• Users Online: 12
  • Home
  • Print this page
  • Email this page
Home About us Editorial board Ahead of print Current issue Search Archives Submit article Instructions Subscribe Contacts Login 

 Table of Contents  
REVIEW ARTICLE
Year : 2015  |  Volume : 8  |  Issue : 1  |  Page : 14-37

Heat shock proteins and parasitic diseases: Part II. Protozoa


Department of Parasitology, Faculty of Medicine, Suez Canal University, Ismailia, Egypt

Date of Submission22-Feb-2015
Date of Acceptance17-Apr-2015
Date of Web Publication24-Aug-2015

Correspondence Address:
Sherif M Abaza
Department of Parasitology, Faculty of Medicine, Suez Canal University, Ismailia
Egypt
Login to access the Email id

Source of Support: None, Conflict of Interest: None


DOI: 10.4103/1687-7942.163407

Rights and Permissions
  Abstract 

List of contents
1. Plasmodium spp.
1.1. Introduction
1.2. Historical background
1.3. Applications
2. Leishmania spp.
2.1. Introduction
2.2. Applications
3. Trypanosoma spp.
3.1. African trypanosomiasis
3.1.1. Applications
3.2. American trypanosomiasis
3.2.1. Introduction
3.2.2. Applications
4. Toxoplasma gondii
4.1. Applications
5. Cryptosporidium spp.
5.1. Applications
6. Other protozoa
6.1. Babesia spp.
6.2. Microsporidium spp.
6.3. Giardia lamblia
6.4. Eimeria spp.
6.5. Trichomonas vaginalis
6.6. Entamoeba histolytica
6.7. Free living amoeba
6.8. Cyclospora cayetanensis
6.9. Blastocystis spp.
6.10. Theileria spp.
Concluding Remarks
References

Abbreviations
ASS: African sleeping sickness; BiP: Binding protein; BS: Bloodstream forms; CL: Cutaneous leishmaniasis; COWP: Cryptosporidium oocyst wall protein; CTL: Cytotoxic T-lymphocyte; Cy: Cytosolic; ER: Endoplasmic reticulum; GA: Geldanamycin; GP60: Glycoprotein 60; HSP: Heat shock protein; ITS: Internal transcribed spacer; KMP-11: Kinetoplastid membrane protein-11 gene; MAb: Monoclonal antibody; MCL: Mucocutaneous leishmaniasis; MHC: Major histocompatibility complexe; Mit: Mitochondrial; NO: Nitric oxide; PS: Procyclic forms; PTEX: Plasmodium translocon of exported proteins; sHSP: small heat shock protein; SSU: Small subunit; TL: Tegumentary leishmaniasis; TLR: Toll like receptor; VL: Visceral leishmaniasis; VSG: Variant surface glycoprotein.

Keywords: apoptosis, drug target, heat shock protein, immune response, parasite survival, protective immunity, protozoa, serodiagnostic marker, vaccine candidate, virulence


How to cite this article:
Abaza SM. Heat shock proteins and parasitic diseases: Part II. Protozoa. Parasitol United J 2015;8:14-37

How to cite this URL:
Abaza SM. Heat shock proteins and parasitic diseases: Part II. Protozoa. Parasitol United J [serial online] 2015 [cited 2023 Sep 26];8:14-37. Available from: http://www.new.puj.eg.net/text.asp?2015/8/1/14/163407


  1. Plasmodium species Top


1.1. Introduction

It is apparent that heat shock proteins (HSPs) play an important role in the survival of Plasmodium spp. against temperature changes associated with its passage from the cold-blooded mosquito vector to the warm-blooded human host. Added to that, there is increased temperature during febrile episodes in malaria. Moreover, it was found that P. falciparum has a proteome with a majority of asparagine (Asn) residues in its amino acid repeats [1] . The presence of Asn-rich sequences in P. falciparum proteins leads to increased aggregate formation that is enhanced under stress conditions with increased temperature. Therefore, P. falciparum has to survive conditions that increase Asn repeat-rich protein aggregation [2] . It was also reported that the capacity of P. falciparum to grow and thrive inside the host bloodstream depends on its ability to export ~5% of its encoded genome (200-300 proteins) into the host cell cytosol [3] . These exported proteins (EXPs) play an important role in virulence by promoting cell rigidity and adhesion of infected cells [4] . To gain access to the host cell cytosol, EXPs must cross the parasite plasma membrane and the parasitophorous vacuole membrane. Transfer across the parasite envelope additionally requires a Plasmodium translocon of exported proteins (PTEX). Discovery of this element was a major advance as it permitted forecasting of the exported proteome of Plasmodium spp. In their studies, Australian investigators identified a translocon that comprised HSP101, PTEX150, and exported protein 2 (EXP2) with two additional accessory proteins PTEX88 and thioredoxin 2 (TRX2). They suggested that this translocon might offer new strategy for development of new drug targets [5] .

At least six P. falciparum HSP70 homologs have been identified across the cytosolic (Cy), endoplasmic reticulum (ER), and mitochondria (Mit) [3] . Of these, only PfHSP70-1 received widespread research attention due to its potential use as a vaccine candidate. A review article discussed its Cy localization, the identical sequence of PfHSP70-1 and PfHSP70-x, and their ability to move intranuclearly when exposed to stress. The authors also proved that PfHSP70-2 originates from ER and interacts with more P. falciparum proteins compared with PfHSP70-1 and PfHSP70-3 (Mit HSP70 homolog) [6] . In another review article also published in 2007, Indian scientists comparatively analyzed structures, complexes, substrate client proteins, and functions of Plasmodium HSPs. Among 92 chaperones encoded by the Plasmodium genomic structure, they discussed the possible applications of HSP90 as a drug target, HSP70 as a vaccine candidate, and HSP40 as the best representative HSP [7] . In 2009, Chiang and his colleagues presented 5 events for the essential functions of HSP70 and HSP40 in P. falciparum life cycle: 1) HSP40 interacts with HSP70 through a conserved bundle known as J domain which enhances HSP70 ATPase activity, 2) HSP70-HSP40 pairs transport across membranes and are essential for parasite's membrane integrity, 3) The formed J domain is required for 'knob' formation to bind parasitized red blood cells (RBCs) to the vascular endothelium, 4) both HSPs are necessary to offset cellular stresses encountered during the P. falciparum life cycle, 5) molecular chaperones of both HSPs help retaining newly synthesized polypeptides in soluble conformations and facilitate protein folding; a function which enables P. falciparum to exhibit protein synthesis when the trophozoite stage starts to initiate several rounds of intracellular division. Based on these events, P. falciparum viability should be sensitive to HSP70-HSP40 inhibition [8] .

Several review articles were published during 2010-2014. Shonhai [9] analyzed and discussed Plasmodium HSP90, HSP70/HSP40 pairing, and small HSPs (sHSPs) structural and functional features, and their potentiality and importance for parasite growth, development, and virulence. He also proposed mechanisms of action of drug targets that inhibit HSPs [9] . Another team from South Africa reviewed also the potential uses of Plasmodium HSPs (90, 70, and 40) as new drug targets against malaria [10] . Rug and Maier [11] reviewed the potential roles of HSP40 in Plasmodium spp. In 2013, a third team from South Africa reviewed Plasmodium HSP70-J protein potentials as a novel antimalarial drug target [12] . Indian reviewers analyzed HSP90 functions in Plasmodium spp. growth and development during febrile malarial attacks and claimed the development of seven antimalarial drugs as well as 30 anticancer therapies under development using research for HSP90 inhibitors [13] . In 2014, three review articles were published. In the first one from Australia, the reviewers discussed PfHSP70-PfHSP40 partnerships. They mentioned that, despite poor and extensive characterization of PfHSP40 and PfHSP70, both interact together and share in several functions, including protein homoeostasis, cytoprotection, and protein trafficking across the parasitophorous vacuole as well as into the infected RBCs [14] . In the second review conducted in Singapore, the reviewers claimed the identification of more than ten cochaperones for HSP90 in the P. falciparum genome of which only five (PfHop, Pfp23, PfAha1, PfPP5, and PfFKBP35) were experimentally proved to interact with HSP90. They also discussed the regulatory roles of PfHSP90's cochaperons as well as their potential influence in the development of new antimalarial drugs targeting PfHSP90 [15] . The last review article, which was also published by Australian reviewers, dealt with exported parasite proteins and their relation to virulence and in generating a new trafficking system in the infected RBCs [16] .

On the other hand, Pavithra et al. [17] carried out a system-level analysis of chaperone networks in P. falciparum, which allowed the scientists to: 1) Predict presumed functions of several hypothetical proteins in P. falciparum, 2) Draw attention to some chaperones that have not been characterized yet, 3) Highlight chaperone systems that are missing in the malarial parasite, 4) Provide a possible basis for the antimalarial activity of drugs affecting the chaperone system; and finally, 5) Determine the functional significance of chaperones during the different developmental stages of the parasite.

They concluded that understanding P. falciparum biology is essential for rational development of antimalarial drugs [17] . Furthermore, new observations regarding P. falciparum biology in vivo were reported. On the basis of the comparison of gene expression profiles of different isolates, parasites were categorized into three clusters of which cluster 1 represented starvation response and cluster 3 represented environmental stress response [18] . In another article, investigators analyzed the transcript levels of 103 chaperones from different P. falciparum clinical isolates. They found that most of the chaperones are highly overexpressed and display a specific pattern. Their results suggested overexpression of Mit and apicoplast chaperones in cluster 1 category, with overexpression of Cy chaperones in cluster 2 category. Cluster 3 consisted of two subclusters, one of which overexpressed Cy chaperones. They also found HSP70 overexpression in cluster 2, whereas HSP90 showed maximum upregulation in cluster 2 and in a specific subpopulation in cluster 3. They concluded that parasites of cluster 3 (representing environmental stress response) are further subclustered on the basis of hsp90 gene expression, and that HSPs gene expression in different P. falciparum clinical isolates is important in understanding host-parasite interactions and, subsequently, treatment of severe malaria [19] .

1.2. Historical background

Ardeshir et al. [20] recognized a 75 kDa surface protein (P75) localized in P. falciparum merozoite. Its sequence proved to be homologous with that of HSP70, suggesting its essential role in P. falciparum life cycle. The investigators also suggested the conservative character of the p75 gene as they found its identical size in nine P. falciparum strains [20] . One year later, another genomic DNA clone isolated from P. falciparum was described to encode a 72 kDa protein whose sequence proved homologous to HSP78 (glucose-regulated protein of rat and hamster) [21] . Indian investigators tested serum samples of patients cured of P. falciparum malaria for detection of antischizont and anti-HSP70 antibodies, and found increased levels of schizont antibodies that significantly inhibited merozoite invasion in vitro. Although they established increased levels of anti-HSP70, no correlation between anti-HSP70 and inhibition of merozoite invasion was detected [22] . In the same year, the same investigators evaluated the human immune response to P. falciparum HSP70. They found that serum samples from more than 50% of patients cured of the disease (with no parasitemia) had no detectable antibodies against PfHSP70, and exhibited antibody response to total schizont antigens. Therefore, they concluded that HSP70 was not a suitable vaccine candidate [23] . Similar results were obtained when Brazilian investigators investigated IgG levels against recombinant protein derived from PfHSP70 in individuals living in areas with unstable and hypoendemic malarial transmission [24] .

In 1992, Sharma [25] identified five HSP genes from P. falciparum, and he discussed their biological roles as stress proteins to protect the parasite from various stresses encountered in the host. In addition, he attributed their antigenicity to the nonhomologous sequences in the C-terminal region [25] . In 1993, American scientists detected expression of proteins with similar sequence to HSP70 in the hepatic stages of malarial parasites. They concluded their potential role in stimulation of immune responses against infected hepatocytes [26] . Later, it was found that the parasitic growth rates among five different P. falciparum isolates were not affected after increasing the temperature to 39°C for 30 min, with marked significant HSP70 expression. The investigators emphasized that the P. falciparum hsp70 gene responded to heat stress by expression of PfHSP70 to protect the parasite from being killed during malarial fever [27] . Moreover, it was shown that monoclonal antibodies (mAbs) generated by mice immunized with P. berghei-infected RBCs were reactive with HSP70 expressed in hepatic and erythrocytic stages. In contrast to previous results, the investigators demonstrated HSP70 expression in malarial sporozoites [28] .

1.3. Applications

1.
Life cycle development: PfHSP90 was shown to have an essential role in parasite development within the erythrocytic stages during the frequent malarial febrile episodes [29] . In another Indian study using the HSP90 inhibitor [geldanamycin (GA)], the investigators observed the significant role of HSP90 in parasite development and growth [30] . In addition, PfHSP40 was found to share in the regulation of transport of exported proteins across several membranes to target RBCs [31] . HSP40 members are divided into four distinct classes on the basis of their encoded domains. Type IV HSP40 protein (PfL2550w), which was named P. falciparum gametocyte erythrocyte Cy protein (PfGECO), was found to have a role in sexual differentiation. Using northern blotting, the investigators demonstrated that PfGECO was transcribed from asexual stages to stage III gametocytes, and it was predominantly expressed in stage I-IV gametocytes. The investigators concluded that gametocytes sustain many exported proteins from the asexual blood stages to facilitate the changes in shape that occur during their five stages of development and/or to provide protection throughout it. Therefore, type IV HSP40 genes do have an essential role during gametocyte sequestration [32] . By using the molecular genetics approach, investigators showed that Plasmodium HSP20 is critical for fast sporozoite locomotion in vitro and for efficient natural malarial transmission in vivo. Based on their observational studies, gliding sporozoites motility was critical for 1) their entry and dispersion into the mosquito salivary glands, 2) intradermal migration from the mosquito to blood of the host, 3) penetration of the endothelial barrier of liver sinusoid and successful invasion and migration through the liver parenchyma. The investigators postulated that HSP20 directly contributes to cell-substrate adhesion by physical interaction with transmembrane parasite invasins. According to their explanation, HSP20 modulates the motor complex that generates the traction forces during sporozoite locomotion, and it regulates the turnover of actin polymerization and competes with actomyosin interaction, or both [33] .

2. Parasite invasion and virulence: HSP70 was identified as one of the five major proteins responsible for regulation of parasite actin polymerization, which provides the mechanism for localized actin filament growth and movement of the parasite into the host cell [34] . In another study, the investigators claimed that initial exposure of the parasites to HSPs provided better survival and improved infectivity. Accordingly, their conclusion was that febrile episodes promoted intraerythrocytic development of the parasite [30] . In 2012, other investigators discovered the ability of Cy HSP110 to prevent Asn repeat-rich protein aggregation in vitro and in vivo. They concluded that HSP110 is vital for proteostasis of P. falciparum proteome, allowing the propagation of these repeats within the parasite proteome, and for parasite growth and survival during its brief exposure to febrile temperatures [2] . Recently, after investigating a conserved Plasmodium protein (named plasmoDJ1) as regards its response to stress and its role in parasite growth and development, it was found that the expression of this particular protein in all intraerythrocytic stages and ookinetes increased upon heat stress as well as oxidative stress due to H 2 O 2 and artemisinin treatment. The study showed that plasmoDJ1 not only reduced H 2 O 2 and significantly decreased protease activities but also attenuated virulence and reduced oocyst production upon plasmoDJ1 gene knockout [35] .

3. Elicitation of immune response: γ δ T cells collected from experimental P. yoelii-infected mice were capable of proliferation in vitro in response to PfHSP60 and PfHSP70, suggesting the immunological role of HSPs in cellular immune response [36] . HSP65 was found to be strongly expressed in splenic cells of mice infected with a nonlethal strain of P. yoelii and slightly expressed in mice infected with a lethal strain. The investigators suggested the important role of HSP65 in protection against malarial infection in mice [37] . Two years later, the same investigators extended their studies to confirm the crucial role of HSP65 in the prevention of macrophage apoptosis in P. yoelii-infected mice, and CD4 + T cells were proven to be essential for HSP65 expression [38] . In another study conducted by the same investigators in Thailand on malaria patients, IgG, IgM, and IgA significantly increased against HSP90, whereas only IgM became significantly elevated in response to HSP70 and IgA in response to HSP65. The investigators concluded that in malarial infection the antigenic potential of HSP90 is higher than that of HSP70 and HSP65 [39] . In Korea, investigators attempted to characterize the molecular properties of HSP70 of P. vivax. They obtained isolates from Korea, Myanmar, Thailand, and Indonesia and amplified the gene encoding PvHSP70. They tested it against patients' sera infected with P. vivax and P. falciparum as well as in patients with other parasitic diseases and in healthy controls. The results strongly suggested high antigenicity of this HSP and the investigators recommended further studies to evaluate the biological significance of immune responses to PvHSP70 [40] .

4. Diagnosis: As a diagnostic marker, IgG mAbs were raised against PfHSP70 conjugate and used in comparison with malarial antigen conjugate and polyclonal IgG antibodies against P. falciparum malarial conjugate in the diagnosis of P. falciparum malaria. Using a rapid agglutination test, results showed 90% sensitivity and 80% specificity. Similar sensitivity and specificity were obtained with polyclonal Abs, while malarial antigen conjugates gave lower results [41] . In Portugal, investigators succeeded in developing a rapid, simple, and sensitive immunoassay to detect malarial antigens in infected blood cultures. The assay utilized labeled recombinant PfHSP70 bound to gold nanoparticles functionalized with anti-HSP70 mAbs. The immunoassay succeeded in detecting malarial antigen at a 3% parasitemia level [42] .

