Open access peer-reviewed chapter

The Biology of Malaria Gametocytes

Written By

Che Julius Ngwa, Thiago F. de A. Rosa and Gabriele Pradel

Submitted: 18 November 2015 Reviewed: 01 September 2016 Published: 30 November 2016

DOI: 10.5772/65464

From the Edited Volume

Current Topics in Malaria

Edited by Alfonso J. Rodriguez-Morales

Chapter metrics overview

3,875 Chapter Downloads

View Full Metrics

Abstract

Gametocytes are sexual precursor cells of the malaria parasite that mediate the transmission of the parasite from its mammalian host to the Anopheles mosquito. Unlike the asexual blood stages, which are responsible for the clinical outcome of malaria, gametocytes cause no clinical manifestations. However, they are very crucial for the transmission of the disease thus represent key targets for transmission-blocking interventions. Despite their essential role in malaria transmission, only in the last decade gametocytes became a hot topic of research and their biology is not well understood. This chapter provides a detailed review on the biology of the human malaria gametocytes with emphasis on aspects such as gametocyte commitment, gametocyte maturation (gametocytogenesis), gametocyte metabolism and gametogenesis. Proper understanding of these processes will deepen our knowledge on the gametocyte biology and therefore open up more avenues for the development of malaria transmission-blocking intervention strategies.

Keywords

  • gametocyte
  • gametocytogenesis
  • transmission
  • gametogenesis
  • sexual reproduction
  • mosquito

1. Introduction

Gametocytes are specialized sexual precursor cells that mediate the transmission of the malaria parasite from the mammalian host to the mosquito. Once these cells have gained maturity, they are picked up by an Anopheles mosquito during a blood meal. In the midgut of the mosquito, they become activated and differentiate into male and female gametes. The male gamete then fertilizes the female gamete resulting in the formation of a zygote. The zygote further develops into a motile ookinete, which penetrates the gut epithelium and subsequently develops into an oocyst. The oocyst then matures releasing sporozoites, which migrate to the salivary gland of the mosquito. The parasite is transmitted to another mammalian host through an infected mosquito bite [1]. Gametocytes therefore provide a link in the transmission of malaria from the mammalian host to the mosquito, thereby making them prime targets for transmission-blocking intervention strategies. These strategies rely on either vaccines or drugs, which target the gametocyte or the mosquito midgut stages to block malaria transmission from the human to the mosquito, thereby preventing the spread of the disease.

Despite their essential role in maintaining the parasite life cycle, research on gametocytes has been very much left out of the limelight. Several reasons could be accounted for this: first, gametocytes cause no clinical manifestation in their host as opposed to the asexual blood stages that are responsible for the clinical features of malaria. Second, it has been a great challenge to culture gametocytes because of the cost and time-consuming cultivation procedure, and finally, it has been a hurdle to obtain pure gametocytes for molecular and metabolic studies as well as drug screening purposes. However, in the last decade, the rapid emergence of drug-resistant parasites and mosquitoes, together with the absence of a malaria vaccine, has rekindled the interest of researchers on alternative measures aimed at interrupting the parasite life cycle by targeting gametocytes. This interest has led to the improvement of gametocyte cultivation and purification methods, which have enabled a greater understanding of the gametocyte biology.

In this chapter, we focus on key aspects of the gametocyte biology such as gametocyte commitment, maturation (gametocytogenesis), metabolism and gametogenesis. This will deepen our understanding on the biology of gametocytes and thereby open up more avenues for the development of malaria transmission-blocking intervention strategies.

Advertisement

2. Gametocyte commitment

The asexual blood stage parasites produce the clinical form of malaria due to destruction of the erythrocytes, while replicating in the mammalian host. However, these asexual blood stages cannot be transmitted to the mosquito to continue the life cycle of the parasite. For this reason, a subset of the asexual blood stages is able to switch to the sexual pathway, resulting in the formation of the sexual precursor cells, the gametocytes. The process of switching from the asexual blood stage to the gametocyte is referred to as gametocyte commitment.

Commitment to gametocyte development is believed to take place at some point before schizogony, in which each individual schizont produces a progeny of merozoites that develop into either sexual forms or asexual blood stage parasites, which continue the erythrocytic cycle [2, 3] (Figure 1). In addition, merozoites released from a single sexually committed schizont can either become male or female gametocytes and the characteristic female-biased sex ratio observed in the malaria parasite is due to the production of a higher percentage of committed female schizonts than their male counterpart [3, 4].

Until recently, gametocyte commitment was mainly assigned to stress factors such as high parasitaemia, anaemia, host immune response or drug treatment, with an overwhelming amount of studies to justify these facts. A study by Chaubey and colleagues suggested that the malaria parasite reacts to endoplasmic reticulum stress by switching to gametocytes [5]. The treatment of mice with phenylhydrazine, which induces hyper-reticulocyte formation, stimulates gametocyte formation when infected with P. chabaudi [6]. Lymphocyte and serum from Gambian children have also been shown to induce P. falciparum gametocyte production [7]. Another study showed that P. falciparum cultures grown in reticulocyte-rich blood from patients with sickle cell anaemia result in enhanced gametocyte formation as compared to normal blood [8]. Haemolysis of infected erythrocytes also triggers the formation of P. falciparum gametocytes [9]. The treatment of the malaria parasite with drugs such as steroid hormones [10], fansidar [11] and chloroquine or sulfadoxine-pyrimethamine [12] also enhances gametocyte production.

Figure 1.

Proposed mechanism of gametocyte commitment. During the erythrocyte cycle, epigenetic and transcriptional regulation work hand in hand to control gametocyte commitment. Two populations of blood stage parasites exist. Non-committed parasites continue the erythrocytic replication of the parasite, and committed parasites are able to differentiate to gametocytes.

However, despite the knowledge that gametocyte commitment can be enhanced by stress factors, the mechanism by which stress induces this process is not well understood. Recent studies have also associated gametocyte commitment to the concerted action of epigenetics and transcriptional regulation.

2.1. Epigenetic mechanism of gametocyte commitment

Epigenetic mechanisms through histone modifications have been shown to regulate gene expression of important proteins in the malaria parasite [13, 14]. In P. falciparum, two master players in gene silencing, P. falciparum heterochromatin protein 1 (PfHP1) and P. falciparum histone deacetylase 2 (PfHda2) were demonstrated to be implicated in the epigenetic regulation of gametocyte commitment [15, 16].

Heterochromatin protein 1 (HP1) was initially described in Drosophila as the main suppressor of position-effect variegation and the major component of heterochromatin gene silencing [17]. Brancucci and colleagues [16] reported that in P. falciparum, PfHP1 binds to histone H3-trimethylated residues on lysine 9 (H3K9me3) to maintain heterochromatin gene silencing, which results in the suppression of gametocyte commitment and variegated expression of var genes. Var genes encode for a family of 60 parasite virulent proteins in P. falciparum named P. falciparum erythrocyte membrane protein 1 (PfEMP1). PfEMP1 is present on surface of the P. falciparum-infected erythrocyte of which only one is expressed at a time and this phenomenon is associated with antigenic variation. Conditional depletion of PfHP1 results in the hyper-induction of viable gametocytes, thereby demonstrating a role of PfHP1 in suppressing gametocyte commitment. Depletion of PfHP1 also results in derepression of the ap2-g gene, which encodes for the AP2-G transcription factor, the only member of the ApiAP2 family shown to associate with PfHP1 [18]. AP2-G was demonstrated to be important in gametocyte commitment in both P. falciparum and the rodent malaria parasite P. berghei by acting as a transcriptional switch that controls gametocyte differentiation through activating the transcription of early gametocyte genes [19, 20].

Histone deacetylases are histone-modifying enzymes that promote transcriptional silencing by removing acetyl groups from histones. The removal of the acetyl groups facilitates histone methylation and HP1 binding [18, 21, 22], thereby resulting in reduced accessibility for transcriptional factors and heterochromatin formation. A study by Coleman and colleagues [15], which complements the results of Brancucci et al. [16], showed an essential role of PfHda2 in the global silencing of var genes as well as controlling gametocyte commitment in P. falciparum. Knockdown of PfHda2 results in up-regulation of gene expression of genes associated with gametocytogenesis leading to an increase in gametocyte production. It also results in derepression of the ap2-g gene. The authors further reported that most of the genes dysregulated following PfHda2 knockdown were bound to PfHP1, and both proteins localize together with PfHda2-regulated genes, which were significantly enriched in H3K9me3, indicating an interaction between PfHda2 and PfHP1. The two studies therefore provide a link between PfHda2 and PfHP1 in controlling gametocyte commitment.

2.2. Transcriptional regulation of gametocyte commitment

AP2-G is a member of the apicomplexan AP2 DNA-binding protein family, characterized by the presence of an Apetala2/ethylene response factor (AP2/ERF) DNA-binding domain [23]. In the malaria parasite, other AP2 proteins have been assigned to the stage-specific transition during parasite development, for example, the development of midgut, sporozoite and liver stages [2427].

Two studies in P. falciparum and P. berghei have shown that AP2-G triggers a transcription cascade that initiates gametocyte commitment [14, 19, 20]. To determine initially, if AP2-G is essential for gametocyte commitment and production, both studies sequenced and analysed the ap2-g gene from specific gametocyte-less producing strains for single nucleotide polymorphisms, insertions or mutations. Their results showed nonsense or missense mutations in the ap2-g gene providing an explanation for the inability of the strains to commit and produce gametocytes. Their initial findings were further confirmed by targeted disruption of the ap2-g locus in both P. falciparum and P. berghei, which resulted in the loss of sexual commitment and gametocyte production. The disruption of the ap2-g locus also resulted in the down-regulation of many genes highly expressed in early gametocytes such as Pfs16 and Pfg25/27. The DNA motif involved in AP2-G-binding was identified as GTAC and shown to be located upstream of many genes associated with gametocytogenesis [19, 20, 28]. Because the ap2-g gene also possesses the GTAC domain on its promoter region, this suggests that it regulates itself through a feedback mechanism. The AP2-G DNA-binding motif was shown to interact with gametocyte-associated gene promoters in a motif-dependent manner. Mutations of the motif in the promoter region result in a blockage of the gene expression of the gametocyte-associated genes, thereby indicating that AP2-G induces the expression of genes containing the AP2-G DNA motif by binding to it at the promoter region [19].

