Open access peer-reviewed chapter

Genome-Based Vaccinology Applied to Bovine Babesiosis

Written By

Juan Mosqueda, Diego Josimar Hernández-Silva and Mario Hidalgo-Ruiz

Submitted: 18 August 2017 Reviewed: 22 November 2017 Published: 20 December 2017

DOI: 10.5772/intechopen.72636

From the Monograph

Farm Animals Diseases, Recent Omic Trends and New Strategies of Treatment

Edited by Rosa Estela Quiroz-Castañeda

Chapter metrics overview

1,385 Chapter Downloads

View Full Metrics

Abstract

Genomics approaches in veterinary research have been a very useful tool to identify candidates with potential to be used in prevention of animal diseases. In Babesia, genome information analysis has elucidated a wide variety of protein families and some members are described in this chapter. Here, we present some of the most recent studies about B. bovis and B. bigemina genomes where some proteins have been identified with potential to prevent infections by these parasites.

Keywords

  • bovine babesiosis
  • bioinformatics
  • vaccines
  • genomics

1. Introduction

Bovine babesiosis is a tick-transmitted disease caused by apicomplexan parasites of the genus Babesia. This disease is caused by Babesia bovis and B. bigemina in the Americas including Mexico, where it is distributed in tropical and subtropical regions, occupying 51.5% of the national territory [1, 2]. This disease was reported for the first time by Viktor Babes in Rumania in 1888. However, it was until 1893 when Smith and Kilborne demonstrated that the disease is transmitted to cattle by infected ticks [3, 4]. In Mexico, the first report of bovine babesiosis occurred in 1905; however, it is believed that it was first introduced to the American continent by the Spaniards during the conquest. To date, measures used to control bovine babesiosis include vector control, an early diagnosis, treatment of sick animals and vaccination. The negative, severe impact that cattle fever tick and babesiosis have in the cattle industry in Mexico and the world has not diminished due mainly to a lack of commercially available, safe and effective vaccines. Vaccines based on approaches using genomics and bioinformatics are a promissory solution to this problem [5]. It has been shown that experimental vaccines based on recombinant antigens have been developed successfully in apicomplexan parasites like Plasmodium, Toxoplasma and Theileria [6, 7, 8]. With the completion of the B. bovis genome [9] and the partial sequencing of the Babesia bigemina genome (http://www.sanger.ac.uk/), it is now possible to study these pathogens to the genomic level, taking advantage of the bioinformatics tools developed for this purpose. This approach is now generating valuable information on the essential characteristics of the genome structure and allows comparative analyses with genomes of other apicomplexan pathogens of importance in human and animal health, as well as the identification of genes with a potential use in diagnostics, vaccines or therapeutics. More specific analyses are also possible with the generation of expressed sequence tags (EST) obtained for B. bovis, which allow the analysis of those genes specifically expressed in the different stages of the parasite’s life cycle [10] and, finally, implementation of methods for genome-wide analysis like microarrays which will be in short available for their use [11]. Additionally, research on this important disease is complemented with all the information generated so far about those genes codifying antigens with a potential as candidates in vaccines, diagnostics or therapeutics, which have been discovered in the last 30 years. Equally important is the knowledge about the life cycle of the parasite, the interaction with the vector tick and the genes involved in this interaction, which are poorly studied so far. In the following sections, we describe the most relevant aspects of the B. bovis and B. bigemina genomes and genes characterized to date.

Advertisement

2. Babesiosis

2.1. Babesia bovis genome

Although Babesia bovis, B. bigemina and B. divergens are causative agents of bovine babesiosis, B. bovis is regarded as the most important and has a bigger impact in the livestock industry due to its virulence and high mortality rate. For this reason, the B. bovis genome was the first to be sequenced. This was done by Washigton State University in collaboration with the Agricultural Research Service and The Institute for Genomic Research (TIGR) in the USA. The genome sequence was obtained from the T2Bo strain, a virulent strain isolated from a clinical case in Texas, USA.

The genome of Babesia bovis has a length of 8.2 Mbp, contains 3671 genes and consists of four chromosomes, three of them are acrocentric: chromosome 1, has a length of 1.25 Mpb, is the smallest and contains a gap, which is estimated to be 150 Kpb long. Chromosome 2 is fully sequenced and contains 1.73 Mbp in length. Chromosome 3 is also fully sequenced and is 2.59 Mpb in length. Finally, chromosome 4, which is the only submetacentric one, it is partially sequenced because it contains an assembly gap that has not been solved and is 2.62 Mbp in length. The structural features of the B. bovis genome are similar to those of Theileria parva but have major differences with Plasmodium falciparum (Table 1), despite the fact that B. bovis and P. falciparum share similar clinical and pathological features [9].

Table 1.

Genome characteristics of B. bovis, T. parva and P. falciparum [9].

B. bovis contains two extrachromosomal genomes: a lineal mitochondrial genome of 6 Kbp and an apicoplast genome that is circular and consists of 33 kpb with 32 genes and 25 sequences of tRNA. The apicoplast, which is an organelle conserved in the phylum apicomplexa, is a nonphotosynthetic plastid essential for survival [12]. The apicoplast genome was first sequenced in 1974 from Plasmodium lophurae and was thought that it could be mitochondrial DNA. In 1994 it was finally related to plastids DNA when a gen coding for a protein of 470 amino acids in length was identified and it contained a 50% identity with a protein only described in the plastome of red algae [13]. It is believed that the plastid is derived from a secondary endosymbiosis from red algae like in dinoflagellates [14]. Furthermore, Chromera velia has a plastid originated from red algae that is closely related with the apicoplast [15].

