<<

NUCLEAR ARCHITECTURE, DYNAMIC SPATIAL POSITION

AND CHROMATIN MODIFICATIONS ASSOCIATED TO

GENE REGULATION IN PLASMODIUM FALCIPARUM

Liliana Mancio Silva

Nuclear Architecture, Dynamic Spatial Position and Chromatin Modifications

associated to Regulation in Plasmodium falciparum

Liliana Mancio Silva

Dissertation to obtain the Ph.D degree submitted to the Faculty of Pharmacy, University of Porto

Porto 2008

Supervisor: Prof. Artur Scherf (Institut Pasteur) Co-Supervisor: Prof. Anabela Cordeiro da Silva (Faculdade de Farmácia, U. Porto)

DECLARATION

Under the terms of the Decree-Law nº 216/92, October 13th, it is stated that the following original articles (published or submitted to publication) are an integral part of this thesis.

I. Figueiredo LM, Rocha EP, Mancio-Silva L, Prevost C, Hernandez-Verdun D, Scherf A. The unusually large Plasmodium reverse-transcriptase localizes in a discrete compartment associated with the nucleolus. Nucleic Acids Res. 2005 Feb 18; 33(3): 1111-22.

II. Mancio-Silva L, Rojas-Meza AP, Vargas M, Scherf A, Hernandez-Rivas R. Differential association of Orc1 and Sir2 to telomeric domains in Plasmodium falciparum. J Cell Sci. 2008 Jun 15; 121(Pt 12): 2046-53.

III. Perez-Toledo K, Rojas-Meza AP, Mancio-Silva L, Hernandez-Cuevas NA, Delgadillo DM, Vargas M, Martinez-Calvillo S, Scherf A, Hernandez-Rivas R. Plasmodium falciparum heterochromatin 1 binds to trimethylated histone H3 lysine 9 and is a new component of the subtelomeric chromatin. Submitted.

IV. Issar N, Ralph S, Mancio-Silva L, Keeling C and Scherf A. Differential sub-nuclear localization of repressive and active histone methyl modifications in P. falciparum. Submitted.

V. Lopez-Rubio JJ, Mancio-Silva L, Scherf A. Genome-wide localization of heterochromatin couples clonally variant gene regulation to perinuclear epigenetic factories in malaria parasites. Submitted.

According to the aforementioned law, I confirm having actively participated in the conceptual design and technical execution of the work, interpretation of the results and manuscript preparation of the listed articles, in association with the co-authors.

v

ABSTRACT

In this thesis, we investigated fundamental aspects of the nuclear and biology associated to gene regulation in the protozoan P. falciparum malaria parasites. Plasmodium have the capacity of recruiting proteins that facilitate the spreading of heterochromatin into subtelomeric DNA regions and contribute to the default silencing of virulence gene families located herein. Here, we identified and characterized two telomere binding proteins, P. falciparum origin-recognition-complex 1 (PfOrc1) and P. falciparum heterochromatin protein 1 (PfHP1), that might work in concert with P. falciparum silent information regulator 2 (PfSir2) in telomeric heterochromatin formation and silencing. PfOrc1, a conserved protein implicated in DNA replication in other eukaryotes, was found to associate mostly to telomeric DNA and to subtelomeric chromatin. In the nucleus, PfOrc1 localizes together with PfSir2 to the nucleolus and telomeric clusters in ring-stage and redistribute into the nucleoplasm and cytoplasm prior to DNA replication in trophozoite-stage. PfHP1, which is the first member of the histone mark reading machinery identified in malaria parasites, binds largely to subtelomeric DNA repeats and localizes to some PfSir2 perinuclear clusters, but not in the nucleolus. It binds specifically to the histone H3 lysine 9 trimethylated (H3K9me3), that in the P. falciparum genome is restricted to telomeric, subtelomeric domains and internal chromosomal regions that similar to subtelomeric regions contain clonally variant multigene families. All enriched in H3K9me3 locate predominantly in the nuclear periphery, suggesting a role for this histone modification in spatial organization of P. falciparum . The P. falciparum H3K9-methyltransferase, PfKMT1, also localizes in the perinuclear region, possibly participating in a specific epigenetic perinuclear silencing complex together with PfSir2, PfOrc1 and PfHP1. Furthermore, we investigated whether the subnuclear position is important for the epigenetic regulation of two differentially transcribed gene families, the var genes and the rRNA genes. We found that all members of var gene family (~60) localize at the nuclear periphery, irrespective of their chromosomal location and activity. The subtelomeric var genes are linked to telomeric clusters, whereas the chromosomal internal var genes can associate to telomeric clusters or form different clusters. However, upon activation of the single var gene, regardless of its chromosomal location, dissociates from the telomeric clusters, suggesting the existence of a particular perinuclear expression zone. Regarding the rRNA genes, we observed that active and silent members, although dispersed in distinct chromosomes, locate in the same compartment, the nucleolus. This subcompartment is placed in the periphery of the nucleus and is excluded from the telomeric clusters. In summary, the work of this thesis has shown various perinuclear subcompartments, defined by several proteins. In addition, we showed that subnuclear compartmentalization is associated to repression and transcription of distinct gene families, demonstrating the importance of spatial nuclear organization in P. falciparum gene regulation.

vii

RÉSUMÉ

Dans cette thèse, nous avons étudié des aspects fondamentaux de la biologie des télomères et du noyau associés à la régulation de gènes du parasite responsable du paludisme, P. falciparum. Les télomères de ce microorganisme ont la capacité de recruter des protéines qui facilitent la propagation de l’hétérochromatine dans les régions subtélomériques, et qui contribuent à la répression par défaut des familles de gènes de virulence situées dans cette région. Nous avons identifié et caractérisé deux protéines télomériques, PfOrc1 (P. falciparum Origin of recognition complex 1) et PfHP1 (P. falciparum Heterochromatin protein 1), qui pourraient travailler de concert avec PfSir2 (P. falciparum Silent information regulator 2) dans la formation de l’hétérochromatine télomérique et dans la répression. PfOrc1, une protéine conservée impliquée dans la réplication de l'ADN dans d'autres eucaryotes, a été trouvé associée à la chromatine télomérique et subtélomérique. Dans le noyau, PfOrc1 colocalise avec PfSir2 dans le nucléole et dans des ‘clusters’ de télomères au stade ‘anneau’, et est redistribuée dans le cytoplasme et le nucléoplasme avant la réplication de l'ADN au stade ‘trophozoite’. PfHP1 est la première protéine pouvant reconnaître des modifications d’histones du parasite du paludisme a avoir été caractérisé. Elle est liée à la chromatine subtélomérique, est localisée dans certains ‘clusters’ de télomères définis par PfSir2 mais pas dans le nucléole. Elle se lie spécifiquement à la lysine 9 triméthylée de l’histone H3 (H3K9me3), qui, dans le génome de P. falciparum, est limitée à des domaines télomériques et subtélomériques, mais aussi à l’intérieur de régions chromosomiques, qui contient des familles multigéniques similaires aux régions subtélomériques. Tout gène enrichit en H3K9me3 est localisé principalement dans la périphérie nucléaire, ce qui suggère un rôle pour cette modification d’histone dans l’organisation des chromosomes de P. falciparum. La méthyltransférase de P. falciparum spécifique de la H3K9, appelée PfKMT1, est également localisée dans la région périnucléaire et peut faire partie, avec PfSir2, PfOrc1 et PfHP1, d’un complexe périnucléaire de protéines associées à la répression épigénétique. Nous avons aussi examiné si la position précise au sein du noyau était importante pour la régulation épigénétique de deux familles de gènes transcrits de manière différentielle: les gènes var et les gènes d'ARNr (ARN ribosomaux). Nous avons constaté que tous les membres de la famille des gènes var sont localisés à la périphérie du noyau, quelque soit leur emplacement chromosomique et leur état transcriptionnel. Les gènes var subtélomériques sont liés à des ‘clusters’ de télomères, tandis que les gènes var situés dans des régions internes des chromosomes peuvent être associés à des ‘clusters’ de télomères ou à des ‘clusters’ indépendants. Toutefois, quand un gène var devient actif, quel que soit son emplacement dans le , il est dissocié des ‘clusters’ de télomères, suggérant l'existence d'une zone d’expression périnucléaire. En ce qui concerne les gènes d'ARNr, nous avons observé que les membres actifs et réprimés, bien que dispersés dans différents chromosomes, sont localisés dans le même compartiment, le nucléole. Ce compartiment est placé dans la périphérie du noyau et est exclu des ‘clusters’ de télomères. Les résultats de cette thèse ont montré plusieurs protéines associées à différentes compartiments périnucléaires. De plus, nous avons montré que le compartimentage du noyau est associé à la répression et à la transcription de différentes familles de gènes, ce qui démontre l'importance de l'organisation nucléaire dans la régulation des gènes chez P. falciparum.

ix

RESUMO

Nesta tese, investigamos vários aspectos fundamentais da biologia nuclear e dos telómeros associados à regulação de genes no parasita agente da malária, P. falciparum. Os telómeros deste parasita têm a capacidade de recrutar proteínas que facilitam a propagação da heterocromatina em regiões subteloméricas, contribuindo para a repressão de famílias multigénicas subteloméricas implicadas na virulência do parasita. Identificamos e caracterizamos duas proteínas teloméricas, PfOrc1 (P. falciparum origin-recognition-complex 1) e PfHP1 (P. falciparum Heterochromatin Protein 1), que poderão actuar juntamente com PfSir2 (P. falciparum Silent information regulator 2) na formação de heterocromatina telomérica e repressão de genes. A proteína PfOrc1, conservada entre os eucariotas e implicada na replicação do DNA, encontra- se associada à cromatina telomérica e subtelomérica. No núcleo, PfOrc1 colocaliza-se com PfSir2 no nucléolo e nos agregados teloméricos em parasitas jovens, e distribuísse no citoplasma e no nucleoplasma antes da replicação do DNA. A proteína PfHP1, a primeira a ser identificada em P. falciparum capaz de reconhecer histonas modificadas, liga-se principalmente à chromatina subtelomérica e está associada com PfSir2 em alguns agregados teloméricos, mas não no nucléolo. Esta proteína liga-se especificamente à lisina 9 trimetilada da histona H3 (H3K9me3). No genoma deste parasita, H3K9me3 está restrita a regiões teloméricas e subteloméricas, assim como regiões internas dos cromossomas que contenham famílias multigénicas semelhantes às que se que encontram em regiões subteloméricas. Todos os genes enriquecidos em H3K9me3 localizam-se predominantemente na periferia do núcleo, sugerindo que H3K9me3 desempenha um papel primordial na organização dos cromossomas no núcleo em P. falciparum. A H3K9- metiltransferase de P. falciparum, PfKMT1 (P. falciparum K-Methyltransferase 1) também está localizada na região perinuclear, estando possivelmente associada a outras proteínas, tais como PfSir2, PfOrc1 e PfHP1, formando um complexo proteico especializado em repressão epigenética. Investigamos também a importância da posição nuclear na regulação epigenética de duas famílias de genes transcritas de maneira diferencial (os genes var e os genes ribossomais). Observamos que todos os membros da família var estão localizados na periferia nuclear, independentemente da sua localização no cromossoma e do seu estado transcricional. Os genes var subteloméricos estão associados aos agregados teloméricos, enquanto que os genes var situados em regiões internas dos cromossomas podem associar-se aos agregados teloméricos ou outro tipo agregados. Uma vez activado, o único gene var (seja ele subtelomérico ou interno) encontra-se dissociado dos agregados teloméricos, o que sugere a existência de uma zona especifica para a transcrição. Em relação aos genes ribossomais, observamos que os membros activos e reprimidos, embora dispersos em cromossomas distintos, estão localizados no mesmo compartimento, o nucléolo. Este compartimento encontra-se na periferia do núcleo, excluído dos agregados teloméricos. Os resultados desta tese mostram que na periferia nuclear existem vários compartimentos, constituídos por diferentes proteínas. Além disso, mostramos que repressão e transcrição de distintas famílias multigénicas está associada à compartimentação do núcleo, o que demonstra a importância da arquitectura nuclear na regulação genética em P. falciparum.

xi

CONTENTS

Page INTRODUCTION……………………………………………………………………………………... 1 The malaria parasite……………………………………………………………………...... 1 Current malaria control strategies……………………………………………………...... 3 P. falciparum genome organization………………………………………………………….... 5 P. falciparum gene regulation………………………………………………………………..... 6 Specific DNA elements and transcription factors………………………………………. 6 Chromatin modifications and transcription……………………………………………… 7 Anti-sense RNAs and post-transcriptional regulation………………………………….. 13 Nuclear organization and telomere position effect……………………………………... 13 var genes: organization and regulation……………………………………………………..... 15 rRNA genes: organization and regulation…………………………………………………..... 18 Aims of this thesis……………………………………………………………………………..... 21

RESULTS……………………………………………………………………………………………... 23 I. The unusually large Plasmodium telomerase reverse-transcriptase localizes in a discrete compartment associated with the nucleolus…………………………………… 25

II. Differential association of Orc1 and Sir2 proteins to telomeric domains in Plasmodium falciparum…………………………………………………………………..... 39

III. P. falciparum heterochromatin protein 1 binds to trimethylated histone H3 lysine 9 and is a new component of the subtelomeric chromatin……………………………….. 49

IV. Differential sub-nuclear localization of repressive and active histone methyl modifications in P. falciparum…………………………………………………………...... 69

V. Genome-wide localization of heterochromatin couples clonally variant gene regulation to perinuclear epigenetic factories in malaria parasites………………...... 83

VI. Nucleolar organization of dispersed rDNA in P. falciparum asexual stages…………. 111

VII. Dynamic chromatin structure of rDNA in P. falciparum depends on the temperature………………………………………………………………………………….. 131

CONCLUDING REMARKS………………………………………………………………………..... 145

ACKNOWLEDGMENTS……………………………………………………………………………... 155

REFERENCES……………………………………………………………………………………….. 157

xiii

ABREVIATIONS

TAREs: Telomere Associated Repetitive Elements TERT: Telomerase Reverse Transcriptase TPE: Telomere Position Effect HP1: Heterochromatin Protein 1 Sir2: Silent information regulator 2 Orc1: Origin-recognition-complex protein 1 HAT: Histone Acetyltransferase HDAC: Histone Deacetylase KMT: Histone lysine (K) Methyltransferase KDM: Histone lysine (K) Demethylase Pol I: RNA I Pol II: RNA Polymerase II Pol III: RNA Polymerase III rRNA: ribosomal RNA rDNA: ribosomal DNA RNAi: RNA interference IFA: Immunofluorescence Assay FISH: Fluorescence in situ hybridization ChIP: Chromatin Immunoprecipitation EMSA: Electrophoretic Mobility Shift Assay RBC: Red Blood Cell AT: Adenine-Thymine GC: Guanidine-Cytosine Mb: Megabase kb: kilobase bp:

xv

INTRODUCTION

The malaria parasite

The term malaria designates an infectious disease produced by protozoan parasites of the genus Plasmodium, phylum Apicomplexa. Four species of plasmodia are infectious to humans: P. falciparum, P. vivax, P. malariae and P. ovale, although only P. falciparum is responsible for fatal disease. Malaria parasites have a complex life cycle, which includes differentiation through several morphologically distinct forms with sexual and asexual replication (Figure 1). Although all four species infectious to the man have a similar life cycle, differences in the duration of the life cycle and in biochemistry are crucial in the pathogenicity of each [reviewed in (Fujioka and Aikawa, 1999)]. Infection of the human host begins with the bite of an infected female Anopheles mosquito that inoculates individual sporozoites. These mobile forms reach the liver and start the exoerythrocytic schizogony. Within the hepatocytes each sporozoite undergoes asexual replication, given rise to tens thousands of haploid merozoites, which enter the blood circulation to invade red blood cells (RBC). After invasion the parasite enlarges its cytoplasm progressively as it develops from a ring-form into a trophozoite. A small proportion of these trophozoites differentiate into micro- or macrogametocytes, the precursors of male and female gametes, respectively. This process is known as gametocytogenesis or gametogony. Mature gametocytes are taken into the mosquito and differentiate into gametes. These gametes fuse and form a diploid zygote, which transforms into ookinete that traverses the epithelial cell of the midgut, and develops into an oocyst. Sporogony within the oocyst produces more than ten thousands sporozoites. The oocyst bursts and sporozoites migrate to the salivary glands for onward transmission into another host.

1

Figure 1. The life cycle of P. falciparum and the approaches that are currently in use or in development for malaria treatment and control. Reduce malaria transmission will require combination of three strategies: drug treatment, vaccination and vector control. In the figure, strategies currently used are indicated in black, in advanced clinical studies are in blue and beginning clinical studies or under research are in red. The currently used antimalarials stem from five drug classes: quinoline derivates (interfere with haemoglobin degradation), antifolates (reduce the folate cofactors essential for nucleotide biosynthesis and amino-acid metabolism), antibiotics (inhibit “prokaryote-like” protein biosynthesis in the apicoplast), artemisinins (seem to inhibit parasite growth though inhibition of PfATP6) and inhibitors of mitochondria electron transport (atovaquone). From the drugs that are in or about to enter clinical development only three compounds have novel mechanisms of action: fosmidomycin (inhibits isoprenoid biosynthesis), T3 (inhibits phospholipid biosynthesis) and pafuramidine (not exactly known) (Ridley, 2002, 2003; Schlitzer, 2008). There are different vaccine strategies against malaria: reduce sporozoite invasion (CSP, TRAP, LSA), prevent high parasitemia (AMA1, MSP1, MSP3), block sequestration (var2CSA, PfEMP1), destruction of infected hepatocytes (GAS) and reduce transmission rate (Pfs25, Pfs28) (Greenwood et al., 2008; Matuschewski and Mueller, 2007; Todryk and Hill, 2007; Waters et al., 2005). In the figure, the vaccine target antigens are in brackets, except for those that are under research. RAS, radiation attenuated sporozoites; GAS, genetically attenuated sporozoites; CSP, circumsporozoite protein; AS02, vaccine adjuvant; FP, fowlpox; MVA, modified vaccina virus Ankara; ME-TRAP, multi-epitope thrombospondin- related adhesive protein; Sad, Simian adenovirus; LSA, liver stage antigen; AdHu35, human adenovirus serotype 35; MSP, merozoite surface protein; AMA, apical membrane antigen; GLURP, glutamate-rich protein; PfEMP1, P. falciparum erythrocyte membrane protein; var2CSA, variant surface antigen 2, chondroitin sulphate A-binding; Pfs25/28, ookinete surface proteins; DDT, dichloro-diphenyl-trichloroethane; GMM, genetically modified mosquitoes (Christophides, 2005). Life cycle modified from ref. (Wirth, 2002).

2

The majority of trophozoites, however, enter the erythrocytic schizogony phase. In this phase multiple rounds of nuclear division without cytokenesis occur to result in the formation of 8 to 32 of haploid merozoites, which are released in the bloodstream and invade new RBC. In P. falciparum, P. vivax and P. ovale this asexual blood stage cycle takes 48 hours, whereas in P. malariae it takes 72 hours. In contrast to the preceding asymptomatic infection of liver cells, the ongoing re-infection of RBC by merozoites is responsible for the characteristic clinical manifestations of the disease, such as fever, shivering, cough, headache and vomiting, that occur at regular intervals. Infection with P. falciparum can result in additional severe clinical symptoms. In contrast to the other three species, P. falciparum adheres to endothelium receptors of brain capillaries, sometimes causing coma and death. P. falciparum also sequesters in the placenta and is associated to severe outcomes for the mother and the baby, such as premature delivery, low birth weight and increased newborn mortality. According to the World Health Organization (WHO), each year a million or more people die from malaria, mostly children under 5 years of age, living in the tropical areas of Africa, Asia and Latin America.

Current malaria control strategies

At the present the tools available to control malaria do not differ much from those used in 1950: an insecticide and classical chemotherapy. At that time, DDT (dichloro-diphenyl- trichloroethane) was successfully used as insecticide, but only until 1970 when resistant- mosquitoes emerged; now we dispose of spray applications and bed nets treated with pyrethroid compounds (which target the same voltage-gated sodium channel as the DDT). For treatment and prevention, earlier it was used chloroquine (a synthetic derivate from quinine, a compound isolated from cinchona bark in 1820) and nowadays WHO recommends artemisinin (a derivate from the Artemisia annua found in 1972) in combination with another antimalarial drug. Resistance to antimalarials has been documented for P. falciparum, P. vivax and P. malariae. In P. falciparum, resistance has been observed for almost all antimalarials available (amodiaquine, chloroquine, mefloquine, quinine and sulfadoxine/pyrimethamine) except for artemisinin and its derivatives. Artemisinin compounds act rapidly killing all blood stages of all four species of human malaria parasites, including gametocytes and therefore decreasing malaria transmission (WHO, 2006, 2008). Although stable resistance to artemisinin has not yet been reported from

3

field studies, increasing concentrations of artemisinin produced resistant P. chaubaudi (rodent malaria model) parasites (Afonso et al., 2006). The vulnerability to the emergence of resistance to artemisinin and pyrethroid insecticides together with the lack of alternatives casts a shadow over malaria control. The main hope for changing this situation of ever-emerging parasite resistance to newly introduced drugs, would be an effective malaria vaccine. To date there are no clinically approved malaria vaccines available but there are a number in development and testing [reviewed in (Greenwood et al., 2008; Matuschewski and Mueller, 2007; Todryk and Hill, 2007)]. The vaccine that is most advanced in development is the RTS,S/AS02, it is currently being evaluated in trials in Mozambique and Gambia, but it appears to be less than 50% effective in preventing infection, although it might reduce the clinical severity of malaria (Alonso et al., 2004; Bojang et al., 2001; Sacarlal et al., 2008). Discovering and development of new drugs, vaccines, and insecticides, as well as improved diagnostics and surveillance methods, are thus important challenges facing malaria control today. Each of the developmental stages discussed above represents a potential target at which the life cycle can be interrupted. The different targets and approaches in recent or current development, for parasite and vector control, are summarized on Figure 1. The recent large-scale projects of parasite genome sequencing (Gardner et al., 2002), proteome and transcriptome analysis (Bozdech et al., 2003; Khan et al., 2005; Le Roch et al., 2004; Le Roch et al., 2003; Young et al., 2005), and the development of improved tools for functional genomics (Ekland and Fidock, 2007) might help to elucidate the molecular mechanisms that underpins parasite biology and its interactions with both host and vector. A detailed understanding of the specific mechanisms of transcriptional and translational control in the malaria parasite is also important and might reveal novel therapeutic targets and strategies.

4

P. falciparum genome organization

The P. falciparum comprises a nuclear genome of 22.8 Mb and two non-nuclear genomes: a compact mitochondrial genome of 6 kb and a plastid-like 35 kb circular genome, that resides in the apicoplast – an organelle that is unique to the Apicomplexan parasites. The haploid nuclear genome is organized in 14 linear chromosomes, varying from 0.64 to 3.3 Mb in size, and harbors ~5,300 protein-encoding genes. Of these predicted genes, only 40% could be assigned functions by comparison with other genomes (Gardner et al., 2002). P. falciparum nuclear genome is unusually AT-rich (~80%), however the genome distribution is not uniform, AT-content in coding regions is 60-70% whereas in introns and intergenic regions it rises to ~90% (Gardner et al., 2002). P. falciparum chromosomes are composed of an internal region, in which housekeeping genes are located, and a chromosome end, displaying a higher order DNA organization. Chromosome ends consist of distinct structural regions: the telomere and a polymorphic subtelomeric region. The telomeres are composed of degenerate G-rich heptamer repeats, in which 5’-GGGTT(T/C)A-3’ is the most frequent. Telomeric repeats are followed by a mosaic of six different telomere associated repetitive elements (TAREs 1-6), which are always found in the same order but are variable in length (15-30 kb) (Figueiredo et al., 2000). Adjacent to the non- coding TAREs are members of multi-gene families that are involved in immune evasion and cytoadhesion (var genes) and putative variant antigens (rif and stevor genes). All chromosomes contain one or more members of these multi-gene families, but the composition and the order may vary (Gardner et al., 2002). P. falciparum putative centromeres were initially identified by sequence analysis as extremely AT-rich (~97%) elements of 2.3-2.5 kb (Bowman et al., 1999; Gardner et al., 2002). These regions are tandem repeated sequences unique to each chromosome. Experimental confirmation on the location and nature of centromeric DNA was obtained only recently (Kelly et al., 2006). Importantly, conserved features of centromeres as well as conventional centromeric proteins described in other organisms are apparently absent in malaria parasites. This indicates that P. falciparum centromeric DNA might be functionally distinct of general centomeres. Interestingly, like most eukaryotic cells, Plasmodium centromeres produce non-coding RNAs, however their function remains unknown (Li et al., 2008).

5

P. falciparum gene regulation

In eukaryotic cells, regulation of involves a number of different levels of organization in the nucleus. The most basic level is the linear DNA sequence that contains regulatory elements, such as promoters and enhancers, which activate genes depending on the presence of transcription factors. The accessibility of transcription factors and RNA depends on the state of chromatin condensation, which in turn is regulated by several epigenetic determinants, such as histone modifications and the activity of chromatin remodeling complexes. Another level of gene regulation is the spatial organization of factors and sequences within the nucleus. Several nuclear components are found compartmentalized in nuclear regions, which are distinct in their composition, morphology and abundance. The chromosomes themselves occupy discrete territories in the nuclear space, which also has influence in regulation of genes [reviewed in (Branco and Pombo, 2007; Dillon, 2006)].

P. falciparum malaria parasite possesses a tight gene expression program during the various developmental stages of its life cycle (Ben Mamoun et al., 2001; Bozdech et al., 2003; Le Roch et al., 2004; Le Roch et al., 2003). Early studies have shown that several gene families appeared to be differentially transcribed in the human host and the mosquito vector (e.g. rRNA genes, actin, and α-tubulin) [reviewed in (Horrocks et al., 1998)]. Later, genome-wide transcriptional analysis demonstrated that during asexual intraerytrocytic development, genes with related function are co-regulated, and are transcribed in a sequential manner, beginning with genes involved in protein synthesis, followed by metabolism-related genes, then adhesion/invasion genes, and lastly protein kinases (Bozdech et al., 2003; Le Roch et al., 2003). However, how the parasite achieves to regulate such coordinated cascade of gene expression remains poorly understood.

Specific DNA elements and transcription factors Gene structure of P. falciparum is similar to that of other eukaryotes, they are monocistronically transcribed, contain 5’ and 3’ untranslated regions, introns and a bipartite promoter [reviewed in (Horrocks et al., 1998)]. It has also been described a number of cis- regulatory elements that are predicted to either activate or repress promoter activity in this parasite (Coleman and Duraisingh, 2008; van Noort and Huynen, 2006; Young et al., 2008). Although RNA polymerases I, II and III, the TATA-binding protein and some general

6

transcription factors have been characterized in P. falciparum (Callebaut et al., 2005; Coulson et al., 2004; Fox et al., 1993; Li et al., 1989; Li et al., 1991; McAndrew et al., 1993), a handful of the canonical transcription factors have not yet been found. A recent bioinformatic search for DNA-binding proteins discovered the Apicomplexan AP2 (ApiAP2) family, which are present not only in Plasmodium spp. but also in other Aplicomplexan parasites. These proteins contain the AP2 domains homologous to a family of plant transcription factors called AP2/ERF DNA-binding proteins. In P. falciparum, this family is composed of 26 members that are differentially transcribed during the life cycle (Balaji et al., 2005). More recently, experiments have shown that two members of this family are able to recognize unique DNA motifs located in upstream regions of distinct set of genes that are coregulated during P. falciparum asexual development (De Silva et al., 2008). Although it remains elusive how ApiAP2 proteins function as transcription regulators, it likely involves protein-protein interactions. Interestingly, the yeast two-hybrid Plasmodium screening study (LaCount et al., 2005) suggests that ApiAP2 proteins interact with each other and with chromatin-related proteins, thus supporting their putative function in transcription regulation. Although it is possible that this parasite may evoke mechanisms of transcriptional control drastically different from those used by other eukaryotic organisms, the extreme AT-rich nature of parasite genome may explain why multiple studies had initially failed to recover general transcription factors or conserved cis-regulatory elements. As pointed by Callebaut et al. (2005), the elevated AT-content in coding regions may introduce large changes in both DNA and amino acid sequences, which may influence the bioinformatic search. It is likely that many of the hypothetical proteins, with no predictable function may actually correspond to ‘hidden’ orthologues. Thus, it is expected that future comprehensive analysis using more sensitive search methods may reveal new factors contributing to transcription regulation.

Chromatin modifications and transcription P. falciparum chromatin, like in the majority of eukaryotes, is organized in fundamental units called nucleosomes (Cary et al., 1994). The nucleosome is composed of an octamer of four histones (two copies of H2A, H2B, H3 and H4) wrapped by 147 base pairs of DNA and a linker histone H1 (Figure 2A). The P. falciparum haploid genome encodes the four canonical core histones H3, H4, H2A, and H2B, with exception of the linker histone H1 (Miao et al., 2006; Sullivan et al., 2006).

7

Post-translational histone modifications. The core histones are predominantly globular except for their N-terminal tails, which are unstructured. Histone tails and globular domains are subject to a number of post-translational modifications (Figure 2A). There are at least eight types of histone modifications: acetylation, methylation, phosphorylation, ubiquitylation, sumoylation, ADP ribosylation, deimination and proline isomerization. In addition, methylation at lysines or aginines may be one of three different forms: mono-, di- or trimethyl for lysines and mono- or di- (asymmetric or symmetric) for arginines. The information stored in this way is considered epigenetic since it is not encoded in the DNA but is still inherited to daughter cells [reviewed in (Kouzarides, 2007)]. Some of these modifications have been recently reported to occur in Plasmodium histones. They have been detected either by specific antibodies or by mass spectrometry and include: acetylation of histone H3 lysine (K) residues (K9, K14, K18, K27) (Cui et al., 2007; Fan et al., 2004; Lopez-Rubio et al., 2007a; Miao et al., 2006) and histone H4 lysines (K5, K8, K12, K16) (Freitas-Junior et al., 2005; Miao et al., 2006); methylation of histone H3 lysines (K4, K9, K36) (Chookajorn et al., 2007; Cui et al., 2008a; Lopez-Rubio et al., 2007a) and H4 lysines (K20) (Cui et al., 2008a; Sautel et al., 2007); and sumoylation of histone H4 (Issar et al., 2008).

Figure 2. A. Schematic representation of the nucleosome structure containing two copies of each histone protein (H2A, H2B, H3 and H4). The DNA that wraps around the octameric disc is shown as a black line. The amino acids on the N- terminal tail of histone H3 and histone H4 are subject to modifications, such as acetylation (ac, blue squares) and methylation (me, red and green circles). Acetylation is linked to active transcription, whereas methylation is associated to activation (green) or repression (red) depending of the residue that is modified. These modifications are specifically recognized by effector proteins (grey oval shapes) via specific domains (bromo, chromo, Tudor, PHD, etc). Although only modifications on lysines (K) are shown, other modifications can occur in distinct residues. Figure adapted from ref. (Kouzarides, 2007). B. Histone modifications help the partition of chromatin into distinct reversible states: the heterochromatin (closed conformation that is inaccessible for transcription) and euchromatin (open conformation that is accessible for transcription).

8

Histone modifying enzymes. In model organisms, such Drosophila and yeast, most of these histone modifications are dynamic, as there are specific enzymes that deposit and enzymes that remove modifications. Malaria parasites apparently have conserved several orthologues to these eukaryotic histone-modifying enzymes (Table 1). So far, it has been identified and characterized one Plasmodial histone acetyltransferase (HAT; PfGCN5) (Cui et al., 2007; Fan et al., 2004) and four histone deacetylases (HDAC; PfHDAC1 – Rpd3 homologue, PfHDAC2, PfHDAC3 and PfSir2) (Andrews et al., 2008; Duraisingh et al., 2005; Freitas-Junior et al., 2005; Joshi et al., 1999; Merrick and Duraisingh, 2007). More recently, bioinformatics analysis has detected nine genes encoding histone lysine methyltransferases (KMTs; PfSet 1-9) and two genes encoding histone lysine demethylases (KDMs; PfJmjC1 and PfJmjC2) (Cui et al., 2008a).

Function of histone modifications. The mechanisms underlying the histone modification function include a direct effect on the stability of nucleosomal arrays and/or the creation of docking sites for the binding of regulatory proteins. It is known that modifications of histone acetylation and phosphorylation change the overall charge of chromatin structure. Acetylation neutralizes positive charges on DNA and phosphorylation adds a negative charge (Turner, 2000). These changes can disrupt the intern-nucleosomal contacts leading to decondensation of the chromatin fiber. Indeed, recent in vitro experiments demonstrated that a single modification (H4K16 acetylation) was sufficient to inhibit the formation of the 30 nm chromatin fiber (Shogren-Knaak et al., 2006). The proteins recruited to modifications typically possess chromatin-remodeling activity (e.g. remodeling ATPases) and bind via specific domains. For example, methylation is recognized by PHD, tudor and chromo domains, and acetylation is recognized by bromodomains. Examples of proteins containing this type of domains are represented on Figure 2A. Although bioinformatics survey of Plasmodium spp. genomes revealed a number of peptide binding domains (Iyer et al., 2008; Templeton et al., 2004), the function of the proteins containing the domains has not yet been determined. It is noteworthy that the architecture of some domains vary considerably from other eukaryotes, indicating that the outcome of effector proteins might be different in this protozoan parasite (Iyer et al., 2008). One functional consequence of histone modifications is the compartmentalization of the genome into distinct chromatin domains, such as euchromatin (where DNA is kept accessible for transcription) and heterochromatin (where chromatin is more condensed and less accessible for transcription) (Figure 2B). Actively transcribed euchromatin is associated with acetylation and methylation of H3K4 and H3K36. By contrast, silent heterochromatin is characterized by

9

hypoacetylation and methylation H3K9, H3K27 and H4K20. Apart from establishing global chromatin environments, histone modifications are involved in orchestration of DNA-based processes, such as gene transcription, DNA repair, DNA replication or chromosome condensation [reviewed in (Kouzarides, 2007)].

Histone modifications and chromatin remodeling during transcription. Regarding gene transcription, modifications of histones together with chromatin remodeling complexes are fundamental to open up the chromatin locally to enable access of transcription factors and RNA polymerases to the DNA. Distinctive histone modifications are associated with initiation and elongation steps of transcription. Moreover, the same modification may produce different effects on transcription, depending on timing and position in the gene (upstream region, the 5’-end or 3’- end of coding region). Transcription of genes also involves histone loss or replacement, by histone chaperones and ATP-dependent chromatin remodelers. This is necessary to enable access of transcription machinery in the promoter region and passage of the RNA polymerases within the transcription units [reviewed in (Li et al., 2007; Williams and Tyler, 2007)].

Table 1. Examples of histone modifications present in P. falciparum with their respective putative histone methyltranferases (Cui et al., 2008a) and acetyltransferases (Fan et al., 2004). The orthologues enzymes in other organisms and the domains that recognize distinct histone marks are indicated. Table adapted from refs. (Allis et al., 2007; Li et al., 2007). Notice that a new nomenclature has been proposed by Allis et al. (2007). For example, PfSet3 should now be called PfKMT1, and PfSet1 is now PfKMT2.

10

In general, following the stimulus for activation, the promoter region is characterized by the presence of HATs and chromatin remodelers, and consequently accumulation of acetylated histones H3 and H4 and removal of nucleosomes. The beginning of the coding region displays the initiated form of Pol II (phosphorylated at serine 5 of its C-terminal) associated to KMT2 (Set1), which results in accumulation of H3K4 methylation. Within the coding region, there is the serine 2 phosphorylated elongating form of Pol II coupled to KMT3 (Set2), which methylates H3K36. Notably, H3K36me serves as the binding site for EAF3 protein (via chromodomain) that recruits a histone deacetylase complex (Rpd3), which in turn removes any acetylation behind the polymerase. This together with the rapid reassembly of the chromatin after polymerase passage resets chromatin into its stable state and prevents internal initiation at cryptic transcription start sites [reviewed in (Li et al., 2007; Mellor, 2006)]. By contrast, inactive promoters of model organisms present high nucleosome density and distinct modified histones: di- and trimethylation at H3K9 (H3K9me2/3), H3K27me2/3 and H4K20me3, and sumoylation of all four core histones [reviewed in (Li et al., 2007)]. H3K27me and H4K20me have been mapped in repressed genes, but their function remains vague. Repression by H3K9me invokes the recruitment of heterochromatin protein HP1 (via chromodomain) and KMTs. Importantly, recent studies have documented the presence of H3K9me2/3 and HP1 at the transcribed regions of mammalian genes (Barski et al., 2007; Squazzo et al., 2006; Vakoc et al., 2005). Similarly, H4K20me1 has been identified as another mark of transcription elongation (Barski et al., 2007; Vakoc et al., 2006). The function of H3K9me2/3 and H4K20me1 in transcribed regions is not clear, however it has been proposed that they might work like H3K36m3 in prevention of inappropriate transcription initiation and thus contributing to genomic integrity. The current understanding of histone modifications in Plasmodium transcriptional control remains poor. So far, chromatin immunoprecipitation (ChIP) experiments have demonstrated that Plasmodium active promoters show colocalization of H3K9 acetylation and PfGNC5 HAT (Cui et al., 2008b; Cui et al., 2007; Komaki-Yasuda et al., 2008), whereas inactive promoters display H3K9 methylation (Chookajorn et al., 2007; Cui et al., 2007; Lopez-Rubio et al., 2007a). Some silent var gene promoters also present a HDAC, PfSir2 (Freitas-Junior et al., 2005). H3K9ac, H3K4me2 and H3K4me3 have also been found at 5’end coding region and H3K9me3 at 3’end coding region of actively transcribed var genes (Lopez-Rubio et al., 2007a). Like in higher eukaryotes (Vakoc et al., 2006; Wang et al., 2008), mutually exclusive acetylation and methylation of lysine H3K9 has been detected in P. falciparum (Lopez-Rubio et al., 2007a). In analogy to other organisms, it is very likely that these Plasmodium histone modifications can

11

cross-talk and that a combinatorial pattern of modifications can determine the chromatin status (Latham and Dent, 2007; Wang et al., 2008). Regarding chromatin assembly and remodeling, a number of chromatin-remodeling SNF2/SWI2 ATPases orthologs have been identified in Plasmodium spp. genomes (Ji and Arnot, 1997; Templeton et al., 2004), but not yet investigated in detail. Plasmodium spp. also encodes members of high-mobility-group B (HMGB) family. HMGB proteins bind DNA in a nonsequence-specific manner and bend it. This activity has been demonstrated in vitro for two of the four members in P. falciparum (Briquet et al., 2006). Two nucleosome assembly proteins have also been recently identified and characterized in P. falciparum parasites, PfNapS and PfNapL (Chandra et al., 2005; Navadgi et al., 2006). Interestingly, most of the Plasmodium chromatin related proteins, described above, show differential expression in asexual and sexual stages of the parasite life cycle. This implies that individual members might function in a stage-specific manner and thus contributing to the tightly coordinated expression system in this parasite. While a remarkable number of chromatin modifiers have been identified in this protozoan parasite, the mechanisms by which the chromatin events are orchestrated remains to be established. Also, it is not yet known how these chromatin factors are recruited to specific genomic loci. Taken the knowledge from other organisms, it has been proposed that targeting of silent loci may rely on not yet identified DNA-binding proteins and/or noncoding RNAs (Hakimi and Deitsch, 2007).

DNA methylation and transcription regulation. Another type of chromatin modification that also controls gene expression epigenetically (i.e. without changes in DNA sequence) is the methylation of DNA cytosine. DNA methylation at promoter regions correlates with gene repression, whereas unmethylated promoters are associated to gene activation [reviewed in (Suzuki and Bird, 2008)]. In Plasmodium spp. genomes, homologues for DNA methyltransferases enzymes and methyl DNA binding proteins have not been found, nor the methylated DNA (5- methyl-2’-deoxycytosine) (Choi et al., 2006). Thus, it appears that epigenetic regulation in this parasite depends largely on reversible modifications of histones.

12

Anti-sense RNAs and post-transcriptional regulation Several studies have reported the presence of natural antisense transcripts in P. falciparum parasites (Gunasekera et al., 2004; Kyes et al., 2002; Militello et al., 2005; Patankar et al., 2001; Ralph et al., 2005a). Typically, antisense RNA transcripts are negative regulatory molecules that control transcription, RNA stability and translation. In fact, artificially introduced antisense molecules have been able to specifically downregulate a gene in P. falciparum (Gardiner et al., 2000). Although it is not clear how antisense transcription works in this parasite, it has been proposed as a general control mechanism for Plasmodium gene expression. Interestingly, P. falciparum encodes a large number of proteins that modulate mRNA decay and translation rates. Given that, some studies have put forward the idea that post-translational regulation may be another level of gene regulation in P. falciparum (Coulson et al., 2004; Shock et al., 2007). At least in gametocyte stages, that appears to be the case (Mair et al., 2006).

Nuclear organization and telomere position effect Like in most eukaryotes, P. falciparum genomic DNA is non-randomly distributed in the nucleus. The chromosome ends in this parasite have been shown to form clusters of four to seven telomeres at the nuclear periphery (Freitas-Junior et al., 2000; Marty et al., 2006). This telomere pairing/clustering appears to be random and is thought to promote recombination between subtelomeric virulence genes located on heterologous chromosomes (Freitas-Junior et al., 2000) (Figure 3).

Figure 3. Nuclear architecture of P. falciparum chromosome-ends. Chromosome termini (right) are composed of telomeres and TAREs (Telomere Associated Repetitive Elements). TAREs 1-5 are closer to the telomeric repeats and span 10-15 Kb. TARE6, which was originally described as rep20, is a large array of variable copies of a 21 bp repeat (6-23 kb) and is a common marker for mapping studies. The subtelomeric coding region contains members of several multi-gene families (var, rifin, stevor, Pf60…) important for parasite virulence. FISH analysis of TARE6 (left) showing that the chromosome ends form clusters (yellow) in the parasite nucleus (blue), during asexual (throphozoites, T) and sexual (gