Next Article in Journal
Pleiotropic Regulator GssR Positively Regulates Autotrophic Growth of Gas-Fermenting Clostridium ljungdahlii
Next Article in Special Issue
Molecular Mechanisms of Persistence in Protozoan Parasites
Previous Article in Journal
Performance of the Disc Diffusion Method, MTS Gradient Tests and Two Commercially Available Microdilution Tests for the Determination of Cefiderocol Susceptibility in Acinetobacter spp.
Previous Article in Special Issue
Zoonotic Parasites in Playgrounds in Southern Spain: A One Health Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plasmodium falciparum Development from Gametocyte to Oocyst: Insight from Functional Studies

1
Malaria Research and Training Center, Faculty of Pharmacy, Faculty of Medicine and Dentistry, University of Sciences, Techniques, and Technologies of Bamako, Bamako 1805, Mali
2
Malaria Research Program, Center for Vaccine Development and Global Health, University of Maryland School of Medicine, Baltimore, MD 21201, USA
*
Author to whom correspondence should be addressed.
Microorganisms 2023, 11(8), 1966; https://doi.org/10.3390/microorganisms11081966
Submission received: 8 June 2023 / Revised: 6 July 2023 / Accepted: 14 July 2023 / Published: 31 July 2023
(This article belongs to the Special Issue Cellular Biology of Protozoan Parasites of Mammals)

Abstract

:
Malaria elimination may never succeed without the implementation of transmission-blocking strategies. The transmission of Plasmodium spp. parasites from the human host to the mosquito vector depends on circulating gametocytes in the peripheral blood of the vertebrate host. Once ingested by the mosquito during blood meals, these sexual forms undergo a series of radical morphological and metabolic changes to survive and progress from the gut to the salivary glands, where they will be waiting to be injected into the vertebrate host. The design of effective transmission-blocking strategies requires a thorough understanding of all the mechanisms that drive the development of gametocytes, gametes, sexual reproduction, and subsequent differentiation within the mosquito. The drastic changes in Plasmodium falciparum shape and function throughout its life cycle rely on the tight regulation of stage-specific gene expression. This review outlines the mechanisms involved in Plasmodium falciparum sexual stage development in both the human and mosquito vector, and zygote to oocyst differentiation. Functional studies unravel mechanisms employed by P. falciparum to orchestrate the expression of stage-specific functional products required to succeed in its complex life cycle, thus providing us with potential targets for developing new therapeutics. These mechanisms are based on studies conducted with various Plasmodium species, including predominantly P. falciparum and the rodent malaria parasites P. berghei. However, the great potential of epigenetics, genomics, transcriptomics, proteomics, and functional genetic studies to improve the understanding of malaria as a disease remains partly untapped because of limitations in studies using human malaria parasites and field isolates.

1. Introduction

Malaria is a mosquito-borne infection caused by an apicomplexan parasite of the genus Plasmodium and is transmitted through the bite of a mosquito vector of the genus Anopheles. In humans, malaria is caused by six Plasmodium species, including Plasmodium falciparum, Plasmodium vivax, Plasmodium malariae, Plasmodium ovale curtisi, Plasmodium ovale wallikeri, and the zoonotic parasite Plasmodium knowlesi. Of the human malaria parasites, P. falciparum causes the most significant morbidity and mortality. In 2021, clinical malaria affected 247,000 million individuals worldwide, leading to 619,000 deaths, with the highest burden observed in sub-Saharan Africa [1]. Over the last two decades, artemisinin-based combination therapies (ACTs), intermittent preventive treatments (IPTs), and vector control approaches have significantly reduced the burden of malaria in some areas [1]. However, despite tremendous efforts to control malaria, the emergence of resistant parasites to current antimalarials and the resistance of mosquitos to insecticides are significant obstacles to the global malaria eradication/elimination program [1].
With the renewed interest in malaria elimination and eradication, it has been recognized that intervention strategies should target all stages of the parasite life cycle with a particular interest in preventing transmission [2]. Plasmodium transmission from the human host to the mosquito vector and the parasite stages within the mosquito are promising targets for successfully interrupting the parasite life cycle and reducing the malaria burden in endemic areas [3].
The life cycle of Plasmodium falciparum is very complex. It includes different hosts, different cell types within the same host, asexual multiplication phases, and a sexual replication phase (Figure 1). In humans, the cycle begins with the inoculation of the sporozoite forms of the parasite by an infected mosquito during its blood meal. These sporozoites migrate through the bloodstream, reach the liver, and enter the hepatocytes, where they multiply asexually. After six to ten days, the rupture of infected hepatocytes releases thousands of merozoites, a new form of the parasite, into the bloodstream. The merozoites invade the red blood cells (RBCs) to initiate the intraerythrocytic cycle. After two days, new merozoites are released, and these offspring enter new RBCs to perpetuate the intraerythrocytic cycle responsible for the clinical manifestations of malaria in humans. A small percentage of merozoites differentiate into sexual forms called gametocytes, which mediate transmission. Gametocytes are morphologically (shape, size, and flexibility) [4,5,6] and functionally different from asexual parasites [6,7,8,9]. Mosquitos ingest male and female gametocytes during a blood meal, which results in a wave of differentiation processes and developmental stages within the mosquito [6,10,11,12]. In the mosquito’s gut lumen, gametocytes transform into male and female gametes. These gametes then fuse to produce a zygote that differentiates into a highly motile and invasive ookinete [6,11,13]. The ookinete invades the mosquito epithelial cell wall to form an oocyst, which in turn will ultimately produce thousands of sporozoites through a process known as sporogony [6,12,14]. The sporozoites will migrate and reside in the mosquito salivary gland [12], waiting to be transmitted to the vertebrate host during a blood meal and perpetuating the parasite’s life cycle.
The drastic changes in both shape and function of the Plasmodium parasite to adapt to distinct environmental niches throughout its life cycle rely on tight programming of stage-specific gene expression, as shown by several functional genomics, transcriptomics [15,16,17,18,19,20], and proteomic analysis [21,22]. Understanding the mechanisms that drive the development of gametocytes, gametogenesis, sexual reproduction, and zygote differentiation within the mosquito will likely reveal new avenues to interrupt the parasite life cycle, thus interrupting malaria transmission. Topics of interest for developing transmission-blocking strategies include factors that trigger asexual parasites to commit to sexual differentiation, the molecular players involved in gametocytogenesis, and gametocyte maturation. A better understanding of the mechanisms involved in gamete formation and fusion, the requirements for establishing a successful sexual stage within the mosquito, and the molecular players controlling zygote to oocyst production are also of interest.
Genomics, transcriptomics, and proteomics provide valuable insight into Plasmodium biology, including parasite differentiation within humans and the mosquito vector [18,21,23,24,25]. Since the first publication of the Plasmodium falciparum reference genome in 2002, a significant number of Plasmodium spp. Whole-genome sequences have been released, allowing in-depth comparative analysis [26,27]. These analyses revealed genes potentially linked to phenotypic differences [28,29]. Advances in sequencing technologies and analytical techniques allow a more detailed exploration of the parasite genomes, improving our understanding of the parasite ecology and epidemiology [30,31,32]. Transcriptomics and proteomics improved our knowledge of RNA and protein expression dynamics and their regulation throughout the parasite life cycle [33,34,35].
The cycle begins with the inoculation of the sporozoite forms of the parasite by an infected mosquito during its blood meal. These sporozoites migrate through the bloodstream and reach the liver. The sporozoites will invade the hepatocytes and initiate the exoerythrocytic developmental cycle. The infected hepatocytes release thousands of merozoites 6–10 days after the initial invasion. This new form of the parasite will enter the bloodstream to invade the red blood cells (RBCs) and initiate the intraerythrocytic cycle. Within the intraerythrocytic developmental cycle, a small percentage of the parasite commits to sexual differentiation, resulting in the development of the gametocyte forms. The gametocytes will be picked up by the mosquito during their blood meal. Within the mosquito, the sporogonic cycle will take place.
Gametocytogenesis and Plasmodium development in the mosquito have been the subject of several comprehensive reviews [36,37,38,39]. However, these reviews often focus either on gametocytogenesis in the human host or parasite development within the mosquito vector. In this review, we aim to cover the Plasmodium falciparum developmental process from the pre-sexual stage in the human host to the ookinete stage of the mosquito vector. The review is based on studies conducted with various Plasmodium species, including predominantly P. falciparum and the rodent malaria parasites P. berghei. Although Plasmodium berghei is a rodent malaria parasite, the availability of fast and efficient experimental genetics techniques to access the complete in vivo life cycle made rodent malaria parasites valuable models for understanding some aspects of P. falciparum biology. The review highlights studies assessing the mechanisms involved in the parasite-stage-specific gene expression, sexual stage development in both the vertebrate and mosquito vector, and zygote to oocyst differentiation. This review starts with an outline of genome organization, general gene expression, and regulation mechanisms in the intraerythrocytic asexual stage. Then, we describe the mechanisms driving the development of gametocytes, gametes, fertilization, and zygote to oocyst differentiation. Finally, the review highlights gaps and potential targets for developing new classes of antimalarial drugs or vaccines.

2. Plasmodium Genome Organization and General Mechanism for Gene Expression Regulation

2.1. Plasmodium Genome Organization

The Plasmodium parasite genome of ~18–30 Mb is organized into 14 chromosomes containing about 5300 protein-coding genes, with many sub-telomeric multigene families [26]. The packaging and organization of genomic DNA into chromatin constitute an essential regulatory mechanism for Plasmodium gene expression [30,32]. As in model organisms, specific nucleosome positions restrict the accessibility of regulatory DNA elements and thus can be predictive of gene transcription state [40,41]. The P. falciparum nucleosome organization consists of approximately 155 DNA base pairs (bp) around a histone octamer [42]. Site-specific analysis of nucleosome positioning has revealed that nucleosomes are best positioned when flanked by AT/TA dinucleotides [41]. Plasmodium histone molecules are characterized by distinct biochemical properties thought to modulate both nucleosome stability and their rapid displacement in a temperature-dependent manner [41]. The downstream regions of transcription start sites are unusually depleted of nucleosomes, presumably to facilitate the recruitment of the basal transcription machinery [43].
A comparative analysis of chromosome organization in different plasmodial species revealed that the two most pathogenic human malaria parasites, P. falciparum and P. vivax, have a distinct genome organization [30]. Indeed, the genome organization of these species appears to be shaped by clusters of parasite-specific gene families linked to pathogenicity, virulence, and parasite differentiation in the mosquito (Figure 2). The parasites employ the clustering of specific gene families to regulate the expression of stage-specific genes [44]. Identifying the molecular players controlling these critical assemblies and phenotypic differences will assist in uncovering new therapeutic targets.

2.2. Epigenetic Regulation of Gene Expression

Histone modifications’ spatial and temporal coordination impacts Plasmodium parasite differentiation by inducing, preventing, or poising transcriptional activation, DNA replication, damage repair, and chromatin condensation. Recent quantitative chromatin proteomic approaches and studies on the genome-wide localization of various epigenetic features have identified stage-specific histone modifications and their potential implication in the different life cycle stages of the malaria parasite [45,46] (Figure 2). Gene activation is associated with histone activation marks that include histone 3 Lysine 9 acetylation (H3K9ac), histone 3 Lysine 18 acetylation (H3K18ac), histone 3 Lysine 27 acetylation (H3K27ac), histone 3 lysine 4 trimethylation (H3K4me3), and the histone variant H2A.Z [43,47,48,49]. A study by Gupta et al. identified seven additional histone activation marks, including histone 4 Lysine 8 acetylation (H4K8ac), histone 4 Lysine 16 acetylation (H4K16ac), histone 4 tetra-acetylation (H4ac4), histone 3 Lysine 56 acetylation (H3K56ac), histone 3 Lysine 9 acetylation (H3K9ac), histone 3 Lysine 14 acetylation (H3K14ac), and histone 4 Lysine 20 methylation (H4K20me1) [50]. In contrast, gene repression is associated with distinct histone methylation and acetylation, including histone 3 lysine 9 trimethylation (H3K9me3) [51], histone 3 lysine 36 trimethylation (H3K36me3) [52,53], and histone 4 lysine 20 trimethylation (H4K20me3) [54].
Reversible gene silencing and activation are mediated by histone-modifying enzymes and histone-associated proteins that will “read, write or erase” histone code. P. falciparum is thought to encode 5 histone deacetylases (HDACs), 10 putative histone acetyltransferases (HATs), 10 methyltransferases (HMTs), and 3 demethylases [55,56]. These enzymes are predicted to add or remove methyl and acetyl groups to the lysine residues on histone molecules. Genetic studies have revealed several proteins implicated in gene silencing and heterochromatin state formation. The histone deacetylase enzyme PfHda2 [57], the sirtuin PfSir2a and PfSir2b [58], the H3K36-specific lysine methyltransferase PfSET2 [52], and the H3K4-specific lysine methyltransferase PfSET10 [59] are involved in the silencing of specific gene families, including sexual differentiation genes [54], and the virulence gene family known as var genes [52,54,57,58,59]. The var genes consist of approximately 60 paralogs of P. falciparum erythrocyte membrane protein 1 (PfEMP1) located in the subtelomeric regions of multiple parasite chromosomes [60]. They are displayed on the surfaces of infected red blood cells and characterized by a monoallelic expression. The switch between antigenically and functionally distinct variants is responsible for antigenic variations and the cytoadherence of infected erythrocytes to the microvasculature [61].
Epigenomic and functional genetic studies have also identified several histone-associated proteins, the heterochromatin protein 1 (PfHP1) [62,63] and some bromodomain-containing proteins (BDP), including PfBDP1 [64], PfBDP2 [64], and PfBDP7 [65]. These proteins are implicated and required for silencing gene families involved in red blood cell invasion and gametocyte differentiation [62,64,65]. PfHP1 is recruited by the H3K9me3 repressive epigenetic mark. The binding of PfHP1 regulates the formation of heterochromatins. On the contrary, the histone post-translational modification histone H3, serine 10 phosphorylation (H3S10ph) impairs the binding of HP1 [66]. The authors suggested that H3 phosphorylation by Aurora B is part of a “methyl/phos switch” mechanism that displaces HP1 and perhaps other proteins from heterochromatin.
The activation of gene expression is mediated by the histone acetyltransferase PfMYST (MOZ, YbF2, Sas2, Tip60-like) [67] and PfGCN5 (General control non-depressible 5) [68]. An in vitro study assessing the activity of P. falciparum-purified MYST proteins revealed the recruitment of PfMYST to acetylated histone 4 gene activation marks, suggesting PfMYST implication in the expression of var genes and genes regulating asexual intraerythrocytic growth [67]. Similarly, functional studies on the divergent PfGCN5 highlighted its association with euchromatin marks and the controlled expression of erythrocyte invasion and virulence genes [68].
Epigenomic studies have demonstrated that while gene activation marks (H3K9ac, H3K4me3) are widespread throughout the genome, the repressive marks (H3K9me3, H3K36me3, H4K20me3, HP1) appear to be specific for distinct gene families adjacent to subtelomeric var genes [51]. The gene subset comprises the antigenic variation gene families (var, rif, stevor), the Maurer’s cleft protein encoding two transmembrane domains (Pfmc-2tm), the nutrient acquisition multigene family (clag genes and acyl-CoA synthase genes), parasite differentiation genes (Apicomplexan Apetalla 2 genes; ApiAP2 genes), and the Plasmodium helical interspersed subtelomeric (PHIST) family.
Most epigenetic regulators and chromatin proteins implicated in gene silencing and activation in Plasmodium appear to be evolutionarily conserved in eukaryotic cells, suggesting a limitation in targeting them for therapeutic intervention. However, an in vitro screening of epigenetic modulator inhibitors against multiple life cycle stages of P. falciparum demonstrates the potential of using HDAC and HMT inhibitors as a new class of antimalarial drugs [69]. Coetzee et al. [69] tested the activity of compounds reported to be active against HDAC, HKM, HAT, and DNA demethylases in cancerous mammalian cells. Of the 95 compounds tested, 5 of them targeting histone acetylation and methylation showed potent multistage activity against asexual parasites, gametocyte stages, and the establishment of a successful infection in the mosquito. Thus, computer-aided drug design might be an alternative for developing a new class of antimalarial drugs targeting Plasmodium histone modification enzymes [70].

2.3. Transcriptional Regulation of Gene Expression

In addition to genomic organization and epigenetics, the regulatory machinery for gene expression in the Plasmodium falciparum includes the canonical TATA box binding proteins and the RNA-polymerase-II-dependent RNA expression [71]. A Hidden Markov Model (HMM) profile search for the P. falciparum genome identified 156 parasite-transcription-associated proteins, including all the 12 subunits of the RNA polymerase II and the general transcription factors (GTFs) for the basal transcription machinery [72]. Coulson et al. revealed that one group of GTFs, characterized by the histone 1 folding motif and their associated transcription factor complex (TFIID), is missing in the P. falciparum genome [72]. The authors suggested an evolutionary divergence of the basal transcriptional machinery in these parasites.
A low number of transcriptional regulation motifs and transcription factors that could account for the tight regulation of stage-specific gene expression patterns have been identified [71,72,73]. The ApiAp2 family, consisting of approximately 27 proteins, represents the main class of transcription factors known to regulate gene expression in Plasmodium [73,74]. These transcription factors contain the APETALA-2 (AP2) DNA binding domains and recognize multiple and distinct palindromic DNA sequences, including the TGCATGCA and GTGCAC motifs [75]. With some exceptions, most ApiAp2 transcription factors appear to regulate the expression of a unique gene set during one particular developmental stage [25,76,77,78,79,80]. Interestingly, the role of Plasmodium AP2 transcription factors is not restricted to regulating stage-specific gene expression. Indeed, the ApiAp2 proteins PfAP2Tel [81] and PfSIP2 [82], characterized by atypical AP2 domains that bind to specific DNA sequences, were shown to regulate telomere organization and promote gene silencing, respectively [81,82]. Additional transcription factors associated with gene expression that have been described include PfMyb1 protein (a transcription factor belonging to the tryptophan cluster family) [83], PfPREBP (a protein with four K-homology domains) [84], and PfNF-YB (a protein containing a histone-fold domain) [85,86]. However, these transcription factors have only been associated with gene expression regulation in the intra-erythrocytic stage of the parasite [83,84,85,86].
Systematic profiling of the genome-wide occupancy of eighteen PfApiAp2 revealed that eight were preferentially associated with heterochromatic regions with differential coverage profiles [87]. Shang et al. [88] suggested that these heterochromatin-associated factors (PfAP2-HFs) are likely part of the machinery recognizing repressive states in a DNA-motif-independent manner. Indeed, one of these PfAP2-HFs was shown to be strictly recruited by PfHP1 and to be a core component of heterochromatin.
The shortage of transcription and associated proteins strongly suggest an essential role for epigenetics, chromatin structure, genome organization, and posttranscriptional mechanisms in Plasmodium gene expression. Since the ApiAP2 protein family originated from the plant lineage and has no homologs in humans, they may be good antimalarial drug targets. An in silico screen to dock thousands of small molecules into the crystal structure of the AP2-EXP AP2 domain identified four compounds that specifically block DNA binding by AP2-EXP [89]. One of the inhibitors was shown to alter the transcriptome of P. falciparum trophozoite stages, characterized by a significant decrease in the abundance of AP2-EXP target genes. Two other DNA-binding inhibitory compounds have multi-stage anti-Plasmodium activity against blood and mosquito-stage parasites. A structure-based drug design combined with in vitro screening, in vivo studies in animal models, and the comprehensive profiling of ApiAp2 genetic polymorphisms in field-isolated parasites from natural human infection are likely to help in the development of a novel class of antimalarial drug targeting the Plasmodium ApiAp2 transcription factor. Furthermore, this compound class can potentially interfere with gene transcription in other Plasmodial species and apicomplexan parasites of medical importance, including Toxoplasma gondii and Cryptosporidium spp. While T. gondii causes many deaths in immunocompromised patients and fetal death when newly acquired during pregnancy [90], Cryptosporidium spp. Is one of the leading causes of infant diarrhea in developing countries [91]. T. gondii and Cryptosporidium parvum are predicted to encode 24 and 19 ApiAP2 genes, respectively, with some cross-conservation among Apicomplexan parasites [73,92].

3. Gametocyte Development

The development of male and female gametocytes within the vertebrate host and the sexual recombination in the mosquito vector are critical steps in the Plasmodium falciparum life cycle. These biological stages ensure parasite virulence and transmission. During the erythrocytic cycle, 1–5% of asexual parasites switch from the asexual multiplication pathway to a sexual differentiation pathway. The differentiation of sexually committed parasites into mature gametocytes is divided into five developmental stages (stage I–V) over 14 days. The differentiation process is characterized by dramatic alterations in parasite size, shape, and flexibility. Only early stages (stage I) and mature stage V are detectable in the peripheral blood, as the other stages are sequestered in the bone marrow and possibly other tissues [93]. The differentiation process underlying asexual to stage V gametocyte formation is governed by specific expression mechanisms extensively described in other reviews [94,95,96]. The following paragraphs summarize gene expression regulation mechanisms involved in asexual to early gametocyte formation (Table 1).
Several environmental and metabolic factors that trigger sexual commitment, including antimalarial drugs [97,98], lysophosphatidylcholine [99,100], host immune response [101,102], and parasite factors [103,104], have been reported. ApiAP2-G was the first transcription factor shown to govern the transcription of gametocyte-associated genes [24,25], and its regulation requires interaction with a second transcription factor, PfAP2-I [105]. The cell surface receptors involved in signal transduction and the signaling cascade leading to the activation of ApiAp2-G transcription factors still need to be characterized.
While the P. falciparum genome is predominantly acetylated during the asexual developmental cycle, ApiAP2-G appears to be repressed, and its locus was shown to be associated with histone heterochromatin marks, including H3K4ac, H3K9me1, H3K36me2, PfHP1, and the gametocyte essential factor PfAP2-G5 [106]. PfHP1-dependent gene silencing is antagonized by a sexual commitment activator protein called PfGDV1 (Plasmodium falciparum gametocyte development 1) [107,108]. PfGDV1 targets heterochromatin and triggers the eviction of PfHP1. Interestingly, PfGDV1 activation is, in turn, controlled by a multi-exon long non-coding gdv1 antisense RNA (Pfgdv1asRNA) that initiates downstream of the gdv1 locus [107].
The gametocyte essential factor AP2-G5 further prevents sexual commitment [108]. Indeed, PfAP2-G5 binds upstream of the Pfap2-g locus and other exogenic regions, hence suppressing ap2-g expression [108]. Although findings strongly suggest that ApiAP2 genes are under epigenetic control, histone post-translational modification linked to parasite conversion remains to be characterized. In addition, the signaling cascade events leading to the removal of the repressive state by PfGDV1 and the activation of PfApiAP2-G remain unknown.
Sexual conversion can occur by two possible routes following specific histone modification marks that promote changes in gene expression and sexual differentiation [105,109]. One route suggests that the decision to undergo gametocyte development occurs during the asexual cycle preceding the erythrocyte reinvasion event that leads to gametocyte formation. In this model, the decision to undergo gametocyte development is linked to PfAP2-G expression in the schizont of the preceding cycle [109]. The second route implies that the decision occurs within the same cycle, during the initial PfAP2-G expression in ring stages [109]. Recently, Li et al. identified an ApiAP2 transcription factor in P. yoelii, AP2-O3, implicated in a gender-specific transcription program [110]. AP2-O3, expressed predominantly in female gametocytes, represses the expression of male-specific genes. Genetic studies have shown that the depletion of either HP1 or the histone deacetylase enzyme 2 (Hda2) resulted in the activation of ap2-g transcription and many heterochromatic genes [110].
Gametocyte development from stage I to stage V is characterized by euchromatic post-translational modifications and an abundance of repressive methylation marks on histone 3. Ealy gametocyte stages (I to III) are characterized by the H3K9me3, H3K27me2, H3K27me3, H3K36me2, H3K37me1, H3R17me1, and H3R17me2 modifications [46,106]. The role of arginine methylation as a key feature for the epigenetic regulation of gametocyte development and maturation was suggested by Von Grüning et al. [106].
Three-dimensional structure analyses of the P. falciparum genome revealed that the localization and interaction of sexual differentiation genes in a repressive center are critical for regulating sexual conversion [44]. Further studies are needed to unravel the machinery regulating genome localization and the signaling pathway governing sexual commitment. The repertoire of environmental triggers for sexual conversion is far from being exhaustive. Characterizing proteins interacting with the histone-associated protein PfHP1 and PfGDV1 will undoubtedly unravel the molecular player involved in parasite conversion.
Table 1. Summary of gene expression mechanisms involved in asexual to early gametocytes formation.
Table 1. Summary of gene expression mechanisms involved in asexual to early gametocytes formation.
Mechanisms Underlying Sexual Conversion: from Asexual Forms to Gametocyte StagesReferencesStudied Species
1. Triggers
Environmental, metabolic, host, and parasite factors
      ● LysophosphatidylcholineBrancucci N.M.B et al., 2017 [98]; Abdi A. et al., 2023 [99] P.f
      ● Host immune responseBruce M.C. et al., 1990 [100];
Nixon C.P. et al., 2018 [101];
P.f
      ● DrugsBarkakaty B.N. et al., 1988 [97];
Buckling A. et al., 1999 [96]
P.f
      ● Parasite factorsAyanful-Torgby R. et al., 2016 [102]; Chawla J. et al., 2023 [103]P.f
2. Genome organization
      ●  
Sexual differentiation genes are clustered in a repressive center
Bunnik J.L. et al., 2018 [29]P.f
3. Transcriptional regulation
      ●  
Transcription factor regulating sexual conversion: ApiAP2-G, ApiAp2-I
Sinha A. et al., 2014 [24];
Josling G.A. et al., 2020 [104]
P.f and P.b
      ●  
Transcription factor regulating gender: AP2-O3 (male-gene repressor expressed in female gametocytes)
Li Z. et al., 2021 [109]P.y
4. Epigenetic regulation
   ❖ Before commitment
        ●  
Histone silencing marks: H3K4ac, H3K9me1, H3K36me2
Jiang L. et al., 2013 [51];
von Gruning H. et al., 2022 [105]
P.f
        ●  
Histone modification enzymes: Hda2
Coleman B.I. et al., 2014 [56] P.f
        ●  
Histone-associated proteins for gene silencing: HP1, AP2-G5
Flueck C. et al., 2009 [61];
Shang X. et al., 2021 [107];
von Gruning H. et al., 2022 [105];
P.f
        ●  
Regulatory antisense RNA: gdv1asRNA
Filarsky M. et al., 2018 [106] P.f
   ❖ Early-stage gametocytes (stage 1 to stage 4)
      ●  
Histone marks for the dissociation of repressive protein HP1: H3S10ph
Hirota T. et al., 2005 [65] -
      ●  
Histone activation marks: H3K9me3, H3K27me2, H3K27me3, H3K36me2, H3K37me1, H3R17me1, and H3R17me2
von Gruning H. et al., 2022 [105]; P.f
      ●  
Histone modification enzymes: presumably MYST and GCN5
Miao J. et al., 2010 [66];
Miao J. et al., 2021 [67]
P.f
      ●  
Histone-associated proteins for: GDV1 (sexual commitment activator)
Filarsky M. et al., 2018 [106]P.f
GAPS:
      ●  
Machinery regulating genome localization remains to be characterized
      ●  
Repertoire of environmental triggers far from being exhaustive
      ●  
Signaling pathway governing sexual commitment remains to be thoroughly characterized
Several environmental and metabolic factors trigger sexual commitment. The differentiation of sexually committed parasites into gametocytes is characterized morphologically by dramatic alterations in parasite size, shape, and flexibility. Specific expression mechanisms govern the differentiation. Histone deacetylase 2 (Hda2); Heterochromatin protein 1 (HP1); General control non-depressible 5 (GCN5); MYST (MOZ, YbF2, Sas2, Tip60-like), Apicomplexan Apetalla 2 (ApiAP2); Gametocyte development 1 (GDV1); antisense RNA (asRNA).
Despite their similitude with human enzymes, histone-modifying enzymes could be potential targets for developing antimalarial drugs with transmission-blocking activities. Stenzel et al. [111] designed, synthesized, and tested the biological activity of thirteen terephthalic-acid-based HDAC inhibitors. A subset of these compounds had moderate activity against P. falciparum gametocytes but showed sub-micromolar transmission-blocking activity against the rodent malaria parasite P. berghei. Hence, histone-modifying enzymes are potential targets to prevent malaria infection in the mosquito.

4. Gamete Development

Establishing a successful malaria infection in the Anopheles mosquito and the subsequent spread of competent parasites depend on the coordinated development of the Plasmodium parasite in the mosquito midgut. The differentiation process in the mosquito midgut, lasting almost 20 h, is defined by the rapid conversion of mature stage V female and male gametocytes into female macrogametes and male microgametes, respectively [6,10,11]. This step is followed by the sexual reproduction between macro- and microgametes and, finally, the conversion of newly formed zygotes into ookinetes [6,11]. The transition in the mosquito midgut constitutes a significant bottleneck in the Plasmodium life cycle as there is a 316-fold loss in parasite abundance from the gametocyte to ookinete stage and a 100-fold loss from the ookinete to oocyst stage [112]. Due to this considerable reduction in the parasite population, the plasmodial midgut stages are attractive targets for developing transmission-blocking interventions. Hence, malaria parasite developmental stages in the mosquito midgut have been extensively studied, and several key mechanisms and proteins orchestrating the differentiation process have been described, providing potential targets for vaccines and drug development [113].
Gamete formation is morphologically characterized by a rounding-up of the cell, followed by the parasitophorous vacuole membrane (PVM) rupture and parasite egress from the erythrocyte [10,114]. The differentiation of male gametocytes into male gametes involves three successive mitotic DNA replications that will produce eight motile microgametes through a process called exflagellation. Upon completion of DNA replication, the microgametes egress from the host cell [10]. In contrast, the differentiation of female gametocytes results in a rounding-up of the parasite with no DNA replication and the emergence of the forming gamete from the infected red blood cell [6]. Specific expression mechanisms govern the differentiation of mature stage V gametocytes into gametes (Table 2).

4.1. Gene Expression Regulation Controlling Gametogenesis

Transcriptomic analysis of P. falciparum gametocytes and gametes revealed that small transcriptome changes characterize gametogenesis compared to other life-stage differentiation processes [115]. Most studies investigating gamete to sporozoite formation rely on Plasmodium berghei, causing rodent malaria. The rodent parasites are not of direct practical concern to humans. However, the easy access to Plasmodium life stages in the mosquito and the practicability for the experimental study of human malaria made them a good model for human malaria.
In Plasmodium berghei, the change in gene expression necessary for female gametogenesis is determined by the translation of a large number of mRNAs maintained in a repressive translational state [116,117]. Indeed, in female gametocytes, many transcripts are synthesized but translationally repressed until needed for macrogamete formation and zygote-to-ookinete transformation. Using the Plasmodium berghei model system, Guerreiro et al. revealed that approximately 50% of the transcriptome is maintained in this translationally repressed state by a messenger ribonucleoprotein (mRNP) composed of 16 major factors, including the RNA helicase DOZI (development of zygote inhibited) and the Sm-like factor CITH (homolog of worm CAR-I and fly Trailer Hitch) [118]. Using an in vitro translation assay, Tarique et al. characterized P. falciparum DOZI (PfDZ50) and demonstrated that the protein inhibits translation [119], suggesting a function similar to that reported for P. berghei female gametogenesis.
The expression of stage-specific genes that are necessary for microgamete formation is regulated by transient and reversible protein phosphorylation followed by the de novo synthesis of genes involved in DNA replication, axoneme assembly and motility, chromatin condensation, cytokinesis, and exflagellation [120,121]. Proteomic studies in P. berghei gametes revealed that proteins implicated in RNA translation, protein biosynthesis, glycolysis, environmental stress response, and tubulin-associated cytoskeleton dynamics are predominantly regulated during gamete formation [122]. Gene expression and protein synthesis regulation during both male and female gametogenesis depend highly on the signaling cascade orchestrated by stage-specific kinases and phosphatases [120,123,124].

