Open access peer-reviewed chapter

Between Armour and Weapons — Cell Death Mechanisms in Trypanosomatid Parasites

Written By

Rubem Figueiredo Sadok Menna-Barreto and Solange Lisboa de Castro

Submitted: 24 November 2014 Reviewed: 10 July 2015 Published: 16 December 2015

DOI: 10.5772/61196

From the Edited Volume

Cell Death - Autophagy, Apoptosis and Necrosis

Edited by Tobias M. Ntuli

Chapter metrics overview

2,095 Chapter Downloads

View Full Metrics

Abstract

Among the pathogenic protozoa, trypanosomatids stand out due to their medical and economic impact, especially for low-income populations in tropical countries. Together, sleeping sickness, Chagas disease and leishmaniasis affect millions of humans and animals worldwide, yet are neglected by the pharmaceutical industry. The current drugs for trypanosomatid infections are limited and unsatisfactory, with severe side effects leading to reduced quality of life and, in several instances, to the abandonment of treatment. An intense search for alternative compounds has been performed, aiming at specific parasite targets by cellular, molecular and biochemical approaches. One interesting strategy could be interference with the protozoan cell death pathways. However, these pathways are poorly understood in unicellular eukaryotes, with the controversial existence and uncertain biological relevance of programmed cell death (PCD). This chapter will discuss apoptosis-like and autophagic cell death and necrosis in Trypanosoma brucei, Trypanosoma cruzi and Leishmania sp. and the possible implications of these pathways for the parasite life cycle and infection persistence. It will also revisit the genomic and proteomic metadata of these trypanosomatids in the literature to rebuild the map of cell death proteins expressed under different conditions. The interaction of leading candidates with parasite-specific molecules, especially with enzymes that regulate key steps in the cell death process, is a rational and attractive alternative for drug development for these neglected diseases.

Keywords

  • Cell death
  • apoptosis-like
  • autophagy
  • necrosis
  • Leishmania sp
  • T. cruzi
  • T. brucei

1. Introduction

Neglected tropical diseases (NTDs) are a group of the seventeen mostly life-threatening infections, which affect more than a billion people worldwide. They affect poor populations, often in underdeveloped and developing countries (low-income countries) [1]. Among NTDs, infections caused by the so-called “protozoan” parasites, such as African trypanosomiasis, Chagas disease and leishmaniasis, are responsible for a high annual death toll among the poor populations of tropical countries. New safe and affordable medicines are urgently needed. These diseases all present therapeutic difficulties by developing resistance to existing therapies and/or by toxic side effects.

1.1. Neglected tropical diseases and trypanosomatids

1.1.1. Sleeping sickness

Human African trypanosomiasis (HAT), or sleeping sickness, is caused by extracellular protozoa belonging to the genus Trypanosoma and the species T. brucei. Two subspecies of T. brucei cause diseases with different epidemiological and clinical patterns: T. b. gambiense, a chronic disease present in western and central Africa accounting for 98% of the cases, and T. b. rhodesiense, an acute zoonosis located in eastern and southern Africa that occasionally infects humans. In 2001, WHO launched a major initiative to reinforce disease control and surveillance. After 10 years, the number of new cases of HAT decreased by 73.4%. Presently, the estimated the number of cases is 30, 000, and 70 million people are at risk [2, 3]. HAT clinically evolves in two stages. In the first stage, parasites are found in the lymphatic system and bloodstream. After a variable period of time, which is much shorter for the rhodesiense form, the second stage begins, with the parasites penetrating the blood-brain barrier and invading the central nervous system, leading to progressive neurological damage [4]. HAT is usually fatal if left untreated. Rhodesiense HAT usually progresses to death within six months, while gambiense HAT has a more chronic progressive course with an average duration of almost three years [5].

T. brucei is transmitted by the tsetse fly Glossina spp when it takes a blood meal. Non-dividing metacyclic forms enter the bloodstream of the mammalian host and differentiate into a rapidly dividing slender form able to evade antibody responses through antigenic variation [6]. Most of these forms undergo cell cycle arrest and develop into short-stumpy forms. When the tsetse fly bites an infected host, only the short-stumpy parasites survive in the insect’s midgut and develop into a procyclic form, which undergoes multiple developmental phases on its way to the salivary gland, finally culminating in the infective metacyclic form [7, 8].

The drug of choice for treatment depends on the infecting species and the stage of infection. In early stages, T. b. gambiense and T. b. rhodesiense infections can be treated with pentamidine and suramin, respectively [9]. If the disease has progressed, treatment relies on melarsoprol or eflornithine. Melarsoprol, an arsenical drug, is extremely toxic. Eflornithine is less toxic, but is expensive, and has a difficult administration than melarsoprol and lacks efficacy against T. b. rhodesiense [2]. Since 2001, this drug has been combined with nifurtimox (NECT) for first-line treatment for CNS-stage T. b. gambiense HAT. It is the most recent breakthrough in anti-trypanosomiasis drug research and was added to the World Health Organisation’s list of essential medicines in 2009. A major problem related to the treatment of HAT is the development of resistance to melarsoprol and the other drugs [10].

1.1.2. Chagas disease

Chagas disease is caused by the intracellular obligatory parasite Trypanosoma cruzi and affects approximately eight million individuals in Latin America [11]. The transmission of this disease occurs through the faeces of sucking triatominae insects, blood transfusions, organ transplantation, oral contamination, laboratory accidents and congenital routes [12, 13]. Current major concerns are the outbreaks of acute Chagas disease associated with the ingestion of contaminated food and its emergence in non-endemic areas, such as North America and Europe, due to the immigration of infected individuals [14-16]. This disease is characterised by two clinical phases. The acute phase appears shortly after infection and is defined by patent parasitaemia. If left untreated, symptomatic chronic disease develops in about one-third of individuals after a long latent period (10-30 years), which is known as the indeterminate form. The main clinical manifestations of Chagas disease include digestive and/or cardiac alterations. The chronic cardiac form of the disease is the most significant clinical manifestation. Consequences include dilated cardiomyopathy, congestive heart failure, arrhythmias, cardioembolism and stroke [17].

The life cycle of T. cruzi involves four major developmental stages during its passage through vertebrate and invertebrate hosts [18]. The infective stage of the parasite, the metacyclic trypomastigote, enters the mammalian host from insect faeces through wound openings or mucous membranes. In the mammalian host, the metacyclic trypomastigote differentiates into the amastigote form. After several rounds of replication in the host cells, the amastigote differentiates into the bloodstream trypomastigote, which can enter new cells and perpetuate the infection. When the insect bites an infected host, the bloodstream trypomastigote differentiates into the replicative epimastigote that lives in the insect’s gut. Finally, in the rectum of the insect, the epimastigote differentiates into the infective metacyclic trypomastigote, which is ready to infect its host again.

The available chemotherapy for this illness includes two nitroheterocyclic agents, nifurtimox and benznidazole, which are effective against acute infections, but show poor activity in the late chronic phase, with severe collateral effects and limited efficacy against different parasitic isolates. These drawbacks justify the urgent need to identify better drugs to treat chagasic patients, and several new compounds are currently in preclinical development involving in vitro parasite phenotype screens and target-based drug discovery [19-21]. Recently, clinical trials with the azoles posaconazole and E1224 (ravuconazole prodrug) led to higher percentages of treatment failure in chronic patients than benznidazole [22, 23], suggesting their potential use in combination therapy [24].

1.1.3. Leishmaniasis

Leishmaniasis, which is caused by different species of Leishmania, is a vector-borne disease, with an estimated 12 million cases worldwide. Infection is caused by the bite of infected female sand flies of the genera Phlebotomus (Europe, Asia, Africa) and Lutzomyia (America) [25]. Leishmania parasites live a digenetic life cycle as either a promastigote flagellar or an amastigote form. The type of clinical manifestation depends on the infecting species and host factors, such as general health and genetic and immune constitution [26]. It is a disease complex with three clinical manifestations, visceral (VL, kala-azar), cutaneous (CL) and muco-cutaneous (MCL), which arise from parasite replication in the mononuclear phagocyte system, dermis and naso-oropharyngeal mucosa, respectively [27]. Some post-treated L. donovani-infected patients develop the diffuse cutaneous form named post-kala-azar dermal leishmaniasis (PKDL) [28, 29]. VL, after initial skin lesions, takes 2-8 months to develop gross inflammatory reactions within the viscera (liver and spleen in particular) and is usually fatal unless treated. CL manifests as an open sore at the site of the insect bite and will frequently self-heal, leaving a scar. The diffuse form of CL is more problematic, causing lepromatous type lesions disseminated across the skin that can be difficult to heal. The MCL form, endemic in parts of Latin America, starts with skin sores that spread to the mucosal membranes of the face. Profound inflammatory damage can lead to the erosion of the nostrils and mouth in particular [29].

In the Leishmania life cycle, there are two principal parasite forms: amastigotes and motile promastigotes. In the alimentary tract of the insect vector, the parasite exists as multiplicative, non-infective procyclic promastigotes and non-multiplicative, infective metacyclic promastigotes [30]. Upon injection into the mammalian host, promastigotes are taken up by macrophages where the metacyclic forms differentiate into small multiplicative, non-motile amastigotes that live in a lysosomal compartment known as the parasitophorous vacuole [31]. These developmental forms are distinguished by their nutritional requirements, their growth rate and ability to divide, the regulated expression of their surface molecules, and their morphology. Metacyclic promastigotes are pre-adapted for survival in the mammalian host, as they are complement-resistant. Amastigotes are intracellular, non-motile forms that have adapted to the low pH of this compartment and have an adapted energy metabolism.

The current drugs are highly toxic, resistance is common and compliance of patients to treatment is low, as the treatment is long and the drug price is high. Although recent initiatives have improved the antileishmanial drug arsenal by combining current medicines or using new formulations of old ones, none are ideal for treatment due to their high toxicity, resistance issues, prohibitive prices, long treatment length and need of intravenous administration [32-34]. Pentavalent antimonials (glucantime and pentostan) are first-line drugs for both VL and CL. However, they present several limitations, including variable efficacy, need for daily injectable administration for approximately one month, and severe side effects. Many patients are unable to complete the treatment, increasing the risk of drug resistance development. Amphotericin B is a systemic antifungal that is used as a second-line drug for VL. It is highly toxic, requiring careful and slow intravenous administration. Lipid formulations of amphotericin B have been developed to improve its bioavailability and pharmacokinetic properties, reducing toxicity [35]. Miltefosine is the most recent antileishmanial drug on the market and the first effective oral treatment against VL [36]. However, it has common gastrointestinal side effects and is also limited by its relatively high cost [34], potential teratogenicity and growing concerns in relation to increases in clinical isolate susceptibility [37]. Paromomycin is an aminoglycoside antibiotic that is used in topical treatment for CL and as a parenteral drug for VL. Pentamidine was used as a second-line drug in antimony-resistant VL treatment. However, its high toxicity combined with decreased efficacy led to the abandonment of this drug to treat VL in India, but it is valuable for combined therapies [38].

Advertisement

2. Cell death: State of art

As used for whole organisms, the term death is employed to describe a sequence of events culminating in the breakdown of all biological functions. However, more than one century after the first citation [39], cell death still represents a crucial gap in our understanding of cellular physiology. It can be triggered by natural processes or induced by extrinsic factors (exposure to chemicals or physical stresses). The consequent tissue injury usually leads to a state of disease [40]. On the other hand, many studies pointed to cell death playing a fundamental role in the physiology of multicellular organisms, especially in processes such as metamorphosis and embryogenesis [41]. In this context, in 1964, the term programmed cell death (PCD) was created, proposing a sequence of well-controlled steps regulating a non-accidental cell death process in the absence of an inflammatory response [42]. Currently, it is known that distinct death mechanisms and phenotypes participate in PCD, with apoptosis and autophagy being the most prominent [43].

2.1. Apoptosis

The apoptotic pathway was first described in the early 1970s as a fundamental step for proper embryo development [44]. This process is crucial during tissue development, especially in immune response regulation and removal of infected or damaged cells [45, 46]. Apoptosis is involved not only in growth regulation in multicellular organisms [47, 48] but also in their defence against viral, bacterial or parasitic infections [49-53] and even against cancer development [54-57]. The removal of non-functional cells by the apoptotic pathway is efficient and prevents the inflammatory response [58].