5. Drug Targets: Indian investigators analyzed the P. falciparum genome and found a high degree of sequence identity between PfHSP90 and yeast HSP90. This result allowed them to derive a 3D structure for PfHsp90 by homology modeling. Accordingly, they claimed that understanding the chemical structure of PfHSP90 would greatly help in the development of new drug targets against malaria [43] . A novel class of antimalarial agents was identified by American scientists in 2009. They evaluated the efficacy of nine pyrimidinone amides (HSP70 modulators) to inhibit the replication of P. falciparum in parasitized RBCs. Their results showed that the compounds altered the ATPase activity of purified PfHSP70 [8] . In addition, intraperitoneal administration of 17-(allylamino)-17-demethoxygeldanamycin (HSP90 inhibitor) to P. berghei-infected mice inhibited parasite growth [44] . The obtained results suggested the possibility of chaperone-targeted therapy for the treatment of a variety of protozoan infections. In another study, 'harman' was identified as an alkaloid extracted from leaves of Guiera senegalensis (herbs found in West and Central Africa). In their studies, Canadian investigators demonstrated that 'harmine' (harman derivative) inhibited P. falciparum HSP90 by competition for the N-terminal ATP-binding domain. They concluded that harmine derivatives displayed preferential binding for PfHSP90 orthologs and provided an ATP-binding domain that confers species specificity [45] . Two years later, the same team of scientists confirmed the specific affinity of harmine to PfHSP90 and unexpectantly discovered a related compound 'harmalol' that showed higher affinity for the human HSP90 ATP-binding pocket better than harmine. The investigators extended their studies to show that harmine had potential synergistic effects on chloroquine and artemisinin in vitro as well as in vivo in P. berghei-infected mice [46] . In the same year, American scientists showed the ability of PfHSP110 for proteostasis of the P. falciparum proteome, which motivated the search for an inhibitor to hamper the proteostasis of the parasite proteome [2] . Very recently, similar results were obtained and the investigators demonstrated the essential and universal roles of PfHSP101 in exporting proteins; they provided strong evidence for PTEX function in protein translocation into the host cell [47] . In addition, Elsworth et al. [48] discussed the functions of PTEX and the possibility of its use as a prime antimalarial drug target. The investigators showed that the major virulence factor PfEMP1 was significantly reduced in PTEX knockdown parasites, and that modest PTEX knockdown had a strong effect on erythrocytic stages both in vitro and in vivo [48] . Finally, a purine analog HSP90 inhibitor (PU-H71) proved to be active against PfHSP90. In addition, the investigators found that PU-H71 had high affinity to the PfHSP90 ATP-binding domain and inhibited ATPase activity. PU-H71 also synergized chloroquine to reduce the parasite load and improve survival rates in P. berghei experimentally infected mice. Therefore, they concluded that PU-H71 exhibits potent antimalarial activity both in vitro and in vivo [49] .

6. Vaccination: In three experiments, American scientists used gene encoding P. yoelii HSP60 for immunization of mice against challenge infection. In the first experiment, 40% protection was observed, whereas instead of protection there was delay in the onset of parasitemia in the second experiment. In the third experiment, there was neither reduction of erythrocytic stages nor development of exoerythrocytic stages. The investigators recommended further evaluation of the use of HSP60 in the development of malarial vaccines [50] . The results obtained from a study conducted in Sweden emphasized the potential role of HSP70 as an adjuvant in a DNA vaccine candidate against P. falciparum [51] . Furthermore, Spanish investigators were able to identify four proteins characterized by high immunogenic blood stage antigens using proteomic analysis of P. yoelii antigens. These were two ER proteins (disulfide isomerase and HSP70), a digestive protease plasmepsin, and a ribosome-associated protein. They claimed that isolation of parasitic antigens using IgG from sera of malaria-protected individuals could be a novel strategy for malaria vaccine development [52] .


  2. Leishmania species Top


2.1. Introduction

It was reported that HSP70 and HSP83 were abundant in Leishmania promastigotes and remained constitutively expressed in amastigotes to protect the organism against temperature change from 22 to 28°C in the sandfly to 33-37°C in the mammalian host [53] . In addition to HSP70 and HSP83, the same investigators characterized sHSPs with lower degree of conservation that shared in the protection of L. mexicana amastigotes against heat stress [54] .

2.2. Applications

1.
Virulence and parasite growth: The role of L. major HSP100 was studied and the investigators found that loss of HSP100 chiefly affected the initial phase of the infection, i.e. development from promastigotes to amastigotes. They also found that the impact of HSP100 is stage specific, which indicated its protective role during the establishment of the parasite within the mammalian host [55] . L. donovani HSP100 was also found to be clustered, with the majority of its molecules localized close to the cytoplasmic membrane [56] . Furthermore, a gene encoding HSP100 (expressed exclusively in amastigotes) was established as a virulence factor by a reverse genetic approach. Gene replacement mutants (lacking this gene) were found avirulent in BALB/c mice, and the virulence was recovered on spontaneous clonal divergence within the mutant population leading to emergence of virulent parasites. In their study, the investigators identified and characterized another virulence gene encoding a 46-kDa protein whose overexpression restored infectivity to the mutant L. major strains [57].

2. Immune response elicitation: L. braziliensis genes encoding HSP70 and HSP83 were cloned, identified, and expressed as recombinant rLbHSP83 and rLbHSP70, and the leukocyte responses and anti-IgG antibody titers in the sera of infected patients with different clinical forms of disease were compared. The investigators found that the differential cytokine elicitation patterns of LbHSP83a and LbHSP83b indicated their potential importance in the development of improved immunotherapeutic or immunoprophylactic targets [58] . Proteins of L. chagasi promastigotes were separated into 67 fractions and tested for their ability to stimulate proliferation of peripheral blood mononuclear cells. The investigators suggested that the 69-kDa protein fraction (with similar sequence identity to HSP70) was able to elicit human T-cell immune responses, an important step toward vaccine development [59] . In addition, both L. infantum HSP70 and HSP83 were proved to be potent mitogens for murine splenocytes and able to induce proliferation on B-cell populations purified from BALB/c spleen [60] . Similar results were obtained in a study conducted in Spain, which revealed that the immune response generated during cutaneous leishmaniasis (CL) and mucocutaneous leishmaniasis (MCL) was elicited specifically by the parasitic histone and HSP70 [61] .

3. Diagnosis: There was much controversy regarding the potential use of HSP70 as a serodiagnostic marker in leishmaniasis. The first description and characterization of L. donovani HSP70 was published in 1990. The investigators demonstrated the presence of antibodies against the carboxy-terminal region of L. donovani HSP70 in more than 50% of visceral leishmaniasis (VL) patients and in all samples of Chagas disease. The investigators suggested at that time its importance during the growth of both parasites [62] . Two years later, recombinant L. donovani HSP70 and HSP90 showed strong humoral immune response, which was evident by strong reaction with antibodies in ELIZA when tested against sera from L. donovani-infected patients, but not with sera from Chagas disease patients. The investigators considered HSP70 as a specific biomarker for VL serodiagnosis [63] . Similar results were obtained when sera from patients with CL, malaria, schistosomiasis, and Chagas disease did not recognize the recombinant L. donovani HSP7O using western blots. The investigators suggested that in VL, an immunodominant anti-HSP7O response was raised against a species-specific epitope in the carboxy-terminal region of HSP70 [64] . However, in another study conducted in India, the investigators identified 20 clones that were reactive with pooled sera from Kala azar-infected patients. The clone with the largest cDNA insert and expressing HSP70 was tested for its suitability in serological assays, but the results revealed that it was unreliable for serodiagnosis of Kala azar [65] . One year later, Spanish investigators reported that recombinant L. infantum HSP70 (rLiHSP70) was not reliable as a disease-specific marker for serodiagnosis as it was recognized by sera from patients with Chagas disease [66] . In 2003, another team from Spain isolated the single-copy gene encoding HSP70 of L. braziliensis (663 amino acids) from the genomic DNA library. They showed that its main antigenic determinant was located in the carboxy-terminal end. Using this fragment for specific serodiagnosis of CL and MCL, their results showed 70% sensitivity and 100% specificity [67] . In Iran, the L. infantum hsp70 gene segment was amplified, cloned, and evaluated for VL serodiagnosis. The ELISA results showed that HSP70 was recognized by 81.1% (30/37) of VL patients and 6.3% (5/63) of controls [68] . From L. infantum cDNA and genomic library, the investigators selected five gene members to evaluate their reactivity against a panel of Leishmania spp.-infected canine and human sera. One of these loci was a gene encoding a protein homolog to members of the Cy HSP70 family. Although it cross-reacted with antibodies from a certain proportion of dogs with other infectious diseases, some serum samples from patients with CL, and serum samples from patients with Chagas disease, its optical density (OD) values were low. However, it showed moderate reactivity with sera from most VL patients. The investigators concluded that none of these selected gene members was efficient for diagnosing all canine or human VL cases [69] . Moreover, in an attempt to identify L. donovani protein antigen with high specificity and sensitivity, and in the same time the test becomes negative on patient recovery, Indian investigators succeeded to identify an immunoreactive BHUP1 antigen from L. donovani HSP70. The antigen proved to be 90% specific among healthy individuals living in the endemic region, and 54% of the cured VL patients turned negative when tested with 1-year follow-up. Thus, the investigators considered L. donovani-specific HSP70 (BHUP1) protein as a potential diagnostic and prognostic marker in VL. They also recommended BHUP1 for a rapid immunochromatographic test [70] . Recently, Brazilian investigators identified specific B-cell epitopes from calpain-like cysteine peptidase, thiol-dependent reductase 1, and HSP70 of L. infantum using in-silico analysis. The detected antigens were suggested to be used as diagnostic biomarkers in VL [71] .

As regards HSP83, research showed that recombinant L. infantum HSP83 (rLiHSP83) elicited strong humoral immune response in 90% of sera from dogs with VL. The results suggested its use in serodiagnostic assays for canine leishmaniasis [72] . The specificity of the antibody response in patients with diffuse CL was evaluated using three immunodominant leishmanial antigens [HSP70, HSP83, and Leishmania eukaryotic initiation factor 4A (LeIF)]. The results revealed an antibody specificity pattern that was predominantly restricted to HSP83. The investigators considered Leishmania HSP83 as an important antigen in the diagnosis of infections caused by L. amazonensis [73] . Recently, rLiHSP83 was evaluated with ELISA for serodiagnosis of CL, MCL, and VL as well as in samples from other diseases such as Chagas disease, toxoplasmosis, and malaria. It was also evaluated in treated patients during a follow-up to test anti-rHSP83 antigen antibody titers. Evaluation of rLiHSP83 in comparison with L. major-like total promastigote antigen showed significantly higher sensitivity and 100% specificity of rLiHSP83. The investigators also did not detect a significant decrease in antibody levels after treatment. Therefore, they concluded that rLiHSP83 is a good biomarker for serodiagnosis of leishmaniasis [74] . Finally, using sandwich ELISA and western blotting, it was concluded that combined L. donovani HSP70 and HSP83 are good serodiagnostic biomarkers in VL. The investigators attributed their results to the induction of strong cell-mediated and humoral immune responses during leishmaniasis [75] .

For tegumentary leishmaniasis (TL), the recombinant L. major hsp60 gene was cloned, sequenced, and screened against sera from patients with TL. The results revealed that rLmHSP60 elicited humoral immune response suggesting its use as a potential antigen for TL serodiagnosis [76] . In another study, recombinant L. infantum (rLiHSP83) was evaluated in comparison with L. major-like total antigen using ELISA. Both antigens were tested against sera from patients with CL and MCL and sera from Chagas disease patients as well as sera from normal blood bank donors. Both antigens showed similar results except for the absence of cross-reactivity that occurred with sera from patients with Chagas disease on use of rLiHSP83. Therefore, the investigators considered it to be a good antigen for routine use in serodiagnosis of TL [77] . On the other hand, the restriction fragment length polymorphism (RFLP) protocol was designed using HSP70 to diagnose and identify Leishmania spp. causing TL in skin scrapings with a sensitivity and specificity of 95 and 100%, respectively [78] . Recently, another group of investigators compared six L. infantum-L. chagasi recombinant proteins, including HSP70 and soluble L. infantum-L. chagasi antigen (SLA), for TL serodiagnosis. All recombinant proteins were recognized by sera from TL patients, and rHSP70 showed the best performance (sensitivity and specificity), which was superior to SLA in the diagnosis of CL. The investigators concluded that rHSP70 and/or in combination with different antigens should be considered a good biomarker in serodiagnosis of TL, especially in endemic areas [79] . In another recent study conducted in Brazil, the investigators examined the ability of L. braziliensis rHSP83.1 in TL serodiagnosis against samples from TL and VL patients and from dogs infected with canine VL. Their results showed the potential use of rHSP83.1 in TL serodiagnosis [80] .

On the other hand, the Leishmania sHSPs family includes the most widespread but most poorly conserved collection of HSPs. In an attempt to characterize L. amazonensis HSP20, investigators utilized the complete bioinformatics of the L. major genome database to identify a protein entry. They found that the 17.5 kDa predicted protein that had 155 amino acids was the only sHSP family member found in all Leishmania spp. On the basis of these data, L. amazonensis HSP20 was expressed and used as a serodiagnostic marker to determine the presence of specific IgG in the sera of Leishmania-infected animals and human patients with VL. Results revealed that HSP20 was recognized by all dog sera, but with limited and moderate antigenicity by 30 and 62% of the assayed human sera and sera from experimentally infected hamsters, respectively. Accordingly, it was concluded that the production of anti-HSP20 antibodies coincided with the onset of disease symptoms. The investigators suggested the use of HSP20 as a useful serodiagnostic marker for active leishmaniasis [81] .

4. Species identification: Leishmania HSP70 genes were found to represent adequate targets for sensitive typing of neotropical Leishmania spp. in host tissues. They encode for several major antigens that may be used for probing of the genetic variability of molecules possibly engaged in immunopathology. In addition, they present a lower rate of genetic variation compared with other markers [e.g. glycoprotein 63 (gp63) or rDNA internal transcribed spacer (ITS)] [82] . On the other hand, a species-specific segment in the L. donovani genome employing the carboxy-terminal region of the rHSP70 was amplified. In their study, the PCR assay identified L. donovani in hepatic and splenic infected tissues, in aspirates from lesions of post-Kala azar dermal leishmaniasis, and in bone marrow aspirates from VL patients [83] .

Due to the multiple Leishmania species in some endemic areas such as Brazil, the investigators analyzed the hsp70 gene sequence in several Leishmania species to identify specific restriction enzymes that could be used for PCR-RFLP-based identification. They suggested that the new identification protocol would correct the limitations of the gold standard technique (multilocus enzyme electrophoresis) used for routine Leishmania typing [84] . In addition, the HSP70 PCR-RFLP assay utilizing BccI as the restriction enzyme could differentiate between L. panamensis and L. guyanensis (species of the subgenus Viannia with very high genomic similarity), which are considered potential causes of CL and MCL. The new restriction enzyme overcame the failure of HaeIII (as restriction enzyme) because it showed the same pattern for both species [85] . In the same year, the same investigators published another work presenting a single PCR-RFLP method based on analysis of 51 HSP70 sequences. The assay showed different patterns for 11 species; L. infantum, L. donovani, L. tropica, L. aethiopica, L. major, L. lainsoni, L. naiffi, L. braziliensis, L. peruviana, L. guyanensis, and L. panamensis by applying two subsequent digests. However, only three species (L. mexicana, L. amazonensis, and L. garnhami) failed to generate species-specific patterns [86] . Two years later, they improved the sensitivity and specificity of the technique using alternative PCR primers and RFLPs, and succeeded in testing it on 114 globally representative Leishmania strains and on various other parasites [87] . Later, another team from Brazil analyzed the Leishmania hsp70 gene to select primers ranging from 234 to 384 bp for use as restriction enzyme for the PCR-RFLP protocol. The investigators tested the ability of these primers to distinguish between Leishmania spp. (70 strains, including reference strains and strains circulating in Brazil) in comparison with other PCR-based assays for CL molecular diagnosis using a panel of clinical samples from several endemic areas in Brazil. The results showed that the 234-bp region was promising, given its utility for detecting and identifying Leishmania spp. in clinical samples. Their ITS1 PCR-RFLP yielded similar results to those of 234-bp HSP70 PCR-RFLP. The investigators concluded that the newly developed assay could detect Leishmania spp. in clinical samples and discriminate between all of the species circulating in Brazil, thus rendering this assay particularly useful in endemic areas [88] .

On the basis of multilocus sequence typing, a recent method for differentiation of Leishmania spp., Chinese investigators used five enzyme-encoding genes and two conserved genes; one of them is HSP70. Results demonstrated that Chinese Leishmania isolates were of two types: clade A1 and clade A2. The first was most closely related to L. donovani strains from India and Africa, whereas the second was found to be related to the European strains of L. infantum. The investigators reported that the study provided a basis for further understanding of the phylogeny and evolutionary history of the Chinese Leishmania spp. [89] . From Belgium, scientists presented a completely validated and globally applicable standardized protocol for the use of HSP70 sequences in Leishmania typing without parasite culture. The method is especially suited for use in clinics in nonendemic regions, which have relatively few cases/year. The described method was developed in the framework of a European consortium of tropical infectious disease clinics called 'LeishMan' (http://www.leishman.eu) to characterize Leishmania spp. using a standardized PCR assay. The investigators recommended extension of their strategy to identify additional strains with associated clinical information, and to establish a global database of L. parasites [90] . Finally, Thailand investigators used PCR amplification and sequence analysis of the L. siamensis hsp70 gene to identify species of infected Sergentomyia (vectors of autochthonous VL). The latter was considered an emerging disease in Thailand caused by L. siamensis resulting in VL and disseminated dermal leishmaniasis. The investigators found that the captured Sergentomyia gemmea was a potential vector of L. siamensis in Thailand [91] .

5. Drug therapy and resistance: It was reported that L. infantum HSP70 is encoded by two genes. Investigators studied the effects of either absence or mutation of one of these genes on promastigotes. Major alteration in the growth of promastigotes and frequent aberrant forms were observed. Extending the study by re-expressing HSP70-II in mutant promastigotes restored growth rate, recovery of normal morphology, and increased macrophage interactions. It was suggested that the role played by HSP70-II expression in Leishmania virulence recommends the use of this gene as a new drug target in leishmaniasis [92] . More recently, two small acidic proteins (Lbp23A and Lbp23B) were identified as cochaperones for LbHSP90. Both recombinant proteins shared similar structures and interacted with LbHSP90 and inhibited its ATPase activity with variable efficiencies, but Lbp23A was more stable than Lbp23B [93] . For HSPs application in drug resistance, the investigators analyzed eight gene expressions in ten resistant and four sensitive clinical isolates of L. donovani using quantitative real-time PCR in an attempt to distinguish antimony-resistant and antimony-sensitive strains. They found significant expression of four of the evaluated genes; HSP83 was included in antimony-resistant strains [94] .

6. Vaccination: HSP70 was considered one of several potential immunogenic antigens. L. major HSP70 showed potential importance in stimulating humoral immune responses [95] , and HSP70 from L. donovani was identified through proteomic analysis inducing Th1 type immune response in cured and/or endemic VL [96],[97],[98] . The efficacy of a DNA vaccine with purified P4 nuclease protein along with the adjuvants HSP70 or interleukin (IL)-12 in immunizing BALB/c mice against L. amazonensis and L. major was evaluated. The results showed that P4/IL-12 produced complete protection against infection with L. amazonensis but not against L. major, whereas P4/HSP70 produced complete protection against L. major and partial protection against L. amazonensis. The investigators concluded that both P4 and HSP70 are vaccine candidates that could potentially be useful in DNA-based vaccines for both L. amazonensis and L. major [99] . The immunogenicity of three L. infantum antigens; HSP70, paraflagellar rod protein-2, and kinetoplastida membrane protein-11 (KMP-11), was evaluated in experimentally infected dogs. The investigators found that these recombinant antigens induced increased production of interferon γ (IFN-γ) and tumor necrosis factor α (TNF-α), and they concluded their importance as vaccine candidates for control of canine VL [100] .