4.2. Signaling Cascade Controlling Gametogenesis

Gametocyte activation is stimulated by environmental stimuli, including a temperature drop by approximately 5 °C, the presence of mosquito-derived xanthurenic acid (XA), and an increase in extracellular pH from 7.2 to about 8 [125,126,127]. Plasmodium receptors responsible for sensing temperature drops and binding to XA have not been identified yet. The exposure of P. falciparum gametocytes to XA has been shown to increase cyclic GMP (cGMP) levels in the gametocyte, suggesting parasite guanylyl cyclase (GC) activation [128]. In the rodent malaria parasites, the increase in cGMP eventually leads to an increase in the cytoplasmic calcium level necessary to trigger a calcium-dependent stage-specific gene expression and regulation pathway [129,130,131].
The Plasmodium falciparum genome encodes two guanylyl cyclase proteins (GCα and GCβ) that are only expressed in sexual-stage parasites. Functional studies in P. falciparum revealed that these proteins are active guanylyl cyclases and likely important in gametocyte activation [132]. Pharmacological and genetic studies investigating the signaling cascade of P. falciparum and P. berghei gamete formation have shown that GC activation leads to the formation of a secondary messenger cyclic GMP (cGMP), which in turn activates the protein kinase G (PKG), the unique effector identified in Plasmodium [128,133]. Taylor et al. revealed that the tight regulation of cGMP concentration is critical for P. falciparum gametocyte conversion and that premature high cGMP levels are deleterious for gamete formation [134]. Bennink et al. suggested that activation of PKG leads to the hydrolysis of phosphatidylinositol-(4,5)-bisphosphate (PIP2) by the phosphoinositide-specific phospholipase C (PI-PLC) and the production of two secondary messengers, diacylglycerol (DAG) and inositol triphosphate (IP3) [135,136]. Although IP3 induces the opening of calcium channels on the Plasmodium endoplasmic reticulum membrane, no ortholog of an IP3 receptor channel has yet been identified. A recent biochemical approach of an IP3 affinity chromatography column combined with bioinformatics has revealed a potential transporter associated with multidrug resistance in P. falciparum [137]. Additional work will elucidate the function of this novel protein.
The rapid release of calcium in mature gametocytes mediates a stage-specific calcium-dependent effector pathway in all gametogenesis steps, including gamete formation, gamete egress, microgamete mitotic maturation, and exflagellation [128,133,135]. In Plasmodium, the intracellular calcium level is sensed by calcium-dependent protein kinases (CDPKs) [138]. Macro- and microgamete-specific CDPKs that regulate the expression and activation of genes necessary for converting gametocytes into gametes have been identified [121,139,140].
In mature female gametocytes, the increase in cytoplasmic Ca2+ is sensed by CDPK1 [139]. CDPK1 has been identified as the key protein that regulates the translation of mRNAs in a temporal and stage-specific manner during macrogamete formation [139].
The differentiation of male gametocytes is regulated by a male-specific calcium-dependent protein kinase CDPK4, which initiates DNA replication, axoneme assembly, and cell motility [121,141]. In P. berghei, the lysis of the host cell membrane(s) surrounding the microgametocyte is mediated by CDPK1 [139]. In addition, CDPK1 translationally activates mRNA species in the developing zygote that remain repressed in macrogametes [139]. The cell-division cycle protein 20 (CDC20) appears to regulate male gametocyte mitotic division but not the one during schizogony [142]. Whether CDC20 works with CDPK1 and anaphase-promoting complex 3 (APC3) to modulate chromosome condensation and cytokinesis for microgamete formation is not yet determined [143]. In P. berghei, a histone chaperone protein termed FACT-L (facilitates chromatin transcription), which facilitates chromatin transcription, was shown to be involved in male gametocyte DNA replication and the production of fertile microgametes [144]. The formation of the flagella requires the formation of basal bodies and the assembly of axonemes.
Upon completion of DNA replication, axonemes become motile, facilitating the egress of the microgametes from the host cell. Genetic studies in P. berghei have revealed that axoneme assembly and motility require several proteins, including the stage-specific Actin II [145], the armadillo-repeat motif protein Pf16 [146], the spindle-assembly-related protein SAS6 [147], the SR protein kinase (SRPK) [148], and a gametocyte-specific mitogen-activated protein kinase 2 (MAP2) [149,150]. The Plasmodium falciparum kinome includes four NIMA-related kinases (PfNEK 1 to 4). While PfNEK-1 is expressed in asexual and sexual stages, the mRNA transcripts of PfNEK-2, PfNEk-3, and PfNEK-4 are exclusively expressed in gametocytes. NEK1 and NEK3 have been shown to activate the atypical MAP2 protein through phosphorylation [151,152]. PfNEK-2 and PfNEK-4 are required for meiosis completion in the ookinete [153,154]. A metallo-dependent protein phosphatase, PPM1, also plays an important role in P. berghei male gametocyte exflagellation [124].
Kinases and phosphatases involved in gamete formation can be considered promising targets for drug development. A high-throughput screening of P. falciparum-cGMP-dependent protein kinase identified a thiazole scaffold that kills erythrocytic and sexual-stage parasites [155]. Since the malarian PKG differs from the mammalian PKGs, this scaffold represents a good starting point for developing a new class of antimalarial drugs. Recently, Xitong et al. [156] assessed the activity of 25 phosphatase inhibitors against Plasmodium berghei sexual development and transmissibility to the mosquito. Two compounds from the panel effectively inhibited different development stages, from gametogenesis to ookinete maturation. These examples highlight that in silico modeling and screening combined with in vivo and ex vivo approaches using mouse models and human malaria could help identify parasite kinase and phosphatase inhibitors. The kinases regulating male gamete exflagellation (CDPK4 and atypical MAP-2) and DNA replication (Nek-2, Nek-4) should be of great interest in the screening.
Although rodent malaria parasites are sound model systems to understand P. falciprum gametogenesis, more studies relying on human malaria parasites are necessary to grasp the complexity of the early stages of malaria transmission. Those studies should consider including field studies to better understand the biology of transmission during natural infection.
Table 2. Summary of gene expression and signaling cascades controlling gametogenesis.
Table 2. Summary of gene expression and signaling cascades controlling gametogenesis.
Mechanisms Underlying the Differentiation from Mature Stage V Gametocytes to Gametes ReferencesStudied Species
1. Gene expression characteristics
    ●
Small transcriptome changes between the gametocyte and the gametes
Ngwa C.J. et al., 2013 [114]P.f
    ●
Gene expression and protein synthesis regulation dependent on signaling cascade orchestrated by kinases and phosphatases
Invergo B.M. et al., 2017 [119] P.b
2. Triggers
    ●
Temperature drop by 5 °C and extracellular pH of ~8
Billker O. et al., 1997 [124];P.b
    ●
Xanthurenic acid
Garcia G.E. et al., 1998 [125] P.f and P.g
3. Signaling pathway
    ●
Stimuli lead to increase in cGMP and the activation of PKG;
Mc Robert L. et al., 2008 [127];
Brochet M. et al., 2021 [128]
P.f, P.b, P.y
    ●
PKG leads to the release of Ca2+ from the parasite endoplasmic reticulum and the activation of CDPKs
Wang P.P. et al., 2022 [132] P.b
4. Female gametogenesis
    ●
mRNAs maintained in a repressive translational state by messenger ribonucleoproteins (DOZI+~16 factors + Sm-like factor CITH)
Mair G.R. et al., 2010 [116];
Tarique M. et al., 2013 [118];
Guerreiro A. et al., 2014 [117];
P.b and P.f
    ●
mRNA translation regulator: CDPK1
Sebastian S. et al., 2012 [138] P.b
    ●
Egress from RBC: CDPK1
Invergo B.M. et al., 2017 [119] P.b
5. Male gametogenesis
    ●
Transient and reversible protein phosphorylation and de novo synthesis of effectors involved in DNA replication, axoneme assembly and motility, chromatin condensation, cytokinesis, and exflagellation
Billker O. et al., 2004 [120];
Invergo B.M. et al., 2017 [119]
P.b
    ●
DNA replication, axonemes assembly, and cell motility regulation: CDPK4
Invergo B.M. et al., 2017 [119];
Kumar S. et al., 2021 [140]
P.b and P.f
    ●
Mitotic division regulation: CDPK1, CDC20, APC3, and FACT-L
Guttery D.C. et al., 2012 [141];
Invergo B.M. et al., 2017 [119];
Wall R.J. et al., 2018 [142]
P.b
    ●
Axoneme assembly regulation: Actin II, Pf16, SAS6, SRPK, MAP2, NEK1, NEK3
Dorin D. et al., 2001 [150];
Lye Y.M. et al., 2006 [151];
Straschil U. et al., 2010 [145];
Deligianni E. et al., 2011 [144];
Marques S.R. et al., 2015 [146];
Invergo B.M. et al., 2017 [119]
P.f and P.b
    ●
Exflagellation: protein phosphatase, PPM1
Guttery D.C. et al., 2014 [123] P.b
    ●
Egress from RBC: CDPK1
Invergo B.M. et al., 2017 [119]P.b
GAPS:
    ●
Plasmodium-sensing receptors at the cell surface are unknown
    ●
IP3 receptor channel at the ER membrane has yet to be identified
    ●
Relatively few studies used P. falciparum
Gene expression and protein synthesis regulation during both male and female gametogenesis depend highly on signaling cascades orchestrated by stage-specific kinases and phosphatases. Protein kinase G (PKG); Calcium-dependent protein kinases (CDPKs); Development of zygote inhibited (DOZI); Cell Division Cycle 20 (CDC20); FACT-L (facilitates chromatin transcription L); Armadillo-repeat motif protein (Pf16); Spindle-assembly-related protein 6 (SAS6); SR protein kinase (SRPK); Mitogen-activated protein kinase2 (MAP2); NIMA-related kinases (NEKs); Metallo-dependent protein phosphatase (PPM1).

5. Zygote to Ookinete Development

5.1. Gamete Fusion and Zygote Formation

The zygote results from the fusion between a fertile male microgamete and a female macrogamete. Fertilization of a macrogamete by a microgamete is mediated by stage- and sex-specific proteins synthesized in the respective gamete before the fusion event (Table 3). The proteins involved in gamete fusion also mediate the prerequisite recognition and attachment steps. Genetic studies of P. falciparum and P. berghei genes have revealed that P48/45 and P230, two members of a protein family defined by a disulfide bonding pattern of six conserved cysteine residues, are essential for male gamete fertility and fusion with macrogametes [157,158,159]. In P. falciparum, P48/45 and P230 are localized on the gamete surface and have been shown to form a complex necessary for the fusion of the microgamete to the macrogamete. In vitro and in vivo studies in the mouse model have demonstrated the critical implication of P48/45 and P230 in fertilization. These findings led to their use as targets for transmission-blocking vaccines [113,160]. In macrogamete Pfs47, a paralog of P48/45 was identified [161]. Functional studies demonstrated the expression of Pfs47 exclusively on the surface of female gametes, but the protein did not appear crucial for fertilization [161].
Table 3. Summary of the regulation mechanism in gamete fusion.
Table 3. Summary of the regulation mechanism in gamete fusion.
Mechanisms Underlying Fertilization ReferencesStudied Species
1. Characteristics
Gamete fusion is mediated by stage- and sex-specific proteins synthesized in the respective gametes before the fusion event
2. Proteins mediating fusion
    ●
Gamete surface proteins: P48/45 (male gamete fertilization factor) and P230
Rener J. et al., 1983 [157];
van Dijk, M.R. et al., 2001 [158];
Williamson K.C. et al., 2003 [156]
P.f. and P.b.
    ●
Other factors: HAP2/GCS1, histone chaperone protein FACT
Lui Y. et al., 2008 [161] P.b.
GAPS:
    ●
Are all the players involved in the fusion event determined?
    ●
Is the quaternary structure of the fusion complex fully determined?
Fertilization of a macrogamete by a microgamete is mediated by stage- and sex-specific proteins synthesized in the respective gamete before the fusion event. Hapless 2/Generative cell-specific 1 (HAP2/GCS1); facilitates chromatin transcription (FACT).
The evolutionarily conserved class II gamete fusogen HAP2/GCS1 (hapless 2/Generative-cell-specific 1) has also been implicated in the fusion of the plasma membranes of two haploid gametes [162]. Gene disruption studies of p48/45 or hap2/gcs1 resulted in sterility due to the inability of the male gamete to either attach or fuse to fertile female gametes [159,162]. Studies investigating gamete fusion in Plasmodium revealed that a histone chaperone protein named FACT (facilitates chromatin transcription) plays an essential role in fertilization. However, the mechanism of action of this nuclear protein is still unclear [144].
The fusion of the two gametes requires the prior disassembly of a peculiar organelle located underneath the parasite plasma membrane and is named the inner membrane complex (IMC) [5]. The IMC will be reassembled later to ensure the pellicle’s integrity and parasite polarity [5,163]. The IMC is a membranous scaffold in the Alveolata, a group of diverse unicellular eukaryotes, including Plasmodium spp., Toxoplasma gondii, the ciliates, and dinoflagellates. This membranous patchwork is anchored to the subpellicular microtubule network. It also interacts with various parasite-specific proteins to coordinate parasite morphological changes, the segmentation of daughter cells during asexual replication, and parasite motility. Interestingly, the localization and interaction of many IMC proteins are regulated by post-translational S-palmitoylation mediated by the palmitoyl-acyl-transferase DHHC2 (the ortholog of PfDHHC1) [164]. Inhibitor studies revealed that DHHC2 palmitoylation is critical to zygote differentiation during the initial mosquito infection with P. berghei [165]. Hence, stage-specific palmitoylation enzymes could be novel targets for disrupting IMC assembly and zygote formation and differentiation into ookinetes.
The concept of transmission-blocking immunity is mostly antibody-mediated. Therefore, the development of transmission-blocking vaccines (TBVs) focuses on inducing potent antibodies sustained at adequate levels. Extensive efforts toward the clinical development of P. falciparum TBVs are undertaken, but the functional activities associated with most antibodies remain modest. Further study must be conducted to determine the players involved in gamete fusion and their association with each other to orchestrate gamete fusion. The disassembly of the IMC is a pre-requisite for gamete fusion, and targeting the enzymes modulating this biological process could be an approach to interfere with the establishment of the infection in the mosquito gut. Plasmodium falciparum is predicted to encode 12 putative palmitoyl acyl-transferases thought to ensure lipid-based palmitoylation of parasite proteins and act as a biological rheostat for protein–protein interactions and subcellular trafficking [166]. Inhibitors of palmitoylation enzymes could constitute a new class of antimalarial drugs targeting multiple parasite life stages and other apicomplexan parasites of medical importance [166,167,168].

5.2. Molecular and Genetic Mechanisms of Fertilization

Gene expression studies in P. berghei revealed that zygote development and differentiation into ookinetes require the transcriptional activation of several maternal silenced mRNAs in the zygote. A study investigating the parental contribution of transcripts implicated in zygote differentiation revealed that while inherited maternal mRNAs are activated to drive the early stage of zygote differentiation, the paternal alleles are initially silenced and then reactivated [19,169].
The derepression of a maternal transcript post fertilization will drive significant morphological changes defined by the formation of the IMC and the secretory organelles, as well as the elongation and apical polarization of the differentiating zygote [5,170,171]. The dynamic organization of the IMC and the parasite cytoskeleton is essentially coordinated post-translationally by reversible phosphorylation [139,148,172] and palmitoylation [165,173]. A systematic analysis of the Plasmodium kinome, combined with genetic studies in P. berghei, revealed that protein kinase 7 (PK7) and cyclin-G-associated protein (GAK) are essential in ookinete formation [148,172]. GAK is predicted to regulate clathrin-mediated vesicle trafficking and membrane fusion, suggesting its involvement in the formation of the secretory organelles or the assembly of the IMC [148]. PK7 mutants were blocked in ookinete formation, but the mechanism by which this kinase regulates ookinete development remains undetermined. Reverse genetics studies revealed that the Plasmodium phosphatase PPKL (protein phosphatase with kelch-like domains) is essential during ookinete differentiation and is involved in defining ookinete polarity, pellicle morphology and integrity, and ookinete motility [174]. Plasmodium falciparum DHHC1, a palmitoyl-S-acyl-transferase (PAT) containing a conserved DH(H/Y)C motif, was shown to be exclusively localized to the IMC [165,173]. DHHC1 is apicomplexan-specific and was implicated in ookinete formation and morphogenesis [165,173].
Subsequent regulation of stage-specific gene expression requires a de novo synthesis of ookinete-specific genes by the transcription factor ApiAp2-O [175]. ApiAp2-O appears to be associated with more than 500 genes involved in ookinete development, motility, midgut invasion, and parasite escape from mosquito immunity [78]. The ookinete-specific genes mediate nuclear fusion in the diploid zygote, DNA replication, and meiosis that will produce a motile tetraploid ookinete [176]. The tetraploid state was shown to persist throughout the ookinete stage until the formation of sporozoites in the oocyst [176,177]. Like in eukaryotic model systems, meiosis and cell cycle progression in the Plasmodium parasite are regulated by NIMA-related kinases (Neks). Gene disruption studies in P. berghei revealed that Nek-4 and Nek-2 are abundantly expressed during the gametocyte stages. Nek-4 and Nek-2 are essential for zygote differentiation into the ookinete stage but not for gamete formation and fertilization [153,154]. Given the success in developing drugs targeting human kinases, Plasmodium kinases are attractive targets for the next generation of antimalarials. Indeed, ongoing efforts attempt to characterize Plasmodium kinases while evaluating them as antimalarial drug targets [178,179].