During the malaria life cycle, therefore, two populations of the malaria parasite exist. Non-committed parasites, which continue the erythrocytic cycle, and committed parasites that form gametocytes. In the case of non-committed parasites, PHP1 binds to H3K9me3 residues thereby maintaining the ap2-g locus at a heterochromatic state resulting in silencing of the ap2-g expression. This silencing leads to the inability to initiate transcription of genes important for gametocyte commitment; therefore, the parasite continues with erythrocytic replication. PHda2 is also involved in the silencing of the ap2-g gene expression in non-committed parasites, probably by the removal of acetylated histone residues allowing for their methylation leading to the binding of PHP1. In committed parasites, stress factors and other unknown factors cause the removal of the histone H3 trimethylation mark, thereby preventing PHP1 from binding to the histone and therefore maintaining the ap2-g locus at an active state, resulting in AP2-G expression. The expression of AP2-G promotes the transcription of early gametocyte genes, which leads to gametocyte commitment and formation (Figure 1).

Advertisement

3. Gametocyte maturation

3.1. Gametocyte morphology

The morphology of gametocytes varies among the Plasmodium species. While mature gametocytes from P. ovale, P. malariae, P. vivax and P. knowlesi possess a round to oval shape, P. falciparum adopts a crescent or falciform shape when matured [2931].

P. falciparum is named from the crescent or falciform shape of its mature stage. During gametocyte development, the parasite undergoes five distinct morphological stages (I–V) (Figure 2). The first observations on the morphology of P. falciparum gametocytes were made in the 1950s, when the five development stages were defined [30]. Later studies using light and electron microscopic approaches have defined the shape and time of appearance [32, 33]. Studies have shown that stages I–IV are sequestered into the bone marrow, while stage V circulates in the peripheral blood vessels [3436]. Stage I gametocyte has a rounded shape, which is indistinguishable from the young asexual trophozoite. Stage IIa, on the other hand, can be distinguished from the trophozoite by its large round format and granular distribution of pigment in several food vacuoles. In stage IIb, the parasite starts to elongate and a D-shape can be observed. Stage III gametocytes adopt an oval shape with the red blood cell slightly distorted due to further elongated of the parasite. In stage IV, it is possible to distinguish between male and female gametocytes. Both genders have an elongate thin format with pointed tips. In male gametocytes, the pigment tends to be scattered while in female it is denser. The stage V gametocyte is the most distinguishable among all stages due to the crescent or falciform shape. While the male gametocyte is thicker and the cytoplasm appears pale blue following Giemsa staining, the female is more elongate and curved with a blue-stained cytoplasm. This stage is visible in the periphery circulation after 10–12 days of maturation [29].

Figure 2.

P. falciparum and P. vivax stage-specific protein expression during gametocyte development. The bar represents high-level expression and full line low-level expression. Only gene products were in included in this figure.

Among the other species of malaria, P. vivax is the most important because of its high impact on human health in South and Central America and South Asia [37]. The gametocytes of P. vivax are different from P. falciparum in their biology, dynamics and their formation, which occur within 2–3 days after asexual blood stage appearance in the circulation [38]. The male and female gametocytes of P. vivax possess a round shape as it is observed in P. malariae, P. ovale and P. knowlesi, but it is only in P. vivax that the parasite fills an enlarged red blood cell (Figure 2). In Giemsa-stained thin blood smears, female gametocytes have a blue-stained cytoplasm, and the chromatin is a purplish mass with a peripheral distribution, while male gametocytes possess a large and loose chromatin mass.

One remarkable morphological feature of gametocytes is the presence of the inner membrane complex (IMC), which underlies the plasmalemma and consists of flat membranous sacs called alveoli that are coupled to a supporting cytoskeletal network. The IMC was originally assigned to the invasive stages of Plasmodium as part of actin-myosin motor complex, and thus, the IMC is important for motility [39]. Recently, it was confirmed that gametocytes also possess an IMC [4042], which is thought to be required for the stability of the crescent shape. The outer alveolar membrane contains complexes of the gliding-associated proteins GAP40, GAP45 and GAP50, which facilitate binding of myosin to this membrane. The gametocyte alveoli are subtended by a subpellicular network of interwoven alveoli filaments linked to the alveoli via intramembranous particles [3941, 43]. Underneath the subpellicular network, an array of longitudinally oriented microtubules and F-actin fibrils are found that become reorganized in stage V gametocytes, leading to rounding-up of the gametocyte tips [44].

3.2. Gametocyte-specific proteins

Approximately 20% of all plasmodial genes are specifically expressed in the sexual stages [45, 46]. In P. falciparum gametocytes, the proteins are expressed at different levels in the five distinct developmental stages (Figure 2). Most of these proteins are associated with gametocyte development, integrity and preparation for fertilization. Furthermore, gender-related proteins are expressed in the later stages in order to support the gametes for fertilization in the mosquito midgut [1, 4749].

In the early stage I–II gametocytes, important proteins such as Pfs16, Pfg27 and the six LCCL-domain proteins are expressed [50]. Pfs16 is present in all gametocyte stages, but it is initially expressed in stage I. It is localized in the parasitophorous vacuole (PV) where it plays a role in gametocyte development [5153]. The main function of Pfs16 is unknown, but its disruption leads to a decrease in gametocyte production and an impairment in the ability of microgametocyte to exflagellate [54]. The protein Pfg27 is a phosphoprotein that starts to be expressed from stage II. It is known that Pfg27 forms a homodimer to act as single-stranded RNA-binding proteins in vitro and its gene deletion leads to the formation of abnormal gametocytes [55, 56]. The six LCCL-domain proteins (termed CCp in P. falciparum and LAP in P. berghei) possess adhesive properties and expression begins from stage II. In P. falciparum gametocytes, CCp proteins are localized within the PV, but after gamete emergence, the proteins are associated with the surface of gametes [47, 48, 57, 58].

Starting from stage IIb, well-known sexual-stage proteins such as the 6-cys proteins Pfs48/45 and Pfs230 and the osmiophilic body protein Pfg377 and Pfmdv-1 or sexual stage-specific cytoskeletal proteins such as α-tubII (tubulin II) and actin II are synthesized [50]. Similar to the six LCCL-domain proteins, the surface proteins Pfs48/45 and Pfs230 have adhesive properties and gene disruption of both proteins impair male gamete fertility [59, 60]. Pfg377 is known to be associated with osmiophilic bodies in female gametocyte of P. falciparum, and gene disruption demonstrates a fundamental role in the formation of osmiophilic bodies [61, 62]. The Pfmdv-1 protein, also known as Pfpeg3, is highly expressed in early gametocyte until stage V afterwards reaching a low level of expression. The function of Pfmdv-1 is still unknown although disruption of the Pfmdv-1 encoding gene leads to a 20-fold decrease in the production of male gametocytes [63]. The cytoskeletal proteins α-tubulin II and actin II are sexual stage-specific, and both are shown to be predominantly present in male gametocytes [6466]. While α-tubulin II has been described to incorporate to axonemes of emerging male gametes, the function of actin II is still unclear, but gene disruption of both proteins leads to an impairment of male gametogenesis [65, 66]. In the mature microgametocyte, the protein PfB0400w, also known as PfMR5, can be observed. Although its function is still unknown, the expression of PfB0400w in male stage V gametocyte indicates that it may play a role in preparing gametocytes for exflagellation [50, 67].

In the gametocytes of P. vivax, three sexual stage-specific proteins have hitherto been identified, i.e. the 6-cys proteins Pvs230, Pvs48/45 and Pvs47, all of which are orthologues of those present in P. falciparum (Figure 2). Pvs230 has been characterized as an antigen on the surface of gametocytes and gametes while Pvs48/45 and Pvs47 have been studied recently as potential candidates for transmission-blocking vaccines. Following direct membrane feeding assays, anti-Pvs48/45 and anti-Pvs47 antibodies significantly reduced oocyst numbers in the mosquito midgut [6870].

Advertisement

4. The gametocyte metabolism

During gametocytogenesis, P. falciparum parasites undergo intense metabolic processes including energy production as well as synthesis and degradation of various molecules such as lipids and proteins. These processes are essentials for the survival of the malaria parasite and ensure gametocyte stage development, drug resistance and transmission to the mosquito vector [30, 71, 72].

As vital molecules, lipids are responsible for many functions such as cell signalling, energy storage and the stability of membranes. Furthermore, during the various stages of the Plasmodium life cycle, lipids play a crucial role in the growth and proliferation of the parasite [73]. Among asexual blood stages and gametocytes, there are more than 300 lipids with known and unknown functions [74]. Although not reflected in all classes of lipids, the amount of lipids would more than double during gametocytogenesis [75]. The lipid composition of all developmental stages differs, including during gametocyte maturation. It is known that sphingolipids involved in ceramide metabolism are enriched during gametocyte maturation and the inhibition of the ceramide biosynthetic pathway might kill gametocytes. In addition, phosphatidylserine is also highly expressed during gametocytogenesis and together with ceramides leads to gametocyte induction [74]. Neutral lipids such as cholesteryl ester and diacylglycerol (DAG) have also been shown to be significantly increased during gametocyte maturation; however, more data to better understand the metabolism and function of these lipids during gametocytogenesis are required [75].

In gametocytes, no DNA-replication occurs [76] and the nucleic acid synthesis is exclusive for RNA production. The production of RNA ceases from around day 6 of gametocytogenesis, thus the RNA and protein synthesis are more important in the early than late gametocyte stages [77, 78]. One interesting feature of malaria parasites is the expression of small subunit ribosomal RNA (SSU rRNA) genes [79], seven of them are synthesized in P. falciparum and one (S-type, formally known as C-type) is exclusively expressed in gametocytes. Proteomics identified 931 different proteins (72.2% of the proteins identified in P. falciparum) in gametocytes and within these proteins include conserved, stage-specific, secreted and membrane-associated proteins [46]. It is also known that protein expression differs between gender, female gametocytes synthetize proteins required for mitochondria and ribosomes, while male gametocytes only need proteins for DNA synthesis, axoneme formation and exflagellation [29, 80].

In eukaryotes, the mitochondria have many vital roles such as ATP synthesis, calcium homeostasis, lipid metabolism and synthesis of ion-sulphate clusters. The malaria parasites only possess a single mitochondrion and during parasite development it is observed that the mitochondria of gametocytes are longer and possess more cristae when compare to the ones of the asexual blood stages. This suggests that ATP synthesis via the electron transport chain is the main mechanism of energy production in gametocytes [29, 33]. Indeed the sexual stages appear to have seven times more cytochrome b than the asexual blood stages [81]. The cytochrome bc1 complex feeds the electron transport chain with more protons, which creates a gradient that drives ATP synthesis [82]. The cytochrome bc1 is a prime target for the antimalarial drug atovaquone that inhibits ubiquinone leading to the inhibition of Plasmodium cytochrome bc1 activity and loss of mitochondrial function [8385]. Other drugs such as primaquine and the primaquine analogue bulaquine have been shown to exhibit gametocytocidal activities and primaquine is suggested to have the site of action in the mitochondria by causing its deformation. Although the mode of action is poorly understood, these drugs reduce gametocytemia in infection by P. falciparum and have an effect against the exo-erythrocytic stages of P. vivax [86].

Despite the fact that the mitochondria have been well characterized in other organisms, many mitochondrion-associated events in P. falciparum are still not well understood. Initially, P. falciparum blood stage parasites were thought to be glucose- and glycolysis-dependent for carbon metabolism, ATP production and survival [87]. However, recent reports have also shown the importance of the tricarboxylic acid (TCA) cycle for energy production, especially in gametocytes [88]. The gametocyte utilizes mitochondrial TCA to catabolize host glucose and glutamine. The carbon skeleton derived from these molecules enters in the TCA cycle via acetyl-CoA to generate proton flux and ATP via the ATP synthase. Due to the role of the TCA cycle in parasite development, inhibition of this metabolic pathway leads to defects in gametocytogenesis. In addition, many mitochondrial enzymes involved in metabolic pathways can be targets for drug development [88].

Advertisement

5. Gametocyte activation and gametogenesis

5.1. The signalling cascade of gametocyte activation

The mature stage V gametocytes, which circulate in the blood of the mammalian host, are picked up by the female Anopheles mosquito during a bite and enter the midgut together with the blood meal. During the process, the gametocytes receive environmental signals, which indicate the switch from the warm-blooded host to the insect vector. The perception of these signals results in the immediate activation of the gametocytes followed by the onset of gametogenesis. Signals inducing gamete formation include a decrease in temperature by approximately 5°C, which is mandatory for gametocyte activation, and the presence of the mosquito-derived molecule xanthurenic acid (XA), a by-product of eye pigment synthesis [8991].

While the plasmodial receptor or receptors important for signal perception are still unknown, the signalling pathway resulting in gametogenesis has partially been deciphered (Figure 3). It has previously been shown that male gametogenesis, during which the activated male microgametocyte forms eight motile flagellar microgametes (a process termed exflagellation), involves a fast increase in intracellular calcium and the second messenger cGMP [92, 93]. Two membrane-integrated guanylyl cyclases (GCα and GCβ) were identified in Plasmodium; however, only GCα appears to be important for cGMP synthesis in activated gametocytes [9497]. The guanylyl cyclase can be stimulated by addition of XA. It has to be clarified, though, if XA acts directly or indirectly on GCα.

While in eukaryotes, cGMP is synthesized by guanylyl cyclases, phosphodiesterases (PDEs), which hydrolyse cGMP to GMP, are in general required for fast down-regulation of the second messenger. In Plasmodium, four PDEs, termed PDEα-δ, were identified, and PDEδ is highly expressed in gametocytes [98]. Zaprinast, a PDE inhibitor, is able to stimulate gametogenesis in the absence of XA due to increased cGMP levels [99]. Disruption of the gene encoding for PDEδ, however, impaired egress of P. falciparum gametocytes, indicating that a tight regulation of cGMP levels is important for gametogenesis [97].

The increase in cGMP activates the cGMP-dependent protein kinase PKG and chemical inhibition of PKG impairs rounding-up of the activated gametocytes. Since rounding-up is a calcium-independent step and cannot be inhibited by calcium chelators (see below), PKG acts prior to calcium increase in the activated gametocytes [97]. Gametocyte activation further leads to the generation of the second messenger molecules DAG and inositol-(1,4,5)-trisphosphate (IP3) from phosphatidylinositol-(4,5)-bisphosphate (PIP2) via the plasmodial phospholipase PI-PLC activity (Figure 3), as shown in P. berghei. IP3 synthesis results in a rapid mobilization of calcium from the endoplasmic reticulum [100, 101]. PI-PLC activity itself appears to be regulated by calcium, since it can be impaired by calcium ionophores, pointing to a calcium-regulated feedback mechanism. The increase in cytoplasmic calcium levels is further dependent on PKG activity and quantitative phosphoproteomics demonstrated that PKG is able to phosphorylate enzymes involved in the inositol phospholipid metabolism such as phosphatidylinositol-4-kinase (PI4K) and phosphoatidlyinositol-4-phosphate-5-kinase (PIP5K), which are involved in the synthesis of PIP2 from phosphatidyl-1D-myo-inositol (PI) via phosphatidylinositol-4-phosphate (PI4P) [102].

Figure 3.

The signalling pathway involved in gametogenesis. Ca2+, calcium ion; CDC20, cell division cycle protein 20; CDPK, calcium-dependent protein kinase; cGMP, cyclic guanosine monophosphate; CITH, homolog of worm CAR-I and fly Trailer Hitch; DAG, diacylglycerol; DNA, deoxyribonucleic acid; DOZI, development of zygote inhibited; E, erythrocyte; EM, erythrocyte membrane; ER endoplasmic reticulum; FACT-L, facilitates chromatin transcription protein L; GC, guanylyl cyclase; 5’-GMP, guanosine monophosphate; GTP, guanosine triphosphate; IP3, inositol triphosphate; Map-2, mitogen-activated protein kinase 2; mRNA, messenger ribonucleic acid; PDE, phosphodiesterase; PI, phosphatidyl-1D-myo-inositol; PI4K, phosphatidylinositol-4-kinase; PI4P, phosphatidylinosito-4-phosphate; PIP2, phosphatidylinositol-4,5-bisphosphate; PIP5K, phosphatidylinositol-5-kinase; PKG, cGMP-dependent protein kinase; PPM1, metallo-dependent protein phosphatase 1; Puf2, pumilio/FBF (fem-3 binding factor) family protein 2; PV, parasitophorous vacuole; PVM, parasitophorous vacuole membrane; R, receptor; T, temperature; XA, xanthurenic acid.

In activated P. berghei microgametocytes, the increase in intracellular calcium was shown to activate the calcium-dependent protein kinases CDPK1 and CDPK4 and parasites lacking CDPK4 are not able to undergo DNA synthesis [103] (Figure 3). Further linked to this process is plasmodial FACT-L (facilitates chromatin transcription) [104]. Downstream of DNA replication two more proteins become activated, the mitogen-activated protein kinase MAP2 and the cell division cycle protein CDC20. Loss of MAP2 or CDC20 interferes with chromatin condensation, axoneme motility and cytokinesis [105107]. Furthermore, depletion of the metallo-dependent protein phosphatase PPM1 or the calcium-dependent phosphatase calcineurin as well as knock-down of CDPK1 result in impaired exflagellation [108110]. While these combined studies pinpoint some key regulators of gametogenesis, the detailed pathway leading to gametocyte-to-gamete conversion is not known yet.

In female macrogametocytes, activation results in a sudden onset of protein synthesis, which is due to the CDPK1-regulated release of translational repression of messenger RNAs (mRNAs) [108]. In P. berghei, a total of 731 mRNAs are associated with two proteins of regulatory ribonucleoprotein complexes, i.e. the RNA helicase DOZI (development of zygote inhibited) and the Sm-like factor CITH (homolog of worm CAR-I and fly Trailer Hitch), which function in translational repression of these transcripts (Figure 3). These include the mRNAs encoding for P25, GAP45 and GAP50 [111113]. For P. falciparum, translational repression of mRNAs was shown to involve the Pumilio/FBF (Puf) family RNA-binding protein Puf2 that is required for controlling a number of transcripts in gametocytes including those encoding for P25 and P28 [114]. The repressed mRNAs mostly encode for components of the later ookinete stage.

Figure 4.

Development of malaria parasites in the mosquito midgut.