During apoptosis in multicellular organisms, the cell activates death machinery that culminates in chromosomal condensation and nuclear DNA fragmentation [59, 60]. Biochemically, apoptosis is orchestrated by the activation of a family of cysteine proteases, named caspases, that are activated by extrinsic and intrinsic factors [45, 46]. The extrinsic pathway is activated by the interaction of death ligands with their respective cell surface receptor (i.e., FasL/Fas, TNF-α/TNFR) [61-63]. Such binding triggers the cleavage of procaspase 8 into active caspase 8, which cleaves procaspase 3. Executioner caspase 3 activates endonuclease G (EndoG), starting the characteristic DNA fragmentation, a distinctive marker of apoptosis [63-65]. On the other hand, the intrinsic pathway can be triggered by two distinct mechanisms with mitochondrion or endoplasmic reticulum (ER) dependency. In the mitochondrial pathway, activation occurs by membrane permeabilization, releasing cytochrome c, apoptosis induction factor (AIF), EndoG and regulators of the B-cell lymphoma 2 (Bcl2) protein family into the cytosol. In the cytosol, the apoptosome is formed by the interaction of released cytochrome c with apoptotic protease activating factor 1 (APAF-1) and procaspase 9, activating caspase 9, which subsequently activates the effector caspase 3 [66-70]. The ER pathway is mainly caspase 12-dependent and occurs in this organelle during stress conditions. Because this pathway was described in the mouse and humans lack functional caspase 12, the relevance of ER-mediated apoptosis is still debatable [71-73].

Undoubtedly, the caspase cascade represents a central point in the apoptotic process. Its regulation is well-controlled by pro- and anti-apoptotic molecules from the Bcl-2 family [74]. The apoptotic morphological and biochemical phenotypes include cell shrinkage, membrane blebbing (formation of apoptotic bodies), chromatin condensation and typical internucleosomal DNA fragmentation, externalization of phosphatidylserine (PS), loss of mitochondrial membrane potential (ΔΨm), and target protein degradation by caspase activation [75-79]. The characterization of apoptosis is experimentally based on the detection of apoptotic markers. The loss of ΔΨm (labelling with rhodamine 123 derivatives, such as TMRE), PS exposure (binding to labelled annexin V), chromatin condensation (DAPI labelling) and DNA fragmentation (TUNEL technique) are usually quantified by fluorescence microscopy or flow cytometry. DNA fragmentation can also be assessed by agarose gel electrophoresis, presenting a laddering pattern that represents internucleosomal cleavage. Analysis of caspase activity using labelled specific substrates and/or inhibitors can be performed by immunotechniques such as ELISA [80].

2.2. Autophagy

In the 1950s, acidic organelles involved in the intracellular degradation of macromolecules were described and termed lysosomes by Dr. Christian de Duve. In a subsequent study [81], he proposed the term autophagy for a self-degrading process [82]. Currently, the autophagic pathway is considered to be the main cellular mechanism for the degradation of non-functional organelles and/or macromolecules and is fundamental for homeostasis in eukaryotic cells [83]. In other words, autophagy is a housekeeping self-digestion mechanism that is crucial for cellular turnover and recycling and occurs by the engulfment of cytosolic portions containing material that should be degraded. Degradation starts immediately after the fusion of autophagosomes to lysosomes in an organelle named the autophagolysosome [84, 85].

In multicellular organisms, autophagy is involved in many physiological situations, including development, cell growth and cell differentiation. Autophagy sustains cell survival under ‘extracellular stress’, such as nutrient starvation, hypoxia, acidic pH and high temperature. It acts as a housekeeping device under ‘intracellular stress’ by removing damaged or redundant cytoplasmic components, including organelles [86]. Increased autophagic activity is observed in pathological states and in host defences against pathogens [87-92]. Despite the relevant role of autophagy for the maintenance of the regular cell cycle, prolonged starvation periods or other strong autophagic stimuli induce a cellular misbalance and promote autophagic cell death [93, 94].

The autophagic molecular machinery was first assessed in the yeast model Saccharomyces cerevisiae, and 30 proteins, called Atgs (AuTophaGy-related), were described and associated with different steps of the pathway [95]. Atg orthologues were identified in all eukaryotes, with Atg8 (LC3 in mammals) being one of the most studied [82]. Autophagy can be a selective or non-selective process, degrading specific or random cellular components. Examples of selective routes are mitophagy, pexophagy or reticulophagy, in which mitochondria (or part of the organelle), peroxisomes and ER are degraded, respectively [82].

Additionally, there are three types of autophagy: macroautophagy, microautophagy and chaperone-mediated autophagy (CMA). The most common is macroautophagy, a process that involves the engulfment of cytosolic portions by a double membrane structure called the phagophore. The double-membrane vesicle formed from phagophore engulfment is named the autophagosome and is directed to lysosomes for degradation by lysosomal hydrolases. These steps are regulated by Atgs [92, 96-98]. The chronological events related to macroautophagy are (a) autophagic induction; (b) cargo selection; (c) phagophore elongation; (d) autophagosome formation; (e) fusion to lysosomes; and (f) cargo degradation [99]. The early steps in this process depend on the serine/threonine protein kinase TOR (target of rapamycin), which is essential for autophagic regulation. TOR complexes 1 and 2 work as sensors of nutritional availability (especially amino acids). The autophagic enzyme Atg6 (Beclin 1 in mammals) is a phosphatidylinositol 3-kinase (PI-3K) and shares its signalling function with other cellular pathways. For autophagy, these kinases present a critical role for autophagosome formation [82].

In contrast, there are no autophagosomes in the microautophagic pathway. Invagination of the lysosomal membrane occurs, resulting in a single-membrane small vesicle inside the lysosomes that will be degraded. Interestingly, both macro- and microautophagy could be selective or non-selective processes. Indeed, CMA appears to be the most selective type of autophagy. The proteins that will be degraded contain pentapeptide motifs (KFERQ, QREFK or VDKFQ), the binding sites of a cytosolic chaperone. Such a chaperone-substrate complex binds to a LAMP-2A receptor in the lysosomal membrane, promoting receptor dimerization. A membrane channel is formed, and the specific protein reaches the lysosomal lumen to be degraded [82, 100].

For many years, electron microscopy was the only tool available for the identification of autophagic morphological features, especially the presence of double-membrane vesicles (autophagosomes). In the last 20 years, advances in the molecular description of autophagy allowed the detection, localization and quantification of Atgs by molecular, biochemical and morphological approaches. Currently, the gold-standard method to monitor autophagy is Atg8/LC3 detection by different techniques: (a) Western blotting (presence of two isoforms); (b) confocal or fluorescence microscopy (identification of LC3 puncta); (c) knock down or knock out (deletion and analysis of the phenotype); and (d) pharmacological induction/inhibition (rapamycin and/or PI-3K inhibitors). These techniques can also be employed in vitro or in vivo for other Atgs, indicating autophagic activity [101].

2.3. Necrosis

Necrosis is a term that is extensively employed as synonymous with cell death. In the Greek aetiology, it signifies the "stage of dying". In this death type, strong cellular damage occurs caused by external stimuli (drugs, infection, mechanical trauma), promoting the random degradation of the whole cell, with plasma membrane disruption. Necrosis is defined as an accidental cell death process, differing from PCD (especially apoptosis) [102]. One of the main differences between apoptosis and necrosis is the induction of the inflammatory response in the latter. The release of intracellular material into the extracellular environment during necrotic cell death triggers intense inflammation in the surrounding cells and tissues [103]. Classical necrotic features are the loss of plasma membrane integrity, cytosolic vacuolization, disruption of calcium homeostasis, general degradation by lysosomal hydrolases and induction of the inflammatory response.

Necrosis can also be a regulated process. Necroptosis is a programmed and non-accidental death pathway. Surprisingly, the activation of this pathway can occur by TNF-α or FasL, classical apoptotic ligands. Necroptosis depends on the participation of the receptor-interacting protein kinases 1 and 3 (RIPK1 and RIPK3), which are kinases that regulate this pathway. RIPK1 is pharmacologically inhibited by a small molecule named necrostatin-1 (Nec-1) [104-106].

2.4. Others

In addition to apoptosis, autophagy and necrosis (accidental or not), other non-canonical death styles can take place in eukaryotic cells. In an inflammatory context, pyroptosis and NETosis are prominent. Pyroptosis, primarily observed in macrophages after bacterial infection, is caspase 1-dependent. This caspase promotes an increase in the inflammatory cytokine levels (IL-1β and IL-18) and the formation of plasma membrane pores, leading to the release of cellular material to the extracellular matrix. The main difference between pyroptosis and apoptosis is the participation of caspase 1, which is only involved in the pyroptotic death pathway, a proinflammatory PCD [106-108]. Another type of cell death that plays a crucial role in the innate immune response is the neutrophil extracellular trap (NETosis), where neutrophilic death leads to the release of a neutrophil DNA network coated with histones and elastase to the extracellular environment to capture pathogens. However, the direct antimicrobial effect of the NETs is still controversial [109, 110]. Currently, DNA release has also been described in other immune cells, such as eosinophils, basophils, macrophages and mast cells, but its precise role deserves further analysis [110-114].

Other cell death types not involved in inflammation have been characterized. Ferroptosis is iron-dependent cell death that has been identified in some mammalian cells and involves oxidative stress induced by a small molecule named erastin, which is inhibited by ferrostatin 1. Despite that lack of complete understanding of the erastin mechanism, the XC−Cys/Glu antiporter system is inhibited in ferroptosis, leading to a misbalance of these amino acids inside the cell [106, 115]. Additionally, there is another non-canonical cell death pathway in cancer cells (in vitro and in vivo models) called autoschizis, which involves oxidative stress induced by treatment with ascorbate and menadione. Autoschizic cell death presents remarkable morphological evidence, with electron microscopy as the best technique for its identification. Among the autoschizic features are cell shrinkage, extrusion of large portions of the cytosol (without any organelles), random DNA fragmentation and the subsequent deterioration of all cellular structures [116, 117]. Interestingly, annexin V (AV) and propidium iodide (PI) assays (gold standards for apoptosis detection in mammals) of cells treated with ascorbate and menadione demonstrate high percentages of AV-/PI+ cells [117], which are not discussed in almost all apoptotic studies, suggesting that these membrane shedding events could occur in a large variety of cell models. Table 1 summarizes the main types of cell death discussed herein.

Cell death Features References
apoptosis cell shrinkage
membrane blebbing
DNA fragmentation
externalization of PS
activity of caspases
regulation by Bcl-2 family proteins
loss of ΔΨ
release of cytochrome c
no inflammatory response
[44, 76, 78]
autophagy presence of autophagosomes
participation of Atgs
regulation by PI-3K and TORC
degradation by lysosomes
presence of KFERQ, QREFK or VDKFQ motifs in the protein to be degraded (only in CMA)
[82, 101]
necrosis disruption of plasma membrane
cytoplasmic vacuolization
imbalance of Ca2+ homeostasis
release of lysosomal enzymes
induction of inflammatory response
[102, 103]
necroptosis participation of RIP1 and RIP3
inhibition by Nec-1
[104, 106]
pyroptosis participation of caspase 1
increase in IL-1β and IL-18 levels
induction of inflammatory response
[106-108]
NETosis formation of NETs
participation of elastase and histones
occurrence in neutrophils, macrophages, mast cells, eosinophils and basophils
[109, 110]
ferroptosis participation of iron
presence of oxidative stress
induction by erastin
blockage of XC−Cys/Glu antiporter system
inhibition by ferrostatin 1
[106, 115]
autoschizis cell shrinkage
random DNA fragmentation
extrusion of large cytosolic portions (without organelles)
degradation of cellular components
increase in the AV-/PI+ population
[116, 117]

Table 1.

Types of cell death

Advertisement

3. Cell death in trypanosomatids: An overview

The term PCD was employed for decades to exclusively describe cell death in metazoans and its involvement in embryogenesis and maintenance of homeostasis. Indeed, the relevance of PCD for lower eukaryotes is unclear. In an evolutionary scenario, these regulated processes could allow clonal selection in the parasite population, guaranteeing the propagation of identical genetic information even in adverse environmental conditions. However, differences in the cell death mechanisms observed between metazoans and protozoans must be considered [78, 118]. In the following sections, we will discuss the role of different death styles described in pathogenic trypanosomatids.

3.1. Apoptosis-like

In trypanosomatids, the first PCD report was published in 1995 by Ameisen and coworkers describing apoptotic characteristics (DNA fragmentation and cytoplasmic and nuclear morphological alterations) in T. cruzi epimastigotes during differentiation to trypomastigotes [119]. In the last two decades, a variety of stimuli were reported to induce the appearance of the apoptotic phenotype in this parasite, including exposure to fresh human serum (FHS), heat shock and drugs [76, 119-128]. Curiously, the apoptosis-like phenotype was also associated with the regulation of the T. cruzi life cycle [129]. These cell death phenotypes in pathogenic trypanosomatids have been characterized by the use of classical apoptotic markers (see item 2.1) [76, 79, 118, 123, 130-133]. Among the apoptotic hallmarks identified, we found (a) loss of ΔΨm, (b) cytochrome c release, (c) PS externalization, and (d) abnormal DNA condensation and fragmentation [76, 119, 129, 130, 134] (Table 3, Figure 1).

Figure 1.

Schematic representation of apoptosis-like PCD in pathogenic trypanosomatids. Apoptotic stimuli induce loss of ΔΨm, release of mitochondrial cytochrome c to the cytosol, PS externalization and DNA fragmentation by EndoG activity. Apoptotic regulators from the Bcl-2 family were not found until now, and the role of metacaspases is controversial, suggesting that apoptosis-like PCD in trypanosomatids is a caspase-like- and Bcl-2-independent pathway. N: nucleus; M: mitochondrion; K: kinetoplast; F: flagellum.

In Leishmania sp., apoptotic features (nuclear condensation, DNA fragmentation, cell shrinkage, loss of ΔΨm, and release of cytochrome c) were also observed in stress conditions induced by heat, starvation, oxidative agents and drugs [118, 134, 136-139, 135]. L. donovani, L. major and L. mexicana stationary phase promastigotes and axenic amastigotes exhibited DNA fragmentation with a laddering electrophoretic profile, suggesting oligonucleosomal cleavage. These data were corroborated by the description of a non-canonical, Ca2+- and Mg2+-independent 45-59 kDa endonuclease [76, 136, 140].

As in other pathogenic trypanosomatids, apoptotic features were also identified in T. brucei under non-physiological conditions, such as incubation with drugs, cytokines or ROS [129, 133, 141-143]. Interestingly, the gene for prohibitin and the receptor for activated protein kinase C have been correlated with the apoptotic process, suggesting convergence between these pathways in protozoa and mammals (Table 2) [129]. Despite several reports about caspase-like activity in trypanosomatids [75, 134, 136, 144], the exact role of these proteases in protozoa is not clear. Metacaspases are structurally similar to mammalian orthologues, but their catalytic activity on caspase substrates is quite controversial [145-147]. Despite their presence in T. cruzi, T. brucei and Leishmania sp., only L. major metacaspase shows in vitro self-proteolytic activity (Table 2) [146]. In fact, the participation of metacaspases cleaving vital substrates in the cell death cascade has not yet been described [148, 149]. Surprisingly, experimental evidence pointed to the involvement of these proteases in cell cycle control and metacyclogenesis, not in death [145, 150-153].

In unicellular organisms, the mitochondrion is a central organelle in cell death pathways, leading to ROS production [125]. In T. brucei procyclic forms, mitochondrial Ca2+ influx misbalance culminates in ROS generation [154]. Additionally, prostaglandin D2-induced ROS production in both the bloodstream and procyclic forms led to the labelling of different apoptotic markers, with the death phenotype reverted by oxidative scavengers, such as N-acetyl cysteine [130, 155, 156]. In L. donovani, hydrogen peroxide induced classical apoptotic features (DNA fragmentation, loss of ΔΨm and caspase-like activity). This phenotype was partially reverted by caspase inhibitors [134, 137]. Oxidative stress plays a crucial role not only in apoptosis-like PCD but also in autophagy and necrosis, as we will discuss later [78, 157].

Molecule Organism References
Prohibitin
RACK
T. brucei [129]
Elongation factor 1 ∝ T. cruzi [161]
Metacaspases 1 L. donovani
T. brucei
[147, 162, 163]
Metacaspases 2 L. donovani
T. brucei
[145, 147, 162]
Metacaspases 3 T. cruzi
T. brucei
[145, 150, 153, 162]
Metacaspases 4 T. brucei [162, 164]
Metacaspases 5 T. cruzi
T. brucei
L. major
[145, 150, 162, 165]
Metacaspase Z-DEVD-FMK -sensitive T. cruzi
L. donovani
[124, 134, 136]
Endonuclease G L. major
T. brucei
L. infantum
L. donovani
[132, 158, 166]
LdFEN-1
LdTatD-like nuclease
L. donovani [158]

Table 2.

Apoptotic molecules described in pathogenic trypanosomatids

The participation of EndoG-like in mitochondrial-mediated cell death has been reported, but the process is metacaspase-independent (Table 2) [132, 158, 159]. L. infantum submitted to heat stress also presents an apoptotic pattern, but without caspase-like activity, which was partially reversed by the expression of the anti-apoptotic mammalian gene Bcl-XL [160]. On the other hand, the overexpression of mammalian anti-apoptotic Bcl-2 in T. brucei caused no reversion of the mitochondrial damage induced by ROS [154]. However, members of the Bcl-2 protein family have not been described in trypanosomatids [129]. More studies regarding the regulation steps of apoptosis-like processes in trypanosomatids need to be performed.

3.2. Autophagy

Almost forty years ago, the first morphological autophagic evidence was described in trypanosomatids by electron microscopy of T. brucei [170]. In the last four decades, many studies have described recurrent autophagosome formation (initially named autophagic vacuoles), multivesicular bodies as well as myelin-like structures in pathogenic trypanosomatids treated with different classes of drugs (Figure 2) [169, 168, 171-178]. Such autophagosomes showed distinct levels of degradation depending on the degree of cellular structure damage inside the organelle. Myelin-like structures are one of the most frequent ultrastructural alterations detected in drug-treated parasites and are suggestive of the cellular recycling of damaged structures. Currently, it is postulated that myelin-like structures are phagophores (or pre-autophagosomal structures, PAS), an early step in the formation of doubled-membrane autophagosomes (Table 3). In T. cruzi, ER profiles were reported as the main origin of phagophores (Figure 2). These profiles usually surround a pre-lysosomal compartment, named the reservosome, suggesting the participation of this organelle in autophagolysosome formation in epimastigote forms [82, 178].

Figure 2.

Schematic representation of autophagy in pathogenic trypanosomatids. Autophagic stimuli induce the formation of phagophores from ER profiles. The phagophore engulfs organelles and molecules, generating autophagosomes. Targeting and engulfment are Atg-dependent processes. These autophagosomes fused with lysosomes generate autophagolysosomes. Continuous autophagic stimuli lead to autophagic cell death, which is inhibited by the pre-treatment of the parasite with autophagic inhibitors (wortmannin or 3-methyladenine).

Cell death Features Organism References
apoptosis-like cell shrinkage
membrane blebbing
DNA fragmentation
PS exposure
loss of the ΔΨm
release of cytochrome c
T. cruzi
T. brucei
Leishmania sp.
[76, 118, 119, 130]
autophagy presence of autophagosomes-like
Golgi and/or ER profiles surrounding organelles
detection of Atg8 and Atg4
T. cruzi
T. brucei
Leishmania sp.
[82, 92, 167]
necrosis cytosolic vacuolization
plasma membrane disruption
T. cruzi
T. brucei
Leishmania sp.
[149, 168, 169]

Table 3.

Types of cell death described in pathogenic trypanosomatids

In the last few years, a functional autophagic pathway was characterized in trypanosomatids and ATG homologues were identified. However, almost half of the yeast Atgs are lacking in these protozoa [167, 179-181]. Currently, in trypanosomatids, twenty autophagic genes have been found to be involved in all of the steps, from vesicle expansion and completion to degradation (Figure 2) [167]. Bioinformatic approaches revealed all four genes of the Atg8 conjugation system (Atg3, Atg4, Atg7 and Atg8). Atg8 is well-characterized in T. cruzi, T. brucei and Leishmania sp. and is located in autophagosomes, as observed in yeast and mammals. Atg8 has four isoforms (Atg8, Atg8A, Atg8B and Atg8C) that are processed by two isoforms of Atg4 (Atg4.1 and Atg4.2) [92, 181-185] (Table 3). On the other hand, the Atg12 conjugation system has poor sequence similarity in trypanosomatids, and Atg5, Atg10 and Atg12 sequences are lacking [167, 180]. Pathogenic trypanosomatids have two TOR kinases (TOR1 and TOR2) that form their respective complexes, TORC1 and TORC2 (Table 4). These two complexes show a distinct molecular behaviour, subcellular localization and susceptibility to rapamycin [186, 187]. The treatment of T. brucei bloodstream forms with rapamycin led to cell cycle arrest and an increase in the number of autophagosomes due to TORC2 inhibition. However, rapamycin had no effect on the parasite TORC1, suggesting another function for this complex [186, 188].

Molecules Organism References
Atg4.1, Atg4.2 L. major [197]
Atg5 L. major [190]
Atg5, Atg10, Atg12 L. major [183]
Atg8A, Atg8B, Atg8.2 T. brucei [183]
Atg8.1, Atg8.2 T. cruzi [92]
Atg8, Atg8A, Atg8B, Atg8C L. major [183]
TOR1
TOR2
T. brucei
L. major
[186, 198, 199]
TOR3 L. major [199]
Vps34 T. cruzi
T. brucei
[200, 201]

Table 4.

Autophagic molecules described in pathogenic trypanosomatids

In addition to the recycling function, autophagy plays a fundamental role in parasite differentiation and survival, mitochondrial function and homeostasis of phospholipids [92, 182, 189, 190]. In metacyclogenesis, the autophagic pathway is triggered by nutritional deprivation, playing an important function in both the infectivity and virulence to the vertebrate host [182]. During the T. cruzi life cycle, epimastigotes are submitted to starvation in the insect rectum, a crucial event for protozoa differentiation. Starved epimastigotes express Atg8.1, but such expression is decreased in metacyclic forms [82, 92]. Reservosomes disappeared during differentiation, most likely due to the cysteine proteinase activity, in particular, cruzipain [159, 191, 192].

In Leishmania sp., autophagy is essential for metacyclogenesis, with several observed autophagosomes during the process [193, 194]. The deletion of Atg4.2 led to an accumulation of Atg8 lipidated isoforms, compromising the autophagic activity. Subsequently, a reduction in the number of differentiating promastigotes was observed [194]. Interestingly, autophagy also participates in the differentiation of L. mexicana metacyclic promastigotes to amastigotes [193]. In the sandfly, the exposure of promastigotes to different stress stimuli, including higher temperature, low pH, and nutritional deprivation, acts as a crucial event for the success of the metacyclogenesis [189, 192]. L. mexicana shows that lysosome-like structures, called megasomes, are involved in parasite differentiation, with the activity of two megasomal cysteine peptidases (CPA and CPB) associated with autophagy. The deletion of these proteases strongly impaired its differentiation into amastigotes, leading to an accumulation of autophagosomes containing multi-vesicular tubules (structures related to endocytosis) [180, 193].

A peculiar role for autophagy was observed in T. brucei. In a selective pathway, glycosomes are degraded during differentiation from bloodstream to procyclic forms. This organelle is a peroxisome-like structure that is also involved in the glycolytic pathway, and its degradation via autophagy led to important changes in the protozoa bioenergetics [195]. This evidence supported the existence of pexophagy in trypanosomes, an essential event for energy balance during the parasite life cycle. Depending on the environmental conditions (distinct hosts), the sources of energetic substrates vary, as does the ATP demand [180]. Recently, it was also reported that T. brucei acidocalcisomes (an acidic compartment that stores ions responsible for polyphosphate metabolism) regulate autophagy by the acidification of this organelle. Moreover, the blockage of acidocalcisome biogenesis also inhibited the autophagic pathway without the impairment of lysosomal biogenesis or function, suggesting the relevance of acidocalcisomes as an autophagic regulator [196].