5.3. Formation and Maturation of Ookinete

Ookinete maturation is completed 19 to 36 h post-gametocyte ingestion in the blood meal. The ookinete will quickly exit the midgut lumen via an intracellular or intercellular route [180]. Ookinete motility is regulated by kinases. The cGMP-dependent protein kinase (PKG) pathway activates the gliding motility apparatus on the IMC, and the calcium-dependent protein kinase 3 (CDPK3) mobilizes the intracellularly stored calcium necessary for signaling [181,182]. The traversal of the midgut wall is mediated by an impressive set of microneme proteins discharged by the mature ookinete [183]. Proteomic [183] and genetics studies identified key proteins for midgut traversal, including the circumsporozoite-TRAP-related protein CTRP [184], the membrane-attack ookinete protein (MAOP) [185], the secreted ookinete adhesive protein SOAP [186], the von-Willebrand-Factor-A-domain-related protein WARP [187], the cell-traversal protein for ookinetes and sporozoites CelTOS [188], and the chitinase 1 (CHT1) [189,190]. A summary of the molecular and genetic mechanisms regulating zygote differentiation into ookinetes is provided in Table 4.
The ookinete stage is an attractive target for transmission-blocking strategies as the mosquito immune system naturally kills many ookinetes. Ookinete proteins involved in the attachment and invasion of the midgut epithelial cells are potential vaccine targets. Two ookinetes surface proteins, P25 and P28, that play a role in ookinete adhesion to the midgut and differentiation into oocyst are candidates for transmission-blocking vaccines [113]. Transmission-blocking strategies comprise gametocytocidal drugs, transmission-blocking vaccines (TBV), and the engineering of genetically modified mosquitos refractory to Plasmodium infection [191]. The rationale for transmission-blocking drugs is to promote gametocyte clearance in the human host and to block onward parasite transmission by targeting blood stage parasites [192]. On the contrary, transmission-blocking vaccines aim to induce, in the human host, the production of antibodies against specific proteins accessible in the mosquito gut to prevent mosquitoes from carrying and spreading the parasites. Targeted proteins include surface proteins of Plasmodium stages found in the mosquito [113] or mosquito-specific effectors implicated in the infection [193]. Antibodies will be ingested by the mosquitoes during their blood meal and are expected to interact with the target in the mosquito gut before the ookinete transversal of the midgut wall. Therefore, essential ookinete surface proteins and potentially micronemal proteins secreted by the parasite at the time of gut lumen exit could be targeted by TBV approaches.
Table 4. Summary of the molecular and genetic mechanisms regulating zygote differentiation into ookinete.
Table 4. Summary of the molecular and genetic mechanisms regulating zygote differentiation into ookinete.
Mechanisms Underlying Zygote to Ookinete Formation ReferencesStudied Species
1. Characteristics
Zygote development and differentiation into ookinete require the transcriptional activation of several maternal silenced mRNAs in the zygote Akinosoglou K.A. et al., 2015 [18];
Ukegbu C.V. et al., 2015 [168]
P.b.
2. Gene expression mechanisms
    ●  
Derepression of inherited maternal mRNAs drive the early stage of zygote differentiation
Akinosoglou K.A. et al., 2015 [18] P.b.
    ●  
De novo synthesis of ookinete-specific genes
Janse C.J. et al., 1986 [175] P.b.
3. Transcriptional activation of maternal mRNAs control zygote morphological changes
    ●  
Assembly of the subpellicular membrane complex (IMC) and the subpellicular cytoskeleton
Kono M. et al., 2012 [4]; Volkmann K. et al., 2012 [170]; Poulin B. et al., 2013 [170]P.b.
    ●  
Formation of secretory organelles
Frenal K. et al., 2013 [165]; Kaneko I. et al., 2015 [77]P.b.
    ●  
Cell polarization and morphology changes
Kono M. et al., 2012 [4]; Volkmann K. et al., 2012 [169]; Poulin B. et al., 2013 [170]P.b.
    ●  
Regulatory proteins coordinating morphological changes: Processes essentially coordinated post-translationally by reversible phosphorylation and palmitoylation:
    ●  
Protein kinase 7
Dorin-Semblat D. et al., 2008 [171]; Tewari R. et al., 2010 [147]P.f. and P.b.
  ●  
Cyclin-G-associated protein (GAK)
Tewari R. et al., 2010 [147] P.b.
  ●  
Phosphatase PPKL
Guttery D.S. et al., 2012 [173] P.b.
  ●  
Palmitoylation enzymes DHHC1
Wetzel J. et al., 2015 [172];
Santos J.M. et al., 2015 [164]
P.f. and P.b.
4. Regulation of zygote-specific gene expression
  ●  
Stage-specific transcription factor: ApiAP2-O
Yuda M. et al., 2009 [174];
Kaneko I. et al., 2015 [77]
P.b.
5. Meiosis and cell cycle progression regulators
  ●  
NIMA-related Kinases (NEK4 and NEK2)
Reininger L. et al., 2005 [153];
Reininger L. et al., 2009 [152]
P.f. and P.b.
6. Ookinete maturation
  ●  
Motility regulation: PKG, CDPK3
Ishino T. et al., 2006 [181]; Moon R.W. et al., 2009 [180]; P.b.
  ●  
Midgut wall traversal via micronemal proteins discharge: CTRP, MAOP, SOAP, WARP, CelTOS, CHT1
Dessens J.T et al., 1999 [183]; Vinetz J.M. et al., 2000 [188]; Yuda M. et al., 2001 [186]; Dessens J.T. et al., 2003 [185]; Kadota K. et al., 2004 [184]; Kariu t. et al., 2006 [187]; Viswanath V.K. et al. 2021 [189] P.b., P.g
7. Ookinete differentiation
  ●  
Triggers: extracellular matrix
Kaslow D.C. et al., 1994 [193]-
  ●  
Midgut wall traversal: CTRP, MAOP, SOAP, WARP, CelTOS, CHT1
P.b., P.g
GAPS:
  ●  
Mechanisms triggering ookinete to oocyst development are still unclear
  ●  
Relatively few transcriptomics and proteomics studies assessing changes associated with the early stage of oocyst development
  ●  
Gene regulation mechanisms are understudied
  ●  
Role of abundant antisense transcripts in ookinetes remains to be elucidated
Zygote development and differentiation into ookinetes first require the transcriptional activation of several maternal silenced mRNAs in the zygote. Subsequent regulation of stage-specific gene expression requires de novo synthesis of ookinete-specific genes. Inner Membrane Complex (IMC); Protein kinase 7 (PK7); Cyclin G-associated protein (GAK); Protein phosphatase with kelch-like domains (PPKL); palmitoyl-S-acyl-transferase with conserved DH(H/Y)C motif (DHHC1); NIMA-related kinases (NEK); Apicomplexan Apetalla 2 (ApiAP2); Protein kinase G (PKG); Calcium-dependent protein kinase 3 (CDPK3); Circumsporozoite-TRAP-related protein CTRP; Membrane-attack ookinete protein (MAOP); secreted ookinete adhesive protein (SOAP); von-Willebrand-Factor-A-domain-related protein (WARP); cell-traversal protein for ookinetes and sporozoites (CelTOS); Chitinase 1 (CHT1).
As mentioned above, parasite-specific kinases, whose functions are required for parasite motility, are also good targets for Plasmodium multi-stage drugs.

6. Ookinete to Oocyst Development

Following the traversal of the midgut epithelium, ookinetes settle in the basal lamina surrounding the gut and differentiate into oocysts. Ookinetes can be found in the basal lamina 18–24 h after an infective blood meal [12]. The mechanisms triggering the differentiation process are still unclear, but the extracellular matrices composed of collagen, fibronectin, laminin, and chondroitin sulfate are thought to play a role [194]. With the current proteomic and genomic tools, future studies should further investigate mosquito factors involved in ookinete differentiation.
Compared with other developmental stages, relatively few transcriptomics and proteomics studies assess changes associated with the early stage of oocyst development. Sequencing strand-specific cDNA libraries of seven Plasmodium stages has revealed an abundance of antisense transcripts in gametocytes and ookinetes [195]. These findings suggest that antisense RNA plays a role in gene expression regulation in sexual stages. However, the roles of these antisense transcripts still have to be investigated [195]. Studies investigating the ookinete to oocyst differentiation are required to understand these processes better.

7. Conclusions

Functional studies have uncovered critical processes associated with Plasmodium sexual differentiation within the vertebrate host and the mosquito vector. During intraerythrocytic development, parasites appear to predominantly regulate the expression of functional gene products in a stage-specific manner via epigenetic and transcriptional regulation. When preparing to change hosts, gametocyte formation necessitates post-translational repression mechanisms and epigenetic and transcriptional regulation. The fast differentiation process within the mosquito gut relies on the post-translational modification of proteins by specific kinases, phosphatases, and lipid modifications.
The list of the molecular players involved in Plasmodium sexual differentiation is far from complete. The development of new approaches, including single-cell technologies, will provide greater resolution on transcriptional, epigenomic, and metabolomic variations that regulate parasite differentiation. Current findings suggest that functional genetics data need to be integrated into studies assessing the function of malaria proteins for a better understanding of the developmental processes. Although rodent malaria parasites are good model systems to understand P. falciparum development within the mosquito, more studies relying on human malaria parasites are necessary to grasp the complexity of malaria transmission and infection in the mosquito. Plasmodium functional studies have been carried out using parasites maintained and propagated in vitro. With environmental factors affecting considerable parasite gene expression, additional epigenomic, proteomics, transcriptomic, and metabolomic studies from field isolates are pertinent to better understand the processes that govern parasite differentiation. Finally, these studies will have to target all six malaria species infecting humans to identify conserved and species-specific processes that could be used to discover new vaccine candidates and novel classes of antimalarial drugs. Computer-aided approaches and in silico modeling might be alternatives for developing new classes of antimalarial drugs targeting parasite enzymes.

Funding

This work was supported through the DELTAS Africa Initiative (DELGEME grant 107740/Z/15/Z). The DELTAS Africa Initiative is an independent funding scheme of the African Academy of Sciences (AAS)’s Alliance for Accelerating Excellence in Science in Africa (AESA) and supported by the New Partnership for Africa’s Development Planning and Coordinating Agency (NEPAD Agency) with funding from the Wellcome Trust (DELGEME grant 107740/Z/15/Z) and the UK government. The views expressed in this publication are those of the author(s) and not necessarily those of AAS, NEPAD Agency, Wellcome Trust, or the UK government. Additional support provided by K01HL140285 to AO.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors are unaware of any affiliations, memberships, funding, or financial holdings that might affect the objectivity of this review. The authors declare that there are no conflicts of interest regarding the publication of this review article.