5.2. Gametocyte egress and gamete formation

Gametocyte activation results in three morphological changes in the parasites, i.e. rounding-up, erythrocyte egress and gamete formation, and these changes occur with approximately 15 min following perception of the environmental signals (Figure 4). Gametocytes exit the erythrocyte via an inside-out mode, during which the parasitophorous vacuolar membrane (PVM) ruptures prior to the opening of the erythrocyte membrane (EM). Rupture of the PVM occurs at multiple sites less than 2 min following signal perception and is accompanied by the rounding-up of the parasite. Both events are calcium-independent [91, 99]. It is suggested that the PVM rupture is linked to the discharge of the osmiophilic body content into the PV, since an accumulation of osmiophilic bodies can be observed underneath the rupture sites [91, 115].

In a second, a calcium-dependent step, another type of vesicle discharges its content into the cytoplasm of the host erythrocyte, which includes the plasmodial pore-forming perforin PPLP2. This perforin is able to permeate the EM, resulting in the release the erythrocyte cytoplasm, eventually leaving the forming gamete enclosed by the EM [115117]. While EM permeabilization occurs approximately 6 min post-activation, its rupture takes place 9 min later. Here, the EM opens at a single pore and releases the fertile gamete [91, 117].

The IMC disintegrates shortly after onset of gametogenesis, eventually resulting in gametes solely confined by the plasmalemma. It was shown that the transmembrane protein GAP50, which in gametocytes, is located in the outer membrane of the alveoli, relocates to the gamete plasmalemma during gametogenesis. During relocation, the N-terminal part of GAP50 becomes extracellularly exposed and once the gametes have freed from the EM binds complement factor H from the blood meal, a step crucial to protect the newly formed gametes from lysis by human complement factors [42]. Coevally with the IMC degradation long membranous tubules, termed tunneling nanotubes (TNTs), form, which protrude from the gamete surface. TNTs represent long-distance cell-to-cell connections, which were proposed to mediate intercellular contacts between gametes as a pre-requisite for mating [118]. While the question of the origin of the membranous tubules remains unanswered, one might stipulate that after IMC disintegration the alveoli membranes are used for TNT formation.

Upon activation, the male microgametocyte replicates its genome three times and mitotically produces eight flagellar microgametes [119] (Figure 4). Flagellum formation requires the axonemal assembly from basal bodies, which involves the centriole/basal body protein SAS-6 [120], while axoneme motility is regulated by the conserved armadillo-repeat protein PF16 [121]. Exflagellating microgametes typically adhere avidly to neighbouring erythrocytes and are hidden within such rosettes. During exflagellation, the microgamete detaches from the residual body and is freely motile in search of a female macrogamete. Once adhering to a macrogamete, fertilization begins by the fusion of the two gamete plasma membranes, and the axoneme and attached male nucleus enter the female cytoplasm. While the exact proteins involved in initial binding of male and female gametes are yet unknown, gamete fusion is mediated by the microgamete-specific protein GCS1 (generative cell specific 1), also termed HAP2 [122, 123]. Over the next 3 hours, meiosis occurs and the zygote becomes tetraploid [119]. The transformation of the zygote into an ookinete is completed at 19–36 hours post-blood meal, and the ookinete then rapidly exits the midgut lumen to settle down at the basal site of the midgut epithelium and to differentiate into an oocyst [124127]. Parasite tetraploidy persists until sporozoite budding in the oocyst restores the haploid state [119].

Advertisement

6. Conclusion

Although our knowledge of the biology of gametocytes is not as advanced as that of the asexual blood stages, the great interest by researchers in recent years has considerably deepened our understanding of the gametocyte biology. Advances in gametocyte cultivation, isolation and purification techniques as well as the impressive amount of data gained from genome-wide analyses, proteomics and microarray studies have led to the identification of some gametocyte stage-specific antigens as well as sex-specific proteins, which may be exploited as targets for malaria transmission-blocking intervention strategies. In addition, the recent discoveries on the molecular mechanism of gametocyte commitment have greatly improved our knowledge on how sexual differentiation is initiated and regulated in the malaria parasite. However, several questions are still to be answered regarding the biology of gametocytes. The mechanism by which stress factors are perceived to enhance gametocyte commitment is not well understood. Some components of the signalling cascade of gametocyte activation are yet unknown such as the receptor or receptors responsible for signal perception following gametocyte activation by XA. How gametocyte sequestration in the bone marrow takes place as well as the molecular mechanism underlying this phenomenon has to been deciphered. Also, a proper understanding of the mechanism by which the gametocyte rapidly adapts once in the mosquito vector and the basis of fertilization is needed. Furthermore, not much is known about the metabolism of gametocytes. Most of the metabolic pathways have not been described. A proper understanding of these aspects in the biology of gametocytes would greatly improve strategies aimed to target gametocytes.