Autophagic cell death occurs when the homeostatic balance is broken [40]. To evaluate whether autophagy participates in the cell death process, the use of the PI-3K inhibitors wortmannin and 3-methyladenine (3-MA) before the autophagic stimulus is provided is an interesting experimental approach. Pre-treatment with these inhibitors totally abolished the trypanocidal activity of naphthoimidazoles in T. cruzi epimastigotes and trypomastigotes. Although the involvement of components of the Atg8 conjugation system was also demonstrated, the molecular mechanisms of cell death regulation in this parasite deserve further examination [82, 178].

3.3. Necrosis

As described for higher eukaryotes, necrosis is poorly studied in protozoa, especially due to its conception as an accidental and uncontrolled process. The most typical necrotic feature is the plasma membrane rupture that leads to the loss of cellular homeostasis and consequent cell lysis as the consequence of a mechanical or chemical stimulus [103]. Necrosis is always the cell death endpoint, culminating in the generation of cellular debris. Thus, independent of the cell death mechanism that is induced, all parasites will lyse in a system without phagocytic cells to clean the microenvironment. In this context, a high percentage of anti-trypanosomatid natural or synthetic drugs present a mechanism of action with a lytic effect [29, 149, 202-205] (Table 3).

Another crucial stress condition that induces trypanosomatid disruption is the activation of the complement pathway. This cascade can be triggered by the binding of lectins to lipophosphoglycans presented on the surface of Leishmania sp. promastigotes and of glycosylated molecules in the T. cruzi metacyclic form [206-209]. Indeed, pathogenic trypanosomatids show different mechanisms to evade the complement pathway. For example, T. brucei expresses a vast number of variant surface glycoproteins (VSG) that change the parasite coat to escape from the host immune system [210]. In relation to programmed necrosis, RIPK-like molecules have not yet been identified in unicellular organisms, and the direct effect of Nec-1 has not been evaluated, suggesting that an orchestrated pathway similar to necroptosis is absent in trypanosomatids.

3.4. Others

Curiously, no studies have been reported about non-canonical PCD pathways in trypanosomatids. Pyroptosis and NETosis are processes that are characterized exclusively in mammalian cells, specifically during an inflammatory response. Such pathways involve the death of immune cells to block the progression of any infection by a well-regulated mechanism [106, 110]. The absence of these PCD types in unicellular organisms is not strange. On the other hand, the existence of specific oxidative stress-related cell death types in trypanosomatids would be reasonable. Continuous exposure of these parasites to ROS under distinct environmental conditions during their life cycles indicates the important role of oxidative stress in the control of protozoa populations. ROS involvement in trypanosomatid apoptosis-like processes and autophagy has been described in different experimental conditions [130, 155, 156, 211, 212], but ferroptosis has not yet been investigated. Further studies about the effect of erastin as well as the inhibition by ferrostatin 1 should be performed in these parasites. Autoschizis was only observed in cancer cells under very specific conditions, but interestingly, an autoschizic phenotype (high percentages of AV-/PI+ cells) was detected in T. cruzi treated with naphthoimidazoles [178]. The AV-/PI+ population is ignored in the majority of the studies, including in pathogenic trypanosomatids [213-215]. A better characterization of this parasite population must be performed to exclude the existence of autoschizis in protozoa.

3.5. Cell death and evasion of host immune response

Trypanosomatids presented a highly sophisticated repertoire to evade mammalian immune systems, including the capacity to prevent the cell death pathways of the infected host cells [188]. This efficient strategy allows host PCD modulation by the parasites to establish the infection. Depending on the protozoan species and the host cell type, PCD exacerbation or inhibition fluctuates. For example, the induction of apoptosis in immune cells increases the parasite persistence and survival in immunocompetent hosts [78]. In T. cruzi infection, apoptosis of lymphocytes and macrophages is essential for the parasite to escape, promoting inflammation reduction by anti-inflammatory cytokines and also amastigote proliferation [78, 216, 217]. The Leishmania strategy is quite different. Promastigotes externalize PS to be recognized by phagocytic cells. The binding of PS to its receptor on the phagocyte surface triggers a signalling cascade that guides TGF-β production and the subsequent anti-inflammatory response. This phenomenon, called apoptotic mimicry, facilitates parasite internalization and increases the success of the infection [218, 219]. Additionally, the intracellular cycle of Leishmania sp. also depends on the impairment of host cell apoptosis. This event is necessary to stop or delay the elimination of infected cells. For example, L. major uses the infected apoptotic granulocytes as "Trojan horses" to invade macrophages, the definitive host cells, avoiding the direct activation of phagocytes via the interaction between host receptors and protozoa [220].

Host autophagy also represents a valuable mechanism for both innate and adaptive responses to stop the infection. Its blockage is a crucial tactic for pathogenic trypanosomatids to evade host defences. Autophagy uses a process to eliminate pathogens, called xenophagy, directing microorganisms to be digested in lysosomes. This strategy is usually employed by protozoa living inside parasitophorous vacuoles to use the autophagic machinery to provide nutrients [82]. However, protozoa, such as Leishmania sp., change the autophagosomal pH and impair vesicular traffic, compromising the fusion to lysosomes. L. amazonensis amastigotes proliferate in starvation or even after treatment with rapamycin, but the proliferation is inhibited by incubation with the autophagic inhibitors wortmannin or 3-methyladenine [221]. The importance of autophagy for the Leishmania infection was corroborated by the observation that this pathway is exacerbated in L. amazonensis-infected mice and in a natural L. donovani infection in humans [222, 223]. Similar data were observed in the in vitro T. cruzi infection, suggesting that the autophagic pathway favours the parasite during its interaction with the host cell [224, 225]. However, the role of host autophagy in this trypanosomatid is still controversial due to the autophagic participation in the control of T. cruzi infection [226-228]. Furthermore, differences among strains and host cells must be considered to clarify whether host autophagy kills T. cruzi or provides nutrients for its survival.

Advertisement

4. Concluding remarks

In spite of the variety of studies about cell death in protozoans, including trypanosomatids, and the evidence of PCD, the detailed aspects of the molecular mechanisms and regulation remain unclear. The absence of key molecules together with the lack of an obvious role for this process in unicellular organisms makes the existence of PCD in these cells a debatable point, and the term “apoptosis-like” is more convenient [130, 172, 229]. In this context, the lack of a strong correlation between the proteolytic properties of caspases and their role in PCD should be highlighted. Currently, there is no description of the participation of trypanosomatid metacaspases in cell death processes, but these proteases have been postulated to function in proliferation and differentiation, which are important events for parasite survival [145, 148, 149, 153, 230]. In the post-genomic era, a rigorous search should be performed in proteomic databases of pathogenic trypanosomatids to correct misannotations in cell death proteins, validating the real role of these molecules for PCD processes.

Nevertheless, PCD was conserved during evolution, suggesting its essential function for the survival and maintenance of these species. However, it has been proposed that these pathways appeared in the phylogenetic tree in the multicellular organism branches, suggesting that the death molecular mechanisms identified in unicellular parasites came from a divergent evolutionary event [48]. This idea is supported by the replacement or complete absence of some PCD molecules, justifying the differences observed in protozoa mechanisms [79]. In addition to being an interesting evolutionary model for PCD, its physiological relevance for protozoa is still the most attractive question.

An altruistic hypothesis has been raised for unicellular organisms, especially for pathogenic trypanosomatids [130]. It was associated with the control of parasite populations, including protozoa density regulation, clonal selection and immune host system evasion, events related to the success of the infection [7, 76, 82, 136, 231]. Trypanosomatid cell death limits parasite colonization in insects in response to scarce resources of nutrients, avoiding invertebrate death [118, 130, 134]. On the other hand, PCD of T. cruzi or L. amazonensis insect forms under mammalian temperatures could evade host immune response derived from parasite lysis, facilitating the infection [76, 119, 135].

Autophagic cell death has been proposed as a PCD pathway, suggesting an active role of autophagy in death processes, but the precise mechanisms of regulation are not yet clear [174, 178, 232]. The majority of the autophagic studies were performed in yeast and mammal models. However, little is known about protozoan pathways. Autophagy is a regulated process that is directly involved in the preservation of cellular homeostasis and survival. Several hypotheses have been raised about the participation of this pathway in cell death in dying cells. The selective autophagic degradation of essential cellular factors, such as cell death regulators, triggers death events, including caspase activation [232, 233]. Another hypothesis suggested that autophagy is not a specific and regulated cell death process but is a consequence of extensive injury. Once such an injury compromises cellular physiology, the damaged structure needs to be degraded for cell survival. This hypothesis also explains the presence of similar phenotypes in parasites after treatment with different compounds with distinct mechanisms of action. Such autophagic phenotypes, detected independent of the stimuli, reinforced this pathway as a desperate attempt of the cells to stay alive [168, 212, 232]. The determination of the connection between the autophagic cell death of pathogens, such as trypanosomatids, could have crucial implications for human health, but further mechanistic studies should be addressed in this field.

Figure 3.

Different pathways of trypanosomatid death. The death stimulus triggers specific mechanisms of action depending on the environmental conditions, time of treatment and dose. Death signals lead to distinct well-known phenotypes from each pathway. Cross-talk could also be observed between apoptosis-like processes, autophagy and necrosis, culminating in protozoa death. The existence of an alternative unknown process cannot be discarded (dashed arrow).

The existence of cross-talk among different cell death pathways, especially autophagy and apoptosis, has been proposed (Figure 3) [93, 234]. In unicellular parasites, different cell death types have been described to be induced by physical and/or chemical stress conditions (drugs, heat shock, and nutritional deprivation, among others), resulting in a non-classical cell death phenotype. The total absence of commercial typical PCD markers, such as antibodies and enzyme activity kits, for protozoa and of key autophagic and apoptotic-like molecules reinforce the hypothesis of an interplay of distinct death mechanisms, suggesting their convergence, leading to necrosis. Likewise, the possibility of the occurrence of other PCD forms cannot be excluded [74, 78, 168, 178]. A better molecular characterization of cell death in pathogenic trypanosomatids is essential for advances in novel alternatives for therapeutic intervention.

References

  1. 1. WHO. Neglected tropical diseases, http://www.who.int/neglected_diseases/en/ accessed in 12/01/2015, 2015a.
  2. 2. Simarro PP, Diarra A, Ruiz Postigo JA, Franco JR, Jannin JG. The Human African Trypanosomiasis Control and Surveillance Programme of the World Health Organization 2000–2009: The Way Forward. PLoS Negl Trop Dis 2011;5(2):e1007.
  3. 3. WHO, http://www.who.int/trypanosomiasis_african/en/index.html, accessed in 12/01/2015, 2015b.
  4. 4. Kennedy PG. Clinical features, diagnosis, and treatment of human African trypanosomiasis (sleeping sickness). Lancet Neurol 2013;12(2):186-94.
  5. 5. Franco JR, Simarro PP, Diarra A, Jannin JG. Epidemiology of human African trypanosomiasis. Clin Epidemiol 2014;6:257-75.
  6. 6. Horn D. Antigenic variation in African trypanosomes. Mol Biochem Parasitol 2014;195(2):123-9.
  7. 7. Vickerman K. Developmental cycles and biology of pathogenic trypanosomes. Br Med Bull 1985;41(2):105-14.
  8. 8. Matthews KR. The developmental cell biology of Trypanosoma brucei. J Cell Sci 2005;118(Pt 2):283-90.
  9. 9. Steverding D. The development of drugs for treatment of sleeping sickness: a historical review. Parasit Vectors 2010;3(1):15.
  10. 10. Barrett MP, Vincent IM, Burchmore RJ, Kazibwe AJ, Matovu E. Drug resistance in human African trypanosomiasis. Future Microbiol 2011;6(9):1037-47.
  11. 11. WHO. Working to overcome the global impact of neglected tropical diseases: first WHO report on neglected tropical diseases. Geneva: WHO; 2010. Available from: http://www.who.int/neglected_diseases/2010report/en/
  12. 12. Steindel M, Kramer Pacheco L, Scholl D, Soares M, de Moraes MH, et al. Characterization of Trypanosoma cruzi isolated from humans, vectors, and animal reservoirs following an outbreak of acute human Chagas disease in Santa Catarina State, Brazil. Diagn Microbiol Infect Dis 2008;60(1):25-32.
  13. 13. Dias JC, Amato Neto V, Luna EJ. Alternative transmission mechanisms of Trypanosoma cruzi in Brazil and proposals for their prevention. Rev Soc Bras Med Trop 2011;44:375-9.
  14. 14. Coura JR, Dias JCP. Epidemiology, control and surveillance of Chagas disease - 100 years after its discovery. Mem Inst Oswaldo Cruz 2009;104 (Suppl 1):31-40.
  15. 15. Gascon J, Bern C, Pinazo MJ. Chagas disease in Spain, the United States and other non-endemic countries. Acta Trop 2010;115:22-7.
  16. 16. Schmunis GA, Yadon ZE. Chagas disease: a Latin American health problem becoming a world health problem. Acta Trop 2010;115(1-2):14-21.
  17. 17. Rassi A Jr, Rassi A, Marcondes de Rezende J. American trypanosomiasis (Chagas disease). Infect Dis Clin North Am 2012;26(2):275-91.
  18. 18. De Souza W. From the cell biology to the development of new chemotherapeutic approaches against trypanosomatids: dreams and reality. Kinetoplastid Biol Dis 2002;1(1):3.
  19. 19. McKerrow JH, Doyle PS, Engel JC, Podust LM, Robertson SA, Ferreira R, et al. Two approaches to discovering and developing new drugs for Chagas disease. Mem Inst Oswaldo Cruz 2009;104 Suppl 1:263-9.
  20. 20. Soeiro MN, De Castro SL. Screening of potential anti-Trypanosoma cruzi candidates: in vitro and in vivo studies. Open Med Chem J 2011;5:21-30.
  21. 21. Moraes CB, Giardini MA, Kim H, Franco CH, Araujo-Junior AM, Schenkman S, et al. Nitroheterocyclic compounds are more efficacious than CYP51 inhibitors against Trypanosoma cruzi: implications for Chagas disease drug discovery and development. Sci Rep 2014;4:4703.
  22. 22. Urbina JA. Recent clinical trials for the etiological treatment of chronic Chagas disease: advances, challenges and perspectives. J Eukaryot Microbiol 2015;62(1):149-56.
  23. 23. Chatelain E. Chagas disease drug discovery: toward a new era. J Biomol Screen 2015;20(1):22-35.
  24. 24. Molina I, Gómez-Prat J, Salvador F, Treviño B, Sulleiro E, Serre Net al. Randomized trial of posaconazole and benznidazole for chronic Chagas' disease. N Engl J Med 2014;370(20):1899-908.
  25. 25. WHO. Global report for research on infectious diseases of poverty, available at http://whqlibdoc.who.int/publications/2012/9789241564489_eng.pdf, 2012.
  26. 26. Murray HW, Berman JD, Davies CR, Saravia NG. Advances in leishmaniasis. Lancet 2005;366:1561-77.
  27. 27. Desjeux P. Leishmaniasis. Nat Rev Microbiol 2004;2(9):692.
  28. 28. Zijlstra EE, Musa AM, Khalil EA, el-Hassan IM, el-Hassan AM. Post-kala-azar dermal leishmaniasis. Lancet Infect Dis 2003;3(2):87-98.
  29. 29. Barrett MP, Croft SL. Management of trypanosomiasis and leishmaniasis. Br Med Bull 2012;104:175-96.
  30. 30. Bates PA, Rogers ME. New insights into the developmental biology and transmission mechanisms of Leishmania. Curr Mol Med 2004;4(6):601-9.
  31. 31. Chang KP, Dwyer DM. Multiplication of a human parasite (Leishmania donovani) in phagolysosomes of hamster macrophages in vitro. Science 1976;193:678-80.
  32. 32. Croft SL, Olliaro P. Leishmaniasis chemotherapy - challenges and opportunities. Clin Microbiol Infect 2011;17(10):1478-83.
  33. 33. Sundar S, Chakravarty J. Leishmaniasis: an update of current pharmacotherapy. Expert Opin Pharmacother 2013;14(1):53-63.
  34. 34. Sundar S, Chakravarty J. Investigational drugs for visceral leishmaniasis. Expert Opin Investig Drugs 2015;24(1):43-59.
  35. 35. Yardley V, Croft SL. Activity of liposomal amphotericin B against experimental cutaneous leishmaniasis. Antimicrob Agents Chemother 1997;41(4):752-6.
  36. 36. Sundar S, Singh A, Rai M, Prajapati VK, Singh AK, Ostyn B, et al. Efficacy of miltefosine in the treatment of visceral leishmaniasis in India after a decade of use. Clin Infect Dis 2012;55(4):543-50.
  37. 37. Prajapati VK, Mehrotra S, Gautam S, Rai M, Sundar S. In vitro antileishmanial drug susceptibility of clinical isolates from patients with Indian visceral leishmaniasis - status of newly introduced drugs. Am J Trop Med Hyg 2012;87(4):655-7.
  38. 38. Olliaro PL. Drug combinations for visceral leishmaniasis. Curr Opin Infect Dis 2010;23(6):595-602.
  39. 39. Crile G, Dolley DH. On the effect of complete anemia of the central nervous system in dogs resuscitated after relative death. J Exp Med 1908;10(6):782-810.
  40. 40. Smirlis D, Duszenko M, Ruiz AJ, Scoulica E, Bastien P, Fasel N, Soteriadou K. Targeting essential pathways in trypanosomatids gives insights into protozoan mechanisms of cell death. Parasit Vectors 2010;3:107.
  41. 41. Lockshin RA, Zakeri Z. Programmed cell death and apoptosis: origins of the theory. Nat Rev Mol Cell Biol 2001;2(7):545-50.
  42. 42. Lockshin RA, Williams CM. Programmed cell death. I. Cytology of degeneration in the intersegmental muscles of the pernyi silkmoth. J Insect Physiol 1965;11:123-33.
  43. 43. Kroemer G, Galluzzi L, Vandenabeele P, Abrams J, Alnemri ES, Baehrecke EH, et al., Nomenclature Committee on Cell Death 2009: Classification of cell death: recommendations of the Nomenclature Committee on Cell Death 2009. Cell Death Differ 2009;16:3-11.
  44. 44. Kerr JF. History of the events leading to the formulation of the apoptosis concept. Toxicology 2002;181/182:471-4.
  45. 45. Green DR. Introduction: apoptosis in the development and function of the immune system. Semin Immunol 2003;15(3):121-3.
  46. 46. Danial NN, Korsmeyer SJ. Cell death: critical control points. Cell 2004;116(2):205-19.
  47. 47. Evan IG. Better dead than red. Ther Immunol 1994;1(6):343-8.
  48. 48. Vaux DL, Haecker G, Strasser A. An evolutionary perspective on apoptosis. Cell 1994;76(5):777-9.
  49. 49. Williams GT. Programmed cell death: a fundamental protective response to pathogens. Trends Microbiol 1994;2(12):463-4.
  50. 50. Shen Y, Shenk TE. Viruses and apoptosis. Curr Opin Genet Dev 1995;5:105-11.
  51. 51. Roulston A, Marcellus RC, Branton PE. Viruses and apoptosis. Annu Rev Microbiol 1999;53:577-628.
  52. 52. Weinrauch Y, Zychlinsky A. The induction of apoptosis by bacterial pathogens. Annu Rev Microbiol 1999;53:155-87.
  53. 53. DosReis GA, Barcinski MA. Apoptosis and parasitism: from the parasite to the host immune response. Adv Parasitol 2001;49:133-61.
  54. 54. Williams GT. Programmed cell death: apoptosis and oncogenesis. Cell 1991;65:1097-98.
  55. 55. Raff MC. Social controls on cell survival and cell death. Nature 1992;356(6368):397-400.
  56. 56. Reed JC. Bcl-2 and the regulation of programmed cell death. J Cell Biol 1994;124(1-2):1-6.
  57. 57. Rich T, Allen RL, Wyllie AH. Defying death after DNA damage. Nature 2000;407(6805):777-83.
  58. 58. Voll RE, Herrmann M, Roth EA, Stach C, Kalden JR, Girkontaite I. Immunosuppressive effects of apoptotic cells. Nature 1997;390:350-1.
  59. 59. Wyllie AH. Glucocorticoid-induced thymocyte apoptosis is associated with endoGenous endonuclease activation. Nature 1980;284:555-556.
  60. 60. Steller H. Mechanisms and genes of cellular suicide. Science 1995;267:1445-1449.
  61. 61. Kischkel FC, Hellbardt S, Behrmann I, Germer M, Pawlita M, Krammer PH, Peter ME. Cytotoxicity-dependent APO-1 (Fas/CD95)-associated proteins form a death-inducing signaling complex (DISC) with the receptor. EMBO J 1995;14(22):5579-88.
  62. 62. Muzio M, Chinnaiyan AM, Kischkel FC, O'Rourke K, Shevchenko A, Ni J, et al. FLICE, a novel FADD-homologous ICE/CED-3-like protease, is recruited to the CD95 (Fas/APO-1) death-inducing signaling complex. Cell 1996;85(6):817-27.
  63. 63. Peter ME, Krammer PH. The CD95(APO-1/Fas) DISC and beyond. Cell Death Differ 2003;10(1):26-35.
  64. 64. Jänicke RU, Sprengart ML, Wati MR, Porter AG. Caspase-3 is required for DNA fragmentation and morphological changes associated with apoptosis. J Biol Chem 1998;273(16):9357-60.
  65. 65. Opferman JT, Korsmeyer SJ. Apoptosis in the development and maintenance of the immune system. Nat Immunol 2003;4(5):410-5.
  66. 66. Newmeyer DD, Farschon DM, Reed JC: Cell-free apoptosis in Xenopus egg extracts: inhibition by Bcl-2 and requirement for an organelle fraction enriched in mitochondria. Cell 1994;79:353-64.
  67. 67. Kluck RM, Bossy-Wetzel E, Green DR, Newmeyer DD. The release of cytochrome c from mitochondria: a primary site for Bcl-2 regulation of apoptosis. Science 1997;275:1132-6.
  68. 68. Susin SA, Lorenzo HK, Zamzami N, Marzo I, Snow BE, Brothers GM, et al. Molecular characterization of mitochondrial apoptosis-inducing factor. Nature 1999;397:441-6.
  69. 69. Kroemer G, Reed JC: Mitochondrial control of cell death. Nat Med 2000;6:513-9.
  70. 70. Parrish J, Li L, Klotz K, Ledwich D, Wang X, Xue D. Mitochondrial endonuclease G is important for apoptosis in Caenorhabditis elegans. Nature 2001;412:90-94.
  71. 71. Nakagawa T, Zhu H, Morishima N, Li E, Xu J, Yankner BA, Yuan J. Caspase-12 mediates endoplasmic-reticulum-specific apoptosis and cytotoxicity by amyloid-β. Nature 2000;403(6765):98-103.
  72. 72. Fischer H, Koenig U, Eckhart L, Tschachler E. Human caspase 12 has acquired deleterious mutations. Biochem Biophys Res Commun 2002;293:722-6.
  73. 73. Momoi T. Caspases involved in ER stress-mediated cell death. J Chem Neuroanat 2004;28(1-2):101-5
  74. 74. Guimaraes CA, Linden R. Programmed cell deaths. Apoptosis and alternative deathstyles. Eur J Biochem 2004;271:1638-50.
  75. 75. Arnoult D, Tatischeff I, Estaquier J, Girard M, Sureau F, Tissier JP, et al. On the evolutionary conservation of the cell death pathway: mitochondrial release of an apoptosis-inducing factor during Dictyostelium discoideum cell death. Mol Biol Cell 2001;12(10):3016-30.
  76. 76. Debrabant A, Lee N, Bertholet S, Duncan R and Nakhasi HL. Programmed cell death in trypanosomatids and other unicellular organisms International J Parasitol 2003, 33:257-267.
  77. 77. Carmen JC, Sinai AP. Suicide prevention: disruption of apoptotic pathways by protozoan parasites. Mol Microbiol 2007;64(4):904-16
  78. 78. Bruchhaus I, Roeder T, Rennenberg A, Heussler VT: Protozoan parasites: programmed cell death as a mechanism of parasitism. Trends Parasitol 2007, 23:376-383.
  79. 79. Kaczanowski S, Sajid M, Reece SE. Evolution of apoptosis-like programmed cell death in unicellular protozoan parasites. Parasit Vectors 2011;4:44.
  80. 80. Deponte M. Programmed cell death in protists. Biochim Biophys Acta 2008;1783(7):1396-405.
  81. 81. De Duve C. The lysosome. Sci Am 1963;208:64-72.
  82. 82. Duszenko M, Ginger ML, Brennand A, Gualdrón-López M, Colombo MI, Coombs GH, et al. Autophagy in protists. Autophagy 2011;7(2):127-58.
  83. 83. Reggiori F, Klionsky DJ. Autophagosomes: biogenesis from scratch? Curr Opin Cell Biol 2005;17(4):415-22.
  84. 84. Yoshimori A, Takasawa R, Tanuma S. A novel method for evaluation and screening of caspase inhibitory peptides by the amino acid positional fitness score. BMC Pharmacol 2004;4:7.
  85. 85. Yorimitsu T, Klionsky DJ. Autophagy: molecular machinery for self-eating. Cell Death Differ 2005;12 Suppl 2:1542-52.
  86. 86. He C, Klionsky DJ. Regulation mechanisms and signaling pathways of autophagy. Annu Rev Genet 2009;43:67-93.
  87. 87. Kirkegaard K, Taylor MP, Jackson WT. Cellular autophagy: surrender, avoidance and subversion by microorganisms. Nat Rev Microbiol 2004;2(4):301-14.
  88. 88. Mizushima N. The pleiotropic role of autophagy: from protein metabolism to bactericide. Cell Death Differ 2005;12 Suppl 2:1535-41.
  89. 89. Ogawa M, Sasakawa C. Shigella and autophagy. Autophagy 2006;2(3):171-4.
  90. 90. Swanson MS. Autophagy: eating for good health. J Immunol 2006;177(8):4945-51.
  91. 91. Colombo MI. Autophagy: a pathogen driven process. IUBMB Life 2007;59(4-5):238-42.
  92. 92. Alvarez VE, Kosec G, Sant Anna C, Turk V, Cazzulo JJ, Turk B. Blocking autophagy to prevent parasite differentiation: a possible new strategy for fighting parasitic infections? Autophagy 2008;4(3):361-3.
  93. 93. Levine B, Yuan J. Autophagy in cell death: an innocent convict? J Clin Invest 2005;115(10):2679-88.
  94. 94. Maiuri MC, Zalckvar E, Kimchi A, Kroemer G: Self-eating and self-killing: crosstalk between autophagy and apoptosis. Nature Reviews 2007;8:741-752.
  95. 95. Klionsky DJ, Cregg JM, Dunn WA Jr, Emr SD, Sakai Y, Sandoval IV, et al. A unified nomenclature for yeast autophagy-related genes. Dev Cell 2003;5:539-45.
  96. 96. Klionsky DJ, Ohsumi Y. Vacuolar import of proteins and organelles from the cytoplasm. Annu Rev Cell Dev Biol 1999;15:1-32.
  97. 97. Shintani T, Klionsky DJ. Autophagy in health and disease: a double-edged sword. Science 2004;306(5698):990-5.
  98. 98. Klionsky DJ. Autophagy. Curr Biol 2005;15(8):R282-3.
  99. 99. Brennand A, Gualdrón-López M, Coppens I, Rigden DJ, Ginger ML, Michels PA. Autophagy in parasitic protists: unique features and drug targets. Mol Biochem Parasitol 2011;177(2):83-99.
  100. 100. Bejarano E, Cuervo AM. Chaperone-mediated autophagy. Proc Am Thorac Soc 2010;7(1):29-39.
  101. 101. Klionsky DJ, Abdalla FC, Abeliovich H, Abraham RT, Acevedo-Arozena A, Adeli K, et al., Guidelines for the use and interpretation of assays for monitoring autophagy. Autophagy 2012;8(4):445-544.
  102. 102. Proskuryakov SY, Konoplyannikov AG, Gabai VL. Necrosis: a specific form of programmed cell death?. Exp Cell Res 2003;283(1):1-16.
  103. 103. Zong WX, Thompson CB. Necrotic death as a cell fate. Genes Dev 2006;20(1):1-15.
  104. 104. Holler N, Zaru R, Micheau O, Thome M, Attinger A, Valitutti S, et al. Fas triggers an alternative, caspase-8-independent cell death pathway using the kinase RIP as effector molecule. Nat Immunol 2000;1(6):489-95.
  105. 105. Christofferson DE, Yuan J. Necroptosis as an alternative form of programmed cell death. Curr Opin Cell Biol 2010;22(2):263-8.
  106. 106. Vanden Berghe T, Linkermann A, Jouan-Lanhouet S, Walczak H, Vandenabeele P. Regulated necrosis: the expanding network of non-apoptotic cell death pathways. Nat Rev Mol Cell Biol 2014;15(2):135-47.
  107. 107. Li P, Allen H, Banerjee S, Franklin S, Herzog L, Johnston C et al. Mice deficient in IL-1 β-converting enzyme are defective in production of mature IL-1 beta and resistant to endotoxic shock. Cell 1995;80:401-11.
  108. 108. Fink SL, Cookson BT. Apoptosis, pyroptosis, and necrosis: mechanistic description of dead and dying eukaryotic cells. Infect Immun 2005;73(4):1907-16.
  109. 109. Brinkmann V, Reichard U, Goosmann C, Fauler B, Uhlemann Y, Weiss DS, Weinrauch Y, Zychlinsky A. Neutrophil extracellular traps kill bacteria. Science 2004;303(5663):1532-5.
  110. 110. Yipp BG, Kubes P. NETosis: how vital is it? Blood 2013;122(16):2784-94.
  111. 111. von Köckritz-Blickwede M, Goldmann O, Thulin P, Heinemann K, Norrby-Teglund A, Rohde M, Medina E. Phagocytosis-independent antimicrobial activity of mast cells by means of extracellular trap formation. Blood 2008;111(6):3070-80.
  112. 112. Aulik NA, Hellenbrand KM, Czuprynski CJ. Mannheimia haemolytica and its leukotoxin cause macrophage extracellular trap formation by bovine macrophages. Infect Immun 2012; 80(5):1923-33.
  113. 113. Schorn C, Janko C, Latzko M, Chaurio R, Schett G, Herrmann M. Monosodium urate crystals induce extracellular DNA traps in neutrophils, eosinophils, and basophils but not in mononuclear cells. Front Immunol 2012;3:277.
  114. 114. Ueki S, Melo RC, Ghiran I, Spencer LA, Dvorak AM, Weller PF. Eosinophil extracellular DNA trap cell death mediates lytic release of free secretion-competent eosinophil granules in humans. Blood 2013;121(11):2074-83.
  115. 115. Dixon SJ, Lemberg KM, Lamprecht MR, Skouta R, Zaitsev EM, Gleason CE, et al. Ferroptosis: an iron-dependent form of nonapoptotic cell death. Cell 2012;149(5):1060-72.
  116. 116. Jamison JM, Gilloteaux J, Taper HS, Calderon PB, Summers JL. Autoschizis: a novel cell death. Biochem Pharmacol 2002;63(10):1773-83.
  117. 117. Gilloteaux J, Jamison JM, Arnold D, Neal DR, Summers JL. Morphology and DNA degeneration during autoschizic cell death in bladder carcinoma T24 cells induced by ascorbate and menadione treatment. Anat Rec A Discov Mol Cell Evol Biol 2006;288(1):58-83.
  118. 118. Nguewa PA, Fuertes MA, Valladares B, Alonso C, Pérez JM. Programmed cell death in trypanosomatids: a way to maximize their biological fitness? Trends Parasitol 2004;20(8):375-80.
  119. 119. Ameisen JC, Idziorek T, Billaut-Mulot O, Loyens M, Tissier JP, Potentier A, Ouaissi A. Apoptosis in a unicellular eukaryote (Trypanosoma cruzi): implications for the evolutionary origin and role of programmed cell death in the control of cell proliferation, differentiation and survival. Cell Death Differ 1995;2(4):285-300.
  120. 120. Barcinski MA, DosReis GA. Apoptosis in parasites and parasite-induced apoptosis in the host immune system: a new approach to parasitic diseases. Braz J Med Biol Res 1999;32(4):395-401.
  121. 121. De Souza EM, Araújo-Jorge TC, Bailly C, Lansiaux A, Batista MM, Oliveira GM, Soeiro MN. Host and parasite apoptosis following Trypanosoma cruzi infection in in vitro and in vivo models. Cell Tissue Res. 2003;314(2):223-35.
  122. 122. De Souza EM, Menna-Barreto R, Araújo-Jorge TC, Kumar A, Hu Q, Boykin DW, Soeiro MN. Antiparasitic activity of aromatic diamidines is related to apoptosis-like death in Trypanosoma cruzi. Parasitology 2006;133(Pt 1):75-9.
  123. 123. Ouaissi A. Apoptosis-like death in trypanosomatids: search for putative pathways and genes involved. Kinetoplastid Biol Dis 2003;2(1):5.
  124. 124. Piacenza L, Peluffo G, Radi R. L-arginine-dependent suppression of apoptosis in Trypanosoma cruzi: contribution of the nitric oxide and polyamine pathways. Proc Natl Acad Sci USA 2001;98(13):7301-6
  125. 125. Piacenza L, Irigoín F, Alvarez MN, Peluffo G, Taylor MC, Kelly JM, Wilkinson SR, Radi R. Mitochondrial superoxide radicals mediate programmed cell death in Trypanosoma cruzi: cytoprotective action of mitochondrial iron superoxide dismutase overexpression. Biochem J 2007;403(2):323-34.
  126. 126. Jimenez V, Paredes R, Sosa MA, Galanti N. Natural programmed cell death in Trypanosoma cruzi epimastigotes maintained in axenic cultures. J Cell Biochem 2008;105(3):688-98.
  127. 127. Benitez D, Pezaroglo H, Martínez V, Casanova G, Cabrera G, Galanti N, González M, Cerecetto H. Study of Trypanosoma cruzi epimastigote cell death by NMR-visible mobile lipid analysis. Parasitology 2012;139(4):506-15.
  128. 128. Veiga-Santos P, Reignault LC, Huber K, Bracher F, De Souza W, De Carvalho TM. Inhibition of NAD+-dependent histone deacetylases (sirtuins) causes growth arrest and activates both apoptosis and autophagy in the pathogenic protozoan Trypanosoma cruzi. Parasitology 2014;141(6):814-25.
  129. 129. Welburn SC, Murphy NB. Prohibitin and RACK homologues are up-regulated in trypanosomes induced to undergo apoptosis and in naturally occurring terminally differentiated forms. Cell Death Differ 1998;5(7):615-22.
  130. 130. Duszenko M, Figarella K, Macleod ET, Welburn SC. Death of a trypanosome: a selfish altruism. Trends Parasitol 2006;22(11):536-42.
  131. 131. Lüder CG, Campos-Salinas J, Gonzalez-Rey E, van Zandbergen G. Impact of protozoan cell death on parasite-host interactions and pathogenesis. Parasit Vectors 2010;3:116.
  132. 132. Gannavaram S, Vedvyas C, Debrabant A. Conservation of the pro-apoptotic nuclease activity of endonuclease G in unicellular trypanosomatid parasites. J Cell Sci 2008;121(Pt 1):99-109.
  133. 133. Gannavaram S, Debrabant A. Programmed cell death in Leishmania: biochemical evidence and role in parasite infectivity. Front Cell Infect Microbiol 2012;2:95
  134. 134. Das M, Mukherjee SB, Shaha C. Hydrogen peroxide induces apoptosis-like death in Leishmania donovani promastigotes. J Cell Sci 2001;114(Pt 13):2461-9.
  135. 135. Moreira ME, Del Portillo HA, Milder RV, Balanco JM, Barcinski MA. Heat shock induction of apoptosis in promastigotes of the unicellular organism Leishmania (Leishmania) amazonensis. J Cell Physiol 1996;167(2):305-13.
  136. 136. Lee N, Bertholet S, Debrabant A, Muller J, Duncan R, Nakhasi HL. Programmed cell death in the unicellular protozoan parasite Leishmania. Cell Death Differ 2002;9(1):53-64.
  137. 137. Mukherjee SB, Das M, Sudhandiran G, Shaha C. Increase in cytosolic Ca2+ levels through the activation of non-selective cation channels induced by oxidative stress causes mitochondrial depolarization leading to apoptosis-like death in Leishmania donovani promastigotes. J Biol Chem 2002;277(27):24717-27.
  138. 138. Marquis JF, Drolet M, Olivier M. Consequence of Hoechst 33342-mediated Leishmania DNA topoisomerase-I inhibition on parasite replication. Parasitology 2003;126(Pt 1):21-30
  139. 139. Sen N, Das BB, Ganguly A, Mukherjee T, Tripathi G, Bandyopadhyay S, et al. Camptothecin induced mitochondrial dysfunction leading to programmed cell death in unicellular hemoflagellate Leishmania donovani. Cell Death Differ 2004;11(8):924-36.
  140. 140. Zangger H, Mottram JC, Fasel N. Cell death in Leishmania induced by stress and differentiation: programmed cell death or necrosis? Cell Death Differ 2002;9(10):1126-39.
  141. 141. Welburn SC, Macleod E, Figarella K, Duzensko M. Programmed cell death in African trypanosomes. Parasitology 2006;132Suppl:S7-S18.
  142. 142. Murphy NB, Welburn SC. Programmed cell death in procyclic Trypanosoma brucei rhodesiense is associated with differential expression of mRNAs. Cell Death Differ 1997;4(5):365-70.
  143. 143. Rosenkranz V, Wink M. Alkaloids induce programmed cell death in bloodstream forms of trypanosomes (Trypanosoma brucei brucei). Molecules 2008;13(10):2462-73.
  144. 144. Sen N, Banerjee B, Das BB, Ganguly A, Sen T, Pramanik S, et al. Apoptosis is induced in leishmanial cells by a novel protein kinase inhibitor withaferin A and is facilitated by apoptotic topoisomerase I-DNA complex. Cell Death Differ 2007;14(2):358-67.
  145. 145. Helms MJ, Ambit A, Appleton P, Tetley L, Coombs GH, Mottram JC. Bloodstream form Trypanosoma brucei depend upon multiple metacaspases associated with RAB11-positive endosomes. J Cell Sci 2006;119(Pt 6):1105-17.
  146. 146. González IJ, Desponds C, Schaff C, Mottram JC, Fasel N. Leishmania major metacaspase can replace yeast metacaspase in programmed cell death and has arginine-specific cysteine peptidase activity. Int J Parasitol 2007;37(2):161-72.
  147. 147. Lee N, Gannavaram S, Selvapandiyan A, Debrabant A. Characterization of metacaspases with trypsin-like activity and their putative role in programmed cell death in the protozoan parasite Leishmania. Eukaryot Cell 2007;6(10):1745-57.
  148. 148. Meslin B, Zalila H, Fasel N, Picot S, Bienvenu AL. Are protozoan metacaspases potential parasite killers? Parasit Vectors 2011;4:26.
  149. 149. Proto WR, Coombs GH, Mottram JC. Cell death in parasitic protozoa: regulated or incidental? Nat Rev Microbiol 2013;11(1):58-66.
  150. 150. Kosec G, Alvarez VE, Agüero F, Sánchez D, Dolinar M, Turk B, et al. Metacaspases of Trypanosoma cruzi: possible candidates for programmed cell death mediators. Mol Biochem Parasitol 2006;145(1):18-28.
  151. 151. Moss CX, Westrop GD, Juliano L, Coombs GH, Mottram JC. Metacaspase 2 of Trypanosoma brucei is a calcium-dependent cysteine peptidase active without processing. FEBS Lett 2007;581(29):5635-9.
  152. 152. González IJ. Metacaspases and their role in the life cycle of human protozoan parasites. Biomedica 2009;29(3):485-93.
  153. 153. Laverrière M, Cazzulo JJ, Alvarez VE. Antagonic activities of Trypanosoma cruzi metacaspases affect the balance between cell proliferation, death and differentiation. Cell Death Differ 2012;19(8):1358-69.
  154. 154. Ridgley EL, Xiong ZH, Ruben L. Reactive oxygen species activate a Ca2+-dependent cell death pathway in the unicellular organism Trypanosoma brucei brucei. Biochem J 1999;340(Pt 1):33-40.
  155. 155. Figarella K, Rawer M, Uzcategui NL, Kubata BK, Lauber K, Madeo F, et al. Prostaglandin D2 induces programmed cell death in Trypanosoma brucei bloodstream form. Cell Death Differ 2005;12(4):335-46.
  156. 156. Figarella K, Uzcategui NL, Beck A, Schoenfeld C, Kubata BK, Lang F, Duszenko M. Prostaglandin-induced programmed cell death in Trypanosoma brucei involves oxidative stress. Cell Death Differ 2006;13(10):1802-14.
  157. 157. Le Bras M, Clément MV, Pervaiz S, Brenner C. Reactive oxygen species and the mitochondrial signaling pathway of cell death. Histol Histopathol 2005;20(1):205-19.
  158. 158. BoseDasgupta S, Das BB, Sengupta S, Ganguly A, Roy A, Dey S, et al. The caspase-independent algorithm of programmed cell death in Leishmania induced by baicalein: the role of LdEndoG, LdFEN-1 and LdTatD as a DNA 'degradesome'. Cell Death Differ 2008;15(10):1629-40.
  159. 159. Alvarez VE, Niemirowicz GT, Cazzulo JJ. The peptidases of Trypanosoma cruzi: digestive enzymes, virulence factors, and mediators of autophagy and programmed cell death. Biochim Biophys Acta 2012;1824(1):195-206
  160. 160. Alzate JF, Arias AA, Moreno-Mateos D, Alvarez-Barrientos A, Jiménez-Ruiz A. Mitochondrial superoxide mediates heat-induced apoptotic-like death in Leishmania infantum. Mol Biochem Parasitol 2007;152(2):192-202.
  161. 161. Billaut-Mulot O, Fernandez-Gomez R, Loyens M, Ouaissi A. Trypanosoma cruzi elongation factor 1-alpha: nuclear localization in parasites undergoing apoptosis. Gene 1996;174(1):19-26.
  162. 162. McLuskey K, Moss CX, Mottram JC. Purification, characterization, and crystallization of Trypanosoma metacaspases. Methods Mol Biol 2014;1133:203-21.
  163. 163. Mottram JC, Helms MJ, Coombs GH, Sajid M. Clan CD cysteine peptidases of parasitic protozoa. Trends Parasitol 2003;19(4):182-7.
  164. 164. Szallies A, Kubata BK, Duszenko M. A metacaspase of Trypanosoma brucei causes loss of respiration competence and clonal death in the yeast Saccharomyces cerevisiae. FEBS Lett 2002;517(1-3):144-50.
  165. 165. Ambit A, Fasel N, Coombs GH, Mottram JC. An essential role for the Leishmania major metacaspase in cell cycle progression. Cell Death Differ 2008;15(1):113-22.
  166. 166. Rico E, Alzate JF, Arias AA, Moreno D, Clos J, Gago F, Moreno I, Domínguez M, Jiménez-Ruiz A. Leishmania infantum expresses a mitochondrial nuclease homologous to EndoG that migrates to the nucleus in response to an apoptotic stimulus. Mol Biochem Parasitol 2009;163(1):28-38.
  167. 167. Herman M, Gillies S, Michels PA, Rigden DJ. Autophagy and related processes in trypanosomatids: insights from genomic and bioinformatic analyses. Autophagy 2006;2(2):107-18.
  168. 168. Menna-Barreto RF, Salomão K, Dantas AP, Santa-Rita RM, Soares MJ, Barbosa HS, De Castro SL. Different cell death pathways induced by drugs in Trypanosoma cruzi: an ultrastructural study. Micron. 2009a Feb;40(2):157-68.
  169. 169. McGwire BS, Kulkarni MM. Interactions of antimicrobial peptides with Leishmania and trypanosomes and their functional role in host parasitism. Exp Parasitol 2010;126(3):397-405.
  170. 170. Vickerman K, Tetley L. Recent ultrastructural studies on trypanosomes. Ann Soc Belg Med Trop 1977;57(4-5):441-57.
  171. 171. Merkel P, Beck A, Muhammad K, Ali SA, Schönfeld C, Voelter W, Duszenko M. Spermine isolated and identified as the major trypanocidal compound from the snake venom of Eristocophis macmahoni causes autophagy in Trypanosoma brucei. Toxicon 2007;50(4):457-69.
  172. 172. Rodrigues JC, Seabra SH, De Souza W. 2006. Apoptosis-like death in parasitic protozoa. Braz J Morphol Sci 20056;23(2):87-98.
  173. 173. Santa-Rita RM, Lira R, Barbosa HS, Urbina JA, de Castro SL. Anti-proliferative synergy of lysophospholipid analogues and ketoconazole against Trypanosoma cruzi (Kinetoplastida: Trypanosomatidae): cellular and ultrastructural analysis. J Antimicrob Chemother 2005;55:780-4.
  174. 174. Bera A, Singh S, Nagaraj R, Vaidya T. Induction of autophagic cell death in Leishmania donovani by antimicrobial peptides. Mol Biochem Parasitol 2003;127(1):23-35.
  175. 175. Delgado M, Anderson P, Garcia-Salcedo JA, Caro M, Gonzalez-Rey E. Neuropeptides kill African trypanosomes by targeting intracellular compartments and inducing autophagic-like cell death. Cell Death Differ 2009;16(3):406-16.
  176. 176. Uzcategui NL, Carmona-Gutierrez D, Denninger V, Schoenfeld C, Lang F, Figarella K, Duszenko M. Antiproliferative effect of dihydroxyacetone on Trypanosoma brucei bloodstream forms: cell cycle progression, subcellular alterations, and cell death. Antimicrob Agents Chemother 2007;51:3960-8.
  177. 177. Menna-Barreto RF, Corrêa JR, Pinto AV, Soares MJ, de Castro SL. Mitochondrial disruption and DNA fragmentation in Trypanosoma cruzi induced by naphthoimidazoles synthesized from β-lapachone. Parasitol Res 2007;101(4):895-905.
  178. 178. Menna-Barreto RF, Corrêa JR, Cascabulho CM, Fernandes MC, Pinto AV, Soares MJ, De Castro SL. Naphthoimidazoles promote different death phenotypes in Trypanosoma cruzi. Parasitology 2009b;136(5):499-510.
  179. 179. Klionsky DJ. What can we learn from trypanosomes? Autophagy 2006;2(2):63-4.
  180. 180. Kiel JA. Autophagy in unicellular eukaryotes. Philos Trans R Soc Lond B Biol Sci 2010;365:819-30.
  181. 181. Rigden DJ, Herman M, Gillies S, Michels PA. Implications of a genomic search for autophagy-related genes in trypanosomatids. Biochem Soc Trans 2005;33(Pt 5):972-4.