References

  1. World Health Organization. World Malaria Report 2022; 9789240064898; World Health Organization: Geneva, Switzerland, 2022. [Google Scholar]
  2. Alonso, P.L.; Brown, G.; Arevalo-Herrera, M.; Binka, F.; Chitnis, C.; Collins, F.; Doumbo, O.K.; Greenwood, B.; Hall, B.F.; Levine, M.M.; et al. A research agenda to underpin malaria eradication. PLoS Med. 2011, 8, e1000406. [Google Scholar] [CrossRef] [PubMed]
  3. Kamiya, T.; Paton, D.G.; Catteruccia, F.; Reece, S.E. Targeting malaria parasites inside mosquitoes: Ecoevolutionary consequences. Trends Parasitol. 2022, 38, 1031–1040. [Google Scholar] [CrossRef] [PubMed]
  4. Aingaran, M.; Zhang, R.; Law, S.K.; Peng, Z.; Undisz, A.; Meyer, E.; Diez-Silva, M.; Burke, T.A.; Spielmann, T.; Lim, C.T.; et al. Host cell deformability is linked to transmission in the human malaria parasite Plasmodium falciparum. Cell. Microbiol. 2012, 14, 983–993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kono, M.; Herrmann, S.; Loughran, N.B.; Cabrera, A.; Engelberg, K.; Lehmann, C.; Sinha, D.; Prinz, B.; Ruch, U.; Heussler, V.; et al. Evolution and architecture of the inner membrane complex in asexual and sexual stages of the malaria parasite. Mol. Biol. Evol. 2012, 29, 2113–2132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Aikawa, M. Plasmodium: The fine structure of malarial parasites. Exp. Parasitol. 1971, 30, 284–320. [Google Scholar] [CrossRef]
  7. Beri, D.; Balan, B.; Tatu, U. Commit, hide and escape: The story of Plasmodium gametocytes. Parasitology 2018, 145, 1772–1782. [Google Scholar] [CrossRef]
  8. Lasonder, E.; Rijpma, S.R.; van Schaijk, B.C.; Hoeijmakers, W.A.; Kensche, P.R.; Gresnigt, M.S.; Italiaander, A.; Vos, M.W.; Woestenenk, R.; Bousema, T.; et al. Integrated transcriptomic and proteomic analyses of P. falciparum gametocytes: Molecular insight into sex-specific processes and translational repression. Nucleic Acids Res. 2016, 44, 6087–6101. [Google Scholar] [CrossRef] [Green Version]
  9. Silvestrini, F.; Bozdech, Z.; Lanfrancotti, A.; Di Giulio, E.; Bultrini, E.; Picci, L.; Derisi, J.L.; Pizzi, E.; Alano, P. Genome-wide identification of genes upregulated at the onset of gametocytogenesis in Plasmodium falciparum. Mol. Biochem. Parasitol. 2005, 143, 100–110. [Google Scholar] [CrossRef] [Green Version]
  10. Yahiya, S.; Jordan, S.; Smith, H.X.; Gaboriau, D.C.A.; Famodimu, M.T.; Dahalan, F.A.; Churchyard, A.; Ashdown, G.W.; Baum, J. Live-cell fluorescence imaging of microgametogenesis in the human malaria parasite Plasmodium falciparum. PLoS Pathog. 2022, 18, e1010276. [Google Scholar] [CrossRef]
  11. Bannister, L.H.; Sinden, R.E. New knowledge of parasite morphology. Br. Med. Bull. 1982, 38, 141–145. [Google Scholar] [CrossRef]
  12. Sinden, R.E.; Strong, K. An ultrastructural study of the sporogonic development of Plasmodium falciparum in Anopheles gambiae. Trans. R. Soc. Trop. Med. Hyg. 1978, 72, 477–491. [Google Scholar] [CrossRef]
  13. Vlachou, D.; Zimmermann, T.; Cantera, R.; Janse, C.J.; Waters, A.P.; Kafatos, F.C. Real-time, in vivo analysis of malaria ookinete locomotion and mosquito midgut invasion. Cell. Microbiol. 2004, 6, 671–685. [Google Scholar] [CrossRef]
  14. Yoshikawa, Y.; Kimura, S.; Soga, A.; Sugiyama, M.; Ueno, A.; Kondo, H.; Zhu, Z.; Ochiai, K.; Nakayama, K.; Hakozaki, J.; et al. Plasmodium berghei Brca2 is required for normal development and differentiation in mice and mosquitoes. Parasites Vectors 2022, 15, 244. [Google Scholar] [CrossRef]
  15. Howick, V.M.; Russell, A.J.C.; Andrews, T.; Heaton, H.; Reid, A.J.; Natarajan, K.; Butungi, H.; Metcalf, T.; Verzier, L.H.; Rayner, J.C.; et al. The Malaria Cell Atlas: Single parasite transcriptomes across the complete Plasmodium life cycle. Science 2019, 365, eaaw2619. [Google Scholar] [CrossRef] [Green Version]
  16. Lu, X.M.; Batugedara, G.; Lee, M.; Prudhomme, J.; Bunnik, E.M.; Le Roch, K.G. Nascent RNA sequencing reveals mechanisms of gene regulation in the human malaria parasite Plasmodium falciparum. Nucleic Acids Res. 2017, 45, 7825–7840. [Google Scholar] [CrossRef] [Green Version]
  17. Bozdech, Z.; Llinás, M.; Pulliam, B.L.; Wong, E.D.; Zhu, J.; DeRisi, J.L. The transcriptome of the intraerythrocytic developmental cycle of Plasmodium falciparum. PLoS Biol. 2003, 1, E5. [Google Scholar] [CrossRef] [Green Version]
  18. Walzer, K.A.; Kubicki, D.M.; Tang, X.; Chi, J.T. Single-cell analysis reveals distinct gene expression and heterogeneity in male and female Plasmodium falciparum gametocytes. mSphere 2018, 3, e00130-18. [Google Scholar] [CrossRef] [Green Version]
  19. Akinosoglou, K.A.; Bushell, E.S.; Ukegbu, C.V.; Schlegelmilch, T.; Cho, J.S.; Redmond, S.; Sala, K.; Christophides, G.K.; Vlachou, D. Characterization of Plasmodium developmental transcriptomes in Anopheles gambiae midgut reveals novel regulators of malaria transmission. Cell. Microbiol. 2015, 17, 254–268. [Google Scholar] [CrossRef] [Green Version]
  20. Tripathi, J.; Zhu, L.; Nayak, S.; Stoklasa, M.; Bozdech, Z. Stochastic expression of invasion genes in Plasmodium falciparum schizonts. Nat. Commun. 2022, 13, 3004. [Google Scholar] [CrossRef]
  21. Miao, J.; Chen, Z.; Wang, Z.; Shrestha, S.; Li, X.; Li, R.; Cui, L. Sex-Specific Biology of the Human Malaria Parasite Revealed from the Proteomes of Mature Male and Female Gametocytes. Mol. Cell. Proteom. 2017, 16, 537–551. [Google Scholar] [CrossRef] [Green Version]
  22. Johnson, N.; Philip, N. Beyond phosphorylation: Putative roles of post-translational modifications in Plasmodium sexual stages. Mol. Biochem. Parasitol. 2021, 245, 111406. [Google Scholar] [CrossRef] [PubMed]
  23. Real, E.; Howick, V.M.; Dahalan, F.A.; Witmer, K.; Cudini, J.; Andradi-Brown, C.; Blight, J.; Davidson, M.S.; Dogga, S.K.; Reid, A.J.; et al. A single-cell atlas of Plasmodium falciparum transmission through the mosquito. Nat. Commun. 2021, 12, 3196. [Google Scholar] [CrossRef] [PubMed]
  24. Kafsack, B.F.; Rovira-Graells, N.; Clark, T.G.; Bancells, C.; Crowley, V.M.; Campino, S.G.; Williams, A.E.; Drought, L.G.; Kwiatkowski, D.P.; Baker, D.A.; et al. A transcriptional switch underlies commitment to sexual development in malaria parasites. Nature 2014, 507, 248–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Sinha, A.; Hughes, K.R.; Modrzynska, K.K.; Otto, T.D.; Pfander, C.; Dickens, N.J.; Religa, A.A.; Bushell, E.; Graham, A.L.; Cameron, R.; et al. A cascade of DNA-binding proteins for sexual commitment and development in Plasmodium. Nature 2014, 507, 253–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Gardner, M.J.; Hall, N.; Fung, E.; White, O.; Berriman, M.; Hyman, R.W.; Carlton, J.M.; Pain, A.; Nelson, K.E.; Bowman, S.; et al. Genome sequence of the human malaria parasite Plasmodium falciparum. Nature 2002, 419, 498–511. [Google Scholar] [CrossRef] [Green Version]
  27. Hall, N.; Karras, M.; Raine, J.D.; Carlton, J.M.; Kooij, T.W.A.; Berriman, M.; Florens, L.; Janssen, C.S.; Pain, A.; Christophides, G.K.; et al. A comprehensive survey of the Plasmodium life cycle by genomic, transcriptomic, and proteomic analyses. Science 2005, 307, 82–86. [Google Scholar] [CrossRef]
  28. Liu, X.; Wang, Y.; Liang, J.; Wang, L.; Qin, N.; Zhao, Y.; Zhao, G. In-depth comparative analysis of malaria parasite genomes reveals protein-coding genes linked to human disease in Plasmodium falciparum genome. BMC Genom. 2018, 19, 312. [Google Scholar] [CrossRef] [Green Version]
  29. Escalante, A.A.; Cepeda, A.S.; Pacheco, M.A. Why Plasmodium vivax and Plasmodium falciparum are so different? A tale of two clades and their species diversities. Malar. J. 2022, 21, 139. [Google Scholar] [CrossRef]
  30. Bunnik, E.M.; Venkat, A.; Shao, J.; McGovern, K.E.; Batugedara, G.; Worth, D.; Prudhomme, J.; Lapp, S.A.; Andolina, C.; Ross, L.S.; et al. Comparative 3D genome organization in apicomplexan parasites. Proc. Natl. Acad. Sci. USA 2019, 116, 3183–3192. [Google Scholar] [CrossRef] [Green Version]
  31. Oresegun, D.R.; Thorpe, P.; Benavente, E.D.; Campino, S.; Muh, F.; Moon, R.W.; Clark, T.G.; Cox-Singh, J. De novo assembly of Plasmodium knowlesi genomes from clinical samples explains the counterintuitive intrachromosomal organization of variant SICAvar and kir multiple gene family members. Front. Genet. 2022, 13, 855052. [Google Scholar] [CrossRef]
  32. Batugedara, G.; Le Roch, K.G. Unraveling the 3D genome of human malaria parasites. Semin. Cell. Dev. Biol. 2019, 90, 144–153. [Google Scholar] [CrossRef]
  33. Swearingen, K.E.; Lindner, S.E. Plasmodium parasites viewed through proteomics. Trends Parasitol. 2018, 34, 945–960. [Google Scholar] [CrossRef]
  34. Lee, H.J.; Georgiadou, A.; Otto, T.D.; Levin, M.; Coin, L.J.; Conway, D.J.; Cunnington, A.J. Transcriptomic Studies of Malaria: A Paradigm for Investigation of Systemic Host-Pathogen Interactions. Microbiol. Mol. Biol. Rev. 2018, 82, e00071-17. [Google Scholar] [CrossRef] [Green Version]
  35. Yin, S.; Fan, Y.; He, X.; Wei, G.; Wen, Y.; Zhao, Y.; Shi, M.; Wei, J.; Chen, H.; Han, J.; et al. The cryptic unstable transcripts are associated with developmentally regulated gene expression in blood-stage Plasmodium falciparum. RNA Biol. 2020, 17, 828–842. [Google Scholar] [CrossRef]
  36. Liu, Z.; Miao, J.; Cui, L. Gametocytogenesis in malaria parasite: Commitment, development and regulation. Future Microbiol. 2011, 6, 1351–1369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Sinden, R.E. Malaria, sexual development and transmission: Retrospect and prospect. Parasitology 2009, 136, 1427–1434. [Google Scholar] [CrossRef] [PubMed]
  38. Singh, M.; Suryanshu; Kanika; Singh, G.; Dubey, A.; Chaitanya, R.K. Plasmodium’s journey through the Anopheles mosquito: A comprehensive review. Biochimie 2021, 181, 176–190. [Google Scholar] [CrossRef] [PubMed]
  39. Usui, M.; Williamson, K.C. Stressed out about Plasmodium falciparum gametocytogenesis. Front. Cell. Infect. Microbiol. 2021, 11, 790067. [Google Scholar] [CrossRef]
  40. Kensche, P.R.; Hoeijmakers, W.A.; Toenhake, C.G.; Bras, M.; Chappell, L.; Berriman, M.; Bartfai, R. The nucleosome landscape of Plasmodium falciparum reveals chromatin architecture and dynamics of regulatory sequences. Nucleic Acids Res. 2016, 44, 2110–2124. [Google Scholar] [CrossRef] [Green Version]
  41. Silberhorn, E.; Schwartz, U.; Loffler, P.; Schmitz, S.; Symelka, A.; de Koning-Ward, T.; Merkl, R.; Langst, G. Plasmodium falciparum nucleosomes exhibit reduced stability and lost sequence dependent nucleosome positioning. PLoS Pathog. 2016, 12, e1006080. [Google Scholar] [CrossRef] [Green Version]
  42. Lanzer, M.; Wertheimer, S.P.; de Bruin, D.; Ravetch, J.V. Chromatin structure determines the sites of chromosome breakages in Plasmodium falciparum. Nucleic Acids Res. 1994, 22, 3099–3103. [Google Scholar] [CrossRef] [Green Version]
  43. Ruiz, J.L.; Tena, J.J.; Bancells, C.; Cortes, A.; Gomez-Skarmeta, J.L.; Gomez-Diaz, E. Characterization of the accessible genome in the human malaria parasite Plasmodium falciparum. Nucleic Acids Res. 2018, 46, 9414–9431. [Google Scholar] [CrossRef] [Green Version]
  44. Bunnik, E.M.; Cook, K.B.; Varoquaux, N.; Batugedara, G.; Prudhomme, J.; Cort, A.; Shi, L.; Andolina, C.; Ross, L.S.; Brady, D.; et al. Changes in genome organization of parasite-specific gene families during the Plasmodium transmission stages. Nat. Commun. 2018, 9, 1910. [Google Scholar] [CrossRef] [Green Version]
  45. Connacher, J.; von Gruning, H.; Birkholtz, L. Histone Modification Landscapes as a Roadmap for Malaria Parasite Development. Front. Cell. Dev. Biol. 2022, 10, 848797. [Google Scholar] [CrossRef]
  46. Coetzee, N.; Sidoli, S.; van Biljon, R.; Painter, H.; Llinas, M.; Garcia, B.A.; Birkholtz, L.M. Quantitative chromatin proteomics reveals a dynamic histone post-translational modification landscape that defines asexual and sexual Plasmodium falciparum parasites. Sci. Rep. 2017, 7, 607. [Google Scholar] [CrossRef] [Green Version]
  47. Witmer, K.; Fraschka, S.A.; Vlachou, D.; Bartfai, R.; Christophides, G.K. An epigenetic map of malaria parasite development from host to vector. Sci. Rep. 2020, 10, 6354. [Google Scholar] [CrossRef] [Green Version]
  48. Toenhake, C.G.; Fraschka, S.A.; Vijayabaskar, M.S.; Westhead, D.R.; van Heeringen, S.J.; Bartfai, R. Chromatin Accessibility-Based Characterization of the Gene Regulatory Network Underlying Plasmodium falciparum Blood-Stage Development. Cell Host Microbe 2018, 23, 557–569. [Google Scholar] [CrossRef] [Green Version]
  49. Tang, J.; Chisholm, S.A.; Yeoh, L.M.; Gilson, P.R.; Papenfuss, A.T.; Day, K.P.; Petter, M.; Duffy, M.F. Histone modifications associated with gene expression and genome accessibility are dynamically enriched at Plasmodium falciparum regulatory sequences. Epigenet. Chromatin 2020, 13, 50. [Google Scholar] [CrossRef]
  50. Gupta, A.P.; Chin, W.H.; Zhu, L.; Mok, S.; Luah, Y.H.; Lim, E.H.; Bozdech, Z. Dynamic epigenetic regulation of gene expression during the life cycle of malaria parasite Plasmodium falciparum. PLoS Pathog. 2013, 9, e1003170. [Google Scholar] [CrossRef]
  51. 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 2009, 5, 179–190. [Google Scholar] [CrossRef] [Green Version]
  52. Jiang, L.; Mu, J.; Zhang, Q.; Ni, T.; Srinivasan, P.; Rayavara, K.; Yang, W.; Turner, L.; Lavstsen, T.; Theander, T.G.; et al. PfSETvs methylation of histone H3K36 represses virulence genes in Plasmodium falciparum. Nature 2013, 499, 223–227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Salcedo-Amaya, A.M.; van Driel, M.A.; Alako, B.T.; Trelle, M.B.; van den Elzen, A.M.G.; Cohen, A.M.; Janssen-Megens, E.M.; van de Vegte-Bolmer, M.; Selzer, R.R.; Iniguez, A.L.; et al. Dynamic histone H3 epigenome marking during the intraerythrocytic cycle of Plasmodium falciparum. Proc. Natl. Acad. Sci. USA 2009, 106, 9655–9660. [Google Scholar] [CrossRef] [PubMed]