Sachdeva and his colleagues [101] presented the first study in which the L. donovani hsp70 gene was used as an adjuvant to a polytope DNA vaccine. The investigators evaluated three vaccine formulations in a mouse model; gp63 DNA vaccine (encoding the region of gp63 gene of L. donovani), polytope DNA vaccine and polytope DNA vaccine fused to L. donovani HSP70 as genetic adjuvant to improve T cell responses as it has both cytokine and chaperone functions. The latter vaccine resulted in significant reduction in the number of amastigotes in the liver and spleen after four weeks of challenge infection. The investigators assumed that the HSP70-fused polytope antigen was internalized by receptor-mediated endocytosis, thus constituting the antigenic HSP-associated peptides [via their cell surface major histocompatibility complex (MHC) class I] presented to CD8 + T cells, consequently improving antigen presenting function. They concluded its efficiency as a preventive strategy for VL as it enhanced the cytolytic activity of splenocytes isolated from vaccinated BALB/c mice and induced strong Th1 responses [101] . Finally, in an Indian study conducted recently, investigators developed recombinant L. donovani HSP70 (rLdHSP70) to evaluate its potentiality to stimulate immune responses in lymphocytes of cured Leishmania-infected hamsters and in the peripheral blood mononuclear cells of cured VL patients. They used rLdHSP70 either individually or in combination with recombinant L. donovani proteins such as elongation factor-2, protein disulfide isomerase, and triose phosphate isomerase (TPI). Results showed that all vaccinated hamsters had enhanced proliferative response compared with control animals. In addition, they noted (1) remarkable nitric oxide (NO) production, (2) significant delayed-type hypersensitivity response, (3) significant upregulation of Th1 type of immune response (high levels of IFN-γ and IL-12), and (4) IgG and IgG1 levels that proportionally increased with very low L. donovani loads in vaccinated animals. Therefore, the investigators postulated that vaccines based on a combination of LdHSP70 with Th1-stimulatory proteins could be an essential strategies for vaccine development against VL [102] .

HSP100 was also reported as an efficient vaccine candidate. It was reported that CL produced self-limiting infection in C57/BL/6 mice and progressive and systemic infection in BALB/c mice. In their experimental studies, German investigators vaccinated both types of mice with hsp100 gene of mutated L. major wild strain. Their results encouraged the idea of generating a safe attenuated vaccine using targeted replacement of single genes required for parasite pathogenicity like the hsp100 gene [103] . On the other hand, combined use of HSP70 and HSP83 as vaccine administered alone or with one of two adjuvants (MPLA and ALD) in L. donovani-infected BALB/c mice was evaluated. The used measures for evaluation included hepatic parasite load and IgG2a, IFN-γ, and IL-2 levels. The three vaccines gave efficient results, but use of adjuvants (especially MPLA) significantly raised the IgG2a, IFN-γ, and IL-2 levels [104] .


  3. Trypanosoma species Top


3.1. African trypanosomiasis

3.1.1. Applications


1. Parasite survival: T. brucei HSP60 (63.7 kDa) was found to be expressed by both bloodstream (BS) and procyclic (PC) forms. It has 51.3 and 94.5% amino acid identity to the mammalian homologs P1 protein and T. cruzi HSP60, respectively [105] . Furthermore, Mit HSP60 and HSP70 demonstrated elevated expression in T. brucei stumpy forms, and they were both markers of morphological transformation from stumpy to PC forms [106] .

2. Immune response elicitation: T. congolense 69 kDa protein is conserved among all developmental stages of African trypanosomes. It was analyzed and proved to have 45-65% identity with HSP70s, and it was responsible for the major immunogenicity in trypanosomes that cause African sleeping sickness (ASS) [107] . In 2000, a study conducted in Belgium compared between the effects of anti-T. brucei HSP60, anti-invariant surface gp70, and anti-variant surface glycoprotein (VSG) in humoral immune responses during ASS in BALB/c mice. The results revealed that HSP60 was specifically recognized during the entire course of infection. Immunoglobulins G2a and G2b (induced mainly in a T-cell-independent manner) were detected during the first peak of parasitemia, whereas G1 and G3 (due to a specific T-cell-mediated response) were observed at the end of the infection. The investigators were convinced that HSP60 causes a significant host humoral immune response of an autoimmune character. They concluded that, during the course of the disease, several switch factors and B-cell activation pathways can determine the humoral immune response, which might mean that these responses are the outcome of mixed antiparasite and antihost immune reactions [108] . The same investigators conducted another study using T. brucei-purified flagellar antigens to generate polyclonal antisera in rats that recognized several proteins other than VSG. Screening of the T. brucei proteome form cDNA library, ~67% of the selected specific cDNA inserts encoded trypanosome HSP60 [109] .

3. Pathogenesis: The role of the hsp70.1 gene in ASS caused by T. congolense was investigated, and the parameters assessed were survival time, levels of parasitemia, anemia, and course of infection. Results revealed that susceptible A/J mice developed only moderate anemia, whereas HSP70.1-deficient C57BL/6 J and resistant wild-type C57BL/6 J mice controlled parasitemia, but developed severe anemia in the late stage of infection [110] .

4. Diagnosis: In 2002, investigators from Kenya reported that T. congolense 69 kDa protein was homologous to mammalian immunoglobulin-binding protein (BiP) and used the latter in an indirect ELISA technique for diagnosis of bovine ASS. Results showed limited sensitivity, and further evaluation was recommended [111] . Later, anti-BiP mAbs were used against sera from T. congolense experimentally infected cattle. The HSP70/BiP-based inhibition ELISA assay showed good sensitivity, with an improved sensitivity in secondary infections [112] . In addition, aiming to find out a diagnostic marker for ASS caused by T. brucei, German scientists selected several proteins including HSP70, histone H2B, histone H3, phosphoglycerate kinase (PGKC), rhodesain, and others for analysis against patient sera. They found that the majority of sera reacted with HPS70, but recommended its use as a component of a multiplex diagnostic assay containing several immunogenic proteins to ASS [113] . On the other hand, the T. congolense Mit hsp70 gene (MTP) was cloned and Japanese investigators determined its entire nucleic acid sequence. They found that the specific antigenic epitope was located in its C-terminal region. No significant differences were detected in transcription and translation levels between BS and PC forms, which indicated that it is similarly expressed in both forms and hence could be used as a diagnostic marker. They examined the MTP antigenic epitope against specific anti-MTP antibody production in hosts with primary infection, and detected MTP-specific antibodies produced in sera from infected mice. This detection was stable irrespective of the route of infection; intraperitoneal or subcutaneous. The investigators recommended the use of specific anti-MTP antibodies against sera from naturally infected animals [114] .

5. Drug therapy: The main function of HSP90, also known as HSP83 or HSP86 because of the variable molecular weight among different orthologs, is to limit protein aggregation during thermal stress [115] . The HSP90 inhibitor (17AAG) was used as a drug target against T. evansi, which causes surra in domestic animals. Targeting HSP90 for the treatment of a variety of human and animal protozoan infections was suggested [44] . Later, HSP90 inhibitors [GA, and GA analogs (17AAG and 17DMAG), radicicol and novobiocin] were investigated as drug targets. The results revealed that 17AAG and 17DMAG were the most effective against T. brucei, as the former caused severe morphological abnormalities and the latter cured infected mice [116] . Another recent study conducted by collaborative teams from the USA, UK, and Canada published a detailed structural and chemical characterization of T. brucei HSP83 as a potential drug target against ASS. The investigators could identify a benzamide derivative compound capable of interacting with TbHSP83 with inhibition of parasitic growth in vitro [117] .

Besides the previous HSPs applications in ASS, T. brucei possesses a large HSP40 complement that consists mostly of type III HSP40. In their work, Louw et al. [118] expressed and purified a novel HSP40 that potentially functions as a novel cochaperone of HSP70 in T. brucei.

3.2. American trypanosomiasis

3.2.1. Introduction

It was reported that different protective functions attributed to HSPs in heart injury patients included the repair of ion channels, the restoration of redox balance, interaction with NO-induced protection, inhibition of proinflammatory cytokines, and prevention of activation of the apoptosis pathway [119] . This was supported by the detection of a significant number of HSPs (40 spots for 13 different HSPs) in the myocardium of chronic Chagas cardiomyopathy [120] . Brazilian investigators identified 12 spots (HSP70 isoforms) and claimed that HSP70 increased in response to ischemic stress. In addition, there was increase in myocardial HSP60 production, which was previously reported to be associated with the development of chronic heart failure [121] .

3.3. Applications

1. Parasite survival: Enhanced synthesis of HSP60, HSP70, and HSP90 was detected in T. cruzi-infected mice exposed to increased temperature (39°C) using antibody immunoprecipitation analysis. The investigators also analyzed the histopathological changes in the cardiac and skeletal muscles of infected mice maintained at room temperature or at 36°C. It was found that transfer of mice from room temperature to 36°C led to decreased numbers of circulating parasites with consequently increased survival rate and reduction in parasitemia [122] .

2. Autoimmunity and pathogenesis: T. cruzi HSP70 was first recognized in 1990, when it was investigated as a specific target of humoral autoimmunity with a role in the pathogenesis of Chagas disease. The results strongly argued against its role in the pathogenesis of Chagas disease [123] . In 1993, Spanish scientists investigated its immunological response in infected sera compared with controls. There was reactivity in both groups but it was significantly higher in patient sera. However, mapping of T. cruzi HSP70 epitopes could identify the particular immunogenic patterns of Chagas disease pathogenesis [124] .

3. Diagnosis: T. cruzi ER HSP7O was found to encode a 78-kDa glucose-regulated protein (grp78). It was cloned and molecularly characterized to identify the immunogenicity of T. cruzi antigens in Chagas disease. The investigators analyzed other members of the HSP70 family in their studies to clarify the importance of HSP70 in the parasite's biology as well as host immune response. They found that grp78 is the most immunogenic protein as it reacted strongly with sera from infected mice [125] . Similar results were obtained in a study conducted in Brazil, when the best sensitivity and specificity were obtained using grp78 (ER T. cruzi HSP70) in comparison with Cy HSP70 and Mit HSP70. In addition, the investigators used grp78 with T. cruzi flagellar calcium-BiP, and the diagnostic sensitivity increased from 90 to 97% but increased leishmanial reactivity only from 3 to 8%. The investigators also concluded that none of the three HSP70 antigens was effective in distinguishing cured and uncured treated patients, which may be considered indicative of effective treatment [126] . Later, the IgG subclass antibody responses against three different T. cruzi antigens; crude antigen, KMP-11 gene, and HSP70 as well as against T. rangeli HSP70 were analyzed. The results showed the ability of T. cruzi KMP-11 and T. rangeli HSP70 to act as a marker for Chagas disease (healthy or infected, acute or chronic) [127] .

4. Species differentiation: In a study conducted recently in Cuba, investigators succeeded in using the hsp70 gene to differentiate between T. cruzi and T. rangeli strains in single and mixed infections using PCR-RFLP [128] .

5. Drug targets: Recently, it was found that activation of CD8 + cytotoxic T lymphocyte (CTL) response is one of the targets for development of immunotherapy against Chagas disease. In this study, 30 peptides were selected, synthesized, and tested for human leukocyte antigen (HLA-A*02:01) binding. Four HSP70 immunodominant epitopes (210-218, 255-263, 316-324, and 345-353) were recognized, two of which (210-218 and 316-324) were also recognized by CTL of HLA-A*02:01 from Chagas patients [129] .

6. Vaccination: HSP70 could act as an adjuvant with T. cruzi KMP11 to immunize mice against Chagas disease. Its combination with the vaccine leads to MHC class I processing pathway and elicits CD8 + response [130] . Its use in combination with T. cruzi paraflagellar rod proteins (PFR2 and PFR3) produced activation of PFR2 antigen-specific CTLs manifested by increase in IL-12 and IFN-γ and decrease in the percentage of cells expressing IL-4. The investigators concluded that PFR2-HSP70 provided a protective response in T. cruzi-infected mice [131] . Moreover, truncated TcHSP70 was shown to induce human dendritic cell (DC) maturation in patients with Chagas disease. This was manifested by increase in the expression levels of CD83, CD86, IL-12, TNF-α, and IL-6 [132] . Finally, plasmids encoding PFR2-HSP70 and PFR3-HSP70 were used to immunize transgenic mice, and the immunodominant CD8 + T-cell epitopes for both proteins were recognized by CTLs from sera of patients. Results showed that some peptides had high binding affinity to the HLA-A*02:01 molecule, and T cells from chronic patients showed proinflammatory cytokine production (IFN-γ, TNF-α, and IL-6) [133] . Interestingly, the chaperone or ATPase domains of TcHSP70 was used as an adjuvant in DNA vaccine with Entamoeba histolytica surface collagen-BiP peroxiredoxin to induce effective immune response and to protect hamsters against amoebic liver abscess development [134] .

7. Apoptosis: T. cruzi HSP65 plays an essential role in preventing macrophage apoptosis and contributes to host resistance against Chagas disease. Its expression mechanism was investigated, and results revealed that macrophages of resistant C57BL/6 and DBA/2 infected mice showed strong HSP65 expression compared with that of susceptible BALB/c mice. In addition, CD4 + T cells were responsible for the induction of HSP65 expression in macrophages [135] .


  4. Toxoplasma gondii Top


4.1. Applications

1. Antigen presentation and autoreactivity: The results obtained from a study conducted in Japan revealed that heatshock cognate protein (HSC71) had a potential role in presenting T. gondii antigen to CD4 + CTL. The investigators infected P36 cell culture with T. gondii and used specific anti-HLA-DR mAbs to show CD4 + CTL lytic activity to the T. gondii-infected cells. Using flow cytometric analysis, HSC71 was found to be expressed on the cell surface of infected and uninfected P36 cells [136] . On the other hand, because of the homologous sequence between TgHSP70 and mouse mHSP70 and human hHSP70, another group of Japanese investigators speculated that it is possible to induce autoreactive immune responses in T. gondii-infected hosts. They were able to demonstrate the production of anti-TgHSP70 antibody that cross-reacted with self mHSP70 and showed that B-1a cells are responsible for this autoreactivity in T. gondii-infected BALB/c (resistant strain) and B6 mice (susceptible strain) [137] .

2. Anaphylactic shock reaction: The expression of TgHSP70 was found to increase rapidly just before the death of infected host cells, indicating its role as a warning signal during acute toxoplasmosis [138] . It was also shown that TgHSP70 caused deterioration of the host defense through downregulation of NO release [139] , production of anti-HSP70 autoantibodies [140] , and activation of B cells and DCs [141],[142] . In Japan, scientists observed that the mechanism of lethal anaphylactic reaction induced by TgHSP70 injection in T. gondii-infected mice was IFN-γ dependent and through platelet-activating factor, not by an IgE-dependent pathway [143] . In 2010, Japanese investigators evaluated the use of T. gondii hsp70 gene vaccination in mice against TgHSP70-induced anaphylactic reaction in toxoplasmosis. They found that the vaccine targeted peripheral DCs, and produced prolonged survival rates in immunized infected mice. Therefore, they concluded its ability to induce protective immunity against toxoplasmosis [144] .

3. Protective immunity: It was shown that mice immunized with Toxoplasma cell homogenates acquired protective immunity only against infection with a low-virulence strain but not against a highly virulent strain [145] . One year later, the same team of investigators reported that HSP65 expressed in peritoneal exudate cells of Toxoplasma-infected mice was capable of generating HSP65 production and expression on the macrophage surface, thus developing effective immunity. However, it was found that HSP65 expression depends on strain virulence; it increased in a relatively avirulent strain and was not expressed in a more virulent strain [146] . Contradicting to these results, Australian investigators observed HSP65 expression on host macrophage surface in all strains in vivo and in vitro. On the other hand, HSP70 was only detected in virulent strains in vivo, and poorly expressed in avirulent strains. It was neither detected in vitro nor in infection of immunocompromised mice, and was poorly expressed in avirulent strains [147] . Furthermore, Japanese investigators extended their studies to demonstrate that HSP65 expression within and on macrophage surfaces markedly increased in the presence of γ δ T cells. Mice depleted of these T cells died in the early stages of infection, whereas mice depleted of α β T cells survived but infection could not be controlled in its late stages. The investigators attributed these results to the increase in the number of T cells bearing the γ δ receptor with thymic (Thy-1 + and Thy-1 - ) phenotypes in the peritoneal cavity and spleen [148] . Similar results were obtained in the same year using low-virulent and high-virulent T. gondii strains [149] . One year later, the same investigators demonstrated that γ δ T cells induced HSP65 expression through secretion of IFN-γ and TNF-α and NO production [150] . They added that extra-Thy rather than intra-Thy γ δ T cells have crucial roles in HSP65 expression in toxoplasmosis [151] . Later, another Japanese team found that natural killer T cells were responsible for suppression of protective immunity against toxoplasmosis through generation of IL4, which interfered with γ δ T cells [152] . Besides HSP65, HSP90 was first identified as an antigen with high ability to elicit IgG production during chronic infection in comparison with the acute stage [153] .

Moreover, TgHSP70 was found to be responsible for inhibition of inducible NO synthase expression and NO production. The investigators observed this effect only in mice infected with virulent strains; i.e. virulent T. gondii is protected from host immune responses by producing HSP70. They attributed their results to the genetic structures of the gene encoding HSP70 in virulent and avirulent strains, which may have important implications for the synthesis and stability of this protein [154] . Another study was conducted to evaluate the effects of toll-like receptors (TLR4, TLR2), and myeloid differentiation primary response gene 88 (myd88) in T. gondii-infected mice (wild-type) and mice deficient of these factors. The investigators observed high mortality rate and high levels of T. gondii load in different organs in myd88-deficient infected mice. Meanwhile, high levels of anti-mouse HSP70 autoantibody and anti-T. gondii HSP70 antibody production were detected in the sera of myd88-deficient mice [155] . Another group of investigators compared between T. gondii HSP70 and lipopolysaccharide (LPS) effects on spleen B and T cells and measured the proliferative responses in vitro. They found that HSP70 induced prominent B-cell proliferation in infected and uninfected mice, which was not significantly inhibited by polymyxin B (a potent inhibitor of LPS). They also found that both HSP70-induced and LPS-induced proliferative responses of spleen cells required TLR4 as a receptor, whereas MyD88 was involved only in LPS-induced proliferative response [141] .