  182. 182. Besteiro S, Williams RA, Morrison LS, Coombs GH, Mottram JC. Endosome sorting and autophagy are essential for differentiation and virulence of Leishmania major. J Biol Chem 2006;281(16):11384-96.
  183. 183. Williams RA, Woods KL, Juliano L, Mottram JC, Coombs GH. Characterization of unusual families of ATG8-like proteins and ATG12 in the protozoan parasite Leishmania major. Autophagy 2009;5(2):159-72.
  184. 184. Koopmann R, Muhammad K, Perbandt M, Betzel C, Duszenko M. Trypanosoma brucei ATG8: structural insights into autophagic-like mechanisms in protozoa. Autophagy 2009;5(8):1085-91.
  185. 185. Li FJ, Shen Q, Wang C, Sun Y, Yuan AY, He CY. A role of autophagy in Trypanosoma brucei cell death. Cell Microbiol 2012;14(8):1242-56.
  186. 186. Barquilla A, Crespo JL, Navarro M. Rapamycin inhibits trypanosome cell growth by preventing TOR complex 2 formation. Proc Natl Acad Sci USA 2008;105(38):14579-84.
  187. 187. Saldivia M, Barquilla A, Bart JM, Diaz-González R, Hall MN, Navarro M. Target of rapamycin (TOR) kinase in Trypanosoma brucei: an extended family. Biochem Soc Trans 2013;41(4):934-8.
  188. 188. Denninger V, Koopmann R, Muhammad K, Barth T, Bassarak B, Schönfeld C, Kilunga BK, Duszenko M. Kinetoplastida: model organisms for simple autophagic pathways? Methods Enzymol 2008;451:373-408.
  189. 189. Besteiro S, Williams RA, Coombs GH, Mottram JC. Protein turnover and differentiation in Leishmania. Int J Parasitol 2007;37(10):1063-75.
  190. 190. Williams RA, Smith TK, Cull B, Mottram JC, Coombs GH. ATG5 is essential for ATG8-dependent autophagy and mitochondrial homeostasis in Leishmania major. PLoS Pathog. 2012;8(5):e1002695.
  191. 191. Cazzulo JJ, Stoka V, Turk V. Cruzipain, the major cysteine proteinase from the protozoan parasite Trypanosoma cruzi. Biol Chem 1997;378(1):1-10.
  192. 192. Brennand A, Rico E, Michels PA. Autophagy in trypanosomatids. Cells 2012;1(3):346-71.
  193. 193. Williams RA, Tetley L, Mottram JC, Coombs GH. Cysteine peptidases CPA and CPB are vital for autophagy and differentiation in Leishmania mexicana. Mol Microbiol 2006;61(3):655-74.
  194. 194. Besteiro S, Coombs GH, Mottram JC. The SNARE protein family of Leishmania major. BMC Genomics 2006b;7:250.
  195. 195. Herman M, Pérez-Morga D, Schtickzelle N, Michels PA. Turnover of glycosomes during life-cycle differentiation of Trypanosoma brucei. Autophagy 2008;4(3):294-308.
  196. 196. Li FJ, He CY. Acidocalcisome is required for autophagy in Trypanosoma brucei. Autophagy 2014;10(11):1978-88.
  197. 197. Williams RA, Mottram JC, Coombs GH. Distinct roles in autophagy and importance in infectivity of the two ATG4 cysteine peptidases of Leishmania major. J Biol Chem 2013;288(5):3678-90.
  198. 198. Barquilla A, Navarro M. Trypanosome TOR as a major regulator of cell growth and autophagy. Autophagy 2009;5(2):256-8.
  199. 199. Madeira-da-Silva L, Beverley SM. Expansion of the target of rapamycin (TOR) kinase family and function in Leishmania shows that TOR3 is required for acidocalcisome biogenesis and animal infectivity. Proc Natl Acad Sci USA 2010;107(26):11965-70.
  200. 200. Bao Y, Weiss LM, Braunstein VL, Huang H. Role of protein kinase A in Trypanosoma cruzi. Infect Immun 2008;76(10):4757-63.
  201. 201. Hall BS, Gabernet-Castello C, Voak A, Goulding D, Natesan SK, Field MC. TbVps34, the trypanosome orthologue of Vps34, is required for Golgi complex segregation. J Biol Chem 2006;281(37):27600-12.
  202. 202. Freitas-Junior LH, Chatelain E, Kim HA, Siqueira-Neto JL. Visceral leishmaniasis treatment: What do we have, what do we need and how to deliver it? Int J Parasitol Drugs Drug Resist 2012;2:11-9.
  203. 203. Mäser P, Wittlin S, Rottmann M, Wenzler T, Kaiser M, Brun R. Antiparasitic agents: new drugs on the horizon. Curr Opin Pharmacol 2012;12(5):562-6.
  204. 204. Bustamante JM, Tarleton RL. Potential new clinical therapies for Chagas disease. Expert Rev Clin Pharmacol 2014;7(3):317-25.
  205. 205. Patterson S, Wyllie S. Nitro drugs for the treatment of trypanosomatid diseases: past, present, and future prospects. Trends Parasitol 2014;30(6):289-98.
  206. 206. Ambrosio AR, De Messias-Reason IJ. Leishmania (Viannia) braziliensis: interaction of mannose-binding lectin with surface glycoconjugates and complement activation. An antibody-independent defence mechanism. Parasite Immunol 2005;27(9):333-40.
  207. 207. Lambris JD, Ricklin D, Geisbrecht BV. Complement evasion by human pathogens. Nat Rev Microbiol 2008;6(2):132-42.
  208. 208. Cestari I, Ramirez MI. Inefficient complement system clearance of Trypanosoma cruzi metacyclic trypomastigotes enables resistant strains to invade eukaryotic cells. PLoS One 2010;5(3):e9721.
  209. 209. Evans-Osses I, de Messias-Reason I, Ramirez MI. The emerging role of complement lectin pathway in trypanosomatids: molecular bases in activation, genetic deficiencies, susceptibility to infection, and complement system-based therapeutics. ScientificWorldJournal 2013;2013:675898.
  210. 210. Rudenko G. African trypanosomes: the genome and adaptations for immune evasion. Essays Biochem 2011;51:47-62.
  211. 211. Fernandes MC, Da Silva EN, Pinto AV, De Castro SL, Menna-Barreto RF. A novel triazolic naphthofuranquinone induces autophagy in reservosomes and impairment of mitosis in Trypanosoma cruzi. Parasitology 2012;139(1):26-36.
  212. 212. Desoti VC, Lazarin-Bidóia D, Sudatti DB, Pereira RC, Alonso A, Ueda-Nakamura T, Dias Filho BP, Nakamura CV, Silva SO. Trypanocidal action of (-)-elatol involves an oxidative stress triggered by mitochondria dysfunction. Mar Drugs 2012;10(8):1631-46.
  213. 213. Paris C, Loiseau PM, Bories C, Bréard J. Miltefosine induces apoptosis-like death in Leishmania donovani promastigotes. Antimicrob Agents Chemother 2004;48(3):852-9.
  214. 214. Mamani-Matsuda M, Rambert J, Malvy D, Lejoly-Boisseau H, Daulouède S, Thiolat D, Coves S, Courtois P, Vincendeau P, Mossalayi MD. Quercetin induces apoptosis of Trypanosoma brucei gambiense and decreases the proinflammatory response of human macrophages. Antimicrob Agents Chemother 2004;48(3):924-9.
  215. 215. Deolindo P, Teixeira-Ferreira AS, Melo EJ, Arnholdt AC, Souza Wd, Alves EW, DaMatta RA. Programmed cell death in Trypanosoma cruzi induced by Bothrops jararaca venom. Mem Inst Oswaldo Cruz 2005;100(1):33-8.
  216. 216. Freire-de-Lima CG, Nascimento DO, Soares MB, Bozza PT, Castro-Faria-Neto HC, de Mello FG, DosReis GA, Lopes MF. Uptake of apoptotic cells drives the growth of a pathogenic trypanosome in macrophages. Nature 2000;403(6766):199-203.
  217. 217. Ribeiro-Gomes FL, Silva MT, Dosreis GA. Neutrophils, apoptosis and phagocytic clearance: an innate sequence of cellular responses regulating intramacrophagic parasite infections. Parasitology 2006;132 Suppl:S61-8.
  218. 218. Balanco JMF, Moreira ME, Bonomo A, Bozza PT, Amarante-Mendes G, Pirmez C, Barcinski MA. Apoptotic mimicry by an obligate intracellular parasite downregulates macrophage microbicidal activity. Curr Biol 2001;11(23):1870-3.
  219. 219. El-Hani CN, Borges VM, Wanderley JL, Barcinski MA. Apoptosis and apoptotic mimicry in Leishmania: an evolutionary perspective. Front Cell Infect Microbiol 2012;2:96.
  220. 220. Van Zandbergen G, Klinger M, Mueller A, Dannenberg S, Gebert A, Solbach W, Laskay T. Cutting edge: neutrophil granulocyte serves as a vector for Leishmania entry into macrophages. J Immunol 2004;173(11):6521-5.
  221. 221. Pinheiro RO, Nunes MP, Pinheiro CS, D'Avila H, Bozza PT, Takiya CM, Côrte-Real S, Freire-de-Lima CG, DosReis GA. Induction of autophagy correlates with increased parasite load of Leishmania amazonensis in BALB/c but not C57BL/6 macrophages. Microbes Infect 2009;11(2):181-90.
  222. 222. Mitroulis I, Kourtzelis I, Papadopoulos VP, Mimidis K, Speletas M, Ritis K. In vivo induction of the autophagic machinery in human bone marrow cells during Leishmania donovani complex infection. Parasitol Intern 2009;58(4):475-7.
  223. 223. Cyrino LT, Araújo AP, Joazeiro PP, Vicente CP, Giorgio S. In vivo and in vitro Leishmania amazonensis infection induces autophagy in macrophages. Tissue Cell 2012;44(6):401-8.
  224. 224. Romano PS, Arboit MA, Vázquez CL, Colombo MI. The autophagic pathway is a key component in the lysosomal dependent entry of Trypanosoma cruzi into the host cell. Autophagy 2009;5(1):6-18.
  225. 225. Romano PS, Cueto JA, Casassa AF, Vanrell MC, Gottlieb RA, Colombo MI. Molecular and cellular mechanisms involved in the Trypanosoma cruzi/host cell interplay. IUBMB Life 2012;64(5):387-96.
  226. 226. Martins RM, Alves RM, Macedo S, Yoshida N. Starvation and rapamycin differentially regulate host cell lysosome exocytosis and invasion by Trypanosoma cruzi metacyclic forms. Cellular Microbiology 2011;13(7): 943-54.
  227. 227. Maeda FY, Alves RM, Cortez C, Lima FM, Yoshida N. Characterization of the infective properties of a new genetic group of Trypanosoma cruzi associated with bats. Acta Tropica 2011;120(3):231-7.
  228. 228. Duque TLA, Souto XM, Andrade-Neto VV, Ennes-Vidal V, Menna-Barreto RFS. Autophagic Balance Between Mammals and Protozoa: A Molecular, Biochemical and Morphological Review of Apicomplexa and Trypanosomatidae Infections. In: Bailly Y, editor. Autophagy - A Double-Edged Sword - Cell Survival or Death? Rijeka, Croatia: Intech;2013. pp. 289-317.
  229. 229. Ameisen JC. The origin of programmed cell death. Science 1996;272(5266):1278-9.
  230. 230. Menna-Barreto RF, De Castro SL. The double-edged sword in pathogenic trypanosomatids: the pivotal role of mitochondria in oxidative stress and bioenergetics. Biomed Res Int 2014;2014:614014.
  231. 231. Welburn SC, Barcinski MA, Williams GT. Programmed cell death in trypanosomatids. Parasitol Today 1997;13(1):22-6.
  232. 232. Baehrecke EH. Autophagy: dual roles in life and death? Nat Rev Mol Cell Biol 2005;6(6):505-10.
  233. 233. Jesenberger V, Jentsch S. Deadly encounter: ubiquitin meets apoptosis. Nature Rev Mol Cell Biol 2002;3:122-21.
  234. 234. Codogno P, Meijer AJ. Autophagy and signaling: their role in cell survival and cell death. Cell Death Differ 2005;12 Suppl 2:1509-18.

Written By

Rubem Figueiredo Sadok Menna-Barreto and Solange Lisboa de Castro

Submitted: 24 November 2014 Reviewed: 10 July 2015 Published: 16 December 2015