Skip to main content
Advertisement
  • Loading metrics

Comparative and functional genomics of the protozoan parasite Babesia divergens highlighting the invasion and egress processes

  • Luis Miguel González ,

    Contributed equally to this work with: Luis Miguel González, Karel Estrada

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Laboratorio de Referencia e Investigación en Parasitología, Centro Nacional de Microbiología, ISCIII Majadahonda, Madrid, Spain

  • Karel Estrada ,

    Contributed equally to this work with: Luis Miguel González, Karel Estrada

    Roles Data curation, Methodology, Software, Validation, Visualization, Writing – review & editing

    Affiliation Unidad Universitaria de Secuenciación Masiva y Bioinformática, Instituto de Biotecnología, Cuernavaca, México

  • Ricardo Grande,

    Roles Data curation, Software, Validation, Writing – review & editing

    Affiliation Unidad Universitaria de Secuenciación Masiva y Bioinformática, Instituto de Biotecnología, Cuernavaca, México

  • Verónica Jiménez-Jacinto,

    Roles Data curation, Software, Visualization, Writing – review & editing

    Affiliation Unidad Universitaria de Secuenciación Masiva y Bioinformática, Instituto de Biotecnología, Cuernavaca, México

  • Leticia Vega-Alvarado,

    Roles Formal analysis, Software, Validation, Writing – review & editing

    Affiliation Instituto de Ciencias Aplicadas y Tecnología, UNAM, Ciudad de México, México

  • Elena Sevilla,

    Roles Data curation, Methodology, Software, Validation, Writing – review & editing

    Affiliation Laboratorio de Referencia e Investigación en Parasitología, Centro Nacional de Microbiología, ISCIII Majadahonda, Madrid, Spain

  • Jorge de la Barrera,

    Roles Formal analysis, Methodology, Software, Writing – review & editing

    Affiliation Unidad de Bioinformática, Área de Unidades Centrales Científico-Técnicas, ISCIII, Majadahonda, Madrid, Spain

  • Isabel Cuesta,

    Roles Methodology, Software, Writing – review & editing

    Affiliation Unidad de Bioinformática, Área de Unidades Centrales Científico-Técnicas, ISCIII, Majadahonda, Madrid, Spain

  • Ángel Zaballos,

    Roles Methodology, Software, Writing – review & editing

    Affiliation Unidad de Genómica, Área de Unidades Centrales Científico-Técnicas, ISCIII, Majadahonda, Madrid, Spain

  • José Manuel Bautista,

    Roles Investigation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of Biochemistry and Molecular Biology & Research Institute Hospital 12 de Octubre, Facultad de Veterinaria, Universidad Complutense de Madrid, Madrid, Spain

  • Cheryl A. Lobo,

    Roles Investigation, Supervision, Validation, Writing – original draft, Writing – review & editing

    Affiliation Blood Borne Parasites, LFKRI, New York Blood Center, New York, New York, United States of America

  • Alejandro Sánchez-Flores ,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Project administration, Resources, Software, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    alexsf@ibt.unam.mx (ASF); estrella.montero@isciii.es (EM)

    Affiliation Unidad Universitaria de Secuenciación Masiva y Bioinformática, Instituto de Biotecnología, Cuernavaca, México

  • Estrella Montero

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    alexsf@ibt.unam.mx (ASF); estrella.montero@isciii.es (EM)

    Affiliation Laboratorio de Referencia e Investigación en Parasitología, Centro Nacional de Microbiología, ISCIII Majadahonda, Madrid, Spain

Abstract

Babesiosis is considered an emerging disease because its incidence has significantly increased in the last 30 years, providing evidence of the expanding range of this rare but potentially life-threatening zoonotic disease. Babesia divergens is a causative agent of babesiosis in humans and cattle in Europe. The recently sequenced genome of B. divergens revealed over 3,741 protein coding-genes and the 10.7-Mb high-quality draft become the first reference tool to study the genome structure of B. divergens. Now, by exploiting this sequence data and using new computational tools and assembly strategies, we have significantly improved the quality of the B. divergens genome. The new assembly shows better continuity and has a higher correspondence to B. bovis chromosomes. Moreover, we present a differential expression analysis using RNA sequencing of the two different stages of the asexual lifecycle of B. divergens: the free merozoite capable of invading erythrocytes and the intraerythrocytic parasite stage that remains within the erythrocyte until egress. Comparison of mRNA levels of both stages identified 1,441 differentially expressed genes. From these, around half were upregulated and the other half downregulated in the intraerythrocytic stage. Orthogonal validation by real-time quantitative reverse transcription PCR confirmed the differential expression. A moderately increased expression level of genes, putatively involved in the invasion and egress processes, were revealed in the intraerythrocytic stage compared with the free merozoite. On the basis of these results and in the absence of molecular models of invasion and egress for B. divergens, we have proposed the identified genes as putative molecular players in the invasion and egress processes. Our results contribute to an understanding of key parasitic strategies and pathogenesis and could be a valuable genomic resource to exploit for the design of diagnostic methods, drugs and vaccines to improve the control of babesiosis.

Author summary

Babesiosis has long been recognized as an economically important disease of cattle, but only in the last 40 years has Babesia been recognized as an important pathogen in humans. Babesiosis in humans is caused by one of several species (B. microti, B. divergens, B. duncani and B. venatorum). The complete Babesia lifecycle requires two hosts, the ixodid ticks and a vertebrate host. It is the parasite's ability to first recognize and then invade host erythrocytes that is central to the pathogenesis of babesiosis. Once inside the cell, the parasite begins a cycle of maturation and growth, resulting in merozoites that egress from the red blood cells (RBCs) and seek new, uninfected RBCs to invade, perpetuating the infection. To better understand this asexual lifecycle, the authors focused on the parasite genome and transcriptome of the asexual erythrocytic forms of B. divergens. Through this functional and comparative genomic approach, the authors have identified genes putatively involved in invasion, gliding motility, moving junction formation and egress, providing new insights into the molecular mechanisms of these processes necessary for B. divergens to survive and propagate during its life cycle.

Introduction

Babesiosis is a worldwide emerging infectious disease [1] caused by a protozoan parasite of the genus Babesia which is naturally transmitted by ixodid ticks and infects vertebrate erythrocytes. Parasite infection of natural hosts such as in cattle causes bovine babesiosis, resulting in high economic losses. Recently, the severity of human infection has also become apparent [2]. The four identified Babesia species confirmed to infect humans to date are B. microti, B. divergens, B. duncani, and B. venatorum [2,3]. B. microti has a worldwide distribution while B. divergens is a causative agent of babesiosis in humans and cattle in Europe [3].

Symptoms range from asymptomatic to flu-like symptoms in healthy individuals. However, in asplenic, immune-compromised patients, children, elderly people and patient who are experiencing B. divergens infection, the disease can be severe. Thus, human babesiosis results in hemolytic anemia, jaundice and hemoglobinuria due to the lytic effects of parasite multiplication in the red blood cells (RBC). B. divergens infections are therefore considered medical emergencies and patients require immediate treatment [3].

As evident from its status as an emerging disease, the number of human babesiosis cases has experienced a global increase [1]. This number may still not fully reflect the actual numbers of infections as many cases go unreported or are misdiagnosed because the non-specific symptoms of babesiosis can hinder accurate diagnosis. European cases in particular may reflect this situation because babesiosis is not a notifiable disease in the continent and several cases likely go unreported.

This is supported by seroprevalence studies that indicate that the percentage of the European population infected with Babesia is much higher than the low clinical incidence and case reports to date [4]. Notably, these anti-babesial antibody studies also indicate that the detected antibodies were acquired by people who were exposed to the Babesia parasites [4]. In fact, antibody testing is important in identifying asymptomatic and mild cases when parasites may not be detectable in a blood smear or by PCR, suggesting the need for active surveillance programs to accurately detect babesiosis in Europe.

Originally endemic in just a few regions on the world, including Europe [3,4] and EEUU [2] the steady increase of cases in other areas of the world (Asia, Africa, Australia, Canada, Egypt and South America) indicates its zoonotic potential. Notably, the presence of B. divergens outside Europe and North Africa was recently confirmed in China and Japan [5,6].

Onset of symptoms in the host results from the continuous invasion and destruction of RBCs by the parasite. The invasion of RBCs by the B. divergens free merozoite is an active process [7] that exploits an original locomotion mechanism called gliding motility, an intimate tight connection between the parasite and the erythrocyte known as a moving junction and a coordinated secretion of proteins from apical organelles (rhoptries, micronemes, dense granules or spherical bodies according to the species) [810]. Once hidden within the host cell, the intraerythrocytic lifestyle contributes to protect parasites from the interactions of anti-babesial antibodies capable of recognized and block the antigens of the free merozoite surface. Parasites gain nutrients and multiply by binary fission resulting in a considerably complex pleomorphic process that yields different intraerythrocytic parasite stages leading to the development of merozoites that exit to invade new RBCs and thus propagate [7,11].

In vitro, this proliferative cycle [12] has been studied to investigate biological processes such as invasion, multiplication and egress [7,1214]. However, the dynamics and regulation of these processes are still largely unexplored at the molecular level in B. divergens.

To gain new insight into this emerging but neglected pathogen, the genome of B. divergens human isolate (Bd 87 strain) was recently sequenced and assembled by our group using reads produced by three next-generation sequencing (NGS) platforms. The draft genome (publicly available under accession number CCSG01000001 to CCSG01000514) was approximately 10.7-Mb in 514 scaffolds and revealed over 3,741 protein coding-genes [15].

In this study, we significantly improve the quality of assembly of the B. divergens genome by exploiting our previous sequence data using new computational tools and assembly strategies. We also report transcriptional data from two different stages of the B. divergens in vitro asexual lifecycle: i) free extra-cellular merozoite and ii) intraerythrocytic stage. Using genomic and transcriptomic data we provide an annotation-directed improved genome sequence assembly [16] which enable us to explore the synteny between B. divergens and other Babesia species. Moreover, this functional and comparative genomics approach alongside the gene expression profile of the two stages, highlight molecular features associated with key processes to the survival of the parasite, including invasion, gliding motility, moving junction formation and egress.

Results and discussion

Annotation-directed improvement of Babesia divergens genome

Previously, we reported a draft genome [15] for B. divergens, assembled using a hybrid approach with three different platforms and sequencing depths: Illumina (~310x); 454 (~20x); Pacific Biosciences (PacBio) (~20x). Reassembly of the genome delivered larger contigs and scaffolds with chromosome lengths. As mentioned, the assembly continuity has been improved here, as shown in Table 1, where the number of contigs and scaffolds has been reduced and their average length has increased. The latest version of the genome can be found under the accession number GCA001077455 (CCSG02000001-CCSG02000141).

thumbnail
Table 1. Assembly and parameters for B. divergens genome.

https://doi.org/10.1371/journal.pntd.0007680.t001

The new assembly showed a better continuity that can be inferred from the reduced number of scaffolds and contigs and their higher average length. The total bases in the assembly was reduced since several redundant contigs that were generated in the previous assembly, were accurately resolved, although the contig N50 value decreased which is attributable to a more uniform length distribution and probably because some contigs were broken in misassembled regions.

Comparative genomics

In the previous assembly, we generated 5,000 gene models, but in this work, the gene prediction was improved by using RNA sequencing (RNAseq) data from both free merozoites and intraerythrocytic parasite stages (details are explained below) and the corroboration of predicted genes by comparing their products against proteins from 17 different Babesia species. From a total of 40 million paired-end reads, 90% of these reads were successfully mapped to the genome, which allowed the validation of the 4,546 gene models. Additionally, genes were corroborated by ensuring that they had a protein ortholog with at least other Babesia species (S1 Table).

Interestingly, from all the 17 Babesia species analyzed in the protein orthologous comparison, B. microti and B. bovis had the most orthologous in common with B. divergens (3,251 and 3,738, respectively). Taking in account this orthology information, we could support the annotation of 3,386 proteins in B. divergens which we considered as high confidence annotation.

We also analyzed the clustering of orthologous proteins between the three species and found 2,660 groups where 1,427 were the core genes in the three species (S1 Table). According to our analysis, B. bovis presented 18 clusters with gene duplications while B. microti and B. divergens presented only 7 clusters with duplicated genes.

As a result of the ortholog comparison from the gene model prediction improvement, we decided to compare our latest assembly against the genomes of B. bovis and B. microti to explore the synteny between organisms. In Fig 1, the one-to-one orthologous connecting the chromosomes from B. divergens to B. bovis and B. microti are depicted, allowing us to observe that some of our scaffolds have similar chromosome resolution reported for the other species [17,18]. Interestingly, the arrangements in B. divergens are more similar to the organization of B. bovis than B. microti, especially between chromosomes 2 and 3 of the former species. Contigs 1 and 2 (~2.2 and 1.4 Mb) of our improved assembly have a high correspondence to B. bovis chromosomes 3 and 2, respectively. Contig 3 (1.12 Mb) has correspondence to the largest contig of chromosome 4, where we can observe a possible inversion and some regions rearranged in Contigs 4 and 6 (1.07 and 0.87 Mb). Notably, we observed some correspondence to chromosomes 1 and 3 from B. microti, with Contigs 3, 4, 5 (1.06 Mb), 6 and 7 (0.63 Mb), suggesting significant rearrangements between the two species. Chromosome 2 from B. microti presented some correspondence to Contig 1, but not as much as observed with B. bovis chromosome 3. It can be inferred that B. microti has more genomic rearrangements in comparison to B. bovis and B. divergens, which presented a higher number of syntenic regions. In fact, previous phylogenetic investigations confirmed that B. microti constitutes a piroplasmid lineage distinct from B. bovis and B. divegens [19]. It is noteworthy that the three genomes were sequenced and assembled using different sequencing technologies and assembly algorithms, which could hinder the synteny comparison. Since none of them have a finished genome quality, there will be limitations due to missing parts in each genome.

thumbnail
Fig 1. Circos plot of B. divergens synteny with B. bovis and B. microti chromosomes.

Lines show the one-to-one ortholog comparison from each contig to chromosomes in other species. B. divergens contig names are indicated in blue, B. bovis chromosome names in red and B. microti chromosome names in green.

https://doi.org/10.1371/journal.pntd.0007680.g001

Additionally, we performed a pulse field gel electrophoresis (PFGE) for the genomic DNA and the results showed 4 probable chromosomes with the following approximate sizes of 1.05 (chromosome 1), ~1.7–1.8 (chromosome 2), ~2.9–3.0 (chromosome 3) and 4 Mb (chromosome 4) (S1 Fig). The size and number of chromosomes are also similar to those reported for the B. bovis genome [20] where chromosomes 1, 2, 3 and 4 are estimated to be 1.4, 2.0, 2.8 and 3.2 million base pairs, respectively. These results suggest that the probable ordering for our assembly, where we could assign contigs/scaffolds into chromosomes, is very similar to that observed in B. bovis. It is important to mention that although the B. bovis genome is the best Babesia reference genome available, only chromosomes 2 and 3 are finished and the rest are reported in several contigs [17]. In these cases, the use of genetic markers or FISH techniques [21] are necessary to corroborate the correspondence of the assembly into chromosomes and have achieve better synteny resolution.

Sequencing and dynamic transcriptome analysis of B. divergens asexual cycle

Free merozoites and intraerythrocytic parasites (Fig 2) represent the result of the cyclic processes of invasion and egress that B. divergens undergoes during its asynchronous asexual life cycle. Hence, we have sequenced the transcriptomes of both free extracellular parasites and intraerythrocytic stages using Illumina sequencing (S2 and S3 Tables).

thumbnail
Fig 2. B. divergens free merozoites and intraerythrocytic stages observed in in vitro cultures at ≈ 40% of parasitemia by Giemsa staining.

Figure shows free merozoites and intraerythrocytic parasite stages: (1) Multiple parasites stage formed by more than four parasites within the RBC, (2) paired pyriform, (3) free merozoite, (4) double trophozoites, (5) tetrad (6) single trophozoite, (7) double paired pyriforms, (8) transient morphological combination. Bar 15 μm.

https://doi.org/10.1371/journal.pntd.0007680.g002

To profile both transcriptomes, we isolated mRNA from two biological replicates to generate two independent RNAseq libraries for each stage.

We compared the RNAseq expression profile of both free merozoite and intraerythrocytic parasite stages using three different algorithms: DEseq, EdgeR and NOIseq. The NOIseq analysis yielded 1,441 differentially expressed genes, 734 of them were up-regulated and 707 genes down-regulated in the intraerythrocytic parasite stage (S4S6 Tables).

Based on these data and considering that molecular studies in term of host cell invasion and egress are limited in Babesia [22], we restricted our results to 45 genes of interest putatively related to both processes. These genes were manually inspected and curated using Artemis. Of note, 33 of the genes were confirmed by at least one of the three different algorithms and their expression levels were validated for free merozoite and intraerythrocytic stages using real-time quantitative reverse transcription PCR (qRT-PCR) (Table 2).

thumbnail
Table 2. List of genes exhibiting differential expression between B. divergens intraerythrocytic parasites and free merozoites.

https://doi.org/10.1371/journal.pntd.0007680.t002

The real-time qRT-PCR analysis revealed that 26 genes putatively related to the components of the molecular invasion machinery and 3 genes related to the egress process (Table 2) [2225] were upregulated between 0.2 and 2-fold in the intraerythrocytic stage. This result showed a moderately increased expression level of these genes in the intraerythrocytic stage compared with the free merozoite stage (Table 2 and Fig 3). However, 4 genes encoding proteases putatively involved in invasion or egress: i) BDIVROU_0278701.t1.2 (rhomboid-like protease 4.1, ROM4.1), ii) BDIVROU_0230400.t1.2 (rhomboid-like protease 4.4, ROM4.4), iii) BDIVROU_0280000.t1.2 (subtilisin-like serine protease, BdSUB1) and iv) BDIVROU_0161200.t1.1 (mac/perforin protein 2, MAC2) [2629] were modestly induced between 0.2 and 0.5-fold in the free merozoite (Table 2 and Fig 3).

thumbnail
Fig 3. Transcriptomic profiling of both intraerythrocytic stages (above) and free merozoite (below) validated by RNAseq and real-time qRT-PCR.

The values above the X-axis indicate the intraerythrocytic parasite up regulated genes and the values below the X-axis indicate free merozoite up regulated genes. Differential expression profile in both stages is denoted with 1, 2, 3 or 4 asterisks.

https://doi.org/10.1371/journal.pntd.0007680.g003

RNASeq and real-time qRT-PCR findings indicated that genes of specialized B. divergens functions could be present in both the free merozoite and intraerythrocytic stages, suggesting that the majority of the genome is utilized throughout the complete asexual lifecycle of this parasite. Notably, the upregulation of some genes in the intraerythrocytic stage could imply a moderately increased need for invasion, gliding motility, and egress protein expression, before these processes occur (Table 2). Alternatively, a moderate or less marked induction of the same genes in the free merozoite would be required to maintain baseline expression level during the processes. Moreover, this expression profile suggests that the intraerythrocytic parasite could harbor proteins in its apical organelles that are required during the rapid processes of invasion and egress, thus ensuring that the complex and specialized mechanisms of invasion, gliding, and egress are primed and function successfully at the appropriate time. Surprisingly, the free merozoites also exhibit qualitatively a similar expression profile, although the transcripts are at lower levels. This may be linked to the short lifecycle of the parasite and may reflect that these transcripts need to be present to accommodate the requirement for the various proteins as soon as the parasite invades. It will be of interest to check if the protein levels are also equivalent between these two parasite stages.

On the other hand, since the asexual life cycle of B. divergens is asynchronous, there are simultaneously intraerythrocytic parasites at different stages of their development, ranging from parasites which have just invaded new erythrocytes to parasites which are dividing or egressing from the host cell. As a result, there are not apparent abrupt changes in the expression profile between intracellular and extracellular stages. This is reflected in the data dispersion observed for free merozoites replicates in the principal component analysis (PCA) and multidimensional scaling (MDS) plots (S2 Fig.) generated by DESeq2 and EdgeR, respectively. However, for those proteins transcribed in two different phases of the lifecycle, i.e. when the parasite is outside the RBC and when it is inside the RBC, regardless their expression level, the research could lead towards the development of vaccines and new treatments capable of targeting several different stages at the same time, as previously suggested for P. falciparum [30].

The composition and assembly of the invasion and egress molecular machineries of B. divergens

Finally, on the basis of this genomic approach and in the absence of models of invasion and egress available for B. divergens, we have proposed the following potential molecular models according to the B. divergens genes identified and validated in this work (S7 and S8 Tables) and following the general prototypes suggested for the apicomplexan parasites Plasmodium and Toxoplasma [2225].

a) B. divergens genes associated with calcium (Ca2+) in invasion

Since Ca2+-dependent proteins and Ca2+ signaling are important for microneme exocytosis and invasion in apicomplexan parasites [22], we identified the following B. divergens candidates genes that encode proteins related to Ca2+: i) BDIVROU_0339410.t1.2 (protein kinase G, PKG), ii) BDIVROU_0124800.t1.2 (phosphoinositide phospholipase C, PI-PLC), iii) BDIVROU_0071500.t1.1 (diacylglycerol kinase 1, DGK1), iv) BDIVROU_0291600.t1.1 (acylated pleckstrin homology domain containing protein, APH), v) BDIVROU_0224310.t1.2 (Ser/Thr phosphatase, PP1 or calcineurin), vi) BDIVROU_0130101.t1.2 (calcium-dependent protein kinase 4, CDPK-4) and vii) BDIVROU_0416400.t1.1 (double C2 domain protein, DOC2) (Table 2 and Fig 4). The identification of these genes suggests that Ca2+ and calcium–dependent proteins are involved in B. divergens invasion in a similar manner to other apicomplexans. Although PKG activates microneme secretion and parasite egress in a calcium-independent manner [25], this protein is associated with the release of Ca2+ from the endoplasmic reticulum (ER) to the cytoplasm and is mediated by cyclic guanosine monophosphate (cGMP). PI-PLC is then activated by cGMP-dependent PKG, thus leading to hydrolysis of phosphatidylinositol 4,5-bisphosphate (PIP2) to diacylglycerol (DAG) and inositol 1,4,5-trisphosphate (IP3), each of which plays a different role [31,32]. DAG is converted to phosphatidic acid by DGK1. Phosphatidic acid is a mediator of microneme secretion and is associated with APH, a surface microneme protein that acts as a phosphatidic acid–detecting protein. Depletion of APH leads to blockade of micronemes and, consequently, to inactivation of the motile machinery of the parasite or glideosome during invasion [33]. Moreover, IP3 stimulates an influx of Ca2+ from the ER surface to the cytoplasm [31,34]. Simultaneously, this influx of Ca2+ activates calcium-dependent proteins as calcineurin, DOC2 and CDPK4 which interacts with PKG, thus favoring attachment of the merozoite to the RBC, the secretion of several micronemal proteins [35] and the control of the actomyosin motor to ensure an efficient gliding motility for the parasite [36].

thumbnail
Fig 4. A molecular invasion model proposed for B. divergens.

The scheme represents a hypothetical parasite-host complex that is translocated from the front of the parasite to the back by an actomyosin system during invasion. The free merozoite shows Ca2+–dependent proteins that could govern the secretion of transmembrane adhesive proteins from the micronemes during invasion in a Ca2+ dependent manner. These adhesins are translocated to the parasite plasma membrane (PM) and may act as ligands for host cell receptors. During invasion, parasite sheddases cleave adhesins to disengage interactions with RBC receptors. The scheme on the left shows a magnified view of the potential candidates related to the actin polymerization and glideosome, which localized between the inner membrane complex (IMC), the parasite PM and the membrane of the red blood cell (RBC-PM) powers the merozoite motile process. The scheme on the right shows a magnified view of the B. divergens potential candidates that facilitate parasite-RBC receptor connection and allow the moving junction formation. FMR, formin; GAP40, gliding associated protein 40; GAP45, gliding associated protein 45; GAP50, gliding associated protein 50; MYOA, myosin motor protein; MLC1, myosin light chain1; GAC glideosome-associated connector; RON2, rhoptry neck protein 2; RON4, rhoptry neck protein 4; RON5, rhoptry neck protein 5.

https://doi.org/10.1371/journal.pntd.0007680.g004

b) B. divergens genes related to actin dynamics, glideosome and moving junction

The actomyosin-based molecular motor or glideosome and the moving junction facilitate the entry of apicomplexan parasites into the host cell. These molecular systems are well characterized in Plasmodium and Toxoplasma [23,25] and we found the following six implicated genes from both within B. divergens genome: i) BDIVROU_0025700.t1.2 (actin); ii) BDIVROU_0303400.t1.1 (formin-a, FMRa); iii) BDIVROU_0036900.t1.1 (formin-b, FMRb); iv) BDIVROU_0096510.t1.2 (actin-depolymerizing factor, ADF or cofilin); v) BDIVROU_0279300.t1.2 (profilin); and vi) BDIVROU_0073701.t1.2 (F-actin–binding protein or coronin) (Table 2 and Fig 4). The identification of these genes suggests that the movement of B. divergens during invasion is mediated by actin in a similar manner to other apicomplexan parasites [37]. Actin is found as a globular monomer or G-actin and polymerizes into filaments of F-actin via a nucleation phase mediated by formins that localize at the apical pole of invading merozoites [38]. In contrast, ADF and profilin help to maintain a high actin content in the form of a monomer (G-actin) and a low concentration of polymerized actin (F-actin) by sequestering free G-actin [3942] while coronin participates in the organization of F-actin filaments, the regulation of the motility [43] or mediating multiple parasites functions [44].

Actin dynamics and glideosome power the parasite locomotion [25], thus we identified six B. divergens genes related to the glideosome machinery: i) BDIVROU_0242901.t1.2 (myosin heavy chain A, MYOA); ii) BDIVROU_0412101.t1.2 (myosin light chain 1, MTIP for Plasmodium spp and MLC1 for T. gondii); iii) BDIVROU_0339200.t1.2 (gliding associated protein 50, GAP50); iv) BDIVROU_0272901.t1.2 (gliding associated protein 45, GAP45); v) BDIVROU_0052401.t1.2 (gliding associated protein 40, GAP40); and vi) BDIVROU_0152200.t1.2 (glideosome-associated connector, GAC) (Table 2 and Fig 4).

Similar to other apicomplexan parasites, the glideosome of B. divergens may be located between the plasma membrane and the inner membrane complex (IMC) of the parasite and comprises a similar molecular structure formed by MYOA, which in association with MLC1 [45,46], GAP45, GAP50 and GAP40 generates the traction force for motility and entry [38,47]. As an alternative to the MYOA motor, we predicted a second gene BDIVROU_0133800.t3.2 that encodes the myosin heavy chain B (MYOB), and it may play a similar role in parasite motility, thus suggesting that the movement in B. divergens could be governed by alternative myosin motors [48].

GAP proteins anchor the B. divergens glideosome to the IMC and the cytoskeleton [47,49,50], while the GAC connector attaches it to the plasma membrane [51]. This last protein specifically connects the glideosome to a transmembrane adhesin (TgMIC2 for T. gondii) that may interact with a host cell receptor to propel the parasite forwards [52]. In the B. divergens genome and transcriptomes, we found at least five adhesins as potential candidates to connect the parasite to the RBC: i) BDIVROU_0156200.t1.2 that encodes the orthologous thrombospondin-related apical membrane protein (TRAP for Plasmodium and TgMIC2 for T. gondii, mentioned above) [24]; ii) BDIVROU_0126000.t1.2 that codes for the apical membrane antigen BdAMA1, that binds to a trypsin- and chymotrypsin-sensitive receptor on the RBC [53]; iii) BDIVROU_0182900.t1.2 that encoded the glycosyl phosphatidyl inositol (GPI) anchored protein Bd37 involved in the penetration of B. divergens using a potential endocytosis mechanism [54,55]; iv) BDIVROU_0183000.t1.2 that codes for a putative 50-kDa surface protein (BdP50) and v) two identical copies (BDIVROU_0113100.t1.2 and BDIVROU_0112800.t1.1) of the genes that code the rhoptry-associated protein 1, RAP-1, a protein that binds to a RBC receptor [56].

In addition, we confirmed two more genes from the RAP family: i) BDIVROU_0112900.t1.2 that encoded a protein of 492 aa (RAP-1b) with an estimated molecular mass of 56.38 kDa, the amino acid sequence exhibited significant percentages of similarity to other known RAP proteins as B. bovis RAP-1 (83%) [57], B. divergens RAP-1 (72%) [56], B. ovata RAP-1 (79%) [58] and B. gibsoni RAP-1b (63%) [59] and ii) BDIVROU_0117600.t1.2 encoding a ropthry protein [60] that showed 34.4% similarity to the RAP-1 related antigen (RRA) of B. bovis, a protein that may contribute to erythrocyte invasion [61]. The amino acid sequence of RAP-1, RAP1-b and RRA from B. divergens exhibited the common features of the RAP family [61] such as a signal peptide, the presence of the strict conserved 4 cysteine residues and the 14 amino-acid motif. The genome sequence showed the B. divergens rap-1 locus extending over about 7.3 Kb and contains two identical rap-1 genes interspaced with the rap-1b gene. Rap-1 and rra genes were located on the same contig but the rra gene was found around 96.8 kb downstream from the RAP-1 locus. None of the genes were flanked by other rap-1 like genes or any gene related to the invasion or egress processes.

Moreover, for successful cell invasion, apicomplexan parasites establish a moving junction that moves from the apical to the posterior end of the parasite [23]. We identified at least three B. divergens genes implicated in the formation of this system: i) BDIVROU_0259100.t1. (rhoptry neck protein 2, RON2 [56]); ii) BDIVROU_0191501.t1.2 (rhoptry neck protein 4, RON 4); and iii) BDIVROU_0142300.t1.2 (rhoptry neck protein 5, RON5) (Fig 4).

Similar to Plasmodium and Toxoplasma, B. divergens RON proteins could constitute a ring-shaped structure in the host cell membrane. This structure together with AMA1 enables a tight connection between P. falciparum or T. gondii and their respective host cells leading to the formation of the moving junction [23,25,6265]. However, as we mentioned above B. divergens AMA1 binds to a specific receptor on the RBC providing an attachment between the merozoite and the host cell [53], thus suggesting that the constitution of the moving junction core could be different between B. divergens and other species and could involve a host cell receptor. Alternative or multiple roles for B. divergens AMA1 and specific adaptation of this protein to a particular B. divergens merozoite-host cell interaction could be also considered [52].

Interestingly, we identified the gene BDIVROU_0165301.t1.2 (Table 2) that belongs to the spherical body protein family (SBP) of the genus Babesia which seem to be involved in stabilizing the environment after invasion and in aiding parasite growth [66]. We also identified the gene BDIVROU_0161200.t1.1 that encodes a subtilisin-like serine protease (BdSUB1) that play a relevant role during invasion [27] and five novel genes: i) BDIVROU_0278701.t1.2 (ROM4.1), ii) BDIVROU_0280300.t1.1 (ROM4.2), iii) BDIVROU_0439300.t1.1 (ROM4.3), iv) BDIVROU_0230400.t1.2 (ROM4.4) and v) BDIVROU_0449600.t1.1 (ROM4.5) that showed homology with rhomboid-like protease ROM4, suggesting the presence of a novel B. divergens ROM family.

The homologous T. gondii ROM4 and the P. falciparum subtilisin-like serine protease PfSUB2 cleaves the transmembrane domain of adhesins (AMA1, and TRAP) to disengage their interactions with the host cell for parasite entry [26,67,68]. Thus, the ROM protease family detected in the B. divergens asexual life cycle could aid in the shedding of surface adhesins as well.

In addition to shedding, P. falciparum AMA1 is phosphorylated by the cAMP-dependent protein kinase (cAMP-PKA) an enzyme that was recently shown to be involved in merozoite invasion [69]. This gene has also been identified for B. divergens in the present study (BDIVROU_0165500.t1.1).

c) B. divergens genes related to egress of the parasite

Egress is an ordered, regulated process assisted by a cascade of proteolytic activities that occur seconds before the parasite exits the host cell [24,70].

For example, perforins, that exhibit calcium-dependent lytic activity, form pores in the parasitophorous vacuole membrane of Plasmodium and in the plasma membrane of the host cell facilitating the egress process [28,29,71]. Interestingly, four B. divergens genes with homology to mac/perforin proteins (MAC) were also found in this work, these are: i) BDIVROU_0064900.t1.1 (MAC1), ii) BDIVROU_0161200.t1.1 (MAC2), iii) BDIVROU_0065210.t1.2 (MAC3), and iv) BDIVROU_0173601.t1.2 (MAC4) (Table 2). As reported in the literature, cysteine proteases are also involved in the asexual life cycle. Specifically, papain proteases play a dual role in haemoglobin digestion and parasite egress [7274]. We identified eight B. divergens genes encoding cysteine proteases. Three of them belong to the papain family: i) BDIVROU_0308600.t1.1 (papain- 2), ii) BDIVROU_0202300.t2.1 (vignain like protease) and iii) BDIVROU_0197200.t1.1 (papain-like cysteine protease).

Interestingly, we also have curated the remaining 5 genes: i) DIVROU_0190900.t1.1 (calpain 7), ii), BDIVROU_0315500.t1.1 (CPC1-like protease), iii) BDIVROU_0028600.t1.1 (CPC2-like protease), iv) BDIVROU_0012300.t1.1 (B. divergens OTU‐like cysteine protease, BdOTU) and v) BDIVROU_0273700.t1.1 (B. divergens autophagy-related peptidase, BdATG4). These genes, have been previously described in P. falciparum, as essential for cell cycle progression [75] apicoplast homeostasis [76] and growth [77]. Therefore, the presence of these homologs suggests a similar role in B. divergens.

Conclusions

Our study provides novel and comprehensive information on the identification of a large panel of proteins involved in the critical processes of invasion and egress of B. divergens. The strategy developed shows how genome sequencing and RNAseq can further our understanding of transcriptional patterns of B. divergens blood stages. Although apicomplexan genomes differ in terms of sequencing data and assembly strategy, orthologous proteins related to biological processes that have been reported for other parasites can be found in B. divergens.

Thus, we have proposed multiple invasion and egress molecular candidates for B. divergens guided by the cross-comparison with the related Apicomplexan parasites, Plasmodium and Toxoplasma. Based on our results, we confirmed significant similarities with both parasites but also some key differences in the composition of the B. divergens invasion machinery (Fig 4). Therefore, future studies are needed for defining conserved versus evolutionary adaptations of the principal molecules involved in these biological processes between Apicomplexan species. Moreover, our transcriptomic analysis revealed the expression profile of the B. divergens genome and showed which genes are active at a particular point in time. However, the genetic expression landscape observed in B. divergens free merozoite and intracellular life stages are pieces of a complex jigsaw puzzle that needs to be solved by using other techniques such as proteomics and metabolomics.

Furthermore, we will extend our analysis to the transcriptome of other clinical samples and/or isolates, especially those from hosts infected by B. divergens. Thus, we will be able to compare differences between the functional expression profile of the parasite in vivo during active infection and in vitro. Finally, we consider that the results generated in this work, can guide other efforts in the improvement of diagnosis and detection, drug target and vaccine development for the control of this emerging and neglected disease.

Methods

Ethics statement

Human A+ blood from healthy volunteer donors was used to maintain B. divergens blood stage cultures. Protocols for the use of this blood were approved by the Blood Transfusion Center, Madrid, Spain. Donors provided informed written consent for use of their blood for research purposes.

In vitro culture of B. divergens

B. divergens (Bd Rouen 1987 strain) was propagated in vitro in human A+ RBCs. High parasitemia asynchronous parasite cultures were prepared at 40% parasitemia according to previously published protocols [9,78].

Pulse field gel electrophoresis (PFGE)

PFGE was used to determine chromosome number and size of the B. divergens genome. Asynchronous cultures were centrifuged at 1224 x g for 5 min and washed with RPMI 1640 (Life Technologies Corporation, Carlsbad, CA). The resultant pellets, containing intact cells, were embedded in 0.8% agarose plugs using the clamped homogeneous electric field (CHEF) mammalian genomic DNA plug Kit and treated with proteinase K according to the manufacturer’s instructions (Bio-Rad Labs Inc., Hercules, CA). The CHEF-DR III system (Bio-Rad labs Inc.) was used to separate the intact chromosomes of B. divergens. The PFGE conditions used for a 0.1–2.0 Mb DNA size range were: 6 V cm-1, pulses of 60–120 s for 24 h and an angle of 120°. The conditions used for a 1.8–4.6 Mb DNA size range were: 4.5 V cm-1, pulses of 200 s for 48 h and an angle of 106°. The gels were stained and photographed under ultraviolet transillumination.

Genome assembly, post-assembly improvement, gene prediction and annotation

The de novo genome assembly was performed using a hybrid approach with three different sequencing technologies. The first de novo assembly was obtained using the Newbler software [79] with 454 sequencing data and default parameters. Then, a second de novo assembly with Illumina paired-end and mate paired sequencing data, was performed using AllPaths-LG assembler [80] using default parameters. Then, 454 and Illumina scaffolds were merged using the genome assembler, reconciliation and merging (GARM) meta-assembler [81] with default parameters to generate a consensus assembly. Subsequently, SSPACE [82] and iterative mapping and assembly for gap elimination (IMAGE) [83] were used to improve scaffolding and gap filling, respectively. Finally, we used PacBio sequencing data that was previously corrected with Iterative Correction of Reference Nucleotides (iCORN) [84] using five iterations with Illumina reads, to increase the continuity of the assembly using the software Patch [80] with default parameters. The latest version of the genome can be found with the accession number GCA001077455 (CCSG02000001-CCSG02000141) and associated to the BioProject PRJEB6536.

For gene prediction, we used Braker v2.0.4 and AUGUSTUS v3.0.3 using a set of proteins that were obtained from the previous version of the assembly [15] and with the support of RNAseq data. We corroborated the predicted gene models by comparing them against proteins from other Babesia species deposited in the NCBI database and filtering them to 90% identity to remove redundancy. Using Proteinrtho v5.15 [85], we retained those models with at least one ortholog in other species. The following list of Babesia species were used: B. bennetti, B. major, B. odocoilei, B. ovata, B. orientalis, B. capreoli, B. ovis, B. canis, B. rossi, B. rodhaini, B. equi, B. caballi, B. divergens, B. bigemina, B. gibsoni, B. microti and B. bovis. Finally, protein annotation was achieved using the Trinotate v3.0.1 pipeline [86].

Comparative genomics

For the comparative genomic analysis, the genomes available at Genbank for B. bovis (GCA_000165395.1) and B. microti (GCA_000691945.2) were used. The Proteinortho v5.15 [85] was used to obtain protein clusters of Babesia genomes analyzed here with the following parameters: E-value for blast: 1e-05, min. percent identity of best blast alignments: 25, min. coverage of best blast alignments in percent: 50, min. similarity for additional hits: 0.95. The results from the protein orthologous was also used to verify the annotation. Additionally, we used the Circos [87] package for circular data visualization and we were able to observe chromosome rearrangements among B. divergens, B. bovis and B. microti.

Isolation of B. divergens parasites and RNA preparation

Free merozoites and intraerythrocytic parasites were collected from two highly parasitized independent asynchronous B. divergens cultures, 75 ml each at parasitemias of ≈ 40%. Cultures were centrifuged at 600 x g and 4°C for 5 min in fixed-angle rotor heads to collect the supernatant containing the free merozoites and pellets containing the intraerythrocytic forms. Free merozoites were isolated [9] and resuspended in Trizol LS Reagent (ThermoFisher). Pellets containing the intraerythrocytic forms were resuspended in RPMI 1640 and washed 3 times by centrifugation at 600 x g for 10 min in order to remove possible traces of free merozoites. Pellets were diluted to 1:3 with RPMI 1460. This suspension was then layered on top of a 1.122 g/ml isosmotic Percoll solution prepared according to the manufacturer´s instructions (GE Healthcare, Freiburg, Germany) and centrifuge (2000 x g, 10 min). The infected erythrocytes collected at the interface were washed 3 times in PBS at 4°C and 600 x g for 10 min and resuspended in Trizol LS Reagent (ThermoFisher). Total RNA from B. divergens free merozoites and intraerythrocytic parasites was prepared using Trizol LS Reagent (ThermoFisher) and chloroform extraction. DNA was removed using the QIAGEN RNase-Free DNase Set (Qiagen Inc. Valencia CA) and cytoplasmic ribosomal RNA was removed using the Ribo-Zero removal kit (Illumina, San Diego, CA, USA). The RNA was measured on an Agilent 2100 Bioanalyzer using an RNA 6000 Pico Chip kit (Agilent Technologies, Inc, Santa Clara, California) and used for libraries preparation and real-time PCR assays (S2 Fig.).

RNA sequencing library preparation

Libraries were prepared using the Illumina TruSeq RNA Sample Prep Kit v2 (Illumina) following the manufacturer’s protocol. High quality RNA samples from free merozoites (FM1 and FM2) and from intraerythrocytic stages (IE1 and IE2), with RIN numbers ranged from 8.8 to 10, were used for NGS library construction (S2 Fig). FM1, FM2, IE1 and IE2 biological replicates were used to prepare two independent libraries for each stage. A third technical replicate was also prepared for each stage mixing FM1 and FM2 and IE1 and IE2 respectively. The libraries were sequenced using the Illumina HiSeq2000 platform with a paired-end configuration using 202 cycles (2x101 reads).

RNAseq data was deposited at NCBI under the Biosample accession number SAMN12187113 and included in the BioProject PRJNA552284.

Differential expression analysis

Differential expression analysis was performed using two biological replicates through the integrated differential expression analysis multi-experiment (IDEAmex) website (http://zazil.ibt.unam.mx/ideamex/) using three different differential expression packages: edgeR [88], DESeq2 [89] and NOISeq [90]. EdgeR and NOISeq were performed by applying trimmed mean of M values (TMM) [91] as the normalization method. To identify differentially expressed genes, genes whose p value was less or equal to 0.01 and log fold change (FC) greater or equal than 1.5, were selected for each method. The genes reported as differentially expressed in Table 2, were inspected and curated manually using Artemis [92] and selected as candidates for further analysis.

Real-time qRT-PCR verification of RNAseq data

Transcriptomic analyses were validated by real-time qRT-PCR using the same RNA extractions that had been prepared previously for RNAseq sequencing. Two μg each of free merozoite and intraerythrocytic parasite RNA was used for cDNA synthesis using the High Capacity RNA-to-cDNA kit (Applied Biosystems by ThermoFisher) following the manufacturer’s instructions. Controls without reverse transcriptase were used to investigate potential gDNA contamination by PCR using actin-specific primers. A set of 33 genes was tested by real-time qRT-PCR. Primers (S9 Table) were designed using Primer3 software [93] and were validated by testing amplification efficiency using 10-fold dilutions of cDNA by real-time qRT-PCR and melt curve analysis. Real-time qRT-PCR measurements were performed using the Qiagen Rotor-Gene-Q Real-Time PCR Detection System. Reactions were prepared in triplicated in a total volume of 20 μl using the SYBR Premix Ex Taq (Takara-Clontech, Otso, Japan), 20 ng of cDNA and primer concentrations of 0.2 mM. Transcript expression levels were calculated with the 2−ΔCt method using the endogenous control gene BDIVROU_0242600.t1.1 encoding for B. divergens endonuclease/exonuclease/phosphatase family domain containing protein, EnExPh as references.

Supporting information

S1 Fig. Pulsed field gel electrophoresis (PFGE) of intact chromosomes using Clamped homogeneous electric field (CHEF) system.

(A) The PFGE assay resolves the separation of 0.1–2.0 Mb DNA fragments and provides the approximate measurement of chromosome lengths in a 1% agarose gel: 1) Saccharomyces cerevisiae strain YNN295 DNA marker (the manufacture´s estimates of the sizes of S. cerevisiae chromosomes are indicated on the left of the picture); 2) B. divergens genomic DNA showing two bands at 1.05 Mb (chromosome 1) and 1.7–1.8 Mb (chromosome 2). (B) The PFGE assay resolves the separation of 1.8–4.6 Mb DNA fragments and provides the approximate measurement of chromosome lengths in a 0.8% agarose gel: 1) Hansenula wingei DNA marker; 2) Schizosaccharomyces pombe marker (the manufacture´s estimates of the sizes of H. wingei and S. pombe chromosomes are indicated on the left of the picture); 3) B. divergens genomic DNA showing two bands at 2.9–3.0 Mb (chromosome 3) and 4.0 Mb (chromosome 4).

https://doi.org/10.1371/journal.pntd.0007680.s001

(TIF)

S2 Fig. Graphics show an analysis of B. divergens total RNA samples with the Agilent 2100 bioanalyzer, a principal component analysis (PCA) plot displaying variation and account of RNAseq samples and a multidimensional scaling (MDS) plot displaying the relative positions of RNAseq samples generated by DESeq2 and EdgeR, respectively.

Panel A: The electrophoresis of B. divergens FM1, FM2, IE 1 and IE 2 RNA samples shows a visual inspection of RNA integrity. Panel B: RNA Integrity Number (RIN) of FM1, FM2, IE 1 and IE 2 RNA samples. Panel C: PCA plot of FM1, FM2, IE 1 and IE 2 RNAseq samples. Panel D: MDS plot of FM1, FM2, IE 1 and IE 2 RNAseq libraries. FM, free merozoite; IE, intraerythrocytic parasites.

https://doi.org/10.1371/journal.pntd.0007680.s002

(TIF)

S1 Table. List of genes identified in B. divergens and their B. bovis and B. microti orthologous.

Gene IDs for B. bovis and B. microti correspond to the NCBI annotation. #Species: 1, the gene was identified in one species; 2, the gene was identified in two species contain in their genomes; 3, the gene was identified in three species. Genes: total number of genes identified for the three species. Alg.Conn.: proteinortho algebraic connectivity. *absent gene.

https://doi.org/10.1371/journal.pntd.0007680.s003

(CSV)

S2 Table. List of annotated B. divergens genes.

The columns contain the BlastP, PFAM HMMER search, SignalP, KEGG, COG and EC number annotation results, respectively. For BlastP and PFAM results, the Evalue is attached with the “^” character, to the database target ID result. NA = not available.

https://doi.org/10.1371/journal.pntd.0007680.s004

(TXT)

S3 Table. Genome sequences of the annotated genes from the B. divergens genome.

https://doi.org/10.1371/journal.pntd.0007680.s005

(ZIP)

S4 Table. RNAseq expression profile of both free merozoite and intraerythrocytic parasite stages using DESeq.

https://doi.org/10.1371/journal.pntd.0007680.s006

(TXT)

S5 Table. RNAseq expression profile of both free merozoite and intraerythrocytic parasite stages using EdgeR.

https://doi.org/10.1371/journal.pntd.0007680.s007

(TXT)

S6 Table. RNAseq expression profile of both free merozoite and intraerythrocytic parasite stages using NOISeq.

https://doi.org/10.1371/journal.pntd.0007680.s008

(TXT)

S7 Table. Nucleotide sequences in FASTA format (ffn).

https://doi.org/10.1371/journal.pntd.0007680.s009

(FFN)

S8 Table. Amino acid sequences in FASTA format (faa).

https://doi.org/10.1371/journal.pntd.0007680.s010

(FAA)

S9 Table. Verification of RNAseq results by real-time qRT-PCR.

Sequence design of primers used in this study. For each gene, an 18–21 sense and antisense complementary oligonucleotide was generated.

https://doi.org/10.1371/journal.pntd.0007680.s011

(DOCX)

Acknowledgments

We are grateful to “Unidad de Secuenciación Masiva y Bioinformática” of the “Laboratorio Nacional de Apoyo Tecnológico a las Ciencias Genómicas,” CONACyT #260481, at the Instituto de Biotecnología/UNAM for sequencing and bioinformatics support. We thank Centro de Transfusiones de la Comunidad de Madrid which provided the human A+ blood from healthy volunteer donors and S. Sánchez and S. Herrera for pulse field gel electrophoresis assistance.

References

  1. 1. Krause PJ. Human babesiosis. Int J Parasitol. 2019;49: 165–174. pmid:30690090
  2. 2. Vannier EG, Diuk-Wasser MA, Ben Mamoun C, Krause PJ. Babesiosis. Infect Dis Clin North Am. 2015;29: 357–370. pmid:25999229
  3. 3. Hildebrandt A, Gray JS, Hunfeld K-P. Human babesiosis in Europe: what clinicians need to know. Infection. 2013;41: 1057–1072. pmid:24104943
  4. 4. Tomassone L, Berriatua E, De Sousa R, Duscher GG, Mihalca AD, Silaghi C, et al. Neglected vector-borne zoonoses in Europe: Into the wild. Vet Parasitol. 2018;251: 17–26. pmid:29426471
  5. 5. Qi C, Zhou D, Liu J, Cheng Z, Zhang L, Wang L, et al. Detection of Babesia divergens using molecular methods in anemic patients in Shandong Province, China. Parasitol Res. 2011;109: 241–245. pmid:21503639
  6. 6. Zamoto-Niikura A, Tsuji M, Qiang W, Morikawa S, Hanaki K-I, Holman PJ, et al. The Babesia divergens Asia Lineage Is Maintained through Enzootic Cycles between Ixodes persulcatus and Sika Deer in Hokkaido, Japan. Appl Environ Microbiol. 2018;84. pmid:29374041
  7. 7. Sevilla E, González LM, Luque D, Gray J, Montero E. Kinetics of the invasion and egress processes of Babesia divergens, observed by time-lapse video microscopy. Sci Rep. 2018;8: 14116. pmid:30237573
  8. 8. Lobo CA, Rodriguez M, Cursino-Santos JR. Babesia and red cell invasion. Curr Opin Hematol. 2012;19: 170–175. pmid:22488304
  9. 9. Del Carmen Terrón M, González-Camacho F, González LM, Luque D, Montero E. Ultrastructure of the Babesia divergens free merozoite. Ticks Tick Borne Dis. 2016;7: 1274–1279. pmid:27430965
  10. 10. Jalovecka M, Hajdusek O, Sojka D, Kopacek P, Malandrin L. The Complexity of Piroplasms Life Cycles. Front Cell Infect Microbiol. 2018;8: 248. pmid:30083518
  11. 11. Gray J, Zintl A, Hildebrandt A, Hunfeld K-P, Weiss L. Zoonotic babesiosis: overview of the disease and novel aspects of pathogen identity. Ticks Tick Borne Dis. 2010;1: 3–10. pmid:21771506
  12. 12. Cursino-Santos JR, Singh M, Pham P, Rodriguez M, Lobo CA. Babesia divergens builds a complex population structure composed of specific ratios of infected cells to ensure a prompt response to changing environmental conditions. Cell Microbiol. 2015; pmid:26663747
  13. 13. Repnik U, Gangopadhyay P, Bietz S, Przyborski JM, Griffiths G, Lingelbach K. The apicomplexan parasite Babesia divergens internalizes band 3, glycophorin A and spectrin during invasion of human red blood cells. Cell Microbiol. 2015;17: 1052–1068. pmid:25628009
  14. 14. Rossouw I, Maritz-Olivier C, Niemand J, van Biljon R, Smit A, Olivier NA, et al. Morphological and Molecular Descriptors of the Developmental Cycle of Babesia divergens Parasites in Human Erythrocytes. PLoS Negl Trop Dis. 2015;9: e0003711. pmid:25955414
  15. 15. Cuesta I, González LM, Estrada K, Grande R, Zaballos A, Lobo CA, et al. High-Quality Draft Genome Sequence of Babesia divergens, the Etiological Agent of Cattle and Human Babesiosis. Genome Announc. 2014;2. pmid:25395649
  16. 16. Chain PSG, Grafham DV, Fulton RS, Fitzgerald MG, Hostetler J, Muzny D, et al. Genomics. Genome project standards in a new era of sequencing. Science. 2009;326: 236–237. pmid:19815760
  17. 17. Brayton KA, Lau AOT, Herndon DR, Hannick L, Kappmeyer LS, Berens SJ, et al. Genome sequence of Babesia bovis and comparative analysis of apicomplexan hemoprotozoa. PLoS Pathog. 2007;3: 1401–1413. pmid:17953480
  18. 18. Cornillot E, Hadj-Kaddour K, Dassouli A, Noel B, Ranwez V, Vacherie B, et al. Sequencing of the smallest Apicomplexan genome from the human pathogen Babesia microti. Nucleic Acids Res. 2012;40: 9102–9114. pmid:22833609
  19. 19. Schneider CA, Rasband WS, Eliceiri KW. NIH Image to ImageJ: 25 years of image analysis. Nat Methods. 2012;9: 671–675. pmid:22930834
  20. 20. Jones SH, Lew AE, Jorgensen WK, Barker SC. Babesia bovis: genome size, number of chromosomes and telomeric probe hybridisation. Int J Parasitol. 1997;27: 1569–1573. pmid:9467743
  21. 21. Langer-Safer PR, Levine M, Ward DC. Immunological method for mapping genes on Drosophila polytene chromosomes. Proc Natl Acad Sci USA. 1982;79: 4381–4385. pmid:6812046
  22. 22. Chakraborty S, Roy S, Mistry HU, Murthy S, George N, Bhandari V, et al. Potential Sabotage of Host Cell Physiology by Apicomplexan Parasites for Their Survival Benefits. Front Immunol. 2017;8: 1261. pmid:29081773
  23. 23. Besteiro S, Dubremetz J-F, Lebrun M. The moving junction of apicomplexan parasites: a key structure for invasion. Cell Microbiol. 2011;13: 797–805. pmid:21535344
  24. 24. Blackman MJ. Malarial proteases and host cell egress: an “emerging” cascade. Cell Microbiol. 2008;10: 1925–1934. pmid:18503638
  25. 25. Frénal K, Dubremetz J-F, Lebrun M, Soldati-Favre D. Gliding motility powers invasion and egress in Apicomplexa. Nat Rev Microbiol. 2017;15: 645–660. pmid:28867819
  26. 26. Buguliskis JS, Brossier F, Shuman J, Sibley LD. Rhomboid 4 (ROM4) affects the processing of surface adhesins and facilitates host cell invasion by Toxoplasma gondii. PLoS Pathog. 2010;6: e1000858. pmid:20421941
  27. 27. Montero E, Gonzalez LM, Rodriguez M, Oksov Y, Blackman MJ, Lobo CA. A conserved subtilisin protease identified in Babesia divergens merozoites. J Biol Chem. 2006;281: 35717–35726. pmid:16982617
  28. 28. Garg S, Agarwal S, Kumar S, Yazdani SS, Chitnis CE, Singh S. Calcium-dependent permeabilization of erythrocytes by a perforin-like protein during egress of malaria parasites. Nat Commun. 2013;4: 1736. pmid:23591903
  29. 29. Wirth CC, Glushakova S, Scheuermayer M, Repnik U, Garg S, Schaack D, et al. Perforin-like protein PPLP2 permeabilizes the red blood cell membrane during egress of Plasmodium falciparum gametocytes. Cell Microbiol. 2014;16: 709–733. pmid:24602217
  30. 30. Bozdech Z, Llinás M, Pulliam BL, Wong ED, Zhu J, DeRisi JL. The transcriptome of the intraerythrocytic developmental cycle of Plasmodium falciparum. PLoS Biol. 2003;1: E5. pmid:12929205
  31. 31. Brochet M, Billker O. Calcium signalling in malaria parasites. Mol Microbiol. 2016;100: 397–408. pmid:26748879
  32. 32. Alam MM, Solyakov L, Bottrill AR, Flueck C, Siddiqui FA, Singh S, et al. Phosphoproteomics reveals malaria parasite Protein Kinase G as a signalling hub regulating egress and invasion. Nat Commun. 2015;6: 7285. pmid:26149123
  33. 33. Bullen HE, Jia Y, Yamaryo-Botté Y, Bisio H, Zhang O, Jemelin NK, et al. Phosphatidic Acid-Mediated Signaling Regulates Microneme Secretion in Toxoplasma. Cell Host Microbe. 2016;19: 349–360. pmid:26962945
  34. 34. Gilson PR, Crabb BS. Morphology and kinetics of the three distinct phases of red blood cell invasion by Plasmodium falciparum merozoites. Int J Parasitol. 2009;39: 91–96. pmid:18952091
  35. 35. Farrell A, Thirugnanam S, Lorestani A, Dvorin JD, Eidell KP, Ferguson DJP, et al. A DOC2 protein identified by mutational profiling is essential for apicomplexan parasite exocytosis. Science. 2012;335: 218–221. pmid:22246776
  36. 36. Fang H, Gomes AR, Klages N, Pino P, Maco B, Walker EM, et al. Epistasis studies reveal redundancy among calcium-dependent protein kinases in motility and invasion of malaria parasites. Nat Commun. 2018;9: 4248. pmid:30315162
  37. 37. Dobrowolski JM, Sibley LD. Toxoplasma invasion of mammalian cells is powered by the actin cytoskeleton of the parasite. Cell. 1996;84: 933–939. pmid:8601316
  38. 38. Baum J, Tonkin CJ, Paul AS, Rug M, Smith BJ, Gould SB, et al. A malaria parasite formin regulates actin polymerization and localizes to the parasite-erythrocyte moving junction during invasion. Cell Host Microbe. 2008;3: 188–198. pmid:18329618
  39. 39. Mehta S, Sibley LD. Toxoplasma gondii actin depolymerizing factor acts primarily to sequester G-actin. J Biol Chem. 2010;285: 6835–6847. pmid:20042603
  40. 40. Skillman KM, Daher W, Ma CI, Soldati-Favre D, Sibley LD. Toxoplasma gondii profilin acts primarily to sequester G-actin while formins efficiently nucleate actin filament formation in vitro. Biochemistry. 2012;51: 2486–2495. pmid:22397711
  41. 41. Mehta S, Sibley LD. Actin depolymerizing factor controls actin turnover and gliding motility in Toxoplasma gondii. Mol Biol Cell. 2011;22: 1290–1299. pmid:21346192
  42. 42. Whitelaw JA, Latorre-Barragan F, Gras S, Pall GS, Leung JM, Heaslip A, et al. Surface attachment, promoted by the actomyosin system of Toxoplasma gondii is important for efficient gliding motility and invasion. BMC Biol. 2017;15: 1. pmid:28100223
  43. 43. Bane KS, Lepper S, Kehrer J, Sattler JM, Singer M, Reinig M, et al. The Actin Filament-Binding Protein Coronin Regulates Motility in Plasmodium Sporozoites. PLoS Pathog. 2016;12: e1005710. pmid:27409081
  44. 44. Salamun J, Kallio JP, Daher W, Soldati-Favre D, Kursula I. Structure of Toxoplasma gondii coronin, an actin-binding protein that relocalizes to the posterior pole of invasive parasites and contributes to invasion and egress. FASEB J. 2014;28: 4729–4747. pmid:25114175
  45. 45. Herm-Götz A, Weiss S, Stratmann R, Fujita-Becker S, Ruff C, Meyhöfer E, et al. Toxoplasma gondii myosin A and its light chain: a fast, single-headed, plus-end-directed motor. EMBO J. 2002;21: 2149–2158. pmid:11980712
  46. 46. Bergman LW, Kaiser K, Fujioka H, Coppens I, Daly TM, Fox S, et al. Myosin A tail domain interacting protein (MTIP) localizes to the inner membrane complex of Plasmodium sporozoites. J Cell Sci. 2003;116: 39–49. pmid:12456714
  47. 47. Gilk SD, Gaskins E, Ward GE, Beckers CJM. GAP45 phosphorylation controls assembly of the Toxoplasma myosin XIV complex. Eukaryotic Cell. 2009;8: 190–196. pmid:19047362
  48. 48. Hernández PC, Wasserman M, Chaparro-Olaya J. In vitro interaction between Plasmodium falciparum myosin B (PfMyoB) and myosin A tail interacting protein (MTIP). Parasitol Res. 2018;117: 3437–3446. pmid:30094538
  49. 49. Baum J, Richard D, Healer J, Rug M, Krnajski Z, Gilberger T-W, et al. A conserved molecular motor drives cell invasion and gliding motility across malaria life cycle stages and other apicomplexan parasites. J Biol Chem. 2006;281: 5197–5208. pmid:16321976
  50. 50. Meissner M, Schlüter D, Soldati D. Role of Toxoplasma gondii myosin A in powering parasite gliding and host cell invasion. Science. 2002;298: 837–840. pmid:12399593
  51. 51. Jacot D, Tosetti N, Pires I, Stock J, Graindorge A, Hung Y-F, et al. An Apicomplexan Actin-Binding Protein Serves as a Connector and Lipid Sensor to Coordinate Motility and Invasion. Cell Host Microbe. 2016;20: 731–743. pmid:27978434
  52. 52. Bargieri D, Lagal V, Andenmatten N, Tardieux I, Meissner M, Ménard R. Host cell invasion by apicomplexan parasites: the junction conundrum. PLoS Pathog. 2014;10: e1004273. pmid:25232721
  53. 53. Montero E, Rodriguez M, Oksov Y, Lobo CA. Babesia divergens apical membrane antigen 1 and its interaction with the human red blood cell. Infect Immun. 2009;77: 4783–4793. pmid:19720759
  54. 54. Delbecq S, Auguin D, Yang Y-S, Löhr F, Arold S, Schetters T, et al. The solution structure of the adhesion protein Bd37 from Babesia divergens reveals structural homology with eukaryotic proteins involved in membrane trafficking. J Mol Biol. 2008;375: 409–424. pmid:18035372
  55. 55. Barthe P, Murciano B, Schetters T, Gorenflot A, Delbecq S, Roumestand C. 1H, 15N and 13C Backbone resonance assignments of a conformational mutant of the adhesion protein Δ-Bd37 from Babesia divergens. Biomol NMR Assign. 2013;7: 241–244. pmid:22899250
  56. 56. Rodriguez M, Alhassan A, Ord RL, Cursino-Santos JR, Singh M, Gray J, et al. Identification and characterization of the RouenBd1987 Babesia divergens Rhopty-Associated Protein 1. PLoS ONE. 2014;9: e107727. pmid:25226276
  57. 57. Suarez CE, Palmer GH, Hötzel I, McElwain TF. Structure, sequence, and transcriptional analysis of the Babesia bovis rap-1 multigene locus. Mol Biochem Parasitol. 1998;93: 215–224. pmid:9662706
  58. 58. Yamagishi J, Asada M, Hakimi H, Tanaka TQ, Sugimoto C, Kawazu S-I. Whole-genome assembly of Babesia ovata and comparative genomics between closely related pathogens. BMC Genomics. 2017;18: 832. pmid:29078748
  59. 59. Suarez CE, Palmer GH, Florin-Christensen M, Hines SA, Hötzel I, McElwain TF. Organization, transcription, and expression of rhoptry associated protein genes in the Babesia bigemina rap-1 locus. Mol Biochem Parasitol. 2003;127: 101–112. pmid:12672519
  60. 60. Dalrymple BP, Peters JM, Böse R, Wright IG. A polymerase chain reaction method for the identification of genes encoding members of the Bv60/p58 family of rhoptry protein homologues in the genus Babesia. Exp Parasitol. 1996;84: 96–100. pmid:8888737
  61. 61. Suarez CE, Laughery JM, Bastos RG, Johnson WC, Norimine J, Asenzo G, et al. A novel neutralization sensitive and subdominant RAP-1-related antigen (RRA) is expressed by Babesia bovis merozoites. Parasitology. 2011;138: 809–818. pmid:21554842
  62. 62. Lamarque M, Besteiro S, Papoin J, Roques M, Vulliez-Le Normand B, Morlon-Guyot J, et al. The RON2-AMA1 interaction is a critical step in moving junction-dependent invasion by apicomplexan parasites. PLoS Pathog. 2011;7: e1001276. pmid:21347343
  63. 63. Tonkin ML, Roques M, Lamarque MH, Pugnière M, Douguet D, Crawford J, et al. Host cell invasion by apicomplexan parasites: insights from the co-structure of AMA1 with a RON2 peptide. Science. 2011;333: 463–467. pmid:21778402
  64. 64. Alexander DL, Mital J, Ward GE, Bradley P, Boothroyd JC. Identification of the moving junction complex of Toxoplasma gondii: a collaboration between distinct secretory organelles. PLoS Pathog. 2005;1: e17. pmid:16244709
  65. 65. Besteiro S, Michelin A, Poncet J, Dubremetz J-F, Lebrun M. Export of a Toxoplasma gondii rhoptry neck protein complex at the host cell membrane to form the moving junction during invasion. PLoS Pathog. 2009;5: e1000309. pmid:19247437
  66. 66. Hines SA, Palmer GH, Brown WC, McElwain TF, Suarez CE, Vidotto O, et al. Genetic and antigenic characterization of Babesia bovis merozoite spherical body protein Bb-1. Mol Biochem Parasitol. 1995;69: 149–159. pmid:7770080
  67. 67. Santos JM, Ferguson DJP, Blackman MJ, Soldati-Favre D. Intramembrane cleavage of AMA1 triggers Toxoplasma to switch from an invasive to a replicative mode. Science. 2011;331: 473–477. pmid:21205639
  68. 68. Harris PK, Yeoh S, Dluzewski AR, O’Donnell RA, Withers-Martinez C, Hackett F, et al. Molecular identification of a malaria merozoite surface sheddase. PLoS Pathog. 2005;1: 241–251. pmid:16322767
  69. 69. Leykauf K, Treeck M, Gilson PR, Nebl T, Braulke T, Cowman AF, et al. Protein kinase a dependent phosphorylation of apical membrane antigen 1 plays an important role in erythrocyte invasion by the malaria parasite. PLoS Pathog. 2010;6: e1000941. pmid:20532217
  70. 70. Millholland MG, Chandramohanadas R, Pizzarro A, Wehr A, Shi H, Darling C, et al. The malaria parasite progressively dismantles the host erythrocyte cytoskeleton for efficient egress. Mol Cell Proteomics. 2011;10: M111.010678. pmid:21903871
  71. 71. De Niz M, Burda P-C, Kaiser G, Del Portillo HA, Spielmann T, Frischknecht F, et al. Progress in imaging methods: insights gained into Plasmodium biology. Nat Rev Microbiol. 2017;15: 37–54. pmid:27890922
  72. 72. Carletti T, Barreto C, Mesplet M, Mira A, Weir W, Shiels B, et al. Characterization of a papain-like cysteine protease essential for the survival of Babesia ovis merozoites. Ticks Tick Borne Dis. 2016;7: 85–93. pmid:26456355
  73. 73. Martins TM, do Rosário VE, Domingos A. Expression and characterization of the Babesia bigemina cysteine protease BbiCPL1. Acta Trop. 2012;121: 1–5. pmid:21986365
  74. 74. Dhawan S, Dua M, Chishti AH, Hanspal M. Ankyrin peptide blocks falcipain-2-mediated malaria parasite release from red blood cells. J Biol Chem. 2003;278: 30180–30186. pmid:12775709
  75. 75. Russo I, Oksman A, Vaupel B, Goldberg DE. A calpain unique to alveolates is essential in Plasmodium falciparum and its knockdown reveals an involvement in pre-S-phase development. Proc Natl Acad Sci USA. 2009;106: 1554–1559. pmid:19164769
  76. 76. Datta G, Hossain ME, Asad M, Rathore S, Mohmmed A. Plasmodium falciparum OTU-like cysteine protease (PfOTU) is essential for apicoplast homeostasis and associates with noncanonical role of Atg8. Cell Microbiol. 2017;19. pmid:28423214
  77. 77. Walker DM, Mahfooz N, Kemme KA, Patel VC, Spangler M, Drew ME. Plasmodium falciparum erythrocytic stage parasites require the putative autophagy protein PfAtg7 for normal growth. PLoS ONE. 2013;8: e67047. pmid:23825614
  78. 78. Gorenflot A, Brasseur P, Precigout E, L’Hostis M, Marchand A, Schrevel J. Cytological and immunological responses to Babesia divergens in different hosts: ox, gerbil, man. Parasitol Res. 1991;77: 3–12. pmid:1994368
  79. 79. Margulies M, Egholm M, Altman WE, Attiya S, Bader JS, Bemben LA, et al. Genome sequencing in microfabricated high-density picolitre reactors. Nature. 2005;437: 376–380. pmid:16056220
  80. 80. Ribeiro FJ, Przybylski D, Yin S, Sharpe T, Gnerre S, Abouelleil A, et al. Finished bacterial genomes from shotgun sequence data. Genome Res. 2012;22: 2270–2277. pmid:22829535
  81. 81. Soto-Jimenez LM, Estrada K, Sanchez-Flores A. GARM: genome assembly, reconciliation and merging pipeline. Curr Top Med Chem. 2014;14: 418–424. pmid:24304318
  82. 82. Boetzer M, Henkel CV, Jansen HJ, Butler D, Pirovano W. Scaffolding pre-assembled contigs using SSPACE. Bioinformatics. 2011;27: 578–579. pmid:21149342
  83. 83. Tsai IJ, Otto TD, Berriman M. Improving draft assemblies by iterative mapping and assembly of short reads to eliminate gaps. Genome Biol. 2010;11: R41. pmid:20388197
  84. 84. Otto TD, Sanders M, Berriman M, Newbold C. Iterative Correction of Reference Nucleotides (iCORN) using second generation sequencing technology. Bioinformatics. 2010;26: 1704–1707. pmid:20562415
  85. 85. Lechner M, Findeiss S, Steiner L, Marz M, Stadler PF, Prohaska SJ. Proteinortho: detection of (co-)orthologs in large-scale analysis. BMC Bioinformatics. 2011;12: 124. pmid:21526987
  86. 86. web documentation for Trinotate. Contribute to Trinotate/Trinotate.github.io development by creating an account on GitHub [Internet]. Trinotate; 2018. Available: https://github.com/Trinotate/Trinotate.github.io
  87. 87. Krzywinski M, Schein J, Birol I, Connors J, Gascoyne R, Horsman D, et al. Circos: an information aesthetic for comparative genomics. Genome Res. 2009;19: 1639–1645. pmid:19541911
  88. 88. Robinson MD, McCarthy DJ, Smyth GK. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26: 139–140. pmid:19910308
  89. 89. Anders S, Huber W. Differential expression analysis for sequence count data. Genome Biol. 2010;11: R106. pmid:20979621
  90. 90. Tarazona S, García-Alcalde F, Dopazo J, Ferrer A, Conesa A. Differential expression in RNA-seq: a matter of depth. Genome Res. 2011;21: 2213–2223. pmid:21903743
  91. 91. Robinson MD, Oshlack A. A scaling normalization method for differential expression analysis of RNA-seq data. Genome Biol. 2010;11: R25. pmid:20196867
  92. 92. Carver T, Berriman M, Tivey A, Patel C, Böhme U, Barrell BG, et al. Artemis and ACT: viewing, annotating and comparing sequences stored in a relational database. Bioinformatics. 2008;24: 2672–2676. pmid:18845581
  93. 93. Untergasser A, Cutcutache I, Koressaar T, Ye J, Faircloth BC, Remm M, et al. Primer3—new capabilities and interfaces. Nucleic Acids Res. 2012;40: e115. pmid:22730293