4. Virulence and pathogenesis: German scientists characterized the bradyzoite specifically expressed gene (hsp30, formally known as bag1) believed to be responsible for tachyzoite-bradyzoite transformation (an important step in the pathogenesis of toxoplasmosis) [156] . In another study, investigators analyzed the roles of IFN-γ and TgHSP70 in bradyzoites/tachyzoites inter-conversion by measuring the mortality rates of wild type and IFN-γ knockout in T. gondii-infected mice. They found that IFN-γ is essential but not sufficient for protective immunity against toxoplasmosis in the wild type mice [138] . Moreover, it was found that the 82-kDa protein secreted by extracellular tachyzoites reacted specifically with mAb (Tg485). Screening mAb with T. gondii cDNA expression library, followed by amplification and sequencing, revealed significant homology to HSP90 of other Apicomplexan species [157] . Furthermore, another study conducted in Australia showed a relationship between T. gondii virulence and HSP70 expression [154] . A recent study from Argentina showed that the absence of palmitoylation of TgHSP20, which is synthesized as a mature protein in the Cy, causes its accumulation in the inner membrane complex and interference with host cell invasion by T. gondii [158] .

Regarding the identification and characterization of T. gondii sHSPs, five of them were first described by the Argentinian group: HSP20, HSP21, HSP28, HSP29, and HSP30. They are located in different cell compartments and share the homologous α-crystallin domain. They are expressed in both tachyzoites and bradyzoites, with the exception of HSP28 (predominant in tachyzoites) and HSP30 (only in bradyzoites). In their study, the researchers found that 1) Cy HSP21 shares the same subcellular localization as HSP30, 2) HSP28 is only mitochondrial and present in both T. gondii stages (tachyzoite and bradyzoite), 3) HSP29 is a membrane-associated protein in both stages, with its role to regulate membrane fluidity and preserve membrane integrity during thermal fluctuations, and 4) HSP20 is shown in T. gondii apical region of both stages. The investigators concluded that sHSPs localized to the apical surface or in the membrane are of particular interest in view of their potential role in host cell invasion [159] . Later, the same investigators showed that HSP20 co-localized with the motor complex at the outer surface of the inner membrane complex. This unique subcellular localization indicated its role in motility [160] .

5. Diagnosis: The potential value of serum IgG anti-TgHSP70.1 was evaluated as a diagnostic biomarker of ocular toxoplasmosis in a case-control study conducted in France. The study included three groups (laboratory-confirmed and clinically suspected cases as well as healthy blood donors). Significant IgG increased levels were detected in the first two groups compared with the third group [161] . Recombinant T. gondii antigens were evaluated for use in the diagnosis of toxoplasmosis. The investigators compared three proteins; HSP20, surface antigen (SAG1), and dense granule (GRA7) in human samples infected with toxoplasmosis using IgG ELISA. They found that the combination of two or three recombinant antigens improved the sensitivity values, whereas none of them reacted against seronegative IgG and IgM human sera. Therefore, they concluded its possible use as a diagnostic biomarker in toxoplasmosis [162] .

6. Drug target: The use of GA (HSP90 inhibitor) significantly disturbed the intracellular growth of T. gondii [157] . Using fluorescence microscopy, investigators found increased levels of HSP90 in tachyzoites Cy and in the nucleus and Cy of mature bradyzoites. These findings suggested the potential role of HSP90 in tachyzoiteͻbradyzoite transformation, which was confirmed using GA [163] .

7. Vaccination: Mice immunized with recombinant TgHSP30/bag1 enhanced protective immunity against toxoplasmosis, whereas recombinant TgHSP70 inhibited it [139] . One year later, the same Japanese investigators published another study to localize the region of HSP70 involved in inhibition of host protective immunity and deterioration of toxoplasmosis. They utilized its full length, its NH 2 and carboxy terminal regions, and they measured IgG levels and counted cyst numbers in the brain tissue. They found that mice immunized by full length or carboxy-terminal region showed deleterious effects and attributed these results to 2 reasons; 1) activation of immunosuppression by down-regulating NO release of macrophages, 2) HSP70-full length and its carboxy-terminal region may function as a source of autoimmunity [164] . Similar results were obtained as the investigators observed deteriorating effects on host immune responses on intraperitoneal injection of TgHSP70. They found increased production of IL4 and IL10 (Th2 cytokines) as well as reduced NO production from peritoneal macrophages indicating the role of T. gondii HSP70 in immunosuppression [140] .

In 2003, three genes were used to vaccinate mice against toxoplasmosis: HSP70 (virulent tachyzoite-specific gene), HSP30 (a bradyzoite-specific gene), and SAG1 (a tachyzoite-specific gene). Results revealed that the hsp70 gene was the most effective, showing the lowest parasite burden in various organs of experimental T. gondii-infected mice, and it lasted for more than 3 months [165] . In 2012, Wei [166] constructed a polyvalent vaccine using the hsp70 gene to immunize BALB/c mice. He evaluated the vaccine effects by assessment of CD4 + , CD8 + , immunoglobulins (IgG, IgM, and IgA), and cytokines (IFN-γ, TNF, IL-2, IL-4, and IL-12). Results showed strong and significant cellular and humoral immune responses in the supernatant of cultured spleen cells and in sera of immunized mice [166] . Investigators in Argentina raised antibodies against TgHSP20 in rabbits and then incubated T. gondii tachyzoites in rabbit serum. Reduced parasite invasion (57%) and tachyzoite gliding activity (49%) were elicited. In addition, the rabbit sera reduced Neospora caninum invasion of host cells [167] .

8. Apoptosis: HSP65 of infected macrophages was shown to prevent apoptosis by T. gondii, contributing to the protective immunity against T. gondii low-virulence strains. When mice were susceptible to high-virulent strains, no HSP65 expression was obtained and macrophages showed marked apoptosis. The investigators claimed that mice infected with high-virulent strains might have mechanism(s) that interfere with HSP65 expression [168] .


  5. Cryptosporidium species Top


5.1. Applications

1. Viability: In a study on C. parvum oocyst viability, German scientists used HSP70 as a marker. The induction ratio of HSP70 mRNA production in oocysts subjected to heat stress was assessed. Results showed its slight increase in viable oocysts that were heat shocked at 45°C for 20 min, indicating that HSP70 mRNA induction ratio varied according to oocyst viability [169] .

2. Host-parasite relationship: Another HSP70 application was attempted to understand host-parasite relationship in cryptosporidiosis. Results of the multi-locus genetic characterization of small subunit (SSU) rRNA, actin and HSP70 genes of 15 Cryptosporidium spp. revealed that 1) host adaptation is a general phenomenon in cryptosporidiosis, 2) specific genotypes infect specific groups of animals, 3) C. parvum bovine genotype and C. meleagridis that infected rodents and mammals, respectively, were able to infect humans [170] .

3. Diagnosis: The utility of HSP70 was first reported in the diagnosis of cryptosporidiosis when the investigators used two pairs of PCR primers for oocyst detection in drinking water supplies to avoid waterborne outbreaks in the USA. One of the used primers recognized all Cryptosporidium spp., whereas the other targeted C. parvum HSP70 [171] . In 2001, Australian investigators suggested the use of HSP70 to monitor water sources as it is a sensitive biomarker for detection of viable oocysts even at low concentration levels [172] . In 2009, a nanotechnology technique was developed to diagnose C. parvum in clinical samples using oligonucleotide-functionalized gold nanoparticles-targeted HSP70 mRNA from C. parvum oocysts. The nanoparticle probes were complementary to two adjacent target regions on the HSP70 mRNA from C. parvum and did not cross-react with other organisms. The investigators concluded that the ability of nanoparticles to identify HSP70 from C. parvum oocysts will promote research for the use of nanotechnology in the diagnosis of parasitic diseases [173] .

4. Species identification, genotyping and phylogenetic studies: The first work using HSP70 in discrimination between Cryptosporidium species/genotypes was published in 2000. The investigators used genes encoding 18S ribosomal DNA, HSP70 and acetyl coenzyme A synthetase in molecular characterization of Cryptosporidium species in HIV patients. They identified 4 genotypes; C. parvum human and cattle genotypes, C. felis and C. meleagridis [174] . In the same year, HSP70 nucleotide sequences and phylogenetic analysis of Cryptosporidium species were identified. The investigators sequenced ~1,950 bp of the hsp70 genes from 40 Cryptosporidium samples isolated from human and various animals (C. parvum, C. baileyi, C. felis, C. meleagridis, C. muris, C. serpentis, C. wrairi, and unknown species from a desert monitor). They concluded that hsp70 gene offered advantages over SSU rRNA for phylogenetic studies of Cryptosporidium species; 1) higher heterogeneity in hsp70 gene sequences to be better target for genotyping, 2) the arms of mutations (deletion and insertions) are limited in hsp70 gene sequences [175] . One year later, PCR analysis of C. parvum hsp70 gene on 10 isolates was performed and they found its ability to detect C. parvum oocysts genotypes 1 and 2 (5 samples each) [176] . In addition, human samples with clinical cryptosporidiosis from UK were subjected to genomic analysis at hsp70 gene. The results revealed that 149 samples were C. parvum either genotype 1 (~49%) or 2 (~46%), and the investigators concluded the importance of HSP70 in screening and monitoring genotype distribution of cryptosporidiosis for the epidemiological studies [177] .

Analysis of DNA sequences of nine genes belonging to HNJ-1 (a Japanese reference human strain) showed high identity to the reported genotype 2 sequence. Of these genes, hsp70, α-tubulin and β-tubulin, 18S rRNA, and Cryptosporidium oocyst wall protein (COWP) were included [178] . Gene sequence analysis of the 18S rRNA and hsp70 gene of Cryptosporidium from infected cows, sheep, and goats showed that C. parvum was the only species infecting cows. Hence, it was concluded that cows were the possible source of maintaining cryptosporidiosis in farms in Spain [179] .

Phylogenetic analyses of three loci, including hsp70, confirmed that C. macropodum, a species isolated from kangaroos, was genetically different from other known species (C. parvum, C. hominis, C. suis, and C. canis). The investigators considered it a new species, although its oocysts are morphologically identical to the other species [180] . In another study, genes encoding HSP70 and gp60 were used to genotype symptomatic Cryptosporidium samples. The Belgian investigators found C. hominis and C. parvum in 54 and 46% of their samples, respectively. They also observed that the gp60 gene sequencing revealed the predominance of IbA10G2 (a C. hominis subgenotype) and IIaA15G2R1 (a C. parvum subgenotype) in the isolated samples [181] . Using DNA sequence analysis of the 18S rRNA, actin, and hsp70 gene, the investigators could identify C. andersoni and C. ryanae as the dominant species affecting cattle in China [182] .

Several studies utilizing HSP70 in the molecular characterization of Cryptosporidium spp. were published during the last few years (2010-2014). Using 18S rRNA, actin, COWP, HSP70, and gp60, the investigators identified five isolates from cockatiel birds; three of them were new genotypes of C. meleagridis. The latter was reported as a significant cause of cryptosporidiosis in man [183] . To improve PCR methods used for molecular species identification in cryptosporidiosis, Canadian investigators used the immunomagnetic separation (IMS) technique, by which fragments of the genes encoding Cryptosporidium 18S rRNA, HSP70, and COWP were amplified to determine the optimal genes used in molecular identification. Using primers directed against HSP70 they were able to diagnose 60 and 58 cases by IMS-PCR and PCR alone, respectively, in comparison with using 18S rRNA (43 and 42 cases) and gene encoding COWP (56 and 45 cases) [184] . In China, investigators isolated a new strain from a calf and they used SSU rRNA and hsp70 gene sequences to identify it as a new isolate; C. serpentis [185] . Using the same markers, the investigators characterized a new Cryptosporidium genotype in seals [186] . Another new Cryptosporidium genotype (C. viatorum n. sp.) was identified in ten diarrheic British patients coming from India during 2007-2010. Sequences of C. viatorum isolates at HSP70, actin, and SSR RNA loci were not identical to those of C. parvum or C. hominis [187] . Moreover, using genetic identification based on the 18S ribosomal RNA and HSP70, investigators succeeded in assigning four genotypes to infected pigs in Canada; pig genotype II (61%), C. suis (36%), C. parvum (2%), and Cryptosporidium mouse genotype (1%) [188] . On the basis of sequence analyses of four genes encoding 18S rRNA, HSP70, actin, and gp60, other investigators found that two Cryptosporidium isolates belonged to C. hominis, subtype IdA21, which is a rare subtype in China [189] . Further, using genetic analysis of partial 18S rRNA, HSP70, COWP, and actin genes, Chinese investigators claimed that samples isolated from giant pandas were different from those identified from Cryptosporidium spp. and genotypes. Accordingly, they considered these isolates as a new genotype [190] . Similar data were obtained on using the same gene sequences to characterize Cryptosporidium isolates from domestic mice in China (C. tyzzeri) [191] and domestic pigs in the Czech Republic (C. scrofarum) [192] . In Central Vietnam, investigators assigned only one genotype of C. avian and two genotypes of C. suis to all infected ostriches [193] and pigs [194] , respectively. HSP70 was also used with 18S rDNA to diagnose cryptosporidiosis in seals, although immunofluorescence microscopy failed to detect Cryptosporidium oocysts in their fecal samples [195] . HSP70 was also used for phylogenetic analysis of seven samples isolated from Mongolian diarrheic patients that were identified as C. parvum bovine genotype [196] .

Apart from C. felis of cats and C. baileyi of birds, the genomic analysis of other Cryptosporidium spp. affecting man revealed that the hsp90 gene had a conserved sequence identity. On the basis of these data, investigators used it in species identification and were able to identify and differentiate between other species (C. hominis, C. parvum, C. meleagridis, C. canis, C. muris, C. suis, and C. andersoni) [197] .

5. Vaccination: C. andersoni HSP70 was cloned and its recombinant protein was identified (rCaHSP70). In their experimental studies on BALB/c and C57BL/6 mice, the investigators showed that purified rCaHSP70 proved to be highly antigenic and may be used as an immunodiagnostic marker, whereas its constructed plasmid could be used as a candidate vaccine [198] .


  6. Other protozoa Top


6.1.  Babesia More Details species

B. divergens HSP70 is a highly conserved cytoplasmic protein identified from five isolates from different hosts (human and bovine). The investigators detected HSP70 as an early antigen during acute human  Babesiosis More Details, a result that encouraged them to use it as a candidate vaccine [199] . Later, American scientists cloned and expressed B. microti HSP70 and evaluated its use in mice immunization. They found that 30% of the mice survived the challenge infection versus none of the controls [200] . Furthermore, when used in phylogenetic analysis of Babesia spp. and Theileria spp., HSP70 genes of different isolates showed nucleotide sequences with more variety than those of 18S rDNA. Accordingly, species of Babesia and Theileria were classified into 3 groups; A) Babesia infecting dogs such as B. divergens, B. odocoilei, B. bovis, B. caballi, and B. ovis, B) Theileria spp. and C) B. microti and B. rodhaini [201] .

Because expression levels of HSP70 from B. gibsoni and B. microti showed a high degree of conservation, investigators were encouraged to compare between the immunoprotective properties of recombinant BgHSP70 and BmHSP70. Mice immunized with both recombinant proteins elicited high antibody levels and significant reductions in peripheral parasitemia. Accordingly, the use of HSP70 as a candidate vaccine was recommended [202] . Besides, HSP70 was also used in the diagnosis of B. canis vogeli in dogs using multiplex quantitative real-time PCR [203] . On the other hand, using quantitative real-time reverse transcription-PCR, Japanese investigators examined the copy number of B. gibsoni HSP70 transcripts in the strains that showed resistance to diminazene aceturate (DA). Results showed significantly lower levels of HSP70 transcripts in the DA-resistant strains in comparison with the wild-type B. gibsoni. Hence, it was suggested that HSP70 transcript levels were reduced because of parasite exposure to low concentrations of DA. Further studies were recommended to explain the relation between HSP70 and the mechanism of DA resistance in babesiosis [204] . B. gibsoni HSP70 was also found to be a highly immunogenic protein expressed on the surface of exoerythrocytic stages in canine babesiosis [205] . In addition to HSP70, investigators attempted to identify potentially protective antigens from B. bovis by analyzing memory CD4 + T lymphocyte recognition of fractionated merozoite antigens. They detected a 20-kDa protein (Bbo20) with a B lymphocyte epitope that was conserved in different strains and showed 86.4% identity with another protein identified in B. bigemina (Bbg20). Blast analysis of amino acid sequences of both 20 kDa proteins showed significant similarity with HSP20 [206] .

6.2. Microsporidium species

Genes of mitochondrion-related chaperon 60 and Mit HSP70 were found to be coded in the nuclear DNA. These include chaperonin 60 (Cpn60) in E. histolytica, T. vaginalis, and G. lamblia and Mit HSP70 in T. vaginalis and three microsporidians, Nosema locustae, Vairimorpha necatrix, and Encephalitozoon cunuculi [207] . In a study conducted in Japan, Mit HSP70 was used in phylogenetic analysis, when the investigators sequenced genes encoding Mit HSP70 from G. lamblia, E. histolytica, and Encephalitozoon hellem. Their results demonstrated that these protozoa are not primitively amitochodrial, but they lost their mitochondria during their evolutionary history [208] . Although microsporidia have no identifiable mitochondria organelle, Mit HSP70 was eventually expressed from an ancient Mit organelle called mitosome. This discovery placed microsporidia in amitochondriates with E. histolytica, G. lamblia, and T. vaginalis [209] . Expression levels of microsporidial HSP70 were shown in infected insects using affinity chromatography and immunoblotting [210] . Using transmission immunoelectron microscopy, Encephalitozoon cuniculi HSP70 was found to be localized in the spore internal structures [211] .

6.3. Giardia lamblia

The first article that reported the 57 kDa antigen as Giardia HSP was published in 1992. The investigators reported that children with acute infection had specific IgG and IgA response to that antigen [212] . One year later, they published another article evaluating IgG and IgA responses to the 57 kDa antigen in Giardia-infected children with persistent diarrhea and severe malnourishment using SDS-PAGE. Results revealed the absence of IgA response to Giardia HSP antigen in chronic giardiasis, suggesting that specific IgA production was an essential factor for determining parasite clearance [213] . In 2009, Belgian investigators sequenced the genes encoding β-giardin, triose phosphate isomerase (TPI), and glutamate dehydrogenase (GDH) to genotype Giardia isolated from symptomatic patients. Assemblage B was detected in majority of samples (74.4%), however, using a novel species-specific PCR based on tpi gene, mixed infections with both assemblage A and B were detected in 32.4% [181] .HSP70 was found to be essential in excystation of Giardia cysts. In their studies to assess the importance of phosphorylated proteins during excystation, investigators found two phosphoproteins; one of them was HSP70 [214] . In an attempt to determine the infectivity and virulence of a Giardia isolate, mRNA expression was evaluated by quantitative measurements of two biomarkers; HSPs and variant-specific protein (VSP). Using reverse transcription-PCR, investigators measured HSPs in cysts and VSP in trophozoites to assess trophozoite attachment to a cell monolayer (Caco-2 cell culture). They found direct linear correlation between the used biomarkers [215] . In addition, evaluating three kits for DNA extraction, investigators found that use of combined markers (genes encoding HSP and β-giardin) gave better results in the detection and genotyping of G. lamblia isolates than using each marker alone. However, when used alone, the HSP marker gave a higher limit of detection and genotyping compared with the β-giardin marker [216] .