References

  1. 1. Kuehn A, Pradel G. The coming-out of malaria gametocytes. J Biomed Biotechnol [Internet]. 2010;2010:976827. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2810480&tool=pmcentrez&rendertype=abstract
  2. 2. Bruce MC, Alano P, Duthie S, Carter R. Commitment of the malaria parasite Plasmodium falciparum to sexual and asexual development. Parasitology [Internet]. 1990 Apr [cited 2015 Dec 29];100(Pt 2):191–200. Available from: http://www.ncbi.nlm.nih.gov/pubmed/2189114
  3. 3. Smith TG, Lourenço P, Carter R, Walliker D, Ranford-Cartwright LC. Commitment to sexual differentiation in the human malaria parasite, Plasmodium falciparum. Parasitology. 2000;121:127–33.
  4. 4. Silvestrini F, Alano P, Williams JL. Commitment to the production of male and female gametocytes in the human malaria parasite Plasmodium falciparum. Parasitology [Internet]. 2000 Nov [cited 2016 Jan 19];121(Pt 5):465–71. Available from: http://www.ncbi.nlm.nih.gov/pubmed/11128797
  5. 5. Chaubey S, Grover M, Tatu U. Endoplasmic reticulum stress triggers gametocytogenesis in the malaria parasite. J Biol Chem. 2014;289(24):16662–74.
  6. 6. Gautret P, Miltgen F, Gantier JC, Chabaud AG, Landau I. Enhanced gametocyte formation by Plasmodium chabaudi in immature erythrocytes: pattern of production, sequestration, and infectivity to mosquitoes. J Parasitol [Internet]. 1996 Dec [cited 2016 Apr 5];82(6):900–6. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8973397
  7. 7. Smalley ME, Brown J. Plasmodium falciparum gametocytogenesis stimulated by lymphocytes and serum from infected Gambian children. Trans R Soc Trop Med Hyg [Internet]. 1981 Jan [cited 2015 Nov 25];75(2):316–7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/7029805
  8. 8. Trager W, Gill GS. Enhanced gametocyte formation in young erythrocytes by Plasmodium falciparum in vitro. J Protozool [Internet]. 1992 Jan [cited 2016 Apr 5];39(3):429–32. Available from: http://www.ncbi.nlm.nih.gov/pubmed/1640389
  9. 9. Schneweis S, Maier WA, Seitz HM. Haemolysis of infected erythrocytes – a trigger for formation of Plasmodium falciparum gametocytes? Parasitol Res [Internet]. 1991 Jan [cited 2016 Feb 4];77(5):458–60. Available from: http://www.ncbi.nlm.nih.gov/pubmed/1891456
  10. 10. Lingnau A, Margos G, Maier WA, Seitz HM. The effects of hormones on the gametocytogenesis of Plasmodium falciparum in vitro. Appl Parasitol [Internet]. 1993 Sep [cited 2016 Feb 4];34(3):153–60. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8220571
  11. 11. Puta C, Manyando C. Enhanced gametocyte production in Fansidar-treated Plasmodium falciparum malaria patients: implications for malaria transmission control programmes. Trop Med Int Health [Internet]. 1997 Mar [cited 2016 Feb 4];2(3):227–9. Available from: http://www.ncbi.nlm.nih.gov/pubmed/9491100
  12. 12. Talman AM, Paul REL, Sokhna CS, Domarle O, Ariey F, Trape J-F, et al. Influence of chemotherapy on the Plasmodium gametocyte sex ratio of mice and humans. Am J Trop Med Hyg [Internet]. 2004 Dec [cited 2016 Feb 4];71(6):739–44. Available from: http://www.ncbi.nlm.nih.gov/pubmed/15642963
  13. 13. Duffy MF, Selvarajah SA, Josling GA, Petter M. Epigenetic regulation of the Plasmodium falciparum genome. Brief Funct Genomics. 2014;13(3):203–16.
  14. 14. Josling GA, Llinás M. Sexual development in Plasmodium parasites: knowing when it’s time to commit. Nat Rev Microbiol [Internet]. 2015;13(9):573–87. Available from: http://www.ncbi.nlm.nih.gov/pubmed/26272409
  15. 15. Coleman BI, Skillman KM, Jiang RHY, Childs LM, Altenhofen LM, Ganter M, et al. A Plasmodium falciparum histone deacetylase regulates antigenic variation and gametocyte conversion. Cell Host Microbe [Internet]. 2014;16(2):177–86. Available from: http://www.sciencedirect.com/science/article/pii/S1931312814002625
  16. 16. Brancucci NMB, Bertschi NL, Zhu L, Niederwieser I, Chin WH, Wampfler R, et al. Heterochromatin protein 1 secures survival and transmission of malaria parasites. Cell Host and Microbe. 2014;16(2):165–76. Available from: http://ac.els-cdn.com/S1931312814002583/1-s2.0-S1931312814002583-main.pdf?_tid=858c21fc-6d44-11e5-9c59-00000aab0f26&acdnat=1444257857_991a4915b9e0c23b300a77dbed1a4edb
  17. 17. Eissenberg JC, James TC, Foster-Hartnett DM, Hartnett T, Ngan V, Elgin SC. Mutation in a heterochromatin-specific chromosomal protein is associated with suppression of position-effect variegation in Drosophila melanogaster. Proc Natl Acad Sci U S A [Internet]. 1990 Dec [cited 2016 Apr 5];87(24):9923–7. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=55286&tool=pmcentrez&rendertype=abstract
  18. 18. Flueck C, Bartfai R, Volz J, Niederwieser I, Salcedo-Amaya AM, Alako BTF, et al. Plasmodium falciparum heterochromatin protein 1 marks genomic loci linked to phenotypic variation of exported virulence factors. PLoS Pathog. 2009;5(9) e1000569.
  19. 19. Kafsack BFC, Rovira-Graells N, Clark TG, Bancells C, Crowley VM, Campino SG, et al. A transcriptional switch underlies commitment to sexual development in malaria parasites. Nature [Internet]. 2014;507(7491):248–52. Available from: http://www.ncbi.nlm.nih.gov/pubmed/24572369
  20. 20. Sinha A, Hughes KR, Modrzynska KK, Otto TD, Pfander C, Dickens NJ, et al. A cascade of DNA-binding proteins for sexual commitment and development in Plasmodium. Nature [Internet]. 2014;507(7491):253–7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/24572359
  21. 21. Lopez-Rubio J-J, Mancio-Silva L, Scherf A. Genome-wide analysis of heterochromatin associates clonally variant gene regulation with perinuclear repressive centers in malaria parasites. Cell Host Microbe [Internet]. 2009 Feb 19 [cited 2016 Apr 5];5(2):179–90. Available from: http://www.ncbi.nlm.nih.gov/pubmed/19218088
  22. 22. Pérez-Toledo K, Rojas-Meza AP, Mancio-Silva L, Hernández-Cuevas NA, Delgadillo DM, Vargas M, et al. Plasmodium falciparum heterochromatin protein 1 binds to tri-methylated histone 3 lysine 9 and is linked to mutually exclusive expression of var genes. Nucleic Acids Res [Internet]. 2009 May [cited 2016 Apr 5];37(8):2596–606. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2677873&tool=pmcentrez&rendertype=abstract
  23. 23. Balaji S, Babu MM, Iyer LM, Aravind L. Discovery of the principal specific transcription factors of Apicomplexa and their implication for the evolution of the AP2-integrase DNA binding domains. Nucleic Acids Res [Internet]. 2005 Jan [cited 2016 Mar 16];33(13):3994–4006. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1178005&tool=pmcentrez&rendertype=abstract
  24. 24. Yuda M, Iwanaga S, Shigenobu S, Mair GR, Janse CJ, Waters AP, et al. Identification of a transcription factor in the mosquito-invasive stage of malaria parasites. Mol Microbiol [Internet]. 2009 Mar [cited 2016 Feb 15];71(6):1402–14. Available from: http://www.ncbi.nlm.nih.gov/pubmed/19220746
  25. 25. Yuda M, Iwanaga S, Shigenobu S, Kato T, Kaneko I. Transcription factor AP2-Sp and its target genes in malarial sporozoites. Mol Microbiol [Internet]. 2010 Feb [cited 2016 Apr 5];75(4):854–63. Available from: http://www.ncbi.nlm.nih.gov/pubmed/20025671
  26. 26. Painter HJ, Campbell TL, Llinás M. The Apicomplexan AP2 family: integral factors regulating Plasmodium development. Mol Biochem Parasitol [Internet]. 2011 Mar [cited 2016 Apr 5];176(1):1–7. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3026892&tool=pmcentrez&rendertype=abstract
  27. 27. Iwanaga S, Kaneko I, Kato T, Yuda M. Identification of an AP2-family protein that is critical for malaria liver stage development. PLoS One [Internet]. 2012 Jan [cited 2016 Feb 15];7(11):e47557. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3492389&tool=pmcentrez&rendertype=abstract
  28. 28. Campbell TL, De Silva EK, Olszewski KL, Elemento O, Llinás M. Identification and genome-wide prediction of DNA binding specificities for the ApiAP2 family of regulators from the malaria parasite. PLoS Pathog [Internet]. 2010 Jan [cited 2016 Apr 5];6(10):e1001165. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2965767&tool=pmcentrez&rendertype=abstract
  29. 29. Sinden RE. Gametocytogenesis of Plasmodium falciparum in vitro: an electron microscopic study. Parasitology [Internet]. 1982 Feb [cited 2016 Apr 6];84(1):1–11. Available from: http://www.ncbi.nlm.nih.gov/pubmed/7038594
  30. 30. Day KP, Hayward RE, Dyer M. The biology of Plasmodium falciparum transmission stages. Parasitology [Internet]. 1998;116(Supplement S1):S95–109. Available from: href="http://dx.doi.org/10.1017/S0031182000084985
  31. 31. Dixon MWA, Thompson J, Gardiner DL, Trenholme KR. Sex in Plasmodium: a sign of commitment. Trends Parasitol. 2008;24(4):168–75.
  32. 32. Hawking F, Wilson ME, Gammage K. Evidence for cyclic development and short-lived maturity in the gametocytes of Plasmodium falciparum. Trans R Soc Trop Med Hyg [Internet]. 1971 Jan [cited 2016 Feb 19];65(5):549–59. Available from: http://www.ncbi.nlm.nih.gov/pubmed/5003557
  33. 33. Langreth SG, Jensen JB, Reese RT, Trager W. Fine structure of human malaria in vitro. J Protozool [Internet]. 1978 Nov [cited 2016 Apr 6];25(4):443–52. Available from: http://www.ncbi.nlm.nih.gov/pubmed/105129
  34. 34. Tibúrcio M, Silvestrini F, Bertuccini L, Sander AF, Turner L, Lavstsen T, et al. Early gametocytes of the malaria parasite Plasmodium falciparum specifically remodel the adhesive properties of infected erythrocyte surface. Cell Microbiol [Internet]. 2013 Apr [cited 2016 Apr 6];15(4):647–59. Available from: http://www.ncbi.nlm.nih.gov/pubmed/23114006