Advertisement

3. Babesia multigenic families

3.1. Variant erythrocyte surface antigen-1

Even though the apicoplast is an apicomplexa organelle, they share a complex of organelles that is characteristic of the apicomplexa: the apical complex. This complex is composed of spherical body, rhoptries and micronemes; in this organelle, different proteins involved in the life cycle are generated, and some of these are secreted to the media or directed to the membrane [16, 17, 18]. In the erythrocyte stage, some Babesia parasites, including B. bovis and bigemina, can invade the host erythrocytes in a directly way without a pre-erythrocyte stage. Antigenically, the surface of infected erythrocytes is different between Babesia strains and the parasitized erythrocytes present in an infected bovine could fluctuate widely over time, in a process called antigenic variation [19, 20, 21]. This process on the infected erythrocytes is carried out by the antigenically variant protein named the variant erythrocyte surface antigen-1 (VESA1) that is constituted by two subunits (a and b) and encoded in a multigene family; the genes ves-1 involved in this family are related too in cytoadherence and are distributed in the four chromosomes of Babesia [19, 22, 23]. Although more of 350 genes ves-1α and more of 80 genes ves-1β were previously reported, in B. bovis genome only 119 were evidenced [9].

3.2. SmORFs

The family of genes ves-1 is associated across all four chromosomes with another multigene family of proteins that are smaller in size than the ves-1 genes (Figure 1), due to small open reading frames (SmORFs). This family is the second largest in the B. bovis genome and comprises 44 genes without significant sequence identity to any protein or gene sequence available in databases. Of 44 genes, 42 are codified in a single exon, but from these 44 proteins that are extracellular just one does not have a signal peptide [9]. Sequence analysis in the T2Bo and Mo7 strains demonstrated that the repertoire varies between strains and has multiple semi-conserved and variable blocks; this family comprises two major branches called SmORFs A and B, and these branches are defined by a large hypervariable insertion in 20 genes [24]. Although the function of these proteins is unknown, it is believed that it could play a functional role in VESA1 biology or contribute to the antigenic variation and immune evasion as a consequence [9, 24].

Figure 1.

Representation of B. bovis chromosomes and the localization of the centromeres, ves1, and smorf genes. The chromosomes are depicted by black lines, with the chromosome number shown on the left. ves1 loci are depicted with boxes: Black boxes represent unclassified ves1 genes; Grey boxes represent at least one ves1α/ves1α pair within the cluster; Shaded boxes represent at least one ves1α/ves1β pair within the cluster. The number of ves1 and smorf genes is shown above or below each locus, respectively. Finally the centromeres are represented as black circles. Modified from Brayton KA, et al., 2007 [9].

3.3. VMSA

In American strains, the variable merozoite surface antigen (VMSA) family contains the proteins MSA-1, MSA-2a1, MSA-2a2, MSA-2b and MSA-2c, while in Australian strains only tree genes were found: msa-1, msa-2c and msa-2a/b [25]. The genes that conform this family reside on chromosome 1, and the four copies of msa2 gene are arranged tandemly in a head-to-tail fashion as long as msa1 gene is located 5 kbp upstream from the msa2 genes [9]. These five proteins have a conserved GPI domain and are involved in the first attachment to the erythrocyte. However, the exposed epitopes are not conserved between these proteins of this family and between different strains around the world [18, 26, 27]. Even though these proteins are variable, some studies have been shown that msa-2c gene is the most conserved of this family. Monoclonal antibodies against this protein can recognize strains from different geographic regions, and polyclonal antibodies have an effect on the invasion process, suggesting its utility as recombinant vaccine antigen or in diagnostic tests [18, 26, 28, 29, 30, 31, 32, 33]. These results have not been observed in the other MSA proteins; the MSA-1 protein is immunogenic and avoids the invasion process in vitro, but the immunogenic response is not protective [30]. It could be due to the fact that msa-1 gene has an important allelic variation in strains from the nearby geographical regions. This variation suggests that the antibodies generated could not have a cross-reaction between different strains [34, 35].

3.4. SBP

The spherical body protein (SBP) constitutes another family in B. bovis, and these proteins that are located in the spherical body of the apical complex are known as SBP1, SBP2, SBP3 and SBP4. In the invasion process, SBP2 is released from the spherical bodies to the cytoplasmic membrane of the erythrocyte [36]. Twelve truncated copies and just one complete copy of sbp2 gene were identified, showing a conserved 3′ region in these copies [9, 37]. The complete copy and one truncated are located in the chromosome 4, the other truncated copies are located in the chromosome 3, and some of these truncated copies are expressed in erythrocytic stages of B. bovis [10].

3.5. Bbo-6cys

A novel family of genes that codify proteins with similarities to 6cys family of Plasmodium were identified in B. bovis genome. This family contains six genes (6cys-A, B, C, D, E and F), and these genes are located in tandem in the chromosome 2 except for 6cys-F that is located in a distal region. To identify this family was employed the sequence of the P. falciparum PFS230 protein as a query that has a higher homology with Bbo-6cys-E gene. Antibodies against this protein have an inhibitory effect on the invasion process, suggesting its importance in control methods against B. bovis infection [38].

3.6. Bovipain

Inhibitors of cysteine proteases have been shown to hamper intraerythrocytic replication of B. bovis, and four papain-like cysteine proteases are found in B. bovis genome. The bovipain-2, which is the orthologous gene of P. falciparum falcipain-2 that is involved in hemoglobin digestion [39], is located in chromosome 4 by an ORF of 1.3 kb without introns, the characterization of this protein shown a molecular weight of 42 kDa, and a transmembrane region and is highly conserved between B. bovis strains of North and South America [40]. The bovipain-2 could be employed as a vaccine or as a target of drugs in the babesiosis control.

Advertisement

4. Vaccine antigens

4.1. RON proteins and AMA-1

In Toxoplasma, the characterization of the invasion process allowed the identification of a complex of proteins generated in the rhoptry neck, called rhoptry neck proteins (RONs), this family of proteins consists of RON2, RON4, RON5 and RON8 that are related to AMA-1 in the formation of the moving junction (MJ) in the invasion process [41]. The RON complex is inserted into the host cell; meanwhile, AMA-1 is released to the parasite membrane; this process is described in P. falciparum, where the specific interaction between the host membrane and the parasite membrane is mediated by AMA-1 and RON2; the disruption of this interaction avoids the invasion process [42, 43]. RON2 was identified in B. divergens and B. microti, has a full-length sequence of 4053 bp that codifies to a protein of 170 kDa and has apical localization; antibodies against this protein are inhibitors of parasite invasion like in other apicomplexan parasites; The B. divergens RON2 protein has a closely related sequence in B. bovis identified by BLASTp [42, 44, 45]. The firts protein described to participate in the invasion process as part of the MJ in apicomplexan parasites was AMA-1. This protein is stored in the micronemes and secreted to the apical end during the invasion process. In Babesia bovis, AMA-1 contains 606 amino acids and it is codified by a 1818 bp-long gene, without introns. AMA-1 is a type I transmembrane protein with a N-terminal ectodomain, which is divided into three subdomains containing 14 cysteins [18, 46, 47].

4.2. RAP-1

One of the most studied proteins of the rhoptries identified in B. bovis is the rhoptry-associated protein 1 (RAP-1). The gene is constituted by only one exon and has a length of 1698 bp with two copies separated by a noncodifying sequence of 1 kbp in B. bovis [48]. However in B. bigemina, rap-1 is represented by three genes: rap-1a, rap-1b, and rap-1c, arranged in tandem, as explained latter [49]. The members of this family have a signal peptide, four cysteine residues and a 14 aa motive and, moreover, contain immunogenic epitopes B and T that elicit a Th1 humoral response in the host that avoids the attachment of the parasite to the erythrocyte; its structure is conserved between different isolates and is expressed in the sporozoite [37, 50, 51, 52, 53, 54, 55].

4.3. MIC

In B. bovis was described a gene that codifies to a protein like to the T. gondii protein 1 of the micronemes (MIC1). This gene is located in the chromosome 3 and is highly conserved between strains of B. bovis. Its function is involved in the cito-adherence process through the binding to sialic acid, and antibodies against the recombinant protein and synthetic peptides designed on antigenic regions of B. bovis MIC-1 avoid the invasion process on the in vitro culture of the parasites [56].

4.4. HSP-20

The heat shock protein 20 (HSP-20) is a protein of 20.2 kDa associated with other small proteins related to heat shock in mammals and plants [57]. The hsp-20 gene consists of 686 bp with an intron of 153 bp that makes a polypeptide of 177 aa [58]. Antibodies against this protein recognize both B. bovis and B. bigemina, suggesting that HSP-20 contains conserved epitopes in these species [59, 60].

Advertisement

5. Babesia bigemina

During the first decade of twenty-first century, the sequencing and description of B. bovis genome [61] have helped to find genes that play important roles during its life cycle, currently. The Sanger Institute is leading the sequencing project of B. bigemina genome, which is estimated in 10 Mb size distributed in four chromosomes [62]. The advances in genomics of both Babesia species that affects cattle in America are allowing researcher to find, analyze, compare and predict proteins involved directly in pathogenesis and its life cycle. For more than five decades, researchers have been trying to develop a vaccine against piroplasmosis; before 1980, several studies were carried out to immunize cattle in an effective way. The first attempts were directed in animals that were infected on purpose and healed from babesiosis as a strategy to avoid nondesirable infections [63]. Some studies were focused on trying to find a way to reduce the virulence of high infective B. bigemina strains through inoculum passages in several animals [64], and some even tried to immunize calves in utero [65]. As we know now, the development of a of an effective and low-cost vaccine is more complex than initially thought. Nowadays some countries produce vaccines against bovine babesiosis. The Queensland Government in Australia offers a vaccine made of parasitized bovine blood [66]; this live attenuated parasite vaccine must be stored at −196°C and during its production is necessary a batch of splenectomized calves that must remain in quarantine three times before the first procedures of manufacturing [67]. It is evident that piroplasmosis vaccination involves long periods of production, surgery, and maintenance of animal infected blood donors and thorough procedures to achieve high standards of bioethical considerations.

Currently, research is focused on developing vaccines that avoid complex production procedures and the use of live animals; new technologies have arrived bringing opportunities to develop a vaccine using high throughput production. For this, certain obstacles must be solved before an effective vaccine is produced.

In vaccination, developing gene polymorphisms and antigenic variation is one of the first problems that researchers must cope with, and the selection of most suitable antigen candidates is a crucial step. With gene databases, the analysis of sequence variations has been made easier to find differences in distant geographical strains. In this sense, several studies have been carried out to find whether some proteins are conserved and how auspicious its election as vaccine candidate would be.

Antigenic variation is used by microorganisms as an evasion mechanism of the immune response, and in Babesia, the vesa family is the most studied group of genes used to “escape” from the host immune system. As described above for B. bovis, the vesa family is composed of two multicopy genes, ves1α and ves1β, which are distributed within the four chromosomes, and it is estimated that there are 72 and 43 copies of them, respectively. Both genes are located in opposite transcription directions and are governed by a bidirectional promoter followed by small sequences that seem to be incomplete fragments of the same recombined gene. The mechanism proposed for the multiple versions of the protein product of ves1α and ves1β has to do with the fact that along the genome are ves pseudogenes that act like reiterative donors of divergent sequences during several rounds of DNA replications, while the genes that are being transcribed are located in a locus of active transcription, which means that during this multigenic conversion segment event new versions of VESA proteins of B. bovis are being generated [68, 69].