4. Eimeria species

HSP90 was found to be essential for proper cell functions in coccidian protozoa. Molecular analysis of HSP90 from E. tenella and E. acervulina showed high identity and similar nucleotide sequences. However, transcripts detected in all developmental stages of E. tenella disappeared in oocysts undergoing sporulation and in fully sporulated oocysts in E. acervulina. The researchers suggested that regulation systems in messenger RNA underwent different expressions in the two species [217] . Investigating the importance of HSP90 in the intracellular growth of E. tenella, investigators showed increase in EtHSP90 expression in the parasitophorous vacuole membrane, indicating its importance for parasite intracellular growth. Moreover, when specific antibodies and GA were used to attenuate EtHSP90 functions, the parasite was unable to invade and grow in the host cell [218] . Recently, the effect of diclazuril as anticoccidial therapy on HSP90 expression of E. tenella merozoites was attempted. The results showed that the drug directly reduced HSP70 in second-generation merozoites. The investigators concluded that inhibition of HSP90 production might be a promising target for new therapy of E. tenella infection [219] .

On the other hand, HSP70 expression was related to E. tenella virulent strains. Quantitative analysis of HSP70 showed significant gradual decrease in sporozoites and the expressed HSP70 in the excysted sporozoites would be involved in the parasite pathogenicity [220] . Using anti-HSP70 MAbs with immunogold labelling, Spanish investigators succeeded to localize and analyze HSP70 in E. tenella oocysts. They also performed ultrastructural study to visualize the synaptonemal complexes (SCs) to evaluate its chromosomal behavior in presence and absence of HSP70 expression. Meanwhile, inhibition of HSP70 synthesis was inhibited by 3 different doses of quercetin. The investigators found that with increasing the quercetin dose, the number of immunogold particles decreased which means reduction in the HSP70 synthesis. When the dose increased, they found defects in SCs formation followed by sporulation failure (in the last high quercetin dose). The investigators concluded the importance of HSP70 in the sporulation of E. tenella oocysts [221] . HSP70 was also used as an adjuvant in vaccinating chicken against avian coccidiosis. The investigators used the parasite antigen (Ag) in the presence and absence of Ag-loaded DCs or Ag-pulsed DC-derived exosomes (HSP70, CD80 and flotillin, MHC I and II). The results revealed that chicken immunized in the presence of Ag-loaded DCs or Ag-pulsed DC-derived exosomes showed better results in the form of reduction in 1) fecal oocyst shedding, 2) intestinal lesions, and 3) mortality rate, while their body weight increased [222] .

6.5. Trichomonas vaginalis

Using immunoprecipitation, American investigators tested the antigenicity of T. vaginalis HSPs against sera from infected patients and controls. While the majority of HSPs were immunoprecipitated by sera from infected patients as well as controls, only the 38-kDa HSP was immunoprecipitated by sera from infected females, indicating its specificity in trichomoniasis [223] . PCR-RFLP employing the heat-inducible Cy HSP70 was utilized in subtyping T. vaginalis isolates by American scientists. That probe was used as it simultaneously measures the polymorphisms present at multiple gene loci in the Cy hsp70 gene family. They found 10 distinct RFLP pattern subtypes in 36 isolates that did not correlate with metronidazole resistance [224] . Using RFLP-PCR with the same probe, Egyptian investigators reported no relation between T. vaginalis subtypes and presence of T. vaginalis virus. The investigators detected six distinct RFLP pattern subtypes in 13 symptomatic and seven asymptomatic T. vaginalis isolates [225] . On the other hand, a study conducted in 2009 revealed a high degree of diversity among T. vaginalis clinical isolates. As reported, 105 different HSP70 RFLP patterns were detected in 129 isolates; 85 isolates produced unique patterns and 20 additional patterns were shared among 44 isolates. The American investigators found identical patterns in isolates from different geographic locations, confirming the theory of clonal propagation proposed for T. vaginalis isolates. The high genetic diversity among T. vaginalis isolates was attributed to the variable clinical manifestations in trichomoniasis [226] .

6.6. Entamoeba histolytica

The first report dealing with E. histolytica HSPs published in 1992 constructed a cDNA library from E. histolytica trophozoites. Proteins from this library were screened against serum IgG from a patient with invasive amebiasis. Most of the immunopositive proteins were ~70% identical to the human HSP70. When the investigators screened them again with 12 serum samples from infected patients, only three samples expressed HSP70 [227] .

6.7. Free living amoeba

Canadian investigators reported Mit HSP60 expression on abundant organelles within Mastigamoeba balamuthi [228] . Another study confirmed that Cy HSP70 was essential in the proliferation and cytotoxicity of Naegleria fowleri [229] . In addition, HSP70 was detected in Acanthamoeba (nine strains) and in free-living freshwater amoebae (15 strains, belonging to genus Amoeba and Trichamoeba) [230] .

6.8. Cyclospora cayetanensis

A nested PCR method employing the hsp70 gene was developed to diagnose C. cayetanensis infections. This technique was tested in 16 isolates from three different endemic regions: Nepal, Mexico, and Peru. The results revealed that the hsp70 gene could be considered a useful genetic marker for rapid detection of C. cayetanensis [231] .

6.9. Blastocystis species

In Japan, determination of nucleotide sequences encoding Cy HSP70 with other factors (SSU rRNA, translation elongation factor 2 and the noncatalytic 'B' subunit of vacuolar ATPase) enabled the investigators to suggest that Blastocystis (stramenopiles) is the closest relative of alveolates [232] .

6.10. Theileria species

Sera from Theileria-infected sheep were used to screen clones of the T. lestoquardi expression library to identify a specific clone that could be used as a diagnostic marker. The investigators found a reacting clone with high identity to HSP70, which proved to be ~94% sensitive and ~90% specific in diagnosis using ELISA, especially in epidemiological surveys [233] . In addition, HSP70 was used as an adjuvant with the major T. sergenti surface protein (p33), and the results indicated that HSP70-p33 is a promising candidate vaccine against theileriosis [234] .


  Concluding remarks Top


  1. HSPs play essential roles in vector-transmitted protozoa such as Plasmodium, Leishmania, Trypanosome, and Babesia. They are responsible for their survival against temperature changes and the febrile conditions in malaria, leishmaniasis, trypanosomiasis, and toxoplasmosis. They are also responsible for their growth and virulence, as well as for their ability to elicit host immune response. Thus, their potential uses as serodiagnostic markers, drug targets, and vaccine candidates have been interesting topics for studies conducted in the last two decades.
  2. In malignant malaria, the P. falciparum proteome contains Asn-rich sequences, which lead to increased aggregate formation that is enhanced under stress conditions during febrile attacks, another threat by P. falciparum when HSPs have to counteract. Besides, HSPs have to help the organism export ~5% of its encoded genome into the host passing the plasma membrane and the parasitophorous vacuole membrane. PTEX is a translocon that helps in the trafficking across the membranes, and HSPs were identified as constitutents of these PTEXs.
  3. There is a partnership between HSP40 and HSP70 controlling the events in the life cycle of P. falciparum, and its viability is sensitive to HSP70-HSP40 inhibition. Small HSPs (HSP20) also have potential roles in parasite growth and development, as well as in virulence. Cy HSP110 showed its ability to prevent Asn repeat-rich protein aggregation in vitro and in vivo. P. falciparum HSP70 was found to be a serodiagnostic marker and vaccine candidate.
  4. HSP100 has a potential role in Leishmania parasite growth and virulence, whereas HSP20 was suggested to be a useful serodiagnostic marker for active leishmaniasis. HSP70 and HSP83 proved to be highly immunogenic and both have prospective importance in the development of improved immunotherapeutic and/or immunoprophylactic targets. Both HSPs were able to elicit human T-cell immune responses and to induce B-cell proliferation. Therefore, both were considered serodiagnostic markers in VL and TL. The gene encoding HSP70 represented a good target for species identification and genotyping, and it was suggested as a new drug target and an adjuvant with a polytope DNA vaccine in leishmaniasis. Furthermore, the use of recombinant HSP70 proved to exhibit protective immunity to all vaccinated mice.
  5. In African trypanosomiasis, HSP60 was expressed by both BS and PC forms indicating its responsibility for their survival. It also caused a significant host humoral immune response during the entire course of the disease. HSP70 was suggested as a serodiagnostic marker in ASS, whereas the Mit hsp70 gene was found to have a specific antigenic epitope located in its C-terminal region. Investigating new drug targets in ASS revealed drugs that inhibit HSP90 (e.g. GA and GA analogs) and that interact with HSP83 (e.g. benzamide derivative), leading to inhibition of parasite growth.
  6. In American trypanosomiasis, mapping of HSP70 epitopes identified the immunogenic motifs in the pathogenesis of Chagas disease. Thus, it was used as a marker for this disease (healthy or infected, acute or chronic). The gene encoding the hsp70 gene was used to differentiate between T. cruzi and T. rangeli strains in single and mixed infections. It was also used as a drug target and/or adjuvant in a plasmide vaccine candidate to immunize mice against challenge infection. Interestingly, its ATPase domains were used as adjuvants in a DNA vaccine with E. histolytica-BiP to protect hamsters against amoebic liver abscess. On the other hand, HSP65 has an essential role in preventing macrophage apoptosis contributing to host resistance against infection.
  7. In toxoplasmosis, a homologous sequence between HSP70s of T. gondii, mouse and human was detected, and the possibility of TgHSP70 to induce autoreactive immune responses in infected hosts with subsequent occurrence of lethal anaphylactic shock reaction was reported. The hsp70 gene was useful for immunizing mice against anaphylactic reaction, whereas its use in a polyvalent vaccine showed strong and significant results in vaccination. There was also a relationship between virulence and HSP70 expression, and its potential use as a diagnostic biomarker of ocular toxoplasmosis was reported.
  8. Besides, HSP65 expression by macrophages was reported to develop effective immunity, but its expression depends on strain virulence. Therefore, HSP65 was found to prevent apoptosis of infected macrophages by T. gondii low-virulence strains. HSP90 has the ability to elicit IgG production during chronic toxoplasmosis, and HSP90 inhibitors showed significant results for use as drug targets. The role of sHSPs in toxoplasmosis requires further evaluation. They are localized in the apical surface, indicating their role in motility, and they are expressed in both tachyzoites and bradyzoites. Reduction of both parasite invasion and tachyzoites gliding activity was observed on incubation of antibodies against HSP20 with T. gondii tachyzoites.
  9. In cryptosporidiosis, HSP70 was the only HSP that proved to have several applications; 1) understanding host-parasite relationship, 2) assessment of oocyst viability, 3) detection of oocysts in drinking water supplies, 4) species identification and genotyping (phylogenetic analysis), and 5) immunization of mice against challenge infection.
  10. Studies on HSP70s constituted the majority of research published on protozoa. It was essential in excystation of Giardia cysts, and in proliferation and cytotoxicity of N. fowleri. The gene encoding HSP70 was used in phylogenetic analysis of G. lamblia, T. vaginalis, and species of Babesia, Theileria, Blastocystis, and Encephalitozoon. It was also used in the diagnosis of C. cayetanensis, Theileria, and in the diagnosis of babesiosis in dogs. In addition, HSP70 proved to be useful as a vaccine candidate in babesiosis, and as an adjuvant against avian coccidiosis.



  Acknowledgements Top


Conflicts of interest

There are no conflicts of interest.

 
  References Top

1.
Aravind L, Iyer LM, Wellems TE, Miller LH. Plasmodium biology: genomic gleanings. Cell 2003; 115:771-785.  Back to cited text no. 1
    
2.
Muralidharan V, Oksman A, Pal P, Lindquist S, Goldberg DE. Plasmodium falciparum heat shock protein 110 stabilizes the asparagine repeat-rich parasite proteome during malarial fevers. Nat Commun 2012; 3:1310.  Back to cited text no. 2
    
3.
Sargeant TJ, Marti M, Caler E, et al. Lineage-specific expansion of proteins exported to erythrocytes in malaria parasites. Genome Biol 2006; 7:R12.  Back to cited text no. 3
    
4.
Maier AG, Rug M, O′Neill MT, et al. Exported proteins required for virulence and rigidity of Plasmodium falciparum-infected human erythrocytes. Cell 2008; 134:48-61.  Back to cited text no. 4
    
5.
De Koning-Ward TF, Gilson PR, Boddey JA, et al. A newly discovered protein export machine in malaria parasites. Nature 2009; 459:945-949.  Back to cited text no. 5
    
6.
Shonhai A, Boshoff A, Blatch GL. The structural and functional diversity of Hsp70 proteins from Plasmodium falciparum. Protein Sci 2007; 16:1803-1818.  Back to cited text no. 6
    
7.
Acharya P, Kumar R, Tatu U. Chaperoning a cellular upheaval in malaria: heat shock proteins in Plasmodium falciparum. Mol Biochem Parasitol 2007; 153:85-94.  Back to cited text no. 7
    
8.
Chiang AN, Valderramos JC, Balachandran R, et al. Select pyrimidinones inhibit the propagation of the malarial parasite, Plasmodium falciparum. Bioorg Med Chem 2009; 17:1527-1533.  Back to cited text no. 8
    
9.
Shonhai A. Plasmodial heat shock proteins: targets for chemotherapy. FEMS Immunol Med Microbiol 2010; 58:61-74.  Back to cited text no. 9
    
10.
Pesce ER, Cockburn IL, Goble JL, Stephens LL, Blatch GL. Malaria heat shock proteins: drug targets that chaperone other drug targets. Infect Disord Drug Targets 2010; 10:147-157.  Back to cited text no. 10
    
11.
Rug M, Maier AG. The heat shock protein 40 family of the malaria parasite Plasmodium falciparum. IUBMB Life 2011; 63:1081-1086.  Back to cited text no. 11
    
12.
Njunge JM, Ludewig MH, Boshoff A, Pesce ER, Blatch GL. Hsp70s and J proteins of Plasmodium parasites infecting rodents and primates: structure, function, clinical relevance, and drug targets. Curr Pharm Des 2013; 19:387-403.  Back to cited text no. 12
    
13.
Ramdhave AS, Patel D, Ramya I, Nandave M, Kharkar PS. Targeting heat shock protein 90 for malaria. Mini Rev Med Chem 2013; 13:1903-1920.  Back to cited text no. 13
    
14.
Pesce ER, Blatch GL. Plasmodial Hsp40 and Hsp70 chaperones: current and future perspectives. Parasitology 2014; 141:1167-1176.  Back to cited text no. 14
    
15.
Chua CS, Low H, Sim TS. Co-chaperones of Hsp90 in Plasmodium falciparum and their concerted roles in cellular regulation. Parasitology 2014; 141:1177-1191.  Back to cited text no. 15
    
16.
Elsworth B, Crabb BS, Gilson PR. Protein export in malaria parasites: an update. Cell Microbiol 2014; 16:355-363.  Back to cited text no. 16
    
17.
Pavithra SR, Kumar R, Tatu U. Systems analysis of chaperone networks in the malarial parasite Plasmodium falciparum. PLoS Comput Biol 2007; 3:1701-1715.  Back to cited text no. 17
    
18.
Daily JP, Scanfeld D, Pochet N, et al. Distinct physiological states of Plasmodium falciparum in malaria-infected patients. Nature 2007; 450:1091-1095.  Back to cited text no. 18
    
19.
Pallavi R, Acharya P, Chandran S, Daily JP, Tatu U. Chaperone expression profiles correlate with distinct physiological states of Plasmodium falciparum in malaria patients. Malar J2010; 9:236.  Back to cited text no. 19
    
20.
Ardeshir F, Flint JE, Richman SJ, Reese RT. A 75 kd merozoite surface protein of Plasmodium falciparum which is related to the 70 kd heat-shock proteins. EMBO J 1987; 6:493-499.  Back to cited text no. 20
    
21.
Kumar N, Syin CA, Carter R, Quakyi I, Miller LH. Plasmodium falciparum gene encoding a protein similar to the 78-kDa rat glucose-regulated stress protein. Proc Natl Acad Sci USA 1988; 85:6277-6281.  Back to cited text no. 21
    
22.
Biswas S, Sharma YD. Lack of correlation between red cell invasion by merozoites and anti-heat shock protein-70 antibody levels in malaria patients′ sera. Int J Parasitol 1991; 21:213-217.  Back to cited text no. 22
    
23.
Biswas S, Sharma YD. Human response to a malaria vaccine candidate antigen. Vaccine 1991; 9:467-469.  Back to cited text no. 23
    
24.
Alexandre CO, Camargo LM, Mattei D, et al. Humoral immune response to the 72 kDa heat shock protein from Plasmodium falciparum in populations at hypoendemic areas of malaria in western Brazilian Amazon. Acta Trop 1997; 64:155-166.  Back to cited text no. 24
    
25.
Sharma YD. Structure and possible function of heat-shock proteins in falciparum malaria. Comp Biochem Physiol B 1992; 102:437-444.  Back to cited text no. 25
    
26.
Kumar N, Nagasawa H, Sacci JBJr, et al. Expression of members of the heat-shock protein 70 family in the exoerythrocytic stages of Plasmodium berghei and Plasmodium falciparum. Parasitol Res 1993; 79:109-113.  Back to cited text no. 26
    
27.
Biswas S, Sharma YD. Enhanced expression of Plasmodium falciparum heat shock protein PFHSP70-I at higher temperatures and parasite survival. FEMS Microbiol Lett 1994; 124:425-429.  Back to cited text no. 27
    
28.
Tsuji M, Mattei D, Nussenzweig RS, Eichinger D, Zavala F. Demonstration of heat-shock protein 70 in the sporozoite stage of malaria parasites. Parasitol Res 1994; 80:16-21.  Back to cited text no. 28
    
29.
Kumar R, Musiyenko A, Barik S. The heat shock protein 90 of Plasmodium falciparum and antimalarial activity of its inhibitor, geldanamycin. Malar J 2003; 2:30.  Back to cited text no. 29
    
30.
Pavithra SR, Banumathy G, Joy O, Singh V, Tatu U. Recurrent fever promotes Plasmodium falciparum development in human erythrocytes. J Biol Chem 2004; 279:46692-46699.  Back to cited text no. 30
    
31.
Przyborski JM, Lanzer M. Protein transport and trafficking in Plasmodium falciparum-infected erythrocytes. Parasitology 2005; 130:373-388.  Back to cited text no. 31
    
32.
Morahan BJ, Strobel C, Hasan U, et al. Functional analysis of the exported type IV HSP40 protein PfGECO in Plasmodium falciparum gametocytes. Eukaryot Cell 2011; 10:1492-1503.  Back to cited text no. 32
    
33.
Montagna GN, Buscaglia CA, Münter S, et al. Critical role for heat shock protein 20 (HSP20) in migration of malarial sporozoites. J Biol Chem 2012; 287:2410-2422.  Back to cited text no. 33
    
34.
Tardieux I, Baines I, Mossakowska M, Ward GE. Actin-binding proteins of invasive malaria parasites and the regulation of actin polymerization by a complex of 32/34-kDa proteins associated with heat shock protein 70kDa. Mol Biochem Parasitol 1998; 93:295-308.  Back to cited text no. 34
    
35.
Singhal N, Atul, Mastan BS, Kumar KA, Sijwali PS. Genetic ablation of plasmoDJ1, a multi-activity enzyme, attenuates parasite virulence and reduces oocyst production. Biochem J 2014; 461:189-203.  Back to cited text no. 35
    
36.
Kopacz J, Kumar N. Murine gamma delta T lymphocytes elicited during Plasmodium yoelii infection respond to Plasmodium heat shock proteins. Infect Immun 1999; 67:57-63.  Back to cited text no. 36
    
37.
Zhang M, Hisaeda H, Sakai T, et al. Macrophages expressing heat-shock protein 65 play an essential role in protection of mice infected with Plasmodium yoelii. Immunology 1999; 97:611-615.  Back to cited text no. 37
    
38.
Zhang M, Hisaeda H, Sakai T, et al. CD4 + T cells are required for HSP65 expression in host macrophages and for protection of mice infected with Plasmodium yoelii. Parasitol Int 2001; 50:201-209.  Back to cited text no. 38
    
39.
Zhang M, Hisaeda H, Kano S, et al. Antibodies specific for heat shock proteins in human and murine malaria. Microbes Infect 2001; 3:363-367.  Back to cited text no. 39
    
40.
Na BK, Park JW, Lee HW, et al. Characterization of Plasmodium vivax heat shock protein 70 and evaluation of its value for serodiagnosis of tertian malaria. Clin Vaccine Immunol 2007; 14:320-322.  Back to cited text no. 40
    
41.
Polpanich D, Tangboriboonrat P, Elaissari A, Udomsangpetch R. Detection of malaria infection via latex agglutination assay. Anal Chem 2007; 79:4690-4695.  Back to cited text no. 41
    
42.
Guirgis BS, Sá e Cunha C, Gomes I, et al. Gold nanoparticle-based fluorescence immunoassay for malaria antigen detection. Anal Bioanal Chem 2012; 402:1019-1027.  Back to cited text no. 42
    
43.
Kumar R, Pavithra SR, Tatu U. Three-dimensional structure of heat shock protein 90 from Plasmodium falciparum: molecular modelling approach to rational drug design against malaria. J Biosci 2007; 32:531-536.  Back to cited text no. 43
    
44.
Pallavi R, Roy N, Nageshan RK, et al. Heat shock protein 90 as a drug target against protozoan infections: biochemical characterization of HSP90 from Plasmodium falciparum and Trypanosoma evansi and evaluation of its inhibitor as a candidate drug. J Biol Chem 2010; 285:37964-37975.  Back to cited text no. 44
    
45.
Shahinas D, Liang M, Datti A, Pillai DR. A repurposing strategy identifies novel synergistic inhibitors of Plasmodium falciparum heat shock protein 90. J Med Chem 2010; 53:3552-3557.  Back to cited text no. 45
    
46.
Shahinas D, Macmullin G, Benedict C, Crandall I, Pillai DR. Harmine is a potent antimalarial targeting Hsp90 and synergizes with chloroquine and artemisinin. Antimicrob Agents Chemother 2012; 56:4207-4213.  Back to cited text no. 46
    
47.
Beck JR, Muralidharan V, Oksman A, Goldberg DE. PTEX component HSP101 mediates export of diverse malaria effectors into host erythrocytes. Nature 2014; 511:592-595.  Back to cited text no. 47
    
48.
Elsworth B, Matthews K, Nie CQ, et al. PTEX is an essential nexus for protein export in malaria parasites. Nature 2014; 511:587-591.  Back to cited text no. 48
    
49.
Shahinas D, Folefoc A, Taldone T, Chiosis G, Crandall I, Pillai DR. A purine analog synergizes with chloroquine (CQ) by targeting Plasmodium falciparum Hsp90 (PfHsp90). PLoS One 2013; 8:e75446.  Back to cited text no. 49
    
50.
Sanchez GI, Sedegah M, Rogers WO, et al. Immunogenicity and protective efficacy of a Plasmodium yoelii Hsp60 DNA vaccine in BALB/c mice. Infect Immun 2001; 69:3897-3905.  Back to cited text no. 50
    
51.
Qazi KR, Wikman M, Vasconcelos NM, Berzins K, Ståhl S, Fernández C. Enhancement of DNA vaccine potency by linkage of Plasmodium falciparum malarial antigen gene fused with a fragment of HSP70 gene. Vaccine 2005; 23:1114-1125.  Back to cited text no. 51
    
52.
Kamali AN, Marín-García P, Azcárate IG, Diez A, Puyet A, Bautista JM. Plasmodium yoelii blood-stage antigens newly identified by immunoaffinity using purified IgG antibodies from malaria-resistant mice. Immunobiology 2012; 217:823-830.  Back to cited text no. 52
    
53.
Shapira M, McEwen JG, Jaffe CL. Temperature effects on molecular processes which lead to stage differentiation in Leishmania. EMBO J 1988; 7:2895-901.  Back to cited text no. 53
    
54.
Pinelli E, Shapira M. Temperature-induced expression of proteins in Leishmania mexicana amazonensis. A 22-kDa protein is possibly localized in the mitochondrion. Eur J Biochem 1990; 194:685-691.  Back to cited text no. 54
    
55.
Hübel A, Krobitsch S, Hörauf A, Clos J. Leishmania major Hsp100 is required chiefly in the mammalian stage of the parasite. Mol Cell Biol 1997; 17:5987-5995.  Back to cited text no. 55
    
56.
Krobitsch S, Brandau S, Hoyer C, Schmetz C, Hübel A, Clos J. Leishmania donovani heat shock protein 100: characterization and function in amastigote stage differentiation. J Biol Chem 1998; 273:6488-6494.  Back to cited text no. 56
    
57.
Reiling L, Chrobak M, Schmetz C, Clos J. Overexpression of a single Leishmania major gene enhances parasite infectivity in vivo and in vitro. Mol Microbiol 2010; 76:1175-1190.  Back to cited text no. 57
    
58.
Skeiky YA, Benson DR, Guderian JA, Whittle JA, Bacelar O, Carvalho EM, Reed SG. Immune responses of leishmaniasis patients to heat shock proteins of Leishmania species and humans. Infect Immun 1995; 63:4105-4114.  Back to cited text no. 58
    
59.
Jeronimo SM, Higgs E, Vedvick T, et al. Identification of Leishmania chagasi antigens recognized by human lymphocytes. J Infect Dis 1995; 172:1055-1060.  Back to cited text no. 59
    
60.
Rico AI, Gironès N, Fresno M, Alonso C, Requena JM. The heat shock proteins, Hsp70 and Hsp83, of Leishmania infantum are mitogens for mouse B cells. Cell Stress Chaperones 2002; 7:339-346.  Back to cited text no. 60
    
61.
Carmelo E, Zurita AI, Martinez E, Valladares B. The sera from individuals suffering from cutaneous leishmaniasis due to Leishmania brazilensis present antibodies against parasitic conserved proteins, but not their human counterparts. Parasite 2006; 13:231-236.  Back to cited text no. 61
    
62.
MacFarlane J, Blaxter ML, Bishop RP, Miles MA, Kelly JM. Identification and characterization of a Leishmania donovani antigen belonging to the 70-kDa heat-shock protein family. Eur J Biochem 1990; 190:377-384.  Back to cited text no. 62
    
63.
de Andrade CR, Kirchhoff LV, Donelson JE, Otsu K. Recombinant Leishmania Hsp90 and Hsp70 are recognized by sera from visceral leishmaniasis patients but not Chagas disease patients. J Clin Microbiol 1992; 30:330-335.  Back to cited text no. 63
    
64.
Wallace GR, Ball AE, MacFarlane J, El Safi SH, Miles MA, Kelly JM. Mapping of a visceral leishmaniasis-specific immunodominant B-cell epitope of Leishmania donovani Hsp70. Infect Immun 1992; 60:2688-2693.  Back to cited text no. 64
    
65.
Arora SK, Melby PC, Sehgal S. Lack of serological specificity of recombinant heat shock protein of Leishmania donovani. Immunol Cell Biol 1995; 73:446-451.  Back to cited text no. 65
    
66.
Quijada L, Requena JM, Soto M, et al. Mapping of the linear antigenic determinants of the Leishmania infantum hsp70 recognized by leishmaniasis sera. Immunol Lett 1996; 52:73-79.  Back to cited text no. 66
    
67.
Zurita AI, Rodríguez J, Piñero JE, et al. Cloning and characterization of the Leishmania (Viannia) braziliensis Hsp70 gene: diagnostic use of the C-terminal fragment rLb70 (513-663). J Parasitol 2003; 89:372-378.  Back to cited text no. 67
    
68.
Rasouli M, Zavaran HA, Kazemi B, Alborzi A, Kiany S. Expression of recombinant heat-shock protein 70 of MCAN/IR/96/LON-49, a tool for diagnosis and future vaccine research. Iran J Immunol 2009; 6:75-86.  Back to cited text no. 68
    
69.
Oliveira GG, Magalhães FB, Teixeira MC, et al. Characterization of novel Leishmania infantum recombinant proteins encoded by genes from five families with distinct capacities for serodiagnosis of canine and human visceral leishmaniasis. Am J Trop Med Hyg 2011; 85:1025-1034.  Back to cited text no. 69
    
70.
Kumar S, Kumar D, Chakravarty J, Rai M, Sundar S. Identification and characterization of a novel Leishmania donovani antigen for serodiagnosis of visceral leishmaniasis. Am J Trop Med Hyg 2012; 86:601-605.  Back to cited text no. 70
    
71.
Assis LM, Sousa JR, Pinto NF, et al. B-cell epitopes of antigenic proteins in Leishmania infantum: an in silico analysis. Parasite Immunol 2014; 36:313-323.  Back to cited text no. 71
    
72.
Angel SO, Requena JM, Soto M, Criado D, Alonso C. During canine leishmaniasis a protein belonging to the 83-kDa heat-shock protein family elicits a strong humoral response. Acta Trop 1996; 62:45-56.  Back to cited text no. 72
    
73.
Skeiky YA, Benson DR, Costa JL, Badaro R, Reed SG. Association of Leishmania heat shock protein 83 antigen and immunoglobulin G4 antibody titers in Brazilian patients with diffuse cutaneous leishmaniasis. Infect Immun 1997; 65:5368-5370.  Back to cited text no. 73
    
74.
Celeste BJ, Arroyo Sanchez MC, Ramos-Sanchez EM, Castro LG, Lima Costa FA, Goto H. Recombinant Leishmania infantum heat shock protein 83 for the serodiagnosis of cutaneous, mucosal, and visceral leishmaniases. Am J Trop Med Hyg 2014; 90:860-865.  Back to cited text no. 74
    
75.
Kaur J, Kaur S. ELISA and western blotting for the detection of Hsp70 and Hsp83 antigens of Leishmania donovani. J Parasit Dis 2013; 37:68-73.  Back to cited text no. 75
    
76.
Rey-Ladino JA, Joshi PB, Singh B, Gupta R, Reiner NE. Leishmania major: molecular cloning, sequencing, and expression of the heat shock protein 60 gene reveals unique carboxy terminal peptide sequences. Exp Parasitol 1997; 85:249-263.  Back to cited text no. 76
    
77.
Celeste BJ, Angel SO, Castro LG, Gidlund M, Goto H. Leishmania infantum heat shock protein 83 for the serodiagnosis of tegumentary leishmaniasis. Braz J Med Biol Res 2004; 37:1591-1593.  Back to cited text no. 77
    
78.
Garcia AL, Parrado R, De Doncker S, Bermudez H, Dujardin JC. American tegumentary leishmaniasis: direct species identification of Leishmania in non-invasive clinical samples. Trans R Soc Trop Med Hyg 2007; 101:368-371.  Back to cited text no. 78
    
79.
Souza AP, Soto M, Costa JM, et al. Towards a more precise serological diagnosis of human tegumentary leishmaniasis using Leishmania recombinant proteins. PLoS One 2013; 8:e66110.  Back to cited text no. 79
    
80.
Menezes-Souza D, Mendes TA, Gomes Mde S, et al. Epitope mapping of the HSP83.1 protein of Leishmania braziliensis discloses novel targets for immunodiagnosis of tegumentary and visceral clinical forms of leishmaniasis. Clin Vaccine Immunol 2014; 21:949-959.  Back to cited text no. 80
    
81.
Montalvo-Alvarez AM, Folgueira C, Carrión J, Monzote-Fidalgo L, Cañavate C, Requena JM. The Leishmania HSP20 is antigenic during natural infections, but, as DNA vaccine, it does not protect BALB/c mice against experimental L. amazonensis infection. J Biomed Biotechnol 2008; 2008:695432.  Back to cited text no. 81
    
82.
Garcia L, Kindt A, Bermudez H, et al. Culture-independent species typing of neotropical Leishmania for clinical validation of a PCR-based assay targeting heat shock protein 70 genes. J Clin Microbiol 2004; 42:2294-2297.  Back to cited text no. 82
    
83.
Arora SK, Gupta S, Bhardwaj S, Sachdeva N, Sharma NL. An epitope-specific PCR test for diagnosis of Leishmania donovani infections. Trans R Soc Trop Med Hyg 2008; 102: 41-45.  Back to cited text no. 83
    
84.
Da Silva LA, de Sousa CS, da Graça GC, Porrozzi R, Cupolillo E. Sequence analysis and PCR-RFLP profiling of the hsp70 gene as a valuable tool for identifying Leishmania species associated with human leishmaniasis in Brazil. Infect Genet Evol 2010; 10:77-83.  Back to cited text no. 84
    
85.
Montalvo-Alvarez AM, Nodarse JF, Goodridge IM, et al. Differentiation of Leishmania (Viannia) panamensis and Leishmania (V.) guyanensis using BccI for hsp70 PCR-RFLP. Trans R Soc Trop Med Hyg 2010; 104:364-367.  Back to cited text no. 85
    
86.
Montalvo-Alvarez AM, Fraga J, Monzote L, et al. Heat-shock protein 70 PCR-RFLP: a universal simple tool for Leishmania species discrimination in the new and old world. Parasitology 2010; 137:1159-1168.  Back to cited text no. 86
    
87.
Montalvo-Alvarez AM, Fraga J, Maes I, Dujardin JC, Van der Auwera G. Three new sensitive and specific heat-shock protein 70 PCRs for global Leishmania species identification. Eur J Clin Microbiol Infect Dis 2012; 31:1453-1461.  Back to cited text no. 87
    
88.
Graça GC, Volpini AC, Romero GA, et al. Development and validation of PCR-based assays for diagnosis of American cutaneous leishmaniasis and identification of the parasite species. Mem Inst Oswaldo Cruz 2012; 107:664-674.  Back to cited text no. 88
    
89.
Zhang CY, Lu XJ, Du XQ, Jian J, Shu L, Ma Y. Phylogenetic and evolutionary analysis of Chinese Leishmania isolates based on multilocus sequence typing. PLoS One 2013; 8:e63124.  Back to cited text no. 89
    
90.
Van der Auwera G, Maes I, De Doncker S, et al. Heat-shock protein 70 gene sequencing for Leishmania species typing in European tropical infectious disease clinics. Euro Surveill 2013; 18:20543.  Back to cited text no. 90
    
91.
Kanjanopas K, Siripattanapipong S, Ninsaeng U, et al. Sergentomyia (Neophlebotomus) gemmea, a potential vector of Leishmania siamensis in southern Thailand. BMC Infect Dis 2013; 13:333.  Back to cited text no. 91
    
92.
Folgueira C, Carrión J, Moreno J, Saugar JM, Cañavate C, Requena JM. Effects of the disruption of the hsp70-II gene on the growth, morphology, and virulence of Leishmania infantum promastigotes. Int Microbiol 2008; 1:81-89.  Back to cited text no. 92
    
93.
Batista FA, Almeida GS, Seraphim TV, et al. Identification of two p23 co-chaperone isoforms in Leishmania braziliensis exhibiting similar structures and Hsp90 interaction properties despite divergent stabilities. FEBS J 2015; 282:388-406.  Back to cited text no. 93
    
94.
Kumar D, Singh R, Bhandari V, Kulshrestha A, Negi NS, Salotra P. Biomarkers of antimony resistance: need for expression analysis of multiple genes to distinguish resistance phenotype in clinical isolates of Leishmania donovani. Parasitol Res 2012; 111:223-230.  Back to cited text no. 94
    
95.
Rafati S, Gholami E, Hassani N, et al. Leishmania major heat shock protein 70 (HSP70) is not protective in murine models of cutaneous leishmaniasis and stimulates strong humoral responses in cutaneous and visceral leishmaniasis patients. Vaccine 2007; 25:4159-4169.  Back to cited text no. 95
    
96.
Gupta SK, Sisodia BS, Sinha S, et al. Proteomic approach for identification and characterization of novel immunostimulatory proteins from soluble antigens of Leishmania donovani promastigotes. Proteomics 2007; 7:816-823.  Back to cited text no. 96
    
97.
Kumari S, Samant M, Khare P, et al. Induction of Th1-type cellular responses in cured/exposed Leishmania-infected patients and hamsters against polyproteins of soluble Leishmania donovani promastigotes ranging from 89.9 to 97.1 kDa. Vaccine 2008; 26:4813-4818.  Back to cited text no. 97
    
98.
Kumari S, Samant M, Misra P, et al. Th1-stimulatory polyproteins of soluble Leishmania donovani promastigotes ranging from 89.9 to 97.1 kDa offers long-lasting protection against experimental visceral leishmaniasis. Vaccine 2008; 26:5700-5711,  Back to cited text no. 98
    
99.
Campbell K, Diao H, Ji J, Soong L DNA immunization with the gene encoding P4 nuclease of Leishmania amazonensis protects mice against cutaneous leishmaniasis. Infect Immun 2003; 71:6270-6278.  Back to cited text no. 99
    
100.
Carrillo E, Crusat M, Nieto J, et al. Immunogenicity of HSP-70, KMP-11 and PFR-2 leishmanial antigens in the experimental model of canine visceral leishmaniasis. Vaccine 2008; 26:1902-1911.  Back to cited text no. 100
    
101.
Sachdeva R, Banerjea AC, Malla N, Dubey ML. Immunogenicity and efficacy of single antigen Gp63, polytope and polytopeHSP70 DNA vaccines against visceral leishmaniasis in experimental mouse model. PLoS One 2009; 4:e7880.  Back to cited text no. 101
    
102.
Jaiswa AK, Khare P, Joshi S, Kushawaha PK, Sundar S, Dube A. Th1 stimulatory proteins of Leishmania donovani: comparative cellular and protective responses of rTriose phosphate isomerase, rProtein disulfide isomerase and rElongation factor-2 in combination with rHSP70 against visceral leishmaniasis. PLoS One 2014; 9:e108556.  Back to cited text no. 102
    
103.
Reiling L, Jacobs T, Kroemer M, Gaworski I, Graefe S, Clos J. Spontaneous recovery of pathogenicity by Leishmania major hsp100-/- alters the immune response in mice. Infect Immun 2006; 74:6027-6036.  Back to cited text no. 103
    
104.
Kaur J, Kaur T, Kaur S. Studies on the protective efficacy and immunogenicity of Hsp70 and Hsp83 based vaccine formulations in Leishmania donovani infected BALB/c mice. Acta Trop 2011; 119:50-56.  Back to cited text no. 104
    
105.
Bringaud F, Peyruchaud S, Baltz D, Giroud CH, Simpson L, Baltz T. Molecular characterization of the mitochondrial heat shock protein 60 gene from Trypanosoma brucei. Mol Biochem Parasitol 1995; 74:119-123.  Back to cited text no. 105
    
106.
Tyler KM, Matthews KR, Gull K. The bloodstream differentiation-division of Trypanosoma brucei studied using mitochondrial markers. Proc Biol Sci 1997; 264:1481-1490.  Back to cited text no. 106
    
107.
Boulangé A, Authié E. A 69 kDa immunodominant antigen of Trypanosoma (Nannomonas) congolense is homologous to immunoglobulin heavy chain binding protein (BiP). Parasitology 1994; 109:163-173.  Back to cited text no. 107
    
108.
Radwanska M, Magez S, Michel A, Stijlemans B, Geuskens M, Pays E. Comparative analysis of antibody responses against HSP60, invariant surface glycoprotein 70, and variant surface glycoprotein reveals a complex antigen-specific pattern of immunoglobulin isotype switching during infection by Trypanosoma brucei. Infect Immun 2000; 68:848-860.  Back to cited text no. 108
    
109.
Radwanska M, Magez S, Dumont N, Pays A, Nolan D, Pays E. Antibodies raised against the flagellar pocket fraction of Trypanosoma brucei preferentially recognize HSP60 in cDNA expression library. Parasite Immunol 2000; 22:639-650.  Back to cited text no. 109
    
110.
Nakamura Y, Naessens J, Takata M, et al. Susceptibility of heat shock protein 70.1-deficient C57BL/6 J, wild-type C57BL/6 J and A/J mice to Trypanosoma congolense infection. Parasitol Res 2003; 90:171-174.  Back to cited text no. 110
    
111.
Boulangé A, Katende J, Authié E. Trypanosoma congolense: expression of a heat shock protein 70 and initial evaluation as a diagnostic antigen for bovine trypanosomosis. Exp Parasitol 2002; 100:6-11.  Back to cited text no. 111
    
112.
Bossard G, Boulange A, Holzmuller P, Thévenon S, Patrel D, Authie E. Serodiagnosis of bovine trypanosomosis based on HSP70/BiP inhibition ELISA. Vet Parasitol 2010; 173:39-47.  Back to cited text no. 112
    
113.
Manful T, Mulindwa J, Frank FM, Clayton CE, Matovu E. A search for Trypanosoma brucei rhodesiense diagnostic antigens by proteomic screening and targeted cloning. PLoS One 2010; 5:e9630.  Back to cited text no. 113
    
114.
Bannai H, Sakurai T, Inoue N, Sugimoto C, Igarashi I. Cloning and expression of mitochondrial heat shock protein 70 of Trypanosoma congolense and potential use as a diagnostic antigen. Clin Diagn Lab Immunol 2003; 10:926-933.  Back to cited text no. 114
    
115.
Hartl FU, Bracher A, Hayer-Hartl M. Molecular chaperones in protein folding and proteostasis. Nature 2011; 475:324-332.  Back to cited text no. 115
    
116.
Meyer KJ, Shapiro TA. Potent anti-trypanosomal activities of heat shock protein 90 inhibitors in vitro and in vivo. J Infect Dis 2013; 208:489-499.  Back to cited text no. 116
    
117.
Pizarro JC, Hills T, Senisterra G, et al. Exploring the Trypanosoma brucei Hsp83 potential as a target for structure guided drug design. PLoS Negl Trop Dis 2013; 7:e2492.  Back to cited text no. 117
    
118.
Louw CA, Ludewig MH, Blatch GL. Overproduction, purification and characterisation of Tbj1, a novel type III Hsp40 from Trypanosoma brucei, the African sleeping sickness parasite. Protein Expr Purif 2010; 69:168-177.  Back to cited text no. 118
    
119.
Delogu G, Signore M, Mechelli A, Famularo G. Heat shock proteins and their role in heart injury. Curr Opin Crit Care 2002; 8:411-416.  Back to cited text no. 119
    
120.
Teixeira PC, Iwai LK, Kuramoto AC, Honorato R, Fiorelli A, Stolf N, et al. Proteomic inventory of myocardial proteins from patients with chronic Chagas cardiomyopathy. Braz J Med Biol Res 2006; 39:1549-1562.  Back to cited text no. 120
    
121.
Tanonaka K, Yoshida H, Toga W, Furuhama K, Takeo S. Myocardial heat shock proteins during the development of heart failure. Biochem Biophys Res Commun 2001; 283:520-525.  Back to cited text no. 121
    
122.
Arif AA, Gao L, Davis CD, Helm DS. Antibody response to heat shock proteins and histopathology in mice infected with Trypanosoma cruzi and maintained at elevated temperature. J Parasitol 1999; 85:1089-1099.  Back to cited text no. 122
    
123.
Engman DM, Dragon EA, Donelson JE. Human humoral immunity to hsp70 during Trypanosoma cruzi infection. J Immunol 1990; 144:3987-3991.  Back to cited text no. 123
    
124.
Requena JM, Soto M, Guzman F, et al. Mapping of antigenic determinants of the T. cruzi hsp70 in chagasic and healthy individuals. Mol Immunol 1993; 30:1115-1121.  Back to cited text no. 124
    
125.
Tibbetts RS, Kim IY, Olson CL, et al. Molecular cloning and characterization of the 78-kilodalton glucose-regulated protein of Trypanosoma cruzi. Infect Immun 1994; 62:2499-2507.  Back to cited text no. 125
    
126.
Krautz GM, Peterson JD, Godsel LM, Krettli AU, Engman DM. Human antibody responses to Trypanosoma cruzi 70-kD heat-shock proteins. Am J Trop Med Hyg 1998; 58:137-143.  Back to cited text no. 126
    
127.
Flechas ID, Cuellar A, Cucunubá ZM, et al. Characterizing the KMP-11 and HSP-70 recombinant antigens′ humoral immune response profile in chagasic patients. BMC Infect Dis 2009; 9:186.  Back to cited text no. 127
    
128.
Fraga J, Fernandez-Calienes A, Montalvo AM, Maes I, Dujardin JC, Van der Auwera G. Differentiation between Trypanosoma cruzi and Trypanosoma rangeli using heat-shock protein 70 polymorphisms. Trop Med Int Health 2014; 19:195-206.  Back to cited text no. 128
    
129.
Marañón C, Egui A, Carrilero B, et al. Identification of HLA-AFNx0102:01-restricted CTL epitopes in Trypanosoma cruzi heat shock protein-70 recognized by Chagas disease patients. Microbes Infect 2011; 13:1025-1032.  Back to cited text no. 129
    
130.
Planelles L, Thomas MC, Alonso C, López MC. DNA immunization with Trypanosoma cruzi HSP70 fused to the KMP11 protein elicits a cytotoxic and humoral immune response against the antigen and leads to protection. Infect Immun 2001; 69:6558-6563.  Back to cited text no. 130
    
131.
Morell M, Thomas MC, Caballero T, Alonso C, López MC. The genetic immunization with paraflagellar rod protein-2 fused to the HSP70 confers protection against late Trypanosoma cruzi infection. Vaccine 2006; 24:7046-7055.  Back to cited text no. 131
    
132.
Cuellar A, Santander SP, Thomas Mdel C, et al. Monocyte-derived dendritic cells from chagasic patients vs healthy donors secrete differential levels of IL-10 and IL-12 when stimulated with a protein fragment of Trypanosoma cruzi heat-shock protein-70. Immunol Cell Biol 2008; 86:255-260.  Back to cited text no. 132
    
133.
Egui A, Thomas MC, Morell M, et al. Trypanosoma cruzi paraflagellar rod proteins 2 and 3 contain immunodominant CD8(+) T-cell epitopes that are recognized by cytotoxic T cells from Chagas disease patients. Mol Immunol 2012; 52:289-298.  Back to cited text no. 133
    
134.
González-Vázquez MC, Carabarin-Lima A, Baylón-Pacheco L, Talamás-Rohana P, Rosales-Encina JL. Obtaining of three recombinant antigens of Entamoeba histolytica and evaluation of their immunogenic ability without adjuvant in a hamster model of immunoprotection. Acta Trop 2012; 122:169-176.  Back to cited text no. 134
    
135.
Sakai T, Hisaeda H, Ishikawa H, et al. Expression and role of heat-shock protein 65 (HSP65) in macrophages during Trypanosoma cruzi infection: involvement of HSP65 in prevention of apoptosis of macrophages. Microbes Infect 1999; 1:419-427.  Back to cited text no. 135
    
136.
Yang TH, Aosai F, Norose K, Mun HS, Yano A. Heat shock cognate protein 71-associated peptides function as an epitope for Toxoplasma gondii-specific CD4 + CTL. Microbiol Immunol 1997; 41:553-561.  Back to cited text no. 136
    
137.
Chen M, Aosai F, Mun HS, Norose K, Hata H, Yano A. Anti-HSP70 autoantibody formation by B-1 cells in Toxoplasma gondii-infected mice. Infect Immun 2000; 68:4893-4899.  Back to cited text no. 137
    
138.
Mun HS, Aosai F, Norose K, et al. Toxoplasma gondii Hsp70 as a danger signal in Toxoplasma gondii-infected mice. Cell Stress Chaperones 2000; 5:328-335.  Back to cited text no. 138
    
139.
Mun HS, Aosai F, Yano A. Role of Toxoplasma gondii HSP70 and Toxoplasma gondii HSP30/bag1 in antibody formation and prophylactic immunity in mice experimentally infected with Toxoplasma gondii. Microbiol Immunol 1999; 43:471-479.  Back to cited text no. 139
    
140.
Ahmed AK, Mun HS, Aosai F, et al. Roles of Toxoplasma gondii-derived heat shock protein 70 in host defense against T. gondii infection. Microbiol Immunol 2004; 48:911-915.  Back to cited text no. 140
    
141.
Aosai F, Chen M, Kang HK, et al. Toxoplasma gondii-derived heat shock protein HSP70 functions as a B cell mitogen. Cell Stress Chaperones 2002; 7:357.  Back to cited text no. 141
    
142.
Aosai F, Rodriguez Pena MS, Mun HS, et al. Toxoplasma gondii-derived heat shock protein 70 stimulates maturation of murine bone marrow-derived dendritic cells via toll-like receptor 4. Cell Stress Chaperones 2006; 11:13.  Back to cited text no. 142
    
143.
Fang H, Aosai F, Mun HS, et al. Anaphylactic reaction induced by Toxoplasma gondii-derived heat shock protein 70. Int Immunol 2006; 18:1487-1497.  Back to cited text no. 143
    
144.
Kikumura A, Fang H, Mun HS, et al. Protective immunity against lethal anaphylactic reaction in Toxoplasma gondii-infected mice by DNA vaccination with T. gondii-derived heat shock protein 70 gene. Parasitol Int 2010; 59:105-111.  Back to cited text no. 144
    
145.
Nagasawa H, Manabe T, Maekawa Y, Oka M, Himeno K. Role of L3T4 + and Lyt-2 + T cell subsets in protective immune responses of mice against infection with a low or high virulent strain of Toxoplasma gondii. Microbiol Immunol 1991; 35:215-222.  Back to cited text no. 145
    
146.
Nagasawa H, Oka M, Maeda K, et al. Induction of heat shock protein closely correlates with protection against Toxoplasma gondii infection. Proc Natl Acad Sci USA 1992; 89:3155-3158.  Back to cited text no. 146
    
147.
Lyons RE, Johnson AM. Heat shock proteins of Toxoplasma gondii. Parasite Immunol 1995; 17:353-359.  Back to cited text no. 147
    
148.
Hisaeda H, Nagasawa H, Maeda K, et al. Gamma delta T cells play an important role in hsp65 expression and in acquiring protective immune responses against infection with Toxoplasma gondii. J Immunol 1995; 155:244-251.  Back to cited text no. 148
    
149.
Himeno K, Hisaeda H. Contribution of 65-kDa heat shock protein induced by gamma and delta T cells to protection against Toxoplasma gondii infection. Immunol Res 1996; 15:258-264.  Back to cited text no. 149
    
150.
Hisaeda H, Sakai T, Maekawa Y, Ishikawa H, Yasutomo K, Himeno K. Mechanisms of HSP65 expression induced by gamma delta T cells in murine Toxoplasma gondii infection. Pathobiology 1996; 64:198-203.  Back to cited text no. 150
    
151.
Hisaeda H, Sakai T, Nagasawa H, et al. Contribution of extrathymic gamma delta T cells to the expression of heat-shock protein and to protective immunity in mice infected with Toxoplasma gondii. Immunology 1996; 88:551-557.  Back to cited text no. 151
    
152.
Nakano Y, Hisaeda H, Sakai T, et al. Roles of NKT cells in resistance against infection with Toxoplasma gondii and in expression of heat shock protein 65 in the host macrophages. Microbes Infect 2002; 4:1-11.  Back to cited text no. 152
    
153.
Rojas PA, Martin V, Nigro M, et al. Expression of a cDNA encoding a Toxoplasma gondii protein belonging to the heat-shock 90 family and analysis of its antigenicity. FEMS Microbiol Lett 2000; 190:209-213.  Back to cited text no. 153
    
154.
Dobbin CA, Smith NC, Johnson AM. Heat shock protein 70 is a potential virulence factor in murine Toxoplasma infection via immunomodulation of host NF-kappa B and nitric oxide. J Immunol 2002; 169:958-965.  Back to cited text no. 154
    
155.
Chen M, Aosai F, Norose K, et al. Involvement of MyD88 in host defense and the down-regulation of anti-heat shock protein 70 autoantibody formation by MyD88 in Toxoplasma gondii-infected mice. J Parasitol 2002; 88:1017-1019.  Back to cited text no. 155
    
156.
Bohne W, Gross U, Ferguson DJ, Heesemann J. Cloning and characterization of a bradyzoite-specifically expressed gene (hsp30/bag1) of Toxoplasma gondii, related to genes encoding small heat-shock proteins of plants. Mol Microbiol 1995; 16:1221-1230.  Back to cited text no. 156
    
157.
Ahn HJ, Kim S, Nam HW. Molecular cloning of the 82-kDa heat shock protein (HSP90) of Toxoplasma gondii associated with the entry into and growth in host cells. Biochem Biophys Res Commun 2003; 311:654-659.  Back to cited text no. 157
    
158.
De Napoli MG, de Miguel N, Lebrun M, Moreno SN, Angel SO, Corvi MM. N-terminal palmitoylation is required for Toxoplasma gondii HSP20 inner membrane complex localization. Biochim Biophys Acta 2013; 1833:1329-1337.  Back to cited text no. 158
    
159.
de Miguel N, Echeverria PC, Angel SO. Differential subcellular localization of members of the Toxoplasma gondii small heat shock protein family. Eukaryot Cell 2005; 4:1990-1997.  Back to cited text no. 159
    
160.
de Miguel N, Lebrun M, Heaslip A, et al. Toxoplasma gondii Hsp20 is a stripe-arranged chaperone-like protein associated with the outer leaflet of the inner membrane complex. Biol Cell 2008; 100:479-489.  Back to cited text no. 160
    
161.
Chumpitazi BF, Bouillet L, Fricker-Hidalgo H, et al. Contribution of anti-Hsp70.1 IgG antibody levels to the diagnostic certainty of clinically suspected ocular toxoplasmosis. Invest Ophthalmol Vis Sci 2010; 51:5530-5536.  Back to cited text no. 161
    
162.
Cóceres VM, Becher ML, De Napoli MG, Corvi MM, Clemente M, Angel SO. Evaluation of the antigenic value of recombinant Toxoplasma gondii HSP20 to detect specific immunoglobulin G antibodies in Toxoplasma infected humans. Exp Parasitol 2010; 126:263-266.  Back to cited text no. 162
    
163.
Echeverria PC, Matrajt M, Harb OS, et al. Toxoplasma gondii Hsp90 is a potential drug target whose expression and subcellular localization are developmentally regulated. J Mol Biol 2005; 350:723-734.  Back to cited text no. 163
    
164.
Mun HS, Norose K, Aosai F, Chen M, Yano A. A role of carboxy-terminal region of Toxoplasma gondii-heat shock protein 70 in enhancement of T. gondii infection in mice. Korean J Parasitol 2000; 38:107-110.  Back to cited text no. 164
    
165.
Mohamed RM, Aosai F, Chen M, et al. Induction of protective immunity by DNA vaccination with Toxoplasma gondii HSP70, HSP30 and SAG1 genes. Vaccine 2003; 21:2852-2861.  Back to cited text no. 165
    
166.
Wei QK. Study on construction and immune protective effect of recombinant nucleic acid vaccine of Toxoplasma gondii. Zhongguo Xue Xi Chong Bing Fang Zhi Za Zhi 2012; 24:173-177.  Back to cited text no. 166
    
167.
Cóceres VM, Alonso AM, Alomar ML, Corvi MM. Rabbit antibodies against Toxoplasma Hsp20 are able to reduce parasite invasion and gliding motility in Toxoplasma gondii and parasite invasion in Neospora caninum. Exp Parasitol 2012; 132:274-281.  Back to cited text no. 167
    
168.
Hisaeda H, Sakai T, Ishikawa H, et al. Heat shock protein 65 induced by gammadelta T cells prevents apoptosis of macrophages and contributes to host defense in mice infected with Toxoplasma gondii. J Immunol 1997; 159:2375-2381.  Back to cited text no. 168
    
169.
Garcés-Sanchez G, Wilderer PA, Horn H, Munch JC, Lebuhn M. Assessment of the viability of Cryptosporidium parvum oocysts with the induction ratio of hsp70 mRNA production in manure. J Microbiol Methods 2013; 94:280-289.  Back to cited text no. 169
    
170.
Xiao L, Sulaiman IM, Ryan UM, et al. Host adaptation and host-parasite co-evolution in Cryptosporidium: implications for taxonomy and public health. Int J Parasitol 2002; 32:1773-1785.  Back to cited text no. 170
    
171.
Rochelle PA, Ferguson DM, Handojo TJ, De Leon R, Stewart MH, Wolfe RL. An assay combining cell culture with reverse transcriptase PCR to detect and determine the infectivity of waterborne Cryptosporidium parvum. Appl Environ Microbiol 1997; 63:2029-2037.  Back to cited text no. 171
    
172.
Gobet P, Toze S. Relevance of Cryptosporidium parvum hsp70 mRNA amplification as a tool to discriminate between viable and dead oocysts. J Parasitol 2001; 87:226-269.  Back to cited text no. 172
    
173.
Javier DJ, Castellanos-Gonzalez A, Weigum SE, White ACJr, Richards-Kortum R. Oligonucleotide-gold nanoparticle networks for detection of Cryptosporidium parvum heat shock protein 70 mRNA. J Clin Microbiol 2009; 47:4060-4066.  Back to cited text no. 173
    
174.
Morgan U, Weber R, Xiao L, et al. Molecular characterization of Cryptosporidium isolates obtained from human immunodeficiency virus-infected individuals living in Switzerland, Kenya, and the United States. J Clin Microbiol 2000; 38:1180-1183.  Back to cited text no. 174
    
175.
Sulaiman IM, Morgan UM, Thompson RC, Lal AA, Xiao L. Phylogenetic relationships of Cryptosporidium parasites based on the 70-kilodalton heat shock protein (HSP70) gene. Appl Environ Microbiol 2000; 66:2385-2391.  Back to cited text no. 175
    
176.
Gobet P, Toze S. Sensitive genotyping of Cryptosporidium parvum by PCR-RFLP analysis of the 70-kilodalton heat shock protein (HSP70) gene. FEMS Microbiol Lett 2001; 200:37-41.  Back to cited text no. 176
    
177.
El-Osta YG, Chalmers RM, Gasser RB. Survey of Cryptosporidium parvum genotypes in humans from the UK by mutation scanning analysis of a heat shock protein gene region. Mol Cell Probes 2003; 17:127-134.  Back to cited text no. 177
    
178.
Satoh M, Kimata I, Iseki M, Nakai Y. Gene analysis of Cryptosporidium parvum HNJ-1 strain isolated in Japan. Parasitol Res 2005; 97:452-457.  Back to cited text no. 178
    
179.
Castro-Hermida JA, Almeida A, González-Warleta M, Correia da Costa JM, Rumbo-Lorenzo C, Mezo M. Occurrence of Cryptosporidium parvum and Giardia duodenalis in healthy adult domestic ruminants. Parasitol Res 2007; 101:1443-1448.  Back to cited text no. 179
    
180.
Power ML, Ryan UM. A new species of Cryptosporidium (Apicomplexa: Cryptosporidiidae) from eastern grey kangaroos (Macropus giganteus). J Parasitol 2008; 94:1114-1117.  Back to cited text no. 180
    
181.
Geurden T, Levecke B, Cacció SM, et al. Multilocus genotyping of Cryptosporidium and Giardia in non-outbreak related cases of diarrhea in human patients in Belgium. Parasitology 2009; 136:1161-1168.  Back to cited text no. 181
    
182.
Liu A, Wang R, Li Y, et al. Prevalence and distribution of Cryptosporidium spp. in dairy cattle in Heilongjiang Province, China. Parasitol Res 2009; 105:797-802.  Back to cited text no. 182
    
183.
Abe N, Makino I. Multilocus genotypic analysis of Cryptosporidium isolates from cockatiels, Japan. Parasitol Res 2010; 106:1491-1497.  Back to cited text no. 183
    
184.
Coklin T, Farber JM, Parrington LJ, Bin Kingombe CI, Ross WH, Dixon BR. Immunomagnetic separation significantly improves the sensitivity of polymerase chain reaction in detecting Giardia duodenalis and Cryptosporidium spp. in dairy cattle. J Vet Diagn Invest 2011; 23:260-267.  Back to cited text no. 184
    
185.
Chen F, Qiu H. Identification and characterization of a Chinese isolate of Cryptosporidium serpentis from dairy cattle. Parasitol Res 2012; 111:1785-1791.  Back to cited text no. 185
    
186.
Bass AL, Wallace CC, Yund PO, Ford TE. Detection of Cryptosporidium sp. in two new seal species, Phoca vitulina and Cystophora cristata, and a novel Cryptosporidium genotype in a third seal species, Pagophilus groenlandicus, from the Gulf of Maine. J Parasitol 2012; 98:316-322.  Back to cited text no. 186
    
187.
Elwin K, Hadfield SJ, Robinson G, Crouch ND, Chalmers RM. Cryptosporidium viatorum n. sp. (Apicomplexa: Cryptosporidiidae) among travellers returning to Great Britain from the Indian subcontinent, 2007-2011. Int J Parasitol 2012; 42:675-682.  Back to cited text no. 187
    
188.
Budu-Amoako E, Greenwood SJ, Dixon BR, et al. Occurrence of Giardia and Cryptosporidium in pigs on Prince Edward Island, Canada. Vet Parasitol 2012; 184:18-24.  Back to cited text no. 188
    
189.
Zhu H, Zhao J, Wang R, Zhang L. Molecular identification of a rare subtype of Cryptosporidium hominis in infants in China. PLoS One 2012; 7:e43682.  Back to cited text no. 189
    
190.
Liu X, He T, Zhong Z, et al. A new genotype of Cryptosporidium from giant panda (Ailuropoda melanoleuca) in China. Parasitol Int 2013; 62:454-458.  Back to cited text no. 190
    
191.
Ren X, Zhao J, Zhang L, et al. Cryptosporidium tyzzeri n. sp. (Apicomplexa: Cryptosporidiidae) in domestic mice (Mus musculus). Exp Parasitol 2012; 130:274-281.  Back to cited text no. 191
    
192.
Kváč M, Kestřánová M, Pinková M, et al. Cryptosporidium scrofarum n. sp. (Apicomplexa: Cryptosporidiidae) in domestic pigs (Sus scrofa). Vet Parasitol 2013; 191:218-227.  Back to cited text no. 192
    
193.
Nguyen ST, Fukuda Y, Tada C, Huynh VV, Nguyen DT, Nakai Y. Prevalence and molecular characterization of Cryptosporidium in ostriches (Struthio camelus) on a farm in central Vietnam. Exp Parasitol 2013; 133:8-11.  Back to cited text no. 193
    
194.
Nguyen ST, Fukuda Y, Tada C, et al. Molecular characterization of Cryptosporidium in pigs in central Vietnam. Parasitol Res 2013; 112:187-192.  Back to cited text no. 194
    
195.
Rengifo-Herrera C, Ortega-Mora LM, Gómez-Bautista M, García-Peña FJ, García-Párraga D, Pedraza-Díaz S. Detection of a novel genotype of Cryptosporidium in Antarctic pinnipeds. Vet Parasitol 2013; 191:112-118.  Back to cited text no. 195
    
196.
Hong SH, Anu D, Jeong YI, Abmed D, Cho SH, Lee WJ, Lee SE. Molecular characterization of Giardia duodenalis and Cryptosporidium parvum in fecal samples of individuals in Mongolia. Am J Trop Med Hyg 2014; 90:43-47.  Back to cited text no. 196
    
197.
Feng Y, Dearen T, Cama V, Xiao L. 90-Kilodalton heat shock protein, Hsp90, as a target for genotyping Cryptosporidium spp. known to infect humans. Eukaryot Cell 2009; 8:478-482.  Back to cited text no. 197
    
198.
Liu HP, Cao JP, Li XH, et al. Cloning, expression and analysis of the heat shock protein of Cryptosporidium andersoni. Zhongguo Ji Sheng Chong Xue Yu Ji Sheng Chong Bing Za Zhi 2007; 25:163-170.  Back to cited text no. 198
    
199.
Carcy B, Précigout E, Valentin A, Gorenflot A, Reese RT, Schrével J. Heat shock response of Babesia divergens and identification of the hsp70 as an immunodominant early antigen during ox, gerbil and human babesiosis. Biol Cell 1991; 72:93-102.  Back to cited text no. 199
    
200.
Erol E, Kumar N, Carson CA. Immunogenicity of recombinant Babesia microti hsp70 homologue in mice. Int J Parasitol 1999; 29:263-266.  Back to cited text no. 200
    
201.
Yamasaki M, Inokuma H, Sugimoto C, et al. Comparison and phylogenetic analysis of the heat shock protein 70 gene of Babesia parasites from dogs. Vet Parasitol 2007; 145:217-227.  Back to cited text no. 201
    
202.
Terkawi MA, Aboge G, Jia H, et al. Molecular and immunological characterization of Babesia gibsoni and Babesia microti heat shock protein-70. Parasite Immunol 2009; 31:328-340.  Back to cited text no. 202
    
203.
Peleg O, Baneth G, Eyal O, Inbar J, Harrus S. Multiplex real-time qPCR for the detection of Ehrlichia canis and Babesia canis vogeli. Vet Parasitol 2010; 173:292-299.  Back to cited text no. 203
    
204.
Hwang SJ, Yamasaki M, Nakamura K, et al. Reduced transcript levels of the heat shock protein 70 gene in diminazene aceturate-resistant Babesia gibsoni variants under low concentrations of diminazene aceturate. Jpn J Vet Res 2010; 58:155-164.  Back to cited text no. 204
    
205.
Yamasaki M, Ishida M, Nakamura K, et al. Detection of anti-Babesia gibsoni heat shock protein 70 antibody and anti-canine heat shock protein 70 antibody in sera from Babesia gibsoni-infected dogs. Vet Parasitol 2011; 180:215-225.  Back to cited text no. 205
    
206.
Brown WC, Ruef BJ, Norimine J, 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. Mol Biochem Parasitol 2001; 118:97-109.  Back to cited text no. 206
    
207.
Peyretaillade E, Broussolle V, Peyret P, Méténier G, Gouy M, Vivarès CP. Microsporidia, amitochondrial protists, possess a 70-kDa heat shock protein gene of mitochondrial evolutionary origin. Mol Biol Evol 1998; 15:683-689.  Back to cited text no. 207
    
208.
Arisue N, Sánchez LB, Weiss LM, Müller M, Hashimoto T. Mitochondrial-type hsp70 genes of the amitochondriate protists, Giardia intestinalis, Entamoeba histolytica and two microsporidians. Parasitol Int 2002; 51:9-16.   Back to cited text no. 208
    
209.
Burri L, Williams BAP, Bursac D, Lithgow T, Keeling PJ. Microsporidian mitosomes retain elements of the general mitochondrial targeting system. Proc Natl Acad Sci USA 2006; 103:15915-15920.  Back to cited text no. 209
    
210.
Seleznev KV, Kinev AV, Margulis BA. Accumulation of heat-shock proteins 70 in insect fat bodies as evidence of stress in Microsporidia-infected insects. Tsitologiia 2003; 45:826-831.  Back to cited text no. 210
    
211.
Jolly CE, Leonard CA, Hayman, JR. Expression and localization of an Hsp70 protein in the microsporidian Encephalitozoon cuniculi. Int J Microbiol 2010; 2010:523654.  Back to cited text no. 211
    
212.
Char S, Cevallos AM, Farthing MJG. An immunodominant antigen of Giardia lamblia is a heat shock protein. Biotechnol Ther 1992; 3:151-157.  Back to cited text no. 212
    
213.
Char S, Cevallos AM, Yamson P, Sullivan PB, Neale G, Farthing MJ. Impaired IgA response to Giardia heat shock antigen in children with persistent diarrhoea and giardiasis. Gut 1993; 34:38-40.  Back to cited text no. 213
    
214.
Alvarado ME, Wasserman M. Analysis of phosphorylated proteins and inhibition of kinase activity during Giardia intestinalis excystation. Parasitol Int 2010; 59:54-61.  Back to cited text no. 214
    
215.
Alum A, Sbai B, Asaad H, Rubino JR, Khalid Ijaz M. ECC-RT-PCR: a new method to determine the viability and infectivity of Giardia cysts. Int J Infect Dis 2012; 16:e350-353.  Back to cited text no. 215
    
216.
Uda-Shimoda CF, Colli CM, Pavanelli MF, Falavigna-Guilherme AL, Gomes ML. Simplified protocol for DNA extraction and amplification of 2 molecular markers to detect and type Giardia duodenalis. Diagn Microbiol Infect Dis 2014; 78:53-58.  Back to cited text no. 216
    
217.
Miska KB, Fetterer RH, Min W, Lillehoj HS. Heat shock protein 90 genes of two species of poultry Eimeria: expression and evolutionary analysis. J Parasitol 2005; 91:300-306.  Back to cited text no. 217
    
218.
Péroval M, Péry P, Labbé M. The heat shock protein 90 of Eimeria tenella is essential for invasion of host cell and schizont growth. Int J Parasitol 2006; 36:1205-1215.  Back to cited text no. 218
    
219.
Shen X, Wang C, Zhu Q, et al. Effect of the diclazuril on Hsp90 in the second-generation merozoites of Eimeria tenella. Vet Parasitol 2012; 185:290-295.  Back to cited text no. 219
    
220.
Del Cacho E, Gallego M, López-Bernad F, Quílez J, Sánchez-Acedo C. Differences in Hsp70 expression in the sporozoites of the original strain and precocious lines of Eimeria tenella. J Parasitol 2005; 91:1127-1131.  Back to cited text no. 220
    
221.
Del Cacho E, Gallego M, Pages M, Monteagudo L, Sánchez-Acedo C. HSP70 is part of the synaptonemal complex in Eimeria tenella. Parasitol Int 2008; 57:454-459.  Back to cited text no. 221
    
222.
Del Cacho E, Gallego M, Lee SH, et al. Induction of protective immunity against Eimeria tenella infection using antigen-loaded dendritic cells (DC) and DC-derived exosomes. Vaccine 2011; 29:3818-3825.  Back to cited text no. 222
    
223.
Davis-Hayman SR, Shah PH, Finley RW, Meade JC, Lushbaugh WB. Antigenicity of Trichomonas vaginalis heat-shock proteins in human infections. Parasitol Res 2000; 86:115-120.  Back to cited text no. 223
    
224.
Stiles JK, Shah PH, Xue L, et al. Molecular typing of Trichomonas vaginalis isolates by HSP70 restriction fragment length polymorphism. Am J Trop Med Hyg 2000; 62:441-445.  Back to cited text no. 224
    
225.
Hussien EM, El-Sayed HZ, El-Moamly AA, Helmy MM, Shaban MM. Molecular characterization of Egyptian Trichomonas vaginalis clinical isolates by HSP70 restriction fragment length polymorphism. J Egypt Soc Parasitol 2005; 35:699-710.  Back to cited text no. 225
    
226.
Meade JC, de Mestral J, Stiles JK, et al. Genetic diversity of Trichomonas vaginalis clinical isolates determined by EcoRI restriction fragment length polymorphism of heat-shock protein 70 genes. Am J Trop Med Hyg 2009; 80:245-251.  Back to cited text no. 226
    
227.
Ortner S, Plaimauer B, Binder M, Wiedermann G, Scheiner O, Duchêne M. Humoral immune response against a 70-kilodalton heat shock protein of Entamoeba histolytica in a group of patients with invasive amoebiasis. Mol Biochem Parasitol 1992; 54:175-183.  Back to cited text no. 227
    
228.
Gill EE, Diaz-Triviño S, Barberà MJ, et al. Novel mitochondrion-related organelles in the anaerobic amoeba Mastigamoeba balamuthi. Mol Microbiol 2007; 66:1306-1320.  Back to cited text no. 228
    
229.
Song KJ, Song KH, Kim JH, et al. Heat shock protein 70 of Naegleria fowleri is important factor for proliferation and in vitro cytotoxicity. Parasitol Res 2008; 103:313-317.  Back to cited text no. 229
    
230.
Podlipaeva IuI, Gudkov AV. Heat shock proteins of 70 kDa family in the cells of free living and amphizoic amoeboid organisms. Tsitologiia 2009; 51:1019-1024.  Back to cited text no. 230
    
231.
Sulaiman IM, Torres P, Simpson S, Kerdahi K, Ortega Y. Sequence characterization of heat shock protein gene of Cyclospora cayetanensis isolates from Nepal, Mexico, and Peru. J Parasitol 2013; 99:379-382.  Back to cited text no. 231
    
232.
Arisue N, Hashimoto T, Yoshikawa H, et al. Phylogenetic position of Blastocystis hominis and of stramenopiles inferred from multiple molecular sequence data. J Eukaryot Microbiol 2002; 49:42-53.  Back to cited text no. 232
    
233.
Miranda J, Bakheit MA, Liu Z, et al. Development of a recombinant indirect ELISA for the diagnosis of Theileria sp. (China) infection in small ruminants. Parasitol Res 2006; 98:561-567.  Back to cited text no. 233
    
234.
Jeong W, Kweon CH, Kang SW, et al. Adjuvant effect of bovine heat shock protein 70 on piroplasm surface protein, p33, of Theileria sergenti. Biologicals 2009; 37:282-287.  Back to cited text no. 234
    



This article has been cited by
1 Immunological characterization of rLdTCP1? for its prophylactic potential against visceral leishmaniasis in hamster model
Apeksha Anand, Deep Chandra Balodi, Karthik Ramalingam, Shailendra Yadav, Neena Goyal
Molecular Immunology. 2022; 141: 33
[Pubmed] | [DOI]
2 Partners in Mischief: Functional Networks of Heat Shock Proteins of Plasmodium falciparum and Their Influence on Parasite Virulence
Michael O. Daniyan,Jude M. Przyborski,Addmore Shonhai
Biomolecules. 2019; 9(7): 295
[Pubmed] | [DOI]
3 Comparative studies of the low-resolution structure of two p23 co-chaperones for Hsp90 identified in Plasmodium falciparum genome
Noeli S.M. Silva,Thiago V. Seraphim,Karine Minari,Leandro R.S. Barbosa,Júlio C. Borges
International Journal of Biological Macromolecules. 2018; 108: 193
[Pubmed] | [DOI]



 

Top
 
 
  Search
 
Similar in PUBMED
   Search Pubmed for
   Search in Google Scholar for
 Related articles
Access Statistics
Email Alert *
Add to My List *
* Registration required (free)

 
  In this article
Abstract
Concluding remarks
Acknowledgements
1. Pla...
2. Leishmania...
3. Trypanosom...
4. Toxoplasma...
5. Cryptospor...
6. Other protozoa
References

 Article Access Statistics
    Viewed4363    
    Printed292    
    Emailed0    
    PDF Downloaded398    
    Comments [Add]    
    Cited by others 3    

Recommend this journal


[TAG2]
[TAG3]
[TAG4]