  35. 35. Aguilar R, Magallon-Tejada A, Achtman AH, Moraleda C, Joice R, Cisteró P, et al. Molecular evidence for the localization of Plasmodium falciparum immature gametocytes in bone marrow. Blood [Internet]. 2014 Feb 13 [cited 2016 Apr 6];123(7):959–66. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4067503&tool=pmcentrez&rendertype=abstract
  36. 36. Joice R, Nilsson SK, Montgomery J, Dankwa S, Egan E, Morahan B, et al. Plasmodium falciparum transmission stages accumulate in the human bone marrow. Sci Transl Med [Internet]. 2014 Jul 9 [cited 2016 Feb 19];6(244):244re5. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4175394&tool=pmcentrez&rendertype=abstract
  37. 37. Guerra CA, Howes RE, Patil AP, Gething PW, Van Boeckel TP, Temperley WH, et al. The international limits and population at risk of Plasmodium vivax transmission in 2009. PLoS Negl Trop Dis [Internet]. 2010 Jan [cited 2016 Feb 26];4(8):e774. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2914753&tool=pmcentrez&rendertype=abstract
  38. 38. Bousema T, Drakeley C. Epidemiology and infectivity of Plasmodium falciparum and Plasmodium vivax gametocytes in relation to malaria control and elimination. Clin Microbiol Rev [Internet]. 2011;24(2):377–410. Available from: http://cmr.asm.org/cgi/doi/10.1128/CMR.00051-10
  39. 39. Harding CR, Meissner M. The inner membrane complex through development of Toxoplasma gondii and Plasmodium. Cell Microbiol [Internet]. 2014 May [cited 2016 Feb 19];16(5):632–41. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4286798&tool=pmcentrez&rendertype=abstract
  40. 40. Dearnley MK, Yeoman JA, Hanssen E, Kenny S, Turnbull L, Whitchurch CB, et al. Origin, composition, organization and function of the inner membrane complex of Plasmodium falciparum gametocytes. J Cell Sci [Internet]. 2012 Apr 15 [cited 2016 Apr 7];125(Pt 8):2053–63. Available from: http://www.ncbi.nlm.nih.gov/pubmed/22328505
  41. 41. Kono M, Herrmann S, Loughran NB, Cabrera A, Engelberg K, Lehmann C, et al. Evolution and architecture of the inner membrane complex in asexual and sexual stages of the malaria parasite. Mol Biol Evol [Internet]. 2012;29(9):2113–32. Available from: http://mbe.oxfordjournals.org/cgi/doi/10.1093/molbev/mss081
  42. 42. Simon N, Lasonder E, Scheuermayer M, Kuehn A, Tews S, Fischer R, et al. Malaria parasites co-opt human factor h to prevent complement-mediated lysis in the mosquito midgut. Cell Host Microbe [Internet]. 2013;13(1):29–41. Available from: http://dx.doi.org/10.1016/j.chom.2012.11.013
  43. 43. Boucher LE, Bosch J. The apicomplexan glideosome and adhesins – structures and function. J Struct Biol [Internet]. 2015 May [cited 2015 Jun 25];190(2):93–114. Available from: http://www.ncbi.nlm.nih.gov/pubmed/25764948
  44. 44. Hliscs M, Millet C, Dixon MW, Siden-Kiamos I, McMillan P, Tilley L. Organization and function of an actin cytoskeleton in Plasmodium falciparum gametocytes. Cell Microbiol [Internet]. 2015 Feb [cited 2016 Apr 7];17(2):207–25. Available from: http://www.ncbi.nlm.nih.gov/pubmed/25224798
  45. 45. Florens L, Washburn MP, Raine JD, Anthony RM, Grainger M, Haynes JD, et al. A proteomic view of the Plasmodium falciparum life cycle. Nature [Internet]. 2002;419(6906):520–6. Available from: http://www.ncbi.nlm.nih.gov/pubmed/12368866
  46. 46. Lasonder E, Ishihama Y, Andersen JS, Vermunt AMW, Pain A, Sauerwein RW, et al. Analysis of the Plasmodium falciparum proteome by high-accuracy mass spectrometry. Nature [Internet]. 2002 Oct 3 [cited 2016 Apr 7];419(6906):537–42. Available from: http://www.ncbi.nlm.nih.gov/pubmed/12368870
  47. 47. Scholz SM, Simon N, Lavazec C, Dude MA, Templeton TJ, Pradel G. PfCCp proteins of Plasmodium falciparum: gametocyte-specific expression and role in complement-mediated inhibition of exflagellation. Int J Parasitol. 2008;38(3–4):327–40.
  48. 48. Simon N, Scholz SM, Moreira CK, Templeton TJ, Kuehn A, Dude M-A, et al. Sexual stage adhesion proteins form multi-protein complexes in the malaria parasite Plasmodium falciparum. J Biol Chem [Internet]. 2009;284(21):14537–46. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2682902&tool=pmcentrez&rendertype=abstract
  49. 49. Kuehn A, Simon N, Pradel G. Family members stick together: multi-protein complexes of malaria parasites. Med Microbiol Immunol. 2010;199(3):209–26.
  50. 50. Pradel G. Proteins of the malaria parasite sexual stages: expression, function and potential for transmission blocking strategies. Parasitology. 2007;134(Pt. 14):1911–29.
  51. 51. Moelans II, Meis JF, Kocken C, Konings RN, Schoenmakers JG. A novel protein antigen of the malaria parasite Plasmodium falciparum, located on the surface of gametes and sporozoites. Mol Biochem Parasitol [Internet]. 1991 Apr [cited 2016 Apr 7];45(2):193–204. Available from: http://www.ncbi.nlm.nih.gov/pubmed/2038355
  52. 52. Bruce MC, Carter RN, Nakamura K, Aikawa M, Carter R. Cellular location and temporal expression of the Plasmodium falciparum sexual stage antigen Pfs16. Mol Biochem Parasitol [Internet]. 1994 May [cited 2016 Apr 7];65(1):11–22. Available from: http://www.ncbi.nlm.nih.gov/pubmed/7935618
  53. 53. Baker DA, Daramola O, McCrossan MV, Harmer J, Targett GA. Subcellular localization of Pfs16, a Plasmodium falciparum gametocyte antigen. Parasitology [Internet]. 1994 Feb [cited 2016 Apr 7];108(Pt 2):129–37. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8159458
  54. 54. Kongkasuriyachai D, Fujioka H, Kumar N. Functional analysis of Plasmodium falciparum parasitophorous vacuole membrane protein (Pfs16) during gametocytogenesis and gametogenesis by targeted gene disruption. Mol Biochem Parasitol [Internet]. 2004 Feb [cited 2016 Apr 7];133(2):275–85. Available from: http://www.ncbi.nlm.nih.gov/pubmed/14698439
  55. 55. Sharma A, Sharma I, Kogkasuriyachai D, Kumar N. Structure of a gametocyte protein essential for sexual development in Plasmodium falciparum. Nat Struct Biol [Internet]. 2003 Mar [cited 2016 Apr 7];10(3):197–203. Available from: http://www.ncbi.nlm.nih.gov/pubmed/12577051
  56. 56. Olivieri A, Camarda G, Bertuccini L, van de Vegte-Bolmer M, Luty AJF, Sauerwein R, et al. The Plasmodium falciparum protein Pfg27 is dispensable for gametocyte and gamete production, but contributes to cell integrity during gametocytogenesis. Mol Microbiol [Internet]. 2009 Jul [cited 2016 Apr 7];73(2):180–93. Available from: http://www.ncbi.nlm.nih.gov/pubmed/19570101
  57. 57. Pradel G. A multidomain adhesion protein family expressed in plasmodium falciparum is essential for transmission to the mosquito. J Exp Med [Internet]. 2004;199(11):1533–44. Available from: http://www.jem.org/cgi/doi/10.1084/jem.20031274
  58. 58. Pradel G, Wagner C, Mejia C, Templeton TJ. Plasmodium falciparum: co-dependent expression and co-localization of the PfCCp multi-adhesion domain proteins. Exp Parasitol [Internet]. 2006;112(4):263–8. Available from: http://www.ncbi.nlm.nih.gov/pubmed/16388802
  59. 59. van Dijk MR, Janse CJ, Thompson J, Waters AP, Braks JA, Dodemont HJ, et al. A central role for P48/45 in malaria parasite male gamete fertility. Cell [Internet]. 2001 Jan 12 [cited 2016 Apr 7];104(1):153–64. Available from: http://www.ncbi.nlm.nih.gov/pubmed/11163248
  60. 60. Eksi S, Czesny B, van Gemert G-J, Sauerwein RW, Eling W, Williamson KC. Malaria transmission-blocking antigen, Pfs230, mediates human red blood cell binding to exflagellating male parasites and oocyst production. Mol Microbiol [Internet]. 2006;61(4):991–8. Available from: http://doi.wiley.com/10.1111/j.1365-2958.2006.05284.x
  61. 61. de Koning-Ward TF, Olivieri A, Bertuccini L, Hood A, Silvestrini F, Charvalias K, et al. The role of osmiophilic bodies and Pfg377 expression in female gametocyte emergence and mosquito infectivity in the human malaria parasite Plasmodium falciparum. Mol Microbiol [Internet]. 2008 Jan [cited 2016 Apr 7];67(2):278–90. Available from: http://www.ncbi.nlm.nih.gov/pubmed/18086189
  62. 62. Suaréz-Cortés P, Silvestrini F, Alano P. A fast, non-invasive, quantitative staining protocol provides insights in Plasmodium falciparum gamete egress and in the role of osmiophilic bodies. Malar J [Internet]. 2014 Jan [cited 2016 Apr 6];13:389. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4194377&tool=pmcentrez&rendertype=abstract
  63. 63. Furuya T, Mu J, Hayton K, Liu A, Duan J, Nkrumah L, et al. Disruption of a Plasmodium falciparum gene linked to male sexual development causes early arrest in gametocytogenesis. Proc Natl Acad Sci U S A [Internet]. 2005 Nov 15 [cited 2016 Apr 6];102(46):16813–8. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1277966&tool=pmcentrez&rendertype=abstract
  64. 64. Rawlings DJ, Fujioka H, Fried M, Keister DB, Aikawa M, Kaslow DC. Alpha-tubulin II is a male-specific protein in Plasmodium falciparum. Mol Biochem Parasitol [Internet]. 1992 Dec [cited 2016 Apr 7];56(2):239–50. Available from: http://www.ncbi.nlm.nih.gov/pubmed/1484548
  65. 65. Guinet F, Dvorak JA, Fujioka H, Keister DB, Muratova O, Kaslow DC, et al. A developmental defect in Plasmodium falciparum male gametogenesis. J Cell Biol [Internet]. 1996 Oct [cited 2016 Apr 7];135(1):269–78. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2121010&tool=pmcentrez&rendertype=abstract
  66. 66. Deligianni E, Morgan RN, Bertuccini L, Kooij TW, Laforge A, Nahar C, et al. Critical role for a stage-specific actin in male exflagellation of the malaria parasite. Cell Microbiol [Internet]. 2011;13(11):1714–30. Available from: http://www.ncbi.nlm.nih.gov/pubmed/21790945
  67. 67. Eksi S, Williamson KC. Male-specific expression of the paralog of malaria transmission-blocking target antigen Pfs230, PfB0400w. Mol Biochem Parasitol [Internet]. 2002 Jul [cited 2016 Apr 7];122(2):127–30. Available from: http://www.ncbi.nlm.nih.gov/pubmed/12106866
  68. 68. Doi M, Tanabe K, Tachibana S-I, Hamai M, Tachibana M, Mita T, et al. Worldwide sequence conservation of transmission-blocking vaccine candidate Pvs230 in Plasmodium vivax. Vaccine [Internet]. 2011 Jun 10 [cited 2016 Apr 26];29(26):4308–15. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3130600&tool=pmcentrez&rendertype=abstract
  69. 69. Tachibana M, Sato C, Otsuki H, Sattabongkot J, Kaneko O, Torii M, et al. Plasmodium vivax gametocyte protein Pvs230 is a transmission-blocking vaccine candidate. Vaccine [Internet]. 2012;30(10):1807–12. Available from: http://linkinghub.elsevier.com/retrieve/pii/S0264410X12000047
  70. 70. Tachibana M, Suwanabun N, Kaneko O, Iriko H, Otsuki H, Sattabongkot J, et al. Plasmodium vivax gametocyte proteins, Pvs48/45 and Pvs47, induce transmission-reducing antibodies by DNA immunization. Vaccine [Internet]. 2015 Apr 15 [cited 2016 Apr 26];33(16):1901–8. Available from: http://www.ncbi.nlm.nih.gov/pubmed/25765968
  71. 71. Talman AM, Domarle O, McKenzie FE, Ariey F, Robert V. Gametocytogenesis: the puberty of Plasmodium falciparum. Malar J [Internet]. 2004 Jul 14 [cited 2016 Apr 7];3:24. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=497046&tool=pmcentrez&rendertype=abstract
  72. 72. Baker DA. Malaria gametocytogenesis. Mol Biochem Parasitol [Internet]. 2010;172(2):57–65. Available from: http://dx.doi.org/10.1016/j.molbiopara.2010.03.019
  73. 73. Ben Mamoun C, Prigge ST, Vial H. Targeting the lipid metabolic pathways for the treatment of malaria. Drug Dev Res [Internet]. 2010 Feb [cited 2016 Apr 7];71(1):44–55. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2886290&tool=pmcentrez&rendertype=abstract
  74. 74. Gulati S, Ekland EH, Ruggles KV, Chan RB, Jayabalasingham B, Zhou B, et al. Profiling the essential nature of lipid metabolism in asexual blood and gametocyte stages of Plasmodium falciparum. Cell Host Microbe [Internet]. 2015 Sep 9 [cited 2016 Mar 27];18(3):371–81. Available from: http://www.ncbi.nlm.nih.gov/pubmed/26355219
  75. 75. Tran PN, Brown SHJ, Rug M, Ridgway MC, Mitchell TW, Maier AG. Changes in lipid composition during sexual development of the malaria parasite Plasmodium falciparum. Malar J [Internet]. 2016 Jan [cited 2016 Apr 7];15(1):73. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4744411&tool=pmcentrez&rendertype=abstract
  76. 76. Raabe AC, Billker O, Vial HJ, Wengelnik K. Quantitative assessment of DNA replication to monitor microgametogenesis in Plasmodium berghei. Mol Biochem Parasitol [Internet]. 2009 Dec [cited 2016 Apr 7];168(2):172–6. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2789244&tool=pmcentrez&rendertype=abstract
  77. 77. Canning EU, Sinden RE. Nuclear organisation in gametocytes of Plasmodium and Hepatocystis: a cytochemical study. Zeitschrift für Parasitenkd (Berlin, Ger [Internet]. 1975 Jul 16 [cited 2016 Apr 7];46(4):297–9. Available from: http://www.ncbi.nlm.nih.gov/pubmed/812282
  78. 78. Sinden RE, Canning EU, Bray RS, Smalley ME. Gametocyte and gamete development in Plasmodium falciparum. Proc R Soc London Ser B, Biol Sci [Internet]. 1978 Jun 5 [cited 2016 Apr 7];201(1145):375–99. Available from: http://www.ncbi.nlm.nih.gov/pubmed/27809
  79. 79. Waters AP, Syin C, McCutchan TF. Developmental regulation of stage-specific ribosome populations in Plasmodium. Nature [Internet]. 1989 Nov 23 [cited 2016 Apr 7];342(6248):438–40. Available from: http://www.ncbi.nlm.nih.gov/pubmed/2586613
  80. 80. Khan SM, Franke-Fayard B, Mair GR, Lasonder E, Janse CJ, Mann M, et al. Proteome analysis of separated male and female gametocytes reveals novel sex-specific Plasmodium biology. Cell [Internet]. 2005;121(5):675–87. Available from: http://linkinghub.elsevier.com/retrieve/pii/S0092867405002990
  81. 81. Petmitr S, Krungkrai J. Mitochondrial cytochrome b gene in two developmental stages of human malarial parasite Plasmodium falciparum. Southeast Asian J Trop Med Public Health [Internet]. 1995 Dec [cited 2016 Apr 7];26(4):600–5. Available from: http://www.ncbi.nlm.nih.gov/pubmed/9139360
  82. 82. Crofts AR. The cytochrome bc1 complex: function in the context of structure. Annu Rev Physiol [Internet]. 2004 Jan [cited 2016 Apr 7];66:689–733. Available from: http://www.ncbi.nlm.nih.gov/pubmed/14977419
  83. 83. Fry M, Pudney M. Site of action of the antimalarial hydroxynaphthoquinone, 2-[trans-4-(4’-chlorophenyl) cyclohexyl]-3-hydroxy-1,4-naphthoquinone (566C80). Biochem Pharmacol [Internet]. 1992 Apr 1 [cited 2016 Apr 7];43(7):1545–53. Available from: http://www.ncbi.nlm.nih.gov/pubmed/1314606
  84. 84. Fisher N, Abd Majid R, Antoine T, Al-Helal M, Warman AJ, Johnson DJ, et al. Cytochrome b mutation Y268S conferring atovaquone resistance phenotype in malaria parasite results in reduced parasite bc1 catalytic turnover and protein expression. J Biol Chem [Internet]. 2012 Mar 23 [cited 2016 Mar 13];287(13):9731–41. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3322985&tool=pmcentrez&rendertype=abstract
  85. 85. Siregar JE, Kurisu G, Kobayashi T, Matsuzaki M, Sakamoto K, Mi-ichi F, et al. Direct evidence for the atovaquone action on the Plasmodium cytochrome bc1 complex. Parasitol Int [Internet]. 2015 Jun [cited 2016 Mar 22];64(3):295–300. Available from: http://www.ncbi.nlm.nih.gov/pubmed/25264100
  86. 86. Butterworth AS, Skinner-Adams TS, Gardiner DL, Trenholme KR. Plasmodium falciparum gametocytes: with a view to a kill. Parasitology [Internet]. 2013 Dec [cited 2016 Apr 7];140(14):1718–34. Available from: http://www.ncbi.nlm.nih.gov/pubmed/23953486
  87. 87. Sherman IW. Biochemistry of Plasmodium (malarial parasites). Microbiol Rev [Internet]. 1979 Dec [cited 2016 Apr 7];43(4):453–95. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=281489&tool=pmcentrez&rendertype=abstract
  88. 88. MacRae JI, Dixon MW, Dearnley MK, Chua HH, Chambers JM, Kenny S, et al. Mitochondrial metabolism of sexual and asexual blood stages of the malaria parasite Plasmodium falciparum. BMC Biol [Internet]. 2013 Jan [cited 2016 Apr 7];11:67. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3704724&tool=pmcentrez&rendertype=abstract
  89. 89. Billker O, Lindo V, Panico M, Etienne AE, Paxton T, Dell A, et al. Identification of xanthurenic acid as the putative inducer of malaria development in the mosquito. Nature [Internet]. 1998 Mar 19 [cited 2016 Apr 5];392(6673):289–92. Available from: http://www.ncbi.nlm.nih.gov/pubmed/9521324
  90. 90. Garcia GE, Wirtz RA, Barr JR, Woolfitt A, Rosenberg R. Xanthurenic acid induces gametogenesis in Plasmodium, the malaria parasite. J Biol Chem [Internet]. 1998 May 15 [cited 2016 Apr 5];273(20):12003–5. Available from: http://www.ncbi.nlm.nih.gov/pubmed/9575140
  91. 91. Sologub L, Kuehn A, Kern S, Przyborski J, Schillig R, Pradel G. Malaria proteases mediate inside-out egress of gametocytes from red blood cells following parasite transmission to the mosquito. Cell Microbiol. 2011;13(6):897–912.
  92. 92. Kawamoto F, Alejo-Blanco R, Fleck SL, Kawamoto Y, Sinden RE. Possible roles of Ca2+ and cGMP as mediators of the exflagellation of Plasmodium berghei and Plasmodium falciparum. Mol Biochem Parasitol [Internet]. 1990 Aug [cited 2016 Apr 5];42(1):101–8. Available from: http://www.ncbi.nlm.nih.gov/pubmed/2172816
  93. 93. Kawamoto F, Fujioka H, Murakami R, Syafruddin, Hagiwara M, Ishikawa T, et al. The roles of Ca2+/calmodulin- and cGMP-dependent pathways in gametogenesis of a rodent malaria parasite, Plasmodium berghei. Eur J Cell Biol [Internet]. 1993 Feb [cited 2016 Apr 5];60(1):101–7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8385016
  94. 94. Carucci DJ, Witney AA, Muhia DK, Warhurst DC, Schaap P, Meima M, et al. Guanylyl cyclase activity associated with putative bifunctional integral membrane proteins in Plasmodium falciparum. J Biol Chem [Internet]. 2000 Jul 21 [cited 2016 Apr 5];275(29):22147–56. Available from: http://www.ncbi.nlm.nih.gov/pubmed/10747978
  95. 95. Muhia DK, Swales CA, Deng W, Kelly JM, Baker DA. The gametocyte-activating factor xanthurenic acid stimulates an increase in membrane-associated guanylyl cyclase activity in the human malaria parasite Plasmodium falciparum. Mol Microbiol [Internet]. 2001 Oct [cited 2016 Apr 5];42(2):553–60. Available from: http://www.ncbi.nlm.nih.gov/pubmed/11703675
  96. 96. Hirai M, Arai M, Kawai S, Matsuoka H. PbGCbeta is essential for Plasmodium ookinete motility to invade midgut cell and for successful completion of parasite life cycle in mosquitoes. J Biochem [Internet]. 2006 Nov [cited 2016 Apr 5];140(5):747–57. Available from: http://www.ncbi.nlm.nih.gov/pubmed/17030505
  97. 97. Taylor CJ, McRobert L, Baker DA. Disruption of a Plasmodium falciparum cyclic nucleotide phosphodiesterase gene causes aberrant gametogenesis. Mol Microbiol [Internet]. 2008 Jul [cited 2016 Apr 5];69(1):110–8. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2615252&tool=pmcentrez&rendertype=abstract
  98. 98. Young JA, Fivelman QL, Blair PL, de la Vega P, Le Roch KG, Zhou Y, et al. The Plasmodium falciparum sexual development transcriptome: a microarray analysis using ontology-based pattern identification. Mol Biochem Parasitol [Internet]. 2005;143(1):67–79. Available from: http://linkinghub.elsevier.com/retrieve/pii/S0166685105001623
  99. 99. McRobert L, Taylor CJ, Deng W, Fivelman QL, Cummings RM, Polley SD, et al. Gametogenesis in malaria parasites is mediated by the cGMP-dependent protein kinase. PLoS Biol [Internet]. 2008;6(6):e139. Available from: http://biology.plosjournals.org/perlserv/?request=get-document&doi=10.1371/journal.pbio.0060139
  100. 100. Martin JV, Nagele RG, Lee HY. Temporal changes in intracellular free calcium levels in the developing neuroepithelium during neurulation in the chick. Comp Biochem Physiol Comp Physiol [Internet]. 1994 Apr [cited 2016 Apr 5];107(4):655–9. Available from: http://www.ncbi.nlm.nih.gov/pubmed/7911410
  101. 101. Raabe A, Berry L, Sollelis L, Cerdan R, Tawk L, Vial HJ, et al. Genetic and transcriptional analysis of phosphoinositide-specific phospholipase C in Plasmodium. Exp Parasitol [Internet]. 2011 Sep [cited 2016 Apr 5];129(1):75–80. Available from: http://www.ncbi.nlm.nih.gov/pubmed/21651909
  102. 102. Brochet M, Collins MO, Smith TK, Thompson E, Sebastian S, Volkmann K, et al. Phosphoinositide metabolism links cGMP-dependent protein kinase G to essential Ca2+ signals at key decision points in the life cycle of malaria parasites. PLoS Biol [Internet]. 2014 Mar [cited 2016 Mar 24];12(3):e1001806. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3942320&tool=pmcentrez&rendertype=abstract
  103. 103. Billker O, Dechamps S, Tewari R, Wenig G, Franke-Fayard B, Brinkmann V. Calcium and a calcium-dependent protein kinase regulate gamete formation and mosquito transmission in a malaria parasite. Cell. 2004;117(4):503–14.
  104. 104. Laurentino EC, Taylor S, Mair GR, Lasonder E, Bartfai R, Stunnenberg HG, et al. Experimentally controlled downregulation of the histone chaperone FACT in Plasmodium berghei reveals that it is critical to male gamete fertility. Cell Microbiol [Internet]. 2011 Dec [cited 2016 Apr 5];13(12):1956–74. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3429858&tool=pmcentrez&rendertype=abstract
  105. 105. Rangarajan R, Bei AK, Jethwaney D, Maldonado P, Dorin D, Sultan AA, et al. A mitogen-activated protein kinase regulates male gametogenesis and transmission of the malaria parasite Plasmodium berghei. EMBO Rep [Internet]. 2005 May [cited 2016 Apr 6];6(5):464–9. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1299310&tool=pmcentrez&rendertype=abstract
  106. 106. Tewari R, Dorin D, Moon R, Doerig C, Billker O. An atypical mitogen-activated protein kinase controls cytokinesis and flagellar motility during male gamete formation in a malaria parasite. Mol Microbiol. 2005;58(5):1253–63.
  107. 107. Guttery DS, Ferguson DJP, Poulin B, Xu Z, Straschil U, Klop O, et al. A putative homologue of CDC20/CDH1 in the malaria parasite is essential for male gamete development. PLoS Pathog [Internet]. 2012 Feb [cited 2016 Apr 6];8(2):e1002554. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3285604&tool=pmcentrez&rendertype=abstract
  108. 108. Sebastian S, Brochet M, Collins MO, Schwach F, Jones ML, Goulding D, et al. A Plasmodium calcium-dependent protein kinase controls zygote development and transmission by translationally activating pepressed mRNAs. Cell Host Microbe. 2012;12(1):9–19.
  109. 109. Guttery DS, Poulin B, Ramaprasad A, Wall RJ, Ferguson DJP, Brady D, et al. Genome-wide functional analysis of Plasmodium protein phosphatases reveals key regulators of parasite development and differentiation. Cell Host Microbe [Internet]. 2014 Jul 9 [cited 2016 Apr 6];16(1):128–40. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4094981&tool=pmcentrez&rendertype=abstract
  110. 110. Philip N, Waters AP. Conditional degradation of Plasmodium calcineurin reveals functions in parasite colonization of both host and vector. Cell Host Microbe [Internet]. 2015 Jul 8 [cited 2015 Dec 29];18(1):122–31. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4509507&tool=pmcentrez&rendertype=abstract
  111. 111. Mair GR. Regulation of sexual development of plasmodium by translational repression. Science (80-) [Internet]. 2006;313(5787):667–9. Available from: http://www.sciencemag.org/cgi/doi/10.1126/science.1125129
  112. 112. Mair GR, Lasonder E, Garver LS, Franke-Fayard BMD, Carret CK, Wiegant JCAG, et al. Universal features of post-transcriptional gene regulation are critical for plasmodium zygote development. PLoS Pathog [Internet]. 2010;6(2):e1000767. Available from: http://dx.plos.org/10.1371/journal.ppat.1000767
  113. 113. Guerreiro A, Deligianni E, Santos JM, Silva PAGC, Louis C, Pain A, et al. Genome-wide RIP-Chip analysis of translational repressor-bound mRNAs in the Plasmodium gametocyte. Genome Biol [Internet]. 2014 Jan [cited 2016 Apr 6];15(11):493. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4234863&tool=pmcentrez&rendertype=abstract
  114. 114. Miao J, Fan Q, Parker D, Li X, Li J, Cui L. Puf mediates translation repression of transmission-blocking vaccine candidates in malaria parasites. PLoS Pathog [Internet]. 2013 Jan [cited 2016 Apr 6];9(4):e1003268. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3630172&tool=pmcentrez&rendertype=abstract
  115. 115. Wirth CC, Pradel G. Molecular mechanisms of host cell egress by malaria parasites. Int J Med Microbiol [Internet]. 2012;302(4–5):172–8. Available from: http://linkinghub.elsevier.com/retrieve/pii/S1438422112000318
  116. 116. Deligianni E, Morgan RN, Bertuccini L, Wirth CC, Silmon de Monerri NC, Spanos L, et al. A perforin-like protein mediates disruption of the erythrocyte membrane during egress of Plasmodium berghei male gametocytes. Cell Microbiol [Internet]. 2013 Aug [cited 2016 Apr 6];15(8):1438–55. Available from: http://www.ncbi.nlm.nih.gov/pubmed/23461714
  117. 117. Wirth CC, Glushakova S, Scheuermayer M, Repnik U, Garg S, Schaack D, et al. Perforin-like protein PPLP2 permeabilizes the red blood cell membrane during egress of Plasmodium falciparum gametocytes. Cell Microbiol. 2014;16(5):709–33.
  118. 118. Rupp I, Sologub L, Williamson KC, Scheuermayer M, Reininger L, Doerig C, et al. Malaria parasites form filamentous cell-to-cell connections during reproduction in the mosquito midgut. Cell Res [Internet]. 2011;21(4):683–96. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3072464&tool=pmcentrez&rendertype=abstract\nhttp://dx.doi.org/10.1038/cr.2010.176
  119. 119. Janse CJ, Van der Klooster PF, Van der Kaay HJ, Van der Ploeg M, Overdulve JP. Rapid repeated DNA replication during microgametogenesis and DNA synthesis in young zygotes of Plasmodium berghei. Trans R Soc Trop Med Hyg [Internet]. 1986 Jan [cited 2016 Apr 6];80(1):154–7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/3088783
  120. 120. Marques SR, Ramakrishnan C, Carzaniga R, Blagborough AM, Delves MJ, Talman AM, et al. An essential role of the basal body protein SAS-6 in Plasmodium male gamete development and malaria transmission. Cell Microbiol [Internet]. 2015 Feb [cited 2016 Apr 6];17(2):191–206. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=4441282&tool=pmcentrez&rendertype=abstract
  121. 121. Straschil U, Talman AM, Ferguson DJP, Bunting KA, Xu Z, Bailes E, et al. The Armadillo repeat protein PF16 is essential for flagellar structure and function in Plasmodium male gametes. PLoS One [Internet]. 2010 Jan [cited 2016 Apr 6];5(9):e12901. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2944832&tool=pmcentrez&rendertype=abstract
  122. 122. Hirai M, Arai M, Mori T, Miyagishima S-Y, Kawai S, Kita K, et al. Male fertility of malaria parasites is determined by GCS1, a plant-type reproduction factor. Curr Biol [Internet]. 2008 Apr 22 [cited 2016 Apr 6];18(8):607–13. Available from: http://www.ncbi.nlm.nih.gov/pubmed/18403203
  123. 123. Liu Y, Tewari R, Ning J, Blagborough AM, Garbom S, Pei J, et al. The conserved plant sterility gene HAP2 functions after attachment of fusogenic membranes in Chlamydomonas and Plasmodium gametes. Genes Dev [Internet]. 2008 Apr 15 [cited 2016 Apr 6];22(8):1051–68. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2335326&tool=pmcentrez&rendertype=abstract
  124. 124. Aikawa M, Schwartz A, Uni S, Nussenzweig R, Hollingdale M. Ultrastructure of in vitro cultured exoerythrocytic stage of Plasmodium berghei in a hepatoma cell line. Am J Trop Med Hyg [Internet]. 1984 Sep [cited 2016 Apr 6];33(5):792–9. Available from: http://www.ncbi.nlm.nih.gov/pubmed/6091467
  125. 125. Sinden RE. Molecular interactions between Plasmodium and its insect vectors. Cell Microbiol [Internet]. 2002 Nov [cited 2016 Apr 6];4(11):713–24. Available from: http://www.ncbi.nlm.nih.gov/pubmed/12427094
  126. 126. Sinden RE, Hartley RH, Winger L. The development of Plasmodium ookinetes in vitro: an ultrastructural study including a description of meiotic division. Parasitology [Internet]. 1985 Oct [cited 2016 Feb 22];91(Pt 2):227–44. Available from: http://www.ncbi.nlm.nih.gov/pubmed/3906519
  127. 127. Vlachou D, Zimmermann T, Cantera R, Janse CJ, Waters AP, Kafatos FC. Real-time, in vivo analysis of malaria ookinete locomotion and mosquito midgut invasion. Cell Microbiol [Internet]. 2004 Jul [cited 2016 Feb 11];6(7):671–85. Available from: http://www.ncbi.nlm.nih.gov/pubmed/15186403

Written By

Che Julius Ngwa, Thiago F. de A. Rosa and Gabriele Pradel

Submitted: 18 November 2015 Reviewed: 01 September 2016 Published: 30 November 2016