Even though in B. bigemina antigenic variation as vesa family in B. bovis is not described, there is information about important genetic differences between strains from diverse geographical locations. On the next lines, we are going to review in general terms some of the genes that are promissory vaccine candidates in B. bigemina.

5.1. AMA-1

The apical merozoite antigen (AMA-1) is a protein that has been related to the tight junction complex formation; during this step in the red blood cell invasion, the protein interacts directly with the Rhoptry proteins to anchor both membranes; this process is well studied in the apicomplexan parasite T. gondii [70]. In Babesia species, AMA-1 as been described as a low-diversity protein; in B. divergens, ama-1 was sequenced from nine isolates from France, in which only two punctual mutations were observed compared with the reference strain [71], in B. bovis, the analysis of Sri Lankan strains showed that ama-1 is a conserved member with about 95% of identity; the msa genes of this strains were mapped and showed variability. As mentioned previously, msa family is highly polymorphic. This last statement proves that the isolates analyzed are different because of their genotype, and among them, there is a highly conserved ama-1 gene [72]. The B. bigemina ama-1 seems to be conserved as well, and Italian strains have a conserved sequence among them and considerable differences in comparison with Australian reference strains; however, when these are compared with Mexican and Argentinian reference strains, the sequence matches in a 99% of identity [73]. The conservation level makes AMA-1 an excellent target for vaccine development.

5.2. RAP-1

The Rhoptry Associated Proteins are part of a multigenic family composed in B. bigemina of five genes arranged in tandem designated as rap-1a; between them, there are two other genes designated as rap-1b which is present in the same number of copies as rap-1a and at the end of the locus a single copy gene called rap-1c. All rap-1 family members are co-transcribed in merozoites, and some members seem to be conserved in geographical strains [49]. It has been demonstrated in B. bovis that specific antibodies are capable of reducing sporozoite invasion to red blood cells in vitro [53]; in B. bigemina, antibodies against RAP-1 reduced parasitemia in comparison with an ovalbumin control in calves inoculated with iRBC [74].

5.3. GP45

The product of the gp45 gene is a glicosylphosphatidylinositol-anchored protein of 45 KDa, which has been related to the adhesion step during the invasion process to red blood cells; it is postulated that play the same role of msa family in B. bovis [26]. Some studies proved that immunization of calves with the purified GP45 reduced significantly the parasitemia when challenged with a Mexican B. bigemina isolate [74]. At this point, GP45 seemed to be an ideal vaccine candidate, however, several years later, other studies demostrated that the gene is not present among all B. bigemina strains. Southern Blot analysis revealed that Puerto Rico and St. Croix isolates do not possess gp45 sequence and also that in Texcoco strain the upstream sequence has polymorphisms and consequently there are nonfunctional promoters. As a result of this, there is a lack of transcription [75]. This last information put in doubt if GP45 would be a good immunological target, not because of its neutralization efficiency but because of that this would not be a good candidate if the purpose of the vaccine is to have a broad-spectrum protection that includes several strains from distant geographical locations.

5.4. Profilin

Profilin is a protein that participates in cytoskeleton ensemble [76]; in Toxoplasma gondii, due to its characteristic gliding motility where the cytoskeleton takes part of, profilin has been involved as an important protein to invade host cells and its antigenicity has been proved for its recognition by Toll receptors [77]. There is new evidence that profilin is present in B. bigemina, B. bovis and B. microti, and more interesting is that sera from infected cattle with B. bovis and B. bigemina are capable to cross-react with recombinant profilin from both species and even with B. microti; the recombinant cattle babesial profilin is capable of conferring immunity in mice against B. microti [78]. Even though there is no information about protective activity in cattle of profilin immunization against B. bigemina and B. bovis, profilin seems to be another promissory target to work on to achieve an effective vaccine.

The genes mentioned are examples of genes with low variability that can be used as a target to prevent babesiosis by B. bigemina; unfortunately, for vaccines developers the variation of sequences and gene products does not follow a high conservation rule. In this sense, experimental strategies have been built up to find a more suitable way to neutralize Babesia infection. Taking advantage of the information available on databases and the sequence analysis is possible to track the most appropriate targets. Alignment tools allow researchers to display the protein sequences from several distant geographical strain similarities among their proteins and find the most suitable candidates.

Advertisement

6. New strategies against apicomplexan parasites

There are several studies on Babesia proteins, which have an important role on the parasite’s life cycle and their immunogenicity. However, even though the protein role on the parasite development has been described in detail, currently there is not a single protein proposed as vaccine candidate against B. bovis or B. bigemina that generates an immunological protective response as effective as the one produced by the live, attenuated vaccines; the reason of it might be with the fact that one single antigen used as immunogen is not enough to display a strong immunological response. In an attempt to achieve protection against microorganisms, new strategies have been raising and one of them is the design of multiepitopic vaccines; in these novel strategies, Plasmodium genus and Toxoplasma gondii are some of the microorganisms within the apicomplexan parasites where these methodologies are being applied.

Researchers have been developing vaccines using more than one antigen. Such is the case of P. falciparum Chimeric Protein 2.9 (PfCP-2.9) composed of two blood-stage antigens, the carboxyl-terminal region of the protein known as Apical Membrane Merozoite Surface Protein 1 (MSP1–19) and domain III of the Apical Membrane Antigen 1 (AMA-1 III). The PfCP-2.9 resulted in being highly immunogenic in rabbits and in primates and is capable of producing an antibody titer 18-fold higher than both antigens administered in a mixture. The neutralization assays demonstrated that the fusion protein reduces substantially the parasite growth [79].

One concern of last decades is that scientists are predicting that some vector-borne diseases will increase as a consequence of expansion of vector habitats due to global warming [80]. In addition to the fact that vectors are acquiring resistance to pesticides that have been using as a mean of eradication and control, today there are reports that malaria vector exhibits multiresistance to diverse chemical families used for its control leaving any suitable choice to reduce mosquito population [81, 83]. As a novel alternative to cope with this situation, multistage vaccines have been designed in an attempt to interrupt the life cycle in both vertebrate and invertebrate hosts. Using antigens from blood stages as the glutamate-rich protein (GLURP) that had been recognized as a natural antigen in acquiring immunity against malaria [82] and by the usage of sexual stage antigens Pfp48/45 that are involved in gamete fusion during sexual reproduction within the mosquito vector [83], the central objective of this alternative multiepitope vaccine is to confer immunity in the people that is at risk to acquire the infection and to reduce in long term the infection in the vector avoiding transmission [84]. The usage of more than two antigens is also an opportunity to confer immunity against parasites and to ensure the success to block the infection. In T. gondii, an alternative vaccine prototype was designed as a chimeric protein with six predicted epitopes from surface proteins all of them bound in a single polypeptide sequence. This synthetic protein proved to be a good immunogen, and to stimulate CD8+ T cells from seropositive patients in comparison with a mixture of the same antigens and tachyzoite lysates, also the survival percentage in murine models infected with parasites increased substantially in the immunized individuals [85]. The list of new vaccine candidates against Plasmodium and T. gondii still grows. The information that is generated in these two well-studied models serves as a starting point to extrapolate the strategies and propose new ones in the research of vaccines against apicomplexan parasites. This last section is a very narrow landscape of a long list of options that have been generated against other parasites that for sure are helping scientists to find the effective vaccines against cattle babesiosis.

References

  1. 1. Tussaint M. Piroplasma bigeminum en Mexico. Bol del Inst Patológico, Estac Agrícola Cent 1. 1905
  2. 2. McCosker PJ. The global importance of babesiosis. In: Babesiosis. Ristic M, Kreier JP, editors. New York: Academic Press; 1981. pp. 1-24
  3. 3. Babes V. Sur l’ hémoglobinurie bactérienne du boeuf. Comptes Rendus. Académie des Sciences. 1888;107:692-694
  4. 4. Smith T, Kilborne FL. Investigations into the nature, causation, and prevention of Texas or southern cattle fever. Bureau of Animal Industry, Bulletin. 1893;1:85-114
  5. 5. Mosqueda J, Falcon A, Ramos JA, Canto GJ, Camacho-Nuez M. Genome and Molecular strategies for Bovine Babesiosis Control. Revista Mexicana de Ciencias Pecuarias. 2012;3(Supl 1):51-59
  6. 6. Musoke A, Rowlands J, Nene V, Nyanjui J, Katende J, Spooner P, 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
  7. 7. Holec L, Gasior A, Brillowska-Dabrowska A, Kur J. Toxoplasma gondii: Enzyme-linked immunosorbent assay using different fragments of recombinant microneme protein 1 (MIC1) for detection of immunoglobulin G antibodies. Experimental Parasitology. 2008;119(1):1-6
  8. 8. Ravi G, Ella K, Lakshmi Narasu M. Development of pilot scale production process and characterization of a recombinant multiepitope malarial vaccine candidate FALVAC-1A expressed in Escherichia coli. Protein Expression and Purification. 2008;61(1):57-64
  9. 9. 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 Pathogens. 2007;3(10):1401-1413
  10. 10. de Vries E, Corton C, Harris B, Cornelissen AWCA, Berriman M. Expressed sequence tag (EST) analysis of the erythrocytic stages of Babesia bovis. Veterinary Parasitology 2006;138(1-2):61–74
  11. 11. Lau AOT, Tibbals DL, McElwain TF. Babesia bovis: The development of an expression oligonucleotide microarray. Experimental Parasitology. 2007;117(1):93-98
  12. 12. Kohler S, Delwiche CF, Denny PW, Tilney LG, Webster P, Wilson RJ, et al. A plastid of probable green algal origin in apicomplexan parasites. Science. 1997;275(5305):1485-1489
  13. 13. Williamson DH, Gardner MJ, Preiser P, Moore DJ, Rangachari K, Wilson RJ. The evolutionary origin of the 35 kb circular DNA of Plasmodium falciparum: New evidence supports a possible rhodophyte ancestry. Molecular & General Genetics. 1994;243(2):249-252
  14. 14. Zhang Z, Green BR, Cavalier-Smith T. Phylogeny of ultra-rapidly evolving dinoflagellate chloroplast genes: A possible common origin for sporozoan and dinoflagellate plastids. Journal of Molecular Evolution. 2000;51(1):26-40
  15. 15. Moore RB, Oborník M, Janouškovec J, Chrudimský T, Vancová M, Green DH, et al. A photosynthetic alveolate closely related to apicomplexan parasites. Nature. 2008;451(7181):959-963
  16. 16. Sam-Yellowe TY. Rhoptry organelles of the apicomplexa: Their role in host cell invasion and intracellular survival. Parasitology Today. 1996;12(8):308-316
  17. 17. Preiser P, Kaviratne M, Khan S, Bannister L, Jarra W. The apical organelles of malaria merozoites: Host cell selection, invasion, host immunity and immune evasion. Microbes and Infection. 2000;2(12):1461-1477
  18. 18. Yokoyama N, Okamura M, Igarashi I. Erythrocyte invasion by Babesia parasites: Current advances in the elucidation of the molecular interactions between the protozoan ligands and host receptors in the invasion stage. Veterinary Parasitology. 2006;138(1-2):22-32
  19. 19. Allred DR, Al-Khedery B. Antigenic variation as an exploitable weakness of babesial parasites. Veterinary Parasitology. 2006;138(1-2):50-60
  20. 20. Curnow JA. Studies on antigenic changes and strain differences in Babesia argentina infections. Australian Veterinary Journal. 1973;49(6):279-283
  21. 21. Curnow JA. In vitro agglutination of bovine erythrocytes infected with Babesia argentina. Nature. 1968;217(5125):267-268
  22. 22. O’Connor RM, Allred DR. Selection of Babesia bovis-infected erythrocytes for adhesion to endothelial cells coselects for altered variant erythrocyte surface antigen isoforms. Journal of Immunology (Baltimore, MD 1950). 2000;164(4):2037-2045
  23. 23. O’Connor RM, Lane TJ, Stroup SE, Allred DR. Characterization of a variant erythrocyte surface antigen (VESA1) expressed by Babesia bovis during antigenic variation. Molecular and Biochemical Parasitology. 1997;89(2):259-270
  24. 24. Ferreri LM, Brayton KA, Sondgeroth KS, Lau AOT, Suarez CE, McElwain TF. Expression and strain variation of the novel “small open reading frame” (smorf) multigene family in Babesia bovis. International Journal for Parasitology. 2012;42(2):131-138
  25. 25. Berens SJ, Brayton KA, Molloy JB, Bock RE, Lew AE, McElwain TF. Merozoite surface antigen 2 proteins of Babesia bovis vaccine breakthrough isolates contain a unique hypervariable region composed of degenerate repeats. Infection and Immunity. 2005;73(11):7180-7189
  26. 26. Carcy B, Précigout E, Schetters T, Gorenflot A. Genetic basis for GPI-anchor merozoite surface antigen polymorphism of Babesia and resulting antigenic diversity. Veterinary Parasitology. 2006;138(1-2):33-49
  27. 27. Lau AOT, Cereceres K, Palmer GH, Fretwell DL, Pedroni MJ, Mosqueda J, et al. Genotypic diversity of merozoite surface antigen 1 of Babesia bovis within an endemic population. Molecular and Biochemical Parasitology. 2010;172(2-2):107-112
  28. 28. Florin-Christensen M, Suarez CE, Hines SA, Palmer GH, Brown WC, McElwain TF. The Babesia bovis merozoite surface antigen 2 locus contains four tandemly arranged and expressed genes encoding immunologically distinct proteins. Infection and Immunity. 2002;70(7):3566-3575
  29. 29. Hines SA, Palmer GH, Jasmer DP, McGuire TC, McElwain TF. Neutralization-sensitive merozoite surface antigens of Babesia bovis encoded by members of a polymorphic gene family. Molecular and Biochemical Parasitology. 1992;55(1-2):85-94
  30. 30. Hines SA, Palmer GH, Jasmer DP, Goff WL, McElwain TF. Immunization of cattle with recombinant Babesia bovis merozoite surface antigen-1. Infection and Immunity. 1995;63(1):349-352
  31. 31. Mosqueda J, McElwain TF, Palmer GH. Babesia bovis merozoite surface antigen 2 proteins are expressed on the merozoite and sporozoite surface, and specific antibodies inhibit attachment and invasion of erythrocytes. Infection and Immunity. 2002;70(11):6448-6455
  32. 32. Suarez CE, Florin-Christensen M, Hines SA, Palmer GH, Brown WC, McElwain TF. Characterization of allelic variation in the Babesia bovis merozoite surface antigen 1 (MSA-1) locus and identification of a cross-reactive inhibition-sensitive MSA-1 epitope. Infection and Immunity. 2000;68(12):6865-6870
  33. 33. Wilkowsky SE, Farber M, Echaide I, Torioni de Echaide S, Zamorano PI, Dominguez M, et al. Babesia bovis merozoite surface protein-2c (MSA-2c) contains highly immunogenic, conserved B-cell epitopes that elicit neutralization-sensitive antibodies in cattle. Molecular and Biochemical Parasitology. 2003;127(2):133-141
  34. 34. Borgonio V, Mosqueda J, Genis AD, Falcon A, Alvarez JA, Camacho M, et al. msa-1 and msa-2c gene analysis and common epitopes assessment in Mexican Babesia bovis isolates. Annals of the New York Academy of Sciences. 2008;1149:145-148
  35. 35. Leroith T, Brayton KA, Molloy JB, Bock RE, Hines SA, Lew AE, et al. Sequence variation and immunologic cross-reactivity among Babesia bovis merozoite surface antigen 1 proteins from vaccine strains and vaccine breakthrough isolates. Infection and Immunity. 2005;73(9):5388-5394
  36. 36. Dowling SC, Perryman LE, Jasmer DPA. Babesia bovis 225-kilodalton spherical-body protein: Localization to the cytoplasmic face of infected erythrocytes after merozoite invasion. Infection and Immunity. 1996;64(7):2618-2626
  37. 37. Dalrymple BP, Peters JM, Goodger BV, Bushell GR, Waltisbuhl DJ, Wright IG. Cloning and characterisation of cDNA clones encoding two Babesia bovis proteins with homologous amino- and carboxy-terminal domains. Molecular and Biochemical Parasitology. 1993;59(2):181-189
  38. 38. Silva MG, Ueti MW, Norimine J, Florin-Christensen M, Bastos RG, Goff WL, et al. Babesia bovis expresses Bbo-6cys-E, a member of a novel gene family that is homologous to the 6-cys family of Plasmodium. Parasitology International. 2011;60(1):13-18
  39. 39. Shenai BR, Sijwali PS, Singh A, Rosenthal PJ. Characterization of native and recombinant falcipain-2, a principal trophozoite cysteine protease and essential hemoglobinase of Plasmodium falciparum. The Journal of Biological Chemistry. 2000;275(37):29000-29010
  40. 40. Mesplet M, Echaide I, Dominguez M, Mosqueda JJ, Suarez CE, Schnittger L, et al. Bovipain-2, the falcipain-2 ortholog, is expressed in intraerythrocytic stages of the tick-transmitted hemoparasite Babesia bovis. Parasites & Vectors. 2010;3:113
  41. 41. Shen B, Sibley LD. The moving junction, a key portal to host cell invasion by apicomplexan parasites. Current Opinion in Microbiology. 2012;15(4):449-455
  42. 42. Besteiro S, Michelin A, Poncet J, Dubremetz JF, Lebrun M. Export of a Toxoplasma gondii Rhoptry Neck Protein Complex at the Host Cell Membrane to Form the Moving Junction during Invasion. PLOS Pathogens. 2009;5(2):1-14
  43. 43. Srinivasan P, Beatty WL, Diouf A, Herrera R, Ambroggio X, Moch JK, et al. Binding of Plasmodium merozoite proteins RON2 and AMA1 triggers commitment to invasion. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(32):13275-13280
  44. 44. Ord RL, Rodriguez M, Cursino-Santos JR, Hong H, Singh M, Gray J, et al. Identification and characterization of the rhoptry neck protein 2 in Babesia divergens and B. microti. Infection and Immunity. 2016;84(5):1574-1584
  45. 45. Srinivasan P, Ekanem E, Diouf A, Tonkin ML, Miura K, Boulanger MJ, et al. Immunization with a functional protein complex required for erythrocyte invasion protects against lethal malaria. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(28):10311-10316
  46. 46. Gaffar FR, Yatsuda AP, Franssen FFJ, 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. Infection and Immunity 2004;72(5):2947–2955
  47. 47. Hodder AN, Crewther PE, Matthew ML, Reid GE, Moritz RL, Simpson RJ, et al. The disulfide bond structure of Plasmodium apical membrane antigen-1. The Journal of Biological Chemistry. 1996;271(46):29446-29452
  48. 48. Suarez CE, Palmer GH, Hötzel I, Hines SA, McElwain TF. Sequence and functional analysis of the intergenic regions separating babesial rhoptry-associated protein-1 (rap-1) genes. Experimental Parasitology. 1998;90(2):189-194
  49. 49. 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. Molecular and Biochemical Parasitology. 2003;127(2):101-112
  50. 50. Suarez CE, Palmer GH, Jasmer DP, Hines SA, Perryman LE, McElwain TF. Characterization of the gene encoding a 60-kilodalton Babesia bovis merozoite protein with conserved and surface exposed epitopes. Molecular and Biochemical Parasitology. 1991;46(1):45-52
  51. 51. Suarez CE, McElwain TF, Stephens EB, Mishra VS, Palmer GH. Sequence conservation among merozoite apical complex proteins of Babesia bovis, Babesia bigemina and other apicomplexa. Molecular and Biochemical Parasitology. 1991;49(2):329-332
  52. 52. Suarez CE, Palmer GH, Hines SA, McElwain TF. Immunogenic B-cell epitopes of Babesia bovis rhoptry-associated protein 1 are distinct from sequences conserved between species. Infection and Immunity. 1993;61(8):3511-3517
  53. 53. Mosqueda J, McElwain TF, Stiller D, Palmer GH. Babesia bovis merozoite surface antigen 1 and rhoptry-associated protein 1 are expressed in sporozoites, and specific antibodies inhibit sporozoite attachment to erythrocytes. Infection and Immunity. 2002;70(3):1599-1603
  54. 54. Suarez CE, McElwain TF, Echaide I, Torioni de Echaide S, Palmer GH. Interstrain conservation of babesial RAP-1 surface-exposed B-cell epitopes despite rap-1 genomic polymorphism. Infection and Immunity. 1994;62(8):3576-3579
  55. 55. Brown WC, McElwain TF, Ruef BJ, Suarez CE, Shkap V, Chitko-McKown CG, et al. Babesia bovis rhoptry-associated protein 1 is immunodominant for T helper cells of immune cattle and contains T-cell epitopes conserved among geographically distant B. bovis strains. Infection and Immunity. 1996;64(8):3341-3350
  56. 56. Silva MG, Ueti MW, Norimine J, Florin-Christensen M, Bastos RG, Goff WL, et al. Babesia bovis expresses a neutralization-sensitive antigen that contains a microneme adhesive repeat (MAR) domain. Parasitology International. 2010;59(2):294-297
  57. 57. Lindquist S, Craig EA. The heat-shock proteins. Annual Review of Genetics. 1988;22:631-677
  58. 58. Brown WC, Ruef BJ, Norimine J, Kegerreis KA, Suarez CE, Conley PG, et al. A novel 20-kilodalton protein conserved in Babesia bovis and B. bigemina stimulates memory CD4(+) T lymphocyte responses in B. bovis-immune cattle. Molecular and Biochemical Parasitology. 2001;118(1):97-109
  59. 59. Montero E, Rodriguez M, Gonzalez L-M, Lobo CA. Babesia divergens: Identification and characterization of BdHSP-20, a small heat shock protein. Experimental Parasitology. 2008;119(2):238-245
  60. 60. Norimine J, Mosqueda J, Palmer GH, Lewin HA, Brown WC. Conservation of Babesia bovis small heat shock protein (Hsp20) among strains and definition of T helper cell epitopes recognized by cattle with diverse major histocompatibility complex class II haplotypes. Infection and Immunity. 2004;72(2):1096-1106
  61. 61. 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 Pathogens. 2007;3(10):e148
  62. 62. Babesia bigemina: Protozoan genomes. The Sanger Institute. [Internet]. 2013. Available from: http://www.sanger.ac.uk/resources/downloads/protozoa/babesia-bigemina.html [Accessed: 2017-03-15]
  63. 63. Mahoney DF, Wright IG, Goodger BV. Immunity in cattle to Babesia bovis after single infections with parasites of various origin. Australian Veterinary Journal. 1979;55(1):10-12
  64. 64. Dalgliesh RJ, Callow LL, Mellors LT, McGregor W. Development of a highly infective Babesia bigemina vaccine of reduced virulence. Australian Veterinary Journal. 1981;57(1):8-11
  65. 65. Hall WTK, Tammemagi L, Johnston LAY. Bovine Babesiosis: The immunity of calves to Babesia bigemina infection. Australian Veterinary Journal. 1968;44(6):259-264
  66. 66. Tick fever vaccines for cattle. The State of Queensland [Internet]. 2015. Available from: https://www.business.qld.gov.au/industries/farms-fishing-forestry/agriculture/livestock/cattle/tick-fever-vaccines [Accessed: 2017-03-15]
  67. 67. Tick fever vaccines for cattle. The State of Queensland [Internet]. 2015. Available from: https://www.business.qld.gov.au/industries/farms-fishing-forestry/agriculture/livestock/cattle/tick-fever-vaccines/vaccine-info [Accessed: 2017-03-15]
  68. 68. Allred DR, Carlton JM-R, Satcher RL, Long JA, Brown WC, Patterson PE, et al. The ves multigene family of B. bovis encodes components of rapid antigenic variation at the infected erythrocyte surface. Molecular Cell. 2000;5(1):153-162
  69. 69. Al-Khedery B, Allred DR. Antigenic variation in Babesia bovis occurs through segmental gene conversion of the ves multigene family, within a bidirectional locus of active transcription. Molecular Microbiology. 2006;59(2):402-414
  70. 70. 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 Pathogens. 2009;5(2):e1000309
  71. 71. Moreau E, Bonsergent C, Al Dybiat I, Gonzalez LM, Lobo CA, Montero E, et al. Babesia divergens apical membrane antigen-1 (BdAMA-1): A poorly polymorphic protein that induces a weak and late immune response. Experimental Parasitology. 2015;155:40-45
  72. 72. Sivakumar T, Okubo K, Igarashi I, de Silva WK, Kothalawala H, Silva SSP, et al. Genetic diversity of merozoite surface antigens in Babesia bovis detected from Sri Lankan cattle. Infection, Genetics and Evolution 2013;19:134-140
  73. 73. Torina A, Agnone A, Sireci G, Mosqueda JJ, Blanda V, Albanese I, et al. Characterization of the apical membrane antigen-1 in Italian strains of Babesia bigemina: Characterization of the AMA-1 in Babesia bigemina. Transboundary and Emerging Diseases. 2010;57(1-2):52-56
  74. 74. McElwain TF, Perryman LE, Musoke AJ, McGuire TC. Molecular characterization and immunogenicity of neutralization-sensitive Babesia bigemina merozoite surface proteins. Molecular and Biochemical Parasitology. 1991;47(2):213-222
  75. 75. Fisher TG, McElwain TF, Palmer GH. Molecular basis for variable expression of merozoite surface antigen gp45 among American isolates of Babesia bigemina. Infection and Immunity. 2001;69(6):3782-3790
  76. 76. Schlüter K, Jockusch BM, Rothkegel M. Profilins as regulators of actin dynamics. Biochimica et Biophysica Acta (BBA) – Molecular Cell Research. 1997;1359(2):97-109
  77. 77. Plattner F, Yarovinsky F, Romero S, Didry D, Carlier M-F, Sher A, et al. Toxoplasma profilin is essential for host cell invasion and TLR11-dependent induction of an interleukin-12 response. Cell Host & Microbe. 2008;3(2):77-87
  78. 78. Munkhjargal T, Aboge GO, Ueno A, Aboulaila M, Yokoyama N, Igarashi I. Identification and characterization of profilin antigen among Babesia species as a common vaccine candidate against babesiosis. Experimental Parasitology. 2016;166:29-36
  79. 79. Pan W, Huang D, Zhang Q, Qu L, Zhang D, Zhang X, et al. Fusion of two malaria vaccine candidate antigens enhances product yield, immunogenicity, and antibody-mediated inhibition of parasite growth in vitro. Journal of Immunology. 2004;172(10):6167-6174
  80. 80. Khasnis AA, Nettleman MD. Global warming and infectious disease. Archives of Medical Research. 2005;36(6):689-696
  81. 81. Edi CVA, Koudou BG, Jones CM, Weetman D, Ranson H. Multiple-insecticide resistance in Anopheles gambiae mosquitoes, Southern Côte d’Ivoire. Emerging Infectious Diseases. 2012;18(9):1508-1511
  82. 82. Soe S, Theisen M, Roussilhon C, Aye K-S, Druilhe P. Association between protection against clinical malaria and antibodies to merozoite surface antigens in an area of hyperendemicity in Myanmar: Complementarity between responses to merozoite surface protein 3 and the 220-kilodalton glutamate-rich protein. Infection and Immunity. 2004;72(1):247-252
  83. 83. van Schaijk BCL, van Dijk MR, van de Vegte-Bolmer M, van Gemert G-J, van Dooren MW, Eksi S, et al. Pfs47, paralog of the male fertility factor Pfs48/45, is a female specific surface protein in Plasmodium falciparum. Molecular and Biochemical Parasitology 2006;149(2):216–222
  84. 84. Theisen M, Roeffen W, Singh SK, Andersen G, Amoah L, van de Vegte-Bolmer M, et al. A multi-stage malaria vaccine candidate targeting both transmission and asexual parasite life-cycle stages. Vaccine 2014;32(22):2623-2630
  85. 85. El Bissati K, Chentoufi AA, Krishack PA, Zhou Y, Woods S, Dubey JP, et al. Adjuvanted multi-epitope vaccines protect HLA-A*11:01 transgenic mice against Toxoplasma gondii. JCI Insight. 2016;1(15):1-18

Written By

Juan Mosqueda, Diego Josimar Hernández-Silva and Mario Hidalgo-Ruiz

Submitted: 18 August 2017 Reviewed: 22 November 2017 Published: 20 December 2017