  54. Fraschka, S.A.; Filarsky, M.; Hoo, R.; Niederwieser, I.; Yam, X.Y.; Brancucci, N.M.B.; Mohring, F.; Mushunje, A.T.; Huang, X.; Christensen, P.R.; et al. Comparative Heterochromatin Profiling Reveals Conserved and Unique Epigenome Signatures Linked to Adaptation and Development of Malaria Parasites. Cell Host Microbe 2018, 23, 407–420.e408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Kanyal, A.; Rawat, M.; Gurung, P.; Choubey, D.; Anamika, K.; Karmodiya, K. Genome-wide survey and phylogenetic analysis of histone acetyltransferases and histone deacetylases of Plasmodium falciparum. FEBS J. 2018, 285, 1767–1782. [Google Scholar] [CrossRef] [Green Version]
  56. Cui, L.; Fan, Q.; Cui, L.; Miao, J. Histone lysine methyltransferases and demethylases in Plasmodium falciparum. Int. J. Parasitol. 2008, 38, 1083–1097. [Google Scholar] [CrossRef] [Green Version]
  57. Coleman, B.I.; Skillman, K.M.; Jiang, R.H.Y.; Childs, L.M.; Altenhofen, L.M.; Ganter, M.; Leung, Y.; Goldowitz, I.; Kafsack, B.F.C.; Marti, M.; et al. A Plasmodium falciparum histone deacetylase regulates antigenic variation and gametocyte conversion. Cell Host Microbe 2014, 16, 177–186. [Google Scholar] [CrossRef] [Green Version]
  58. Tonkin, C.J.; Carret, C.K.; Duraisingh, M.T.; Voss, T.S.; Ralph, S.A.; Hommel, M.; Duffy, M.F.; Silva, L.M.; Scherf, A.; Ivens, A.; et al. Sir2 paralogues cooperate to regulate virulence genes and antigenic variation in Plasmodium falciparum. PLoS Biol. 2009, 7, e84. [Google Scholar] [CrossRef] [Green Version]
  59. Volz, J.C.; Bartfai, R.; Petter, M.; Langer, C.; Josling, G.A.; Tsuboi, T.; Schwach, F.; Baum, J.; Rayner, J.C.; Stunnenberg, H.G.; et al. PfSET10, a Plasmodium falciparum methyltransferase, maintains the active var gene in a poised state during parasite division. Cell Host Microbe 2012, 11, 7–18. [Google Scholar] [CrossRef] [Green Version]
  60. Rubio, J.P.; Thompson, J.K.; Cowman, A.F. The var genes of Plasmodium falciparum are located in the subtelomeric region of most chromosomes. EMBO J. 1996, 15, 4069–4077. [Google Scholar] [CrossRef]
  61. Su, X.Z.; Heatwole, V.M.; Wertheimer, S.P.; Guinet, F.; Herrfeldt, J.A.; Peterson, D.S.; Ravetch, J.A.; Wellems, T.E. The large diverse gene family var encodes proteins involved in cytoadherence and antigenic variation of Plasmodium falciparum-infected erythrocytes. Cell 1995, 82, 89–100. [Google Scholar] [CrossRef] [Green Version]
  62. Flueck, C.; Bartfai, R.; Volz, J.; Niederwieser, I.; Salcedo-Amaya, A.M.; Alako, B.T.; Ehlgen, F.; Ralph, S.A.; Cowman, A.F.; Bozdech, Z.; et al. Plasmodium falciparum heterochromatin protein 1 marks genomic loci linked to phenotypic variation of exported virulence factors. PLoS Pathog. 2009, 5, e1000569. [Google Scholar] [CrossRef] [Green Version]
  63. Perez-Toledo, K.; Rojas-Meza, A.P.; Mancio-Silva, L.; Hernandez-Cuevas, N.A.; Delgadillo, D.M.; Vargas, M.; Martinez-Calvillo, S.; Scherf, A.; Hernandez-Rivas, R. 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. 2009, 37, 2596–2606. [Google Scholar] [CrossRef]
  64. Josling, G.A.; Petter, M.; Oehring, S.C.; Gupta, A.P.; Dietz, O.; Wilson, D.W.; Schubert, T.; Langst, G.; Gilson, P.R.; Crabb, B.S.; et al. A Plasmodium falciparum Bromodomain Protein Regulates Invasion Gene Expression. Cell Host Microbe 2015, 17, 741–751. [Google Scholar] [CrossRef] [Green Version]
  65. Quinn, J.E.; Jeninga, M.D.; Limm, K.; Pareek, K.; Meissgeier, T.; Bachmann, A.; Duffy, M.F.; Petter, M. The Putative Bromodomain Protein PfBDP7 of the Human Malaria Parasite Plasmodium falciparum Cooperates with PfBDP1 in the Silencing of Variant Surface Antigen Expression. Front. Cell. Dev. Biol. 2022, 10, 816558. [Google Scholar] [CrossRef]
  66. Hirota, T.; Lipp, J.J.; Toh, B.H.; Peters, J.M. Histone H3 serine 10 phosphorylation by Aurora B causes HP1 dissociation from heterochromatin. Nature 2005, 438, 1176–1180. [Google Scholar] [CrossRef]
  67. Miao, J.; Fan, Q.; Cui, L.; Li, X.; Wang, H.; Ning, G.; Reese, J.C.; Cui, L. The MYST family histone acetyltransferase regulates gene expression and cell cycle in malaria parasite Plasmodium falciparum. Mol. Microbiol. 2010, 78, 883–902. [Google Scholar] [CrossRef] [Green Version]
  68. Miao, J.; Wang, C.; Lucky, A.B.; Liang, X.; Min, H.; Adapa, S.R.; Jiang, R.; Kim, K.; Cui, L. A unique GCN5 histone acetyltransferase complex controls erythrocyte invasion and virulence in the malaria parasite Plasmodium falciparum. PLoS Pathog. 2021, 17, e1009351. [Google Scholar] [CrossRef]
  69. Coetzee, N.; von Gruning, H.; Opperman, D.; van der Watt, M.; Reader, J.; Birkholtz, L.M. Epigenetic inhibitors target multiple stages of Plasmodium falciparum parasites. Sci. Rep. 2020, 10, 2355. [Google Scholar] [CrossRef] [Green Version]
  70. Brogi, S.; Ramalho, T.C.; Kuca, K.; Medina-Franco, J.L.; Valko, M. In silico Methods for Drug Design and Discovery. Front. Chem. 2020, 8, 612. [Google Scholar] [CrossRef]
  71. Callebaut, I.; Prat, K.; Meurice, E.; Mornon, J.P.; Tomavo, S. Prediction of the general transcription factors associated with RNA polymerase II in Plasmodium falciparum: Conserved features and differences relative to other eukaryotes. BMC Genom. 2005, 6, 100. [Google Scholar] [CrossRef] [Green Version]
  72. Coulson, R.M.; Hall, N.; Ouzounis, C.A. Comparative genomics of transcriptional control in the human malaria parasite Plasmodium falciparum. Genome Res. 2004, 14, 1548–1554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Balaji, S.; Babu, M.M.; Iyer, L.M.; 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. 2005, 33, 3994–4006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Zhang, C.; Li, Z.; Cui, H.; Jiang, Y.; Yang, Z.; Wang, X.; Gao, H.; Liu, C.; Zhang, S.; Su, X.Z.; et al. Systematic CRISPR-Cas9-mediated modifications of Plasmodium yoelii ApiAP2 genes reveal functional insights into parasite development. MBio 2017, 8, e01986-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Campbell, T.L.; De Silva, E.K.; Olszewski, K.L.; Elemento, O.; Llinas, M. Identification and genome-wide prediction of DNA binding specificities for the ApiAP2 family of regulators from the malaria parasite. PLoS Pathog. 2010, 6, e1001165. [Google Scholar] [CrossRef]
  76. Helm, S.; Lehmann, C.; Nagel, A.; Stanway, R.R.; Horstmann, S.; Llinas, M.; Heussler, V.T. Identification and characterization of a liver stage-specific promoter region of the malaria parasite Plasmodium. PLoS ONE 2010, 5, e13653. [Google Scholar] [CrossRef] [Green Version]
  77. Modrzynska, K.; Pfander, C.; Chappell, L.; Yu, L.; Suarez, C.; Dundas, K.; Gomes, A.R.; Goulding, D.; Rayner, J.C.; Choudhary, J.; et al. A Knockout Screen of ApiAP2 Genes Reveals Networks of Interacting Transcriptional Regulators Controlling the Plasmodium Life Cycle. Cell Host Microbe 2017, 21, 11–22. [Google Scholar] [CrossRef] [Green Version]
  78. Kaneko, I.; Iwanaga, S.; Kato, T.; Kobayashi, I.; Yuda, M. Genome-Wide Identification of the Target Genes of AP2-O, a Plasmodium AP2-Family Transcription Factor. PLoS Pathog. 2015, 11, e1004905. [Google Scholar] [CrossRef] [Green Version]
  79. Santos, J.M.; Josling, G.; Ross, P.; Joshi, P.; Orchard, L.; Campbell, T.; Schieler, A.; Cristea, I.M.; Llinas, M. Red Blood Cell Invasion by the Malaria Parasite is Coordinated by the PfAP2-I Transcription Factor. Cell Host Microbe 2017, 21, 731–741.e710. [Google Scholar] [CrossRef] [Green Version]
  80. Martins, R.M.; Macpherson, C.R.; Claes, A.; Scheidig-Benatar, C.; Sakamoto, H.; Yam, X.Y.; Preiser, P.; Goel, S.; Wahlgren, M.; Sismeiro, O.; et al. An ApiAP2 member regulates expression of clonally variant genes of the human malaria parasite Plasmodium falciparum. Sci. Rep. 2017, 7, 14042. [Google Scholar] [CrossRef] [Green Version]
  81. Sierra-Miranda, M.; Vembar, S.S.; Delgadillo, D.M.; Avila-Lopez, P.A.; Herrera-Solorio, A.M.; Lozano Amado, D.; Vargas, M.; Hernandez-Rivas, R. PfAP2Tel, harbouring a non-canonical DNA-binding AP2 domain, binds to Plasmodium falciparum telomeres. Cell. Microbiol. 2017, 19, e12742. [Google Scholar] [CrossRef] [Green Version]
  82. Flueck, C.; Bartfai, R.; Niederwieser, I.; Witmer, K.; Alako, B.T.; Moes, S.; Bozdech, Z.; Jenoe, P.; Stunnenberg, H.G.; Voss, T.S. A major role for the Plasmodium falciparum ApiAP2 protein PfSIP2 in chromosome end biology. PLoS Pathog. 2010, 6, e1000784. [Google Scholar] [CrossRef] [Green Version]
  83. Gissot, M.; Briquet, S.; Refour, P.; Boschet, C.; Vaquero, C. PfMyb1, a Plasmodium falciparum transcription factor, is required for intra-erythrocytic growth and controls key genes for cell cycle regulation. J. Mol. Biol. 2005, 346, 29–42. [Google Scholar] [CrossRef]
  84. Komaki-Yasuda, K.; Okuwaki, M.; Nagata, K.; Kawazu, S.; Kano, S. Identification of a novel and unique transcription factor in the intraerythrocytic stage of Plasmodium falciparum. PLoS ONE 2013, 8, e74701. [Google Scholar] [CrossRef]
  85. Lima, W.R.; Martins, D.C.; Parreira, K.S.; Scarpelli, P.; Santos de Moraes, M.; Topalis, P.; Hashimoto, R.F.; Garcia, C.R.S. Genome-wide analysis of the human malaria parasite Plasmodium falciparum transcription factor PfNF-YB shows interaction with a CCAAT motif. Oncotarget 2017, 8, 113987–114001. [Google Scholar] [CrossRef]
  86. Lima, W.R.; Moraes, M.; Alves, E.; Azevedo, M.F.; Passos, D.O.; Garcia, C.R. The PfNF-YB transcription factor is a downstream target of melatonin and cAMP signalling in the human malaria parasite Plasmodium falciparum. J. Pineal. Res. 2013, 54, 145–153. [Google Scholar] [CrossRef]
  87. Shang, X.; Wang, C.; Fan, Y.; Guo, G.; Wang, F.; Zhao, Y.; Sheng, F.; Tang, J.; He, X.; Yu, X.; et al. Genome-wide landscape of ApiAP2 transcription factors reveals a heterochromatin-associated regulatory network during Plasmodium falciparum blood-stage development. Nucleic Acids Res. 2022, 50, 3413–3431. [Google Scholar] [CrossRef]
  88. Carrington, E.; Cooijmans, R.H.M.; Keller, D.; Toenhake, C.G.; Bartfai, R.; Voss, T.S. The ApiAP2 factor PfAP2-HC is an integral component of heterochromatin in the malaria parasite Plasmodium falciparum. iScience 2021, 24, 102444. [Google Scholar] [CrossRef]
  89. Russell, T.J.; De Silva, E.K.; Crowley, V.M.; Shaw-Saliba, K.; Dube, N.; Josling, G.; Pasaje, C.F.A.; Kouskoumvekaki, I.; Panagiotou, G.; Niles, J.C.; et al. Inhibitors of ApiAP2 protein DNA binding exhibit multistage activity against Plasmodium parasites. PLoS Pathog. 2022, 18, e1010887. [Google Scholar] [CrossRef]
  90. Elsheikha, H.M.; Marra, C.M.; Zhu, X.Q. Epidemiology, Pathophysiology, Diagnosis, and Management of Cerebral Toxoplasmosis. Clin. Microbiol. Rev. 2021, 34, e01986-17. [Google Scholar] [CrossRef]
  91. Gerace, E.; Lo Presti, V.D.M.; Biondo, C. Cryptosporidium Infection: Epidemiology, Pathogenesis, and Differential Diagnosis. Eur. J. Microbiol. Immunol. 2019, 9, 119–123. [Google Scholar] [CrossRef]
  92. Behnke, M.S.; Wootton, J.C.; Lehmann, M.M.; Radke, J.B.; Lucas, O.; Nawas, J.; Sibley, L.D.; White, M.W. Coordinated progression through two subtranscriptomes underlies the tachyzoite cycle of Toxoplasma Gondii. PLoS ONE 2010, 5, e12354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Joice, R.; Nilsson, S.K.; Montgomery, J.; Dankwa, S.; Egan, E.; Morahan, B.; Seydel, K.B.; Bertuccini, L.; Alano, P.; Williamson, K.C.; et al. Plasmodium falciparum transmission stages accumulate in the human bone marrow. Sci. Transl. Med. 2014, 6, 244re245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Josling, G.A.; Llinas, M. Sexual development in Plasmodium parasites: Knowing when it’s time to commit. Nat. Rev. Microbiol. 2015, 13, 573–587. [Google Scholar] [CrossRef] [PubMed]
  95. Nilsson, S.K.; Childs, L.M.; Buckee, C.; Marti, M. Targeting Human Transmission Biology for Malaria Elimination. PLoS Pathog. 2015, 11, e1004871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Meibalan, E.; Marti, M. Biology of Malaria Transmission. Cold Spring Harb. Perspect. Med. 2017, 7, a025452. [Google Scholar] [CrossRef] [Green Version]
  97. Buckling, A.; Ranford-Cartwright, L.C.; Miles, A.; Read, A.F. Chloroquine increases Plasmodium falciparum gametocytogenesis in vitro. Parasitology 1999, 118, 339–346. [Google Scholar] [CrossRef] [Green Version]
  98. Barkakaty, B.N.; Sharma, G.K.; Chakravorty, N.K. Studies on efficacy of treatment with sulfamethoxazole + trimethoprim and sulfalene + pyrimethamine combinations in Plasmodium falciparum malaria of known and unknown resistant status. J. Commun. Dis. 1988, 20, 165–174. [Google Scholar]
  99. Brancucci, N.M.B.; Gerdt, J.P.; Wang, C.; De Niz, M.; Philip, N.; Adapa, S.R.; Zhang, M.; Hitz, E.; Niederwieser, I.; Boltryk, S.D.; et al. Lysophosphatidylcholine Regulates Sexual Stage Differentiation in the Human Malaria Parasite Plasmodium falciparum. Cell 2017, 171, 1532–1544.e15. [Google Scholar] [CrossRef] [Green Version]
  100. Abdi, A.I.; Achcar, F.; Sollelis, L.; Silva-Filho, J.L.; Mwikali, K.; Muthui, M.; Mwangi, S.; Kimingi, H.W.; Orindi, B.; Andisi Kivisi, C.; et al. Plasmodium falciparum adapts its investment into replication versus transmission according to the host environment. Elife 2023, 12, e85140. [Google Scholar] [CrossRef]
  101. Bruce, M.C.; Alano, P.; Duthie, S.; Carter, R. Commitment of the malaria parasite Plasmodium falciparum to sexual and asexual development. Parasitology 1990, 100 Pt 2, 191–200. [Google Scholar] [CrossRef]
  102. Nixon, C.P.; Nixon, C.E.; Michelow, I.C.; Silva-Viera, R.A.; Colantuono, B.; Obeidallah, A.S.; Jha, A.; Dockery, D.; Raj, D.; Park, S.; et al. Antibodies to PfsEGXP, an Early Gametocyte-Enriched Phosphoprotein, Predict Decreased Plasmodium falciparum Gametocyte Density in Humans. J. Infect. Dis. 2018, 218, 1792–1801. [Google Scholar] [CrossRef]
  103. Ayanful-Torgby, R.; Oppong, A.; Abankwa, J.; Acquah, F.; Williamson, K.C.; Amoah, L.E. Plasmodium falciparum genotype and gametocyte prevalence in children with uncomplicated malaria in coastal Ghana. Malar. J. 2016, 15, 592. [Google Scholar] [CrossRef] [Green Version]
  104. Chawla, J.; Goldowitz, I.; Oberstaller, J.; Zhang, M.; Pires, C.V.; Navarro, F.; Sollelis, L.; Wang, C.C.Q.; Seyfang, A.; Dvorin, J.; et al. Phenotypic Screens Identify Genetic Factors Associated with Gametocyte Development in the Human Malaria Parasite Plasmodium falciparum. Microbiol. Spectr. 2023, 11, e0416422. [Google Scholar] [CrossRef]
  105. Josling, G.A.; Russell, T.J.; Venezia, J.; Orchard, L.; van Biljon, R.; Painter, H.J.; Llinas, M. Dissecting the role of PfAP2-G in malaria gametocytogenesis. Nat. Commun. 2020, 11, 1503. [Google Scholar] [CrossRef] [Green Version]
  106. von Gruning, H.; Coradin, M.; Mendoza, M.R.; Reader, J.; Sidoli, S.; Garcia, B.A.; Birkholtz, L.M. A Dynamic and Combinatorial Histone Code Drives Malaria Parasite Asexual and Sexual Development. Mol. Cell. Proteom. 2022, 21, 100199. [Google Scholar] [CrossRef]
  107. Filarsky, M.; Fraschka, S.A.; Niederwieser, I.; Brancucci, N.M.B.; Carrington, E.; Carrio, E.; Moes, S.; Jenoe, P.; Bartfai, R.; Voss, T.S. GDV1 induces sexual commitment of malaria parasites by antagonizing HP1-dependent gene silencing. Science 2018, 359, 1259–1263. [Google Scholar] [CrossRef] [Green Version]
  108. Shang, X.; Shen, S.; Tang, J.; He, X.; Zhao, Y.; Wang, C.; He, X.; Guo, G.; Liu, M.; Wang, L.; et al. A cascade of transcriptional repression determines sexual commitment and development in Plasmodium falciparum. Nucleic Acids Res. 2021, 49, 9264–9279. [Google Scholar] [CrossRef]
  109. Bancells, C.; Llora-Batlle, O.; Poran, A.; Notzel, C.; Rovira-Graells, N.; Elemento, O.; Kafsack, B.F.C.; Cortes, A. Revisiting the initial steps of sexual development in the malaria parasite Plasmodium falciparum. Nat. Microbiol. 2019, 4, 144–154. [Google Scholar] [CrossRef]
  110. Li, Z.; Cui, H.; Guan, J.; Liu, C.; Yang, Z.; Yuan, J. Plasmodium transcription repressor AP2-O3 regulates sex-specific identity of gene expression in female gametocytes. EMBO Rep. 2021, 22, e51660. [Google Scholar] [CrossRef]
  111. Stenzel, K.; Chua, M.J.; Duffy, S.; Antonova-Koch, Y.; Meister, S.; Hamacher, A.; Kassack, M.U.; Winzeler, E.; Avery, V.M.; Kurz, T.; et al. Design and synthesis of terephthalic acid-based histone deacetylase inhibitors with dual-stage anti-Plasmodium activity. Chem. Med. Chem. 2017, 12, 1627–1636. [Google Scholar] [CrossRef]
  112. Vaughan, J.A.; Noden, B.H.; Beier, J.C. Sporogonic development of cultured Plasmodium falciparum in six species of laboratory-reared Anopheles mosquitoes. Am. J. Trop. Med. Hyg. 1994, 51, 233–243. [Google Scholar] [CrossRef] [PubMed]
  113. Sauerwein, R.W.; Bousema, T. Transmission blocking malaria vaccines: Assays and candidates in clinical development. Vaccine 2015, 33, 7476–7482. [Google Scholar] [CrossRef] [Green Version]
  114. Wirth, C.C.; Pradel, G. Molecular mechanisms of host cell egress by malaria parasites. Int. J. Med. Microbiol. 2012, 302, 172–178. [Google Scholar] [CrossRef] [PubMed]
  115. Ngwa, C.J.; Scheuermayer, M.; Mair, G.R.; Kern, S.; Brügl, T.; Wirth, C.C.; Aminake, M.N.; Wiesner, J.; Fischer, R.; Vilcinskas, A.; et al. Changes in the transcriptome of the malaria parasite Plasmodium falciparum during the initial phase of transmission from the human to the mosquito. BMC Genom. 2013, 14, 256. [Google Scholar] [CrossRef] [PubMed]
  116. Mair, G.R.; Braks, J.A.M.; Garver, L.S.; Wiegant, J.C.A.G.; Hall, N.; Dirks, R.W.; Khan, S.M.; Dimopoulos, G.; Janse, C.J.; Waters, A.P. Regulation of sexual development of Plasmodium by translational repression. Science 2006, 313, 667–669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Mair, G.R.; Lasonder, E.; Garver, L.S.; Franke-Fayard, B.M.; Carret, C.K.; Wiegant, J.C.; Dirks, R.W.; Dimopoulos, G.; Janse, C.J.; Waters, A.P. Universal features of post-transcriptional gene regulation are critical for Plasmodium zygote development. PLoS Pathog. 2010, 6, e1000767. [Google Scholar] [CrossRef] [Green Version]
  118. Guerreiro, A.; Deligianni, E.; Santos, J.M.; Silva, P.A.G.C.; Louis, C.; Pain, A.; Janse, C.J.; Franke-Fayard, B.; Carret, C.K.; Siden-Kiamos, I.; et al. Genome-wide RIP-Chip analysis of translational repressor-bound mRNAs in the Plasmodium gametocyte. Genome Biol. 2014, 15, 493. [Google Scholar] [CrossRef] [Green Version]
  119. Tarique, M.; Ahmad, M.; Ansari, A.; Tuteja, R. Plasmodium falciparum DOZI, an RNA helicase interacts with eIF4E. Gene 2013, 522, 46–59. [Google Scholar] [CrossRef]
  120. Invergo, B.M.; Brochet, M.; Yu, L.; Choudhary, J.; Beltrao, P.; Billker, O. sub-minute phosphoregulation of cell cycle systems during Plasmodium gamete formation. Cell. Rep. 2017, 21, 2017–2029. [Google Scholar] [CrossRef] [Green Version]
  121. 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, 503–514. [Google Scholar] [CrossRef] [Green Version]
  122. Garcia, C.H.S.; Depoix, D.; Queiroz, R.M.L.; Souza, J.M.F.; Fontes, W.; de Sousa, M.V.; Santos, M.D.M.; Carvalho, P.C.; Grellier, P.; Charneau, S. Dynamic molecular events associated to Plasmodium berghei gametogenesis through proteomic approach. J. Proteom. 2018, 180, 88–98. [Google Scholar] [CrossRef]
  123. Alonso-Morales, A.; Gonzalez-Lopez, L.; Cazares-Raga, F.E.; Cortes-Martinez, L.; Torres-Monzon, J.A.; Gallegos-Perez, J.L.; Rodriguez, M.H.; James, A.A.; Hernandez-Hernandez Fde, L. Protein phosphorylation during Plasmodium berghei gametogenesis. Exp. Parasitol. 2015, 156, 49–60. [Google Scholar] [CrossRef] [Green Version]
  124. Guttery, D.S.; Poulin, B.; Ramaprasad, A.; Wall, R.J.; Ferguson, D.J.; Brady, D.; Patzewitz, E.M.; Whipple, S.; Straschil, U.; Wright, M.H.; et al. Genome-wide functional analysis of Plasmodium protein phosphatases reveals key regulators of parasite development and differentiation. Cell Host Microbe 2014, 16, 128–140. [Google Scholar] [CrossRef] [Green Version]
  125. Billker, O.; Shaw, M.K.; Margos, G.; Sinden, R.E. The roles of temperature, pH and mosquito factors as triggers of male and female gametogenesis of Plasmodium berghei in vitro. Parasitology 1997, 115 Pt 1, 1–7. [Google Scholar] [CrossRef]
  126. Garcia, G.E.; Wirtz, R.A.; Barr, J.R.; Woolfitt, A.; Rosenberg, R. Xanthurenic acid induces gametogenesis in Plasmodium, the malaria parasite. J. Biol. Chem. 1998, 273, 12003–12005. [Google Scholar] [CrossRef] [Green Version]
  127. Kawamoto, F.; Alejo-Blanco, R.; Fleck, S.L.; Sinden, R.E. Plasmodium berghei: Ionic regulation and the induction of gametogenesis. Exp. Parasitol. 1991, 72, 33–42. [Google Scholar] [CrossRef]
  128. McRobert, L.; Taylor, C.J.; Deng, W.; Fivelman, Q.L.; Cummings, R.M.; Polley, S.D.; Billker, O.; Baker, D.A. Gametogenesis in malaria parasites is mediated by the cGMP-dependent protein kinase. PLoS Biol. 2008, 6, e139. [Google Scholar] [CrossRef] [Green Version]
  129. Brochet, M.; Balestra, A.C.; Brusini, L. cGMP homeostasis in malaria parasites-The key to perceiving and integrating environmental changes during transmission to the mosquito. Mol. Microbiol. 2021, 115, 829–838. [Google Scholar] [CrossRef]
  130. Muhia, D.K.; Swales, C.A.; Deng, W.; Kelly, J.M.; Baker, D.A. The gametocyte-activating factor xanthurenic acid stimulates an increase in membrane-associated guanylyl cyclase activity in the human malaria parasite Plasmodium falciparum. Mol. Microbiol. 2001, 42, 553–560. [Google Scholar] [CrossRef]
  131. Brochet, M.; Collins, M.O.; Smith, T.K.; Thompson, E.; Sebastian, S.; Volkmann, K.; Schwach, F.; Chappell, L.; Gomes, A.R.; Berriman, M.; 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. 2014, 12, e1001806. [Google Scholar] [CrossRef] [Green Version]
  132. Carucci, D.J.; Witney, A.A.; Muhia, D.K.; Warhurst, D.C.; Schaap, P.; Meima, M.; Li, J.L.; Taylor, M.C.; Kelly, J.M.; Baker, D.A. Guanylyl cyclase activity associated with putative bifunctional integral membrane proteins in Plasmodium falciparum. J. Biol. Chem. 2000, 275, 22147–22156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Wang, P.P.; Jiang, X.; Zhu, L.; Zhou, D.; Hong, M.; He, L.; Chen, L.; Yao, S.; Zhao, Y.; Chen, G.; et al. A G-Protein-coupled receptor modulates gametogenesis via PKG-mediated signaling cascade in Plasmodium berghei. Microbiol. Spectr. 2022, 10, e0015022. [Google Scholar] [CrossRef] [PubMed]
  134. Taylor, C.J.; McRobert, L.; Baker, D.A. Disruption of a Plasmodium falciparum cyclic nucleotide phosphodiesterase gene causes aberrant gametogenesis. Mol. Microbiol. 2008, 69, 110–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Bennink, S.; Kiesow, M.J.; Pradel, G. The development of malaria parasites in the mosquito midgut. Cell. Microbiol. 2016, 18, 905–918. [Google Scholar] [CrossRef] [Green Version]
  136. Martin, S.K.; Jett, M.; Schneider, I. Correlation of phosphoinositide hydrolysis with exflagellation in the malaria microgametocyte. J. Parasitol. 1994, 80, 371–378. [Google Scholar] [CrossRef]
  137. Alves, E.; Nakaya, H.; Guimarães, E.; Garcia, C.R.S. Combining IP3 affinity chromatography and bioinformatics reveals a novel protein-IP3 binding site on Plasmodium falciparum MDR1 transporter. Curr. Res. Microb. Sci. 2022, 4, 100179. [Google Scholar] [CrossRef]
  138. Holder, A.A.; Mohd Ridzuan, M.A.; Green, J.L. Calcium dependent protein kinase 1 and calcium fluxes in the malaria parasite. Microbes Infect. 2012, 14, 825–830. [Google Scholar] [CrossRef]
  139. Sebastian, S.; Brochet, M.; Collins, M.O.; Schwach, F.; Jones, M.L.; Goulding, D.; Rayner, J.C.; Choudhary, J.S.; Billker, O. A Plasmodium calcium-dependent protein kinase controls zygote development and transmission by translationally activating repressed mRNAs. Cell Host Microbe 2012, 12, 9–19. [Google Scholar] [CrossRef] [Green Version]
  140. Bansal, A.; Molina-Cruz, A.; Brzostowski, J.; Liu, P.; Luo, Y.; Gunalan, K.; Li, Y.; Ribeiro, J.M.C.; Miller, L.H. PfCDPK1 is critical for malaria parasite gametogenesis and mosquito infection. Proc. Natl. Acad. Sci. USA 2018, 115, 774–779. [Google Scholar] [CrossRef] [Green Version]
  141. Kumar, S.; Haile, M.T.; Hoopmann, M.R.; Tran, L.T.; Michaels, S.A.; Morrone, S.R.; Ojo, K.K.; Reynolds, L.M.; Kusebauch, U.; Vaughan, A.M.; et al. Plasmodium falciparum Calcium-Dependent Protein Kinase 4 is Critical for Male Gametogenesis and Transmission to the Mosquito Vector. mBio 2021, 12, e0257521. [Google Scholar] [CrossRef]
  142. Guttery, D.S.; Ferguson, D.J.; Poulin, B.; Xu, Z.; Straschil, U.; Klop, O.; Solyakov, L.; Sandrini, S.M.; Brady, D.; Nieduszynski, C.A.; et al. A putative homologue of CDC20/CDH1 in the malaria parasite is essential for male gamete development. PLoS Pathog. 2012, 8, e1002554. [Google Scholar] [CrossRef]
  143. Wall, R.J.; Ferguson, D.J.P.; Freville, A.; Franke-Fayard, B.; Brady, D.; Zeeshan, M.; Bottrill, A.R.; Wheatley, S.; Fry, A.M.; Janse, C.J.; et al. Plasmodium APC3 mediates chromosome condensation and cytokinesis during atypical mitosis in male gametogenesis. Sci. Rep. 2018, 8, 5610. [Google Scholar] [CrossRef] [Green Version]
  144. Laurentino, E.C.; Taylor, S.; Mair, G.R.; Lasonder, E.; Bartfai, R.; Stunnenberg, H.G.; Kroeze, H.; Ramesar, J.; Franke-Fayard, B.; Khan, S.M.; et al. Experimentally controlled downregulation of the histone chaperone FACT in Plasmodium berghei reveals that it is critical to male gamete fertility. Cell Microbiol. 2011, 13, 1956–1974. [Google Scholar] [CrossRef] [Green Version]
  145. Deligianni, E.; Morgan, R.N.; Bertuccini, L.; Kooij, T.W.; Laforge, A.; Nahar, C.; Poulakakis, N.; Schuler, H.; Louis, C.; Matuschewski, K.; et al. Critical role for a stage-specific actin in male exflagellation of the malaria parasite. Cell. Microbiol. 2011, 13, 1714–1730. [Google Scholar] [CrossRef]
  146. Straschil, U.; Talman, A.M.; Ferguson, D.J.; Bunting, K.A.; Xu, Z.; Bailes, E.; Sinden, R.E.; Holder, A.A.; Smith, E.F.; Coates, J.C.; et al. The Armadillo repeat protein PF16 is essential for flagellar structure and function in Plasmodium male gametes. PLoS ONE 2010, 5, e12901. [Google Scholar] [CrossRef] [Green Version]
  147. Marques, S.R.; Ramakrishnan, C.; Carzaniga, R.; Blagborough, A.M.; Delves, M.J.; Talman, A.M.; Sinden, R.E. An essential role of the basal body protein SAS-6 in Plasmodium male gamete development and malaria transmission. Cell Microbiol. 2015, 17, 191–206. [Google Scholar] [CrossRef] [Green Version]
  148. Tewari, R.; Straschil, U.; Bateman, A.; Bohme, U.; Cherevach, I.; Gong, P.; Pain, A.; Billker, O. The systematic functional analysis of Plasmodium protein kinases identifies essential regulators of mosquito transmission. Cell Host Microbe 2010, 8, 377–387. [Google Scholar] [CrossRef]
  149. Rangarajan, R.; Bei, A.K.; Jethwaney, D.; Maldonado, P.; Dorin, D.; Sultan, A.A.; Doerig, C. A mitogen-activated protein kinase regulates male gametogenesis and transmission of the malaria parasite Plasmodium berghei. EMBO Rep. 2005, 6, 464–469. [Google Scholar] [CrossRef] [Green Version]
  150. 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, 1253–1263. [Google Scholar] [CrossRef]
  151. Dorin, D.; Le Roch, K.; Sallicandro, P.; Alano, P.; Parzy, D.; Poullet, P.; Meijer, L.; Doerig, C. Pfnek-1, a NIMA-related kinase from the human malaria parasite Plasmodium falciparum. Eur. J. Biochem. 2001, 268, 2600–2608. [Google Scholar] [CrossRef] [Green Version]
  152. Lye, Y.M.; Chan, M.; Sim, T.S. Pfnek3: An atypical activator of a MAP kinase in Plasmodium falciparum. FEBS Lett. 2006, 580, 6083–6092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Reininger, L.; Tewari, R.; Fennell, C.; Holland, Z.; Goldring, D.; Ranford-Cartwright, L.; Billker, O.; Doerig, C. An essential role for the Plasmodium Nek-2 Nima-related protein kinase in the sexual development of malaria parasites. J. Biol. Chem. 2009, 284, 20858–20868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Reininger, L.; Billker, O.; Tewari, R.; Mukhopadhyay, A.; Fennell, C.; Dorin-Semblat, D.; Doerig, C.; Goldring, D.; Harmse, L.; Ranford-Cartwright, L.; et al. A NIMA-related protein kinase is essential for completion of the sexual cycle of malaria parasites. J. Biol. Chem. 2005, 280, 31957–31964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Penzo, M.; de Las Heras-Duena, L.; Mata-Cantero, L.; Diaz-Hernandez, B.; Vazquez-Muniz, M.J.; Ghidelli-Disse, S.; Drewes, G.; Fernandez-Alvaro, E.; Baker, D.A. High-throughput screening of the Plasmodium falciparum cGMP-dependent protein kinase identified a thiazole scaffold which kills erythrocytic and sexual stage parasites. Sci. Rep. 2019, 9, 7005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Jia, X.; Liu, F.; Bai, J.; Zhang, Y.; Cui, L.; Cao, Y.; Luo, E. Phosphatase inhibitors BVT-948 and alexidine dihydrochloride inhibit sexual development of the malaria parasite Plasmodium berghei. Int. J. Parasitol. Drugs Drug Resist. 2022, 19, 81–88. [Google Scholar] [CrossRef] [PubMed]
  157. Williamson, K.C. Pfs230: From malaria transmission-blocking vaccine candidate toward function. Parasite Immunol. 2003, 25, 351–359. [Google Scholar] [CrossRef]
  158. Rener, J.; Graves, P.M.; Carter, R.; Williams, J.L.; Burkot, T.R. Target antigens of transmission-blocking immunity on gametes of Plasmodium falciparum. J. Exp. Med. 1983, 158, 976–981. [Google Scholar] [CrossRef]
  159. van Dijk, M.R.; Janse, C.J.; Thompson, J.; Waters, A.P.; Braks, J.A.; Dodemont, H.J.; Stunnenberg, H.G.; van Gemert, G.J.; Sauerwein, R.W.; Eling, W. A central role for P48/45 in malaria parasite male gamete fertility. Cell 2001, 104, 153–164. [Google Scholar] [CrossRef] [Green Version]
  160. Ko, K.T.; Lennartz, F.; Mekhaiel, D.; Guloglu, B.; Marini, A.; Deuker, D.J.; Long, C.A.; Jore, M.M.; Miura, K.; Biswas, S.; et al. Structure of the malaria vaccine candidate Pfs48/45 and its recognition by transmission blocking antibodies. Nat. Commun. 2022, 13, 5603. [Google Scholar] [CrossRef]
  161. van Schaijk, B.C.; van Dijk, M.R.; van de Vegte-Bolmer, M.; van Gemert, G.J.; van Dooren, M.W.; Eksi, S.; Roeffen, W.F.; Janse, C.J.; Waters, A.P.; Sauerwein, R.W. Pfs47, paralog of the male fertility factor Pfs48/45, is a female specific surface protein in Plasmodium falciparum. Mol. Biochem. Parasitol. 2006, 149, 216–222. [Google Scholar] [CrossRef]
  162. Liu, Y.; Tewari, R.; Ning, J.; Blagborough, A.M.; Garbom, S.; Pei, J.; Grishin, N.V.; Steele, R.E.; Sinden, R.E.; Snell, W.J.; et al. The conserved plant sterility gene HAP2 functions after attachment of fusogenic membranes in Chlamydomonas and Plasmodium gametes. Genes Dev. 2008, 22, 1051–1068. [Google Scholar] [CrossRef] [Green Version]
  163. Patil, H.; Hughes, K.R.; Lemgruber, L.; Philip, N.; Dickens, N.; Starnes, G.L.; Waters, A.P. Zygote morphogenesis but not the establishment of cell polarity in Plasmodium berghei is controlled by the small GTPase, RAB11A. PLoS Pathog. 2020, 16, e1008091. [Google Scholar] [CrossRef]
  164. Tremp, A.Z.; Al-Khattaf, F.S.; Dessens, J.T. Palmitoylation of Plasmodium alveolins promotes cytoskeletal function. Mol. Biochem. Parasitol. 2017, 213, 16–21. [Google Scholar] [CrossRef]
  165. Santos, J.M.; Kehrer, J.; Franke-Fayard, B.; Frischknecht, F.; Janse, C.J.; Mair, G.R. The Plasmodium palmitoyl-S-acyl-transferase DHHC2 is essential for ookinete morphogenesis and malaria transmission. Sci. Rep. 2015, 5, 16034. [Google Scholar] [CrossRef] [Green Version]
  166. Frenal, K.; Tay, C.L.; Mueller, C.; Bushell, E.S.; Jia, Y.; Graindorge, A.; Billker, O.; Rayner, J.C.; Soldati-Favre, D. Global analysis of apicomplexan protein S-acyl transferases reveals an enzyme essential for invasion. Traffic 2013, 14, 895–911. [Google Scholar] [CrossRef]
  167. Tay, C.L.; Jones, M.L.; Hodson, N.; Theron, M.; Choudhary, J.S.; Rayner, J.C. Study of Plasmodium falciparum DHHC palmitoyl transferases identifies a role for PfDHHC9 in gametocytogenesis. Cell. Microbiol. 2016, 18, 1596–1610. [Google Scholar] [CrossRef] [Green Version]
  168. Yadav, P.; Ayana, R.; Garg, S.; Jain, R.; Sah, R.; Joshi, N.; Pati, S.; Singh, S. Plasmodium palmitoylation machinery engineered in E. coli for high-throughput screening of palmitoyl acyl-transferase inhibitors. FEBS Open Bio 2019, 9, 248–264. [Google Scholar] [CrossRef] [Green Version]
  169. Ukegbu, C.V.; Cho, J.-S.; Christophides, G.K.; Vlachou, D. Transcriptional silencing and activation of paternal DNA during Plasmodium berghei zygotic development and transformation to oocyst. Cell. Microbiol. 2015, 17, 1230–1240. [Google Scholar] [CrossRef] [Green Version]
  170. Volkmann, K.; Pfander, C.; Burstroem, C.; Ahras, M.; Goulding, D.; Rayner, J.C.; Frischknecht, F.; Billker, O.; Brochet, M. The alveolin IMC1h is required for normal ookinete and sporozoite motility behaviour and host colonisation in Plasmodium berghei. PLoS ONE 2012, 7, e41409. [Google Scholar] [CrossRef]
  171. Poulin, B.; Patzewitz, E.M.; Brady, D.; Silvie, O.; Wright, M.H.; Ferguson, D.J.; Wall, R.J.; Whipple, S.; Guttery, D.S.; Tate, E.W.; et al. Unique apicomplexan IMC sub-compartment proteins are early markers for apical polarity in the malaria parasite. Biol. Open 2013, 2, 1160–1170. [Google Scholar] [CrossRef] [Green Version]
  172. Dorin-Semblat, D.; Sicard, A.; Doerig, C.; Ranford-Cartwright, L.; Doerig, C. Disruption of the PfPK7 gene impairs schizogony and sporogony in the human malaria parasite Plasmodium falciparum. Eukaryot. Cell 2008, 7, 279–285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Wetzel, J.; Herrmann, S.; Swapna, L.S.; Prusty, D.; John Peter, A.T.; Kono, M.; Saini, S.; Nellimarla, S.; Wong, T.W.; Wilcke, L.; et al. The role of palmitoylation for protein recruitment to the inner membrane complex of the malaria parasite. J. Biol. Chem. 2015, 290, 1712–1728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Guttery, D.S.; Poulin, B.; Ferguson, D.J.; Szoor, B.; Wickstead, B.; Carroll, P.L.; Ramakrishnan, C.; Brady, D.; Patzewitz, E.M.; Straschil, U.; et al. A unique protein phosphatase with kelch-like domains (PPKL) in Plasmodium modulates ookinete differentiation, motility and invasion. PLoS Pathog. 2012, 8, e1002948. [Google Scholar] [CrossRef] [PubMed]
  175. Yuda, M.; Iwanaga, S.; Shigenobu, S.; Mair, G.R.; Janse, C.J.; Waters, A.P.; Kato, T.; Kaneko, I. Identification of a transcription factor in the mosquito-invasive stage of malaria parasites. Mol. Microbiol. 2009, 71, 1402–1414. [Google Scholar] [CrossRef]
  176. Janse, C.J.; Van der Klooster, P.F.; Van der Kaay, H.J.; Van der Ploeg, M.; Overdulve, J.P. Rapid repeated DNA replication during microgametogenesis and DNA synthesis in young zygotes of Plasmodium berghei. Trans. R. Soc. Trop. Med. Hyg. 1986, 80, 154–157. [Google Scholar] [CrossRef]
  177. Janse, C.J.; van der Klooster, P.F.; van der Kaay, H.J.; van der Ploeg, M.; Overdulve, J.P. DNA synthesis in Plasmodium berghei during asexual and sexual development. Mol. Biochem. Parasitol. 1986, 20, 173–182. [Google Scholar] [CrossRef]
  178. Zhang, V.M.; Chavchich, M.; Waters, N.C. Targeting protein kinases in the malaria parasite: Update of an antimalarial drug target. Curr. Top. Med. Chem. 2012, 12, 456–472. [Google Scholar] [CrossRef]
  179. Moolman, C.; Sluis, R.V.; Beteck, R.M.; Legoabe, L.J. An Update on Development of Small-Molecule Plasmodial Kinase Inhibitors. Molecules 2020, 25, 5182. [Google Scholar] [CrossRef]
  180. Baton, L.A.; Ranford-Cartwright, L.C. How do malaria ookinetes cross the mosquito midgut wall? Trends Parasitol. 2005, 21, 22–28. [Google Scholar] [CrossRef]
  181. Moon, R.W.; Taylor, C.J.; Bex, C.; Schepers, R.; Goulding, D.; Janse, C.J.; Waters, A.P.; Baker, D.A.; Billker, O. A cyclic GMP signalling module that regulates gliding motility in a malaria parasite. PLoS Pathog. 2009, 5, e1000599. [Google Scholar] [CrossRef] [Green Version]
  182. Ishino, T.; Orito, Y.; Chinzei, Y.; Yuda, M. A calcium-dependent protein kinase regulates Plasmodium ookinete access to the midgut epithelial cell. Mol. Microbiol. 2006, 59, 1175–1184. [Google Scholar] [CrossRef]
  183. Lal, K.; Prieto, J.H.; Bromley, E.; Sanderson, S.J.; Yates, J.R., 3rd; Wastling, J.M.; Tomley, F.M.; Sinden, R.E. Characterisation of Plasmodium invasive organelles; an ookinete microneme proteome. Proteomics 2009, 9, 1142–1151. [Google Scholar] [CrossRef] [Green Version]
  184. Dessens, J.T.; Beetsma, A.L.; Dimopoulos, G.; Wengelnik, K.; Crisanti, A.; Kafatos, F.C.; SINDEN, R.E. CTRP is essential for mosquito infection by malaria ookinetes. EMBO J. 1999, 18, 6221–6227. [Google Scholar] [CrossRef] [Green Version]
  185. Kadota, K.; Ishino, T.; Matsuyama, T.; Chinzei, Y.; Yuda, M. Essential role of membrane-attack protein in malarial transmission to mosquito host. Proc. Natl. Acad. Sci. USA 2004, 101, 16310–16315. [Google Scholar] [CrossRef]
  186. Dessens, J.T.; Siden-Kiamos, I.; Mendoza, J.; Mahairaki, V.; Khater, E.; Vlachou, D.; Xu, X.J.; Kafatos, F.C.; Louis, C.; Dimopoulos, G.; et al. SOAP, a novel malaria ookinete protein involved in mosquito midgut invasion and oocyst development. Mol. Microbiol. 2003, 49, 319–329. [Google Scholar] [CrossRef]
  187. Yuda, M.; Yano, K.; Tsuboi, T.; Torii, M.; Chinzei, Y. von Willebrand Factor A domain-related protein, a novel microneme protein of the malaria ookinete highly conserved throughout Plasmodium parasites. Mol. Biochem. Parasitol. 2001, 116, 65–72. [Google Scholar] [CrossRef]
  188. Kariu, T.; Ishino, T.; Yano, K.; Chinzei, Y.; Yuda, M. CelTOS, a novel malarial protein that mediates transmission to mosquito and vertebrate hosts. Mol. Microbiol. 2006, 59, 1369–1379. [Google Scholar] [CrossRef]
  189. Vinetz, J.M.; Valenzuela, J.G.; Specht, C.A.; Aravind, L.; Langer, R.C.; Ribeiro, J.M.; Kaslow, D.C. Chitinases of the avian malaria parasite Plasmodium gallinaceum, a class of enzymes necessary for parasite invasion of the mosquito midgut. J. Biol. Chem. 2000, 275, 10331–10341. [Google Scholar] [CrossRef] [Green Version]
  190. Viswanath, V.K.; Gore, S.T.; Valiyaparambil, A.; Mukherjee, S.; Lakshminarasimhan, A. Plasmodium chitinases: Revisiting a target of transmission-blockade against malaria. Protein. Sci. 2021, 30, 1493–1501. [Google Scholar] [CrossRef]
  191. Hoermann, A.; Habtewold, T.; Selvaraj, P.; Del Corsano, G.; Capriotti, P.; Inghilterra, M.G.; Kebede, T.M.; Christophides, G.K.; Windbichler, N. Gene drive mosquitoes can aid malaria elimination by retarding Plasmodium sporogonic development. Sci. Adv. 2022, 8, eabo1733. [Google Scholar] [CrossRef]
  192. Burrows, J.N.; Duparc, S.; Gutteridge, W.E.; Hooft van Huijsduijnen, R.; Kaszubska, W.; Macintyre, F.; Mazzuri, S.; Mohrle, J.J.; Wells, T.N.C. New developments in anti-malarial target candidate and product profiles. Malar. J. 2017, 16, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Dinglasan, R.R.; Kalume, D.E.; Kanzok, S.M.; Ghosh, A.K.; Muratova, O.; Pandey, A.; Jacobs-Lorena, M. Disruption of Plasmodium falciparum development by antibodies against a conserved mosquito midgut antigen. Proc. Natl. Acad. Sci. USA 2007, 104, 13461–13466. [Google Scholar] [CrossRef] [PubMed]
  194. Kaslow, D.C.; Nussenzweig, V.; Miller, L. Meeting on Parasites and the invertebrate vector. John D and Catherine T MacArthur Foundation, 18–21 November, 1993. Mem. Inst. Oswaldo. Cruz. 1994, 89, 279–295. [Google Scholar] [CrossRef]
  195. López-Barragán, M.J.; Lemieux, J.; Quiñones, M.; Williamson, K.C.; Molina-Cruz, A.; Cui, K.; Barillas-Mury, C.; Zhao, K.; Su, X.-Z. Directional gene expression and antisense transcripts in sexual and asexual stages of Plasmodium falciparum. BMC Genom. 2011, 12, 587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The life cycle of Plasmodium falciparum parasites.
Figure 1. The life cycle of Plasmodium falciparum parasites.
Microorganisms 11 01966 g001
Figure 2. General nuclear organization, epigenetic, and transcriptional mechanisms employed by Plasmodium to regulate gene expression.
Figure 2. General nuclear organization, epigenetic, and transcriptional mechanisms employed by Plasmodium to regulate gene expression.
Microorganisms 11 01966 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ouologuem, D.T.; Dara, A.; Kone, A.; Ouattara, A.; Djimde, A.A. Plasmodium falciparum Development from Gametocyte to Oocyst: Insight from Functional Studies. Microorganisms 2023, 11, 1966. https://doi.org/10.3390/microorganisms11081966

AMA Style

Ouologuem DT, Dara A, Kone A, Ouattara A, Djimde AA. Plasmodium falciparum Development from Gametocyte to Oocyst: Insight from Functional Studies. Microorganisms. 2023; 11(8):1966. https://doi.org/10.3390/microorganisms11081966

Chicago/Turabian Style

Ouologuem, Dinkorma T., Antoine Dara, Aminatou Kone, Amed Ouattara, and Abdoulaye A. Djimde. 2023. "Plasmodium falciparum Development from Gametocyte to Oocyst: Insight from Functional Studies" Microorganisms 11, no. 8: 1966. https://doi.org/10.3390/microorganisms11081966

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop