JL, Grause JF, Hennager SG, et al: Outbreak of equine piroplasmosis in Florida. J Am Vet Med Assoc. 2012, 240 (5): 588-595. 10.2460/javma.240.5.588.

    Article  PubMed  Google Scholar 

  5. Scoles GA, Hutcheson HJ, Schlater JL, Hennager SG, Pelzel AM, Knowles DP: Equine piroplasmosis associated with Amblyomma cajennense Ticks, Texas, USA. Emerg Infect Dis. 2011, 17 (10): 1903-1905. 10.3201/eid1710.101182.

    Article  PubMed Central  PubMed  Google Scholar 

  6. Traub-Dargatz JL, Short MA, Pelzel AM, Norman TE, Knowles DP: Panel on equine piroplasmosis. American Association of Equine Practitioners: Dec. 4, 2010 2010. 2010

    Google Scholar 

  7. ProMed-mail: Equine piroplasmosis - USA (11): multi-state. ProMed-mail - 2009. 2009, 3: 2009-20091203.4128.

    Google Scholar 

  8. Mehlhorn H, Schein E: Redescription of Babesia equi Laveran, 1901 as Theileria equi Mehlhorn, Schein 1998. Parasitol Res. 1998, 84 (6): 467-475. 10.1007/s004360050431.

    Article  CAS  PubMed  Google Scholar 

  9. Allsopp MT, Cavalier-Smith T, De Waal DT, Allsopp BA: Phylogeny and evolution of the piroplasms. Parasitology. 1994, 108 (Pt 2): 147-152.

    Article  CAS  PubMed  Google Scholar 

  10. Allsopp MT, Allsopp BA: Molecular sequence evidence for the reclassification of some Babesia species. Ann N Y Acad Sci. 2006, 1081: 509-517. 10.1196/annals.1373.076.

    Article  CAS  PubMed  Google Scholar 

  11. Musoke AJ, Nene V: Development of recombinant antigen vaccines for the control of theileriosis. Parassitologia. 1990, 32 (1): 73-85.

    CAS  PubMed  Google Scholar 

  12. Morrison WI: Progress towards understanding the immunobiology of Theileria parasites. Parasitology. 2009, 136 (12): 1415-1426. 10.1017/S0031182009990916.

    Article  PubMed  Google Scholar 

  13. Schein E, Rehbein G, Voigt WP, Zweygarth E: Babesia equi (Laveran 1901) 1. Development in horses and in lymphocyte culture. Tropenmed Parasitol. 1981, 32 (4): 223-227.

    CAS  PubMed  Google Scholar 

  14. Knowles DP, Kappmeyer LS, Perryman LE: Specific immune responses are required to control parasitemia in Babesia equi infection. Infect Immun. 1994, 62 (5): 1909-1913.

    PubMed Central  CAS  PubMed  Google Scholar 

  15. Schein E: Equine babesiosis. Babesiosis of domestic animals and man. Edited by: Ristic M. 1988, CRC Press, Inc, Boca Raton, 197-208.

    Google Scholar 

  16. Ueti MW, Palmer GH, Kappmeyer LS, Statdfield M, Scoles GA, Knowles DP: Ability of the vector tick Boophilus microplus to acquire and transmit Babesia equi following feeding on chronically infected horses with low-level parasitemia. J Clin Microbiol. 2005, 43 (8): 3755-3759. 10.1128/JCM.43.8.3755-3759.2005.

    Article  PubMed Central  PubMed  Google Scholar 

  17. Allred DR, Carlton JM, Satcher RL, Long JA, Brown WC, Patterson PE, O'Connor RM, Stroup SE: The ves multigene family of B. bovis encodes components of rapid antigenic variation at the infected erythrocyte surface. Mol Cell. 2000, 5 (1): 153-162. 10.1016/S1097-2765(00)80411-6.

    Article  CAS  PubMed  Google Scholar 

  18. Le Scanf C, Fandeur T, Morales-Betoulle ME, Mercereau-Puijalon O: Plasmodium falciparum: altered expressions of erythrocyte membrane-associated antigens during antigenic variation. Exp Parasitol. 1997, 85 (2): 135-148. 10.1006/expr.1996.4121.

    Article  CAS  PubMed  Google Scholar 

  19. Hajduk SL: Antigenic variation during the developmental cycle of Trypanosoma brucei. J Protozool. 1984, 31 (1): 41-47.

    Article  CAS  PubMed  Google Scholar 

  20. Lau AO, McElwain TF, Brayton KA, Knowles DP, Roalson EH: Babesia bovis: a comprehensive phylogenetic analysis of plastid-encoded genes supports green algal origin of apicoplasts. Exp Parasitol. 2009, 123 (3): 236-243. 10.1016/j.exppara.2009.07.007.

    Article  CAS  PubMed  Google Scholar 

  21. Waller RF, McFadden GI: The apicoplast: a review of the derived plastid of apicomplexan parasites. Curr Issues Mol Biol. 2005, 7 (1): 57-79.

    PubMed  Google Scholar 

  22. Obornik M, Janouskovec J, Chrudimsky T, Lukes J: Evolution of the apicoplast and its hosts: from heterotrophy to autotrophy and back again. Int J Parasitol. 2009, 39 (1): 1-12. 10.1016/j.ijpara.2008.07.010.

    Article  CAS  PubMed  Google Scholar 

  23. Foth BJ, Ralph SA, Tonkin CJ, Struck NS, Fraunholz M, Roos DS, Cowman AF, McFadden GI: Dissecting apicoplast targeting in the malaria parasite Plasmodium falciparum. Science. 2003, 299 (5607): 705-708. 10.1126/science.1078599.

    Article  CAS  PubMed  Google Scholar 

  24. Cilingir G, Broschat SL, Lau AO: ApicoAP: The first computational model for identifying apicoplast-targeted proteins in multiple species of Apicomplexa. PLoS One. 2012, 7 (5): e36598-10.1371/journal.pone.0036598.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  25. Hikosaka K, Watanabe Y, Tsuji N, Kita K, Kishine H, Arisue N, Palacpac NM, Kawazu S, Sawai H, Horii T, et al: Divergence of the mitochondrial genome structure in the apicomplexan parasites, Babesia and Theileria. Mol Biol Evol. 2010, 27 (5): 1107-1116. 10.1093/molbev/msp320.

    Article  CAS  PubMed  Google Scholar 

  26. Gardner MJ, Hall N, Fung E, White O, Berriman M, Hyman RW, Carlton JM, Pain A, Nelson KE, Bowman S, et al: Genome sequence of the human malaria parasite Plasmodium falciparum. Nature. 2002, 419 (6906): 498-511. 10.1038/nature01097.

    Article  CAS  PubMed  Google Scholar 

  27. Pain A, Renauld H, Berriman M, Murphy L, Yeats CA, Weir W, Kerhornou A, Aslett M, Bishop R, Bouchier C, et al: Genome of the host-cell transforming parasite Theileria annulata compared with T. parva. Science. 2005, 309 (5731): 131-133. 10.1126/science.1110418.

    Article  CAS  PubMed  Google Scholar 

  28. Gardner MJ, Bishop R, Shah T, de Villiers EP, Carlton JM, Hall N, Ren Q, Paulsen IT, Pain A, Berriman M, et al: Genome sequence of Theileria parva, a bovine pathogen that transforms lymphocytes. Science. 2005, 309 (5731): 134-137. 10.1126/science.1110439.

    Article  CAS  PubMed  Google Scholar 

  29. Brayton KA, Lau AO, Herndon DR, Hannick L, Kappmeyer LS, Berens SJ, Bidwell SL, Brown WC, Crabtree J, Fadrosh D, et al: Genome sequence of Babesia bovis and comparative analysis of apicomplexan hemoprotozoa. PLoS Pathog. 2007, 3 (10): 1401-1413.

    CAS  PubMed  Google Scholar 

  30. Nagai A, Yokoyama N, Matsuo T, Bork S, Hirata H, Xuan X, Zhu Y, Claveria FG, Fujisaki K, Igarashi I: Growth-inhibitory effects of artesunate, pyrimethamine, and pamaquine against Babesia equi and Babesia caballi in in vitro cultures. Antimicrob Agents Chemother. 2003, 47 (2): 800-803. 10.1128/AAC.47.2.800-803.2003.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  31. Gaffar FR, Wilschut K, Franssen FF, de Vries E: An amino acid substitution in the Babesia bovis dihydrofolate reductase-thymidylate synthase gene is correlated to cross-resistance against pyrimethamine and WR99210. Mol Biochem Parasitol. 2004, 133 (2): 209-219. 10.1016/j.molbiopara.2003.08.013.

    Article  CAS  PubMed  Google Scholar 

  32. Florin-Christensen J, Suarez CE, Florin-Christensen M, Hines SA, McElwain TF, Palmer GH: Phosphatidylcholine formation is the predominant lipid biosynthetic event in the hemoparasite Babesia bovis. Mol Biochem Parasitol. 2000, 106 (1): 147-156. 10.1016/S0166-6851(99)00209-1.

    Article  CAS  PubMed  Google Scholar 

  33. Florin-Christensen J, Suarez CE, Florin-Christensen M, Wainszelbaum M, Brown WC, McElwain TF, Palmer GH: A unique phospholipid organization in bovine erythrocyte membranes. Proc Natl Acad Sci U S A. 2001, 98 (14): 7736-7741. 10.1073/pnas.131580998.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  34. Knowles DP, Kappmeyer LS, Perryman LE: Genetic and biochemical analysis of erythrocyte-stage surface antigens belonging to a family of highly conserved proteins of Babesia equi and Theileria species. Mol Biochem Parasitol. 1997, 90 (1): 69-79. 10.1016/S0166-6851(97)00150-3.

    Article  CAS  PubMed  Google Scholar 

  35. Koenderink JB, Kavishe RA, Rijpma SR, Russel FG: The ABCs of multidrug resistance in malaria. Trends Parasitol. 2010, 26 (9): 440-446. 10.1016/j.pt.2010.05.002.

    Article  CAS  PubMed  Google Scholar 

  36. Pao SS, Paulsen IT, Saier MH: Major facilitator superfamily. Microbiol Mol Biol Rev. 1998, 62 (1): 1-34.

    PubMed Central  CAS  PubMed  Google Scholar 

  37. DeBarry JD, Kissinger JC: Jumbled genomes: missing Apicomplexan synteny. Mol Biol Evol. 2011, 28 (10): 2855-2871. 10.1093/molbev/msr103.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  38. Bishop R, Musoke A, Morzaria S, Sohanpal B, Gobright E: Concerted evolution at a multicopy locus in the protozoan parasite Theileria parva: extreme divergence of potential protein-coding sequences. Mol Cell Biol. 1997, 17 (3): 1666-1673.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  39. Knowles DP, Perryman LE, Kappmeyer LS, Hennager SG: Detection of equine antibody to Babesia equi merozoite proteins by a monoclonal antibody-based competitive inhibition enzyme-linked immunosorbent assay. J Clin Microbiol. 1991, 29 (9): 2056-2058.

    PubMed Central  CAS  PubMed  Google Scholar 

  40. Ikadai H, Ishida H, Sasaki M, Taniguchi K, Miyata N, Koda M, Igarashi I, Oyamada T: Molecular cloning and partial characterization of Babesia equi EMA-3. Mol Biochem Parasitol. 2006, 150 (2): 371-373. 10.1016/j.molbiopara.2006.08.004.

    Article  CAS  PubMed  Google Scholar 

  41. Kumar S, Yokoyama N, Kim JY, Huang X, Inoue N, Xuan X, Igarashi I, Sugimoto C: Expression of Babesia equi EMA-1 and EMA-2 during merozoite developmental stages in erythrocyte and their interaction with erythrocytic membrane skeleton. Mol Biochem Parasitol. 2004, 133 (2): 221-227. 10.1016/j.molbiopara.2003.10.010.

    Article  CAS  PubMed  Google Scholar 

  42. Skilton RA, Musoke AJ, Wells CW, Yagi Y, Nene V, Spooner PR, Gachanja J, Osaso J, Bishop RP, Morzaria SP: A 32 kDa surface antigen of Theileria parva: characterization and immunization studies. Parasitology. 2000, 120 (Pt 6): 553-564.

    Article  CAS  PubMed  Google Scholar 

  43. Gubbels MJ, Katzer F, Hide G, Jongejan F, Shiels BR: Generation of a mosaic pattern of diversity in the major merozoite-piroplasm surface antigen of Theileria annulata. Mol Biochem Parasitol. 2000, 110 (1): 23-32. 10.1016/S0166-6851(00)00253-X.

    Article  CAS  PubMed  Google Scholar 

  44. Cunha CW, Kappmeyer LS, McGuire TC, Dellagostin OA, Knowles DP: Conformational dependence and conservation of an immunodominant epitope within the Babesia equi erythrocyte-stage surface protein equi merozoite antigen 1. Clin Diagn Lab Immunol. 2002, 9 (6): 1301-1306.

    PubMed Central  CAS  PubMed  Google Scholar 

  45. Allred DR, Cinque RM, Lane TJ, Ahrens KP: Antigenic variation of parasite-derived antigens on the surface of Babesia bovis-infected erythrocytes. Infect Immun. 1994, 62 (1): 91-98.

    PubMed Central  CAS  PubMed  Google Scholar 

  46. Shiels B, Langsley G, Weir W, Pain A, McKellar S, Dobbelaere D: Alteration of host cell phenotype by Theileria annulata and Theileria parva: mining for manipulators in the parasite genomes. Int J Parasitol. 2006, 36 (1): 9-21. 10.1016/j.ijpara.2005.09.002.

    Article  CAS  PubMed  Google Scholar 

  47. Wright IG, Casu R, Commins MA, Dalrymple BP, Gale KR, Goodger BV, Riddles PW, Waltisbuhl DJ, Abetz I, Berrie DA, et al: The development of a recombinant Babesia vaccine. Vet Parasitol. 1992, 44 (1–2): 3-13.

    Article  CAS  PubMed  Google Scholar 

  48. Gaffar FR, Yatsuda AP, Franssen FF, de Vries E: Erythrocyte invasion by Babesia bovis merozoites is inhibited by polyclonal antisera directed against peptides derived from a homologue of Plasmodium falciparum apical membrane antigen 1. Infect Immun. 2004, 72 (5): 2947-2955. 10.1128/IAI.72.5.2947-2955.2004.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  49. Kocken CH, Withers-Martinez C, Dubbeld MA, van der Wel A, Hackett F, Valderrama A, Blackman MJ, Thomas AW: High-level expression of the malaria blood-stage vaccine candidate Plasmodium falciparum apical membrane antigen 1 and induction of antibodies that inhibit erythrocyte invasion. Infect Immun. 2002, 70 (8): 4471-4476. 10.1128/IAI.70.8.4471-4476.2002.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  50. Anders RF, Crewther PE, Edwards S, Margetts M, Matthew ML, Pollock B, Pye D: Immunisation with recombinant AMA-1 protects mice against infection with Plasmodium chabaudi. Vaccine. 1998, 16 (2–3): 240-247.

    Article  CAS  PubMed  Google Scholar 

  51. Kocken CH, Dubbeld MA, Van Der Wel A, Pronk JT, Waters AP, Langermans JA, Thomas AW: High-level expression of Plasmodium vivax apical membrane antigen 1 (AMA-1) in Pichia pastoris: strong immunogenicity in Macaca mulatta immunized with P. vivax AMA-1 and adjuvant SBAS2. Infect Immun. 1999, 67 (1): 43-49.

    PubMed Central  CAS  PubMed  Google Scholar 

  52. Healer J, Murphy V, Hodder AN, Masciantonio R, Gemmill AW, Anders RF, Cowman AF, Batchelor A: Allelic polymorphisms in apical membrane antigen-1 are responsible for evasion of antibody-mediated inhibition in Plasmodium falciparum. Mol Microbiol. 2004, 52 (1): 159-168. 10.1111/j.1365-2958.2003.03974.x.

    Article  CAS  PubMed  Google Scholar 

  53. Gaffar FR, Yatsuda AP, Franssen FF, de Vries E: A Babesia bovis merozoite protein with a domain architecture highly similar to the thrombospondin-related anonymous protein (TRAP) present in Plasmodium sporozoites. Mol Biochem Parasitol. 2004, 136 (1): 25-34. 10.1016/j.molbiopara.2004.02.006.

    Article  CAS  PubMed  Google Scholar 

  54. Dolo A, Modiano D, Doumbo O, Bosman A, Sidibe T, Keita MM, Naitza S, Robson KJ, Crisanti A: Thrombospondin related adhesive protein (TRAP), a potential malaria vaccine candidate. Parassitologia. 1999, 41 (1–3): 425-428.

    CAS  PubMed  Google Scholar 

  55. Blagborough AM, Sinden RE: Plasmodium berghei HAP2 induces strong malaria transmission-blocking immunity in vivo and in vitro. Vaccine. 2009, 27 (38): 5187-5194. 10.1016/j.vaccine.2009.06.069.

    Article  CAS  PubMed  Google Scholar 

  56. Ferguson DJ, Sahoo N, Pinches RA, Bumstead JM, Tomley FM, Gubbels MJ: MORN1 has a conserved role in asexual and sexual development across the apicomplexa. Eukaryot Cell. 2008, 7 (4): 698-711. 10.1128/EC.00021-08.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  57. Jani D, Nagarkatti R, Beatty W, Angel R, Slebodnick C, Andersen J, Kumar S, Rathore D: HDP-a novel heme detoxification protein from the malaria parasite. PLoS Pathog. 2008, 4 (4): e1000053-10.1371/journal.ppat.1000053.

    Article  PubMed Central  PubMed  Google Scholar 

  58. Musoke A, Rowlands J, Nene V, Nyanjui J, Katende J, Spooner P, Mwaura S, Odongo D, Nkonge C, Mbogo S, et al: Subunit vaccine based on the p67 major surface protein of Theileria parva sporozoites reduces severity of infection derived from field tick challenge. Vaccine. 2005, 23 (23): 3084-3095. 10.1016/j.vaccine.2004.09.039.

    Article  CAS  PubMed  Google Scholar 

  59. Freeman JM, Kappmeyer LS, Ueti MW, McElwain TF, Baszler TV, Echaide I, Nene VM, Knowles DP: A Babesia bovis gene syntenic to Theileria parva p67 is expressed in blood and tick stage parasites. Vet Parasitol. 2010, 173: 211-218. 10.1016/j.vetpar.2010.06.024.

    Article  CAS  PubMed  Google Scholar 

  60. Graham SP, Pelle R, Honda Y, Mwangi DM, Tonukari NJ, Yamage M, Glew EJ, de Villiers EP, Shah T, Bishop R, et al: Theileria parva candidate vaccine antigens recognized by immune bovine cytotoxic T lymphocytes. Proc Natl Acad Sci U S A. 2006, 103 (9): 3286-3291. 10.1073/pnas.0511273103.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  61. Criado-Fornelio A, Martinez-Marcos A, Buling-Sarana A, Barba-Carretero JC: Molecular studies on Babesia, Theileria and Hepatozoon in southern Europe. Part II. Phylogenetic analysis and evolutionary history. Vet Parasitol. 2003, 114 (3): 173-194. 10.1016/S0304-4017(03)00141-9.

    Article  CAS  PubMed  Google Scholar 

  62. Schnittger L, Rodriguez AE, Florin-Christensen M, Morrison DA: Babesia: A world emerging. Infect Genet Evol. 2012, 12 (8): 1788-1809. 10.1016/j.meegid.2012.07.004.

    Article  PubMed  Google Scholar 

  63. Haas BJ, Delcher AL, Mount SM, Wortman JR, Smith RK, Hannick LI, Maiti R, Ronning CM, Rusch DB, Town CD, et al: Improving the Arabidopsis genome annotation using maximal transcript alignment assemblies. Nucleic Acids Res. 2003, 31 (19): 5654-5666. 10.1093/nar/gkg770.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  64. Lowe TM, Eddy SR: tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 1997, 25 (5): 955-964.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  65. Allen JE, Majoros WH, Pertea M, Salzberg SL: JIGSAW, GeneZilla, and GlimmerHMM: puzzling out the features of human genes in the ENCODE regions. Genome Biol. 2006, 7 (Suppl 1): 1-13. 10.1186/gb-2006-7-s1-s1.

    Article  PubMed  Google Scholar 

  66. Cawley SE, Wirth AI, Speed TP: Phat–a gene finding program for Plasmodium falciparum. Mol Biochem Parasitol. 2001, 118 (2): 167-174. 10.1016/S0166-6851(01)00363-2.

    Article  CAS  PubMed  Google Scholar 

  67. Korf I: Gene finding in novel genomes. BMC Bioinforma. 2004, 5: 59-10.1186/1471-2105-5-59.

    Article  Google Scholar 

  68. Huang X, Adams MD, Zhou H, Kerlavage AR: A tool for analyzing and annotating genomic sequences. Genomics. 1997, 46 (1): 37-45. 10.1006/geno.1997.4984.

    Article  CAS  PubMed  Google Scholar 

  69. Liang C, Liu L, Ji G: WebGMAP: a web service for mapping and aligning cDNA sequences to genomes. Nucleic Acids Res. 2009, 37 (Web Server issue): W77-W83.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  70. Haas BJ, Salzberg SL, Zhu W, Pertea M, Allen JE, Orvis J, White O, Buell CR, Wortman J: Automated eukaryotic gene structure annotation using EVidenceModeler and the Program to Assemble Spliced Alignments. Genome Biol. 2008, 9: R7-10.1186/gb-2008-9-1-r7.

    Article  PubMed Central  PubMed  Google Scholar 

  71. Haas BJ, Wortman JR, Ronning CM, Hannick LI, Smith RK, Maiti R, Chan AP, Yu C, Farzad M, Wu D, et al: Complete reannotation of the Arabidopsis genome: methods, tools, protocols and the final release. BMC Biol. 2005, 3: 7-10.1186/1741-7007-3-7.

    Article  PubMed Central  PubMed  Google Scholar 

  72. Claudel-Renard C, Chevalet C, Faraut T, Kahn D: Enzyme-specific profiles for genome annotation: PRIAM. Nucleic Acids Res. 2003, 31 (22): 6633-6639. 10.1093/nar/gkg847.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  73. Karp PD, Paley S, Romero P: The Pathway Tools software. Bioinformatics. 2002, 18 (Suppl 1): S225-S232. 10.1093/bioinformatics/18.suppl_1.S225.

    Article  PubMed  Google Scholar 

  74. Ren Q, Chen K, Paulsen IT: TransportDB: a comprehensive database resource for cytoplasmic membrane transport systems and outer membrane channels. Nucleic Acids Res. 2007, 35 (Database issue): D274-D279.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  75. Enright AJ, Van Dongen S, Ouzounis CA: An efficient algorithm for large-scale detection of protein families. Nucleic Acids Res. 2002, 30 (7): 1575-1584. 10.1093/nar/30.7.1575.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  76. Thompson JD, Gibson TJ, Higgins DG: Multiple sequence alignment using ClustalW and ClustalX. Curr Protoc Bioinformatics. 2002, Chapter 2: 2-3.

    Google Scholar 

  77. Wilgenbusch JC, Swofford D: Inferring evolutionary trees with PAUP*. Curr Protoc Bioinformatics. 2003, Chapter 6: Unit 6 4-

    PubMed  Google Scholar 

  78. Ronquist F, Huelsenbeck JP: MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 2003, 19 (12): 1572-1574. 10.1093/bioinformatics/btg180.

    Article  CAS  PubMed  Google Scholar 

  79. Stamatakis A, Hoover P, Rougemont J: A rapid bootstrap algorithm for the RAxML Web servers. Syst Biol. 2008, 57 (5): 758-771. 10.1080/10635150802429642.

    Article  PubMed  Google Scholar 

  80. Hayashida K, Hara Y, Abe T, Yamasaki C, Toyoda A, Kosuge T, Suzuki Y, Sato Y, Kawashima S, Katayama T, et al: Comparative genome analysis of three eukaryotic parasites with differing abilities to transform leukocytes reveals key mediators of theileria-induced leukocyte transformation. MBio. 2012, 3 (5): e00204-e00212.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  81. Cornillot E, Hadj-Kaddour K, Dassouli A, Noel B, Ranwez V, Vacherie B, Augagneur Y, Bres V, Duclos A, Randazzo S, et al: Sequencing of the smallest Apicomplexan genome from the human pathogen Babesia microti. Nucleic Acids Res. 2012, 40 (18): 9102-9114. 10.1093/nar/gks700.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We would like to acknowledge the excellent technical assistance of Ralph Horn, James Allison, Melissa Flatt, and Shelby Beckner. David Schneider kindly assisted with microscopy. This research was supported by USDA-ARS-CWP 5348-32000-028–00D. JJG acknowledges support from NIAID contract HHSN272200900040C awarded to Bruno Sobral (Virginia Bioinformatics Institute) and NIH awards R01AI017828 and R01AI59118 (Abdu Azad, University of Maryland).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kelly A Brayton.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

LSK, MT, DRH, AD, JJG, JCS, VMN, RHM, DPK and KAB conceived and designed the experiments. LSK, MT, DRH, JDR, EC, AD, JJG, MGS, AOTL, HER, JCS, CES, and MWU performed the experiments. LSK, MT, DRH, JDR, EC, AD, JJG, MGS, AOTL, HER, JCS, CES, MWU, RHM, DPK, and KAB analyzed the data. LSK, DRH, DPK, and KAB wrote the paper. All authors read and approved the final manuscript.

Electronic supplementary material

12864_2012_4403_MOESM1_ESM.pdf

Additional file 1: Figure. Depiction of the T. equi apicoplast genome gene arrangement, showing unidirectional coding of genes. Known enzymes shown in red, ribosomal proteins in green, rRNA sequences in yellow, and groupings of tRNA molecules in blue. Location of conserved hypothetical (gray), and hypothetical (black) protein-encoding genes are shown by arrows or bars. Members of the three expanded gene families are marked with either “*”, “¡” or “^” to indicate similar genes. The molecule is depicted as linear, though not experimentally demonstrated to be either circular or linear. (PDF 131 KB)

Additional file 2: Table. Nuclear encoded genes potentially targeted to the apicoplast. (PDF 160 KB)

Additional file 3: Table. GPI anchored proteins predicted by GPI-SOM. (PDF 37 KB)

Additional file 4: Table. Transporter comparison B. bovis, T. equi, and T. parva.(PDF 76 KB)

Additional file 5: Table. 30 largest protein families. (PDF 22 KB)

12864_2012_4403_MOESM6_ESM.pdf

Additional file 6: Figure. Alignment of EMA family sequences. Residues highlighted in yellow are conserved among all family members, and those in blue conserved among the majority of family members. Dashes represent gaps introduced to accommodate non-conserved stretches of sequence. (PDF 32 KB)

Additional file 7: Table. Proteins used in phylogenetic analysis. (PDF 119 KB)

12864_2012_4403_MOESM8_ESM.pdf

Additional file 8: Figure. Phylogenetic trees. A: Most frequently recovered trees from maximum parsimony analysis of 150 polypeptides conserved among eight taxa, showing the number of times tree recovered out of 210 total topologies. Value above the line represents the number times that branch was recovered out of total times whole topology was recovered , and value below the line is percentage bootstrap support for that branch out of 1000 replicates. The bootstrap support for the individual MP trees was calculated for each individual dataset, and averages across all of the individual trees are presented. B: Single most parsimonious tree estimated from the concatenated dataset of the 150 polypeptides. Taxon codes are Cp: Cryptosporidium parvum, Tg: Toxoplasma gondii, Pf: Plasmodium falciparum, Pv: Plasmodium vivax, Bb: Babesia bovis, Te: Thieleria equi, Ta: Theileria annulata, Tp: Theileria parva. C: Trees estimated with maximum likelihood using six different models of amino acid substitution. Branch support was assessed with 1000 bootstrap pseudoreplications. (PDF 673 KB)

Authors’ original submitted files for images

Below are the links to the authors’ original submitted files for images.

Authors’ original file for figure 1

Authors’ original file for figure 2

Authors’ original file for figure 3

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Kappmeyer, L.S., Thiagarajan, M., Herndon, D.R. et al. Comparative genomic analysis and phylogenetic position of Theileria equi. BMC Genomics 13, 603 (2012). https://doi.org/10.1186/1471-2164-13-603

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2164-13-603

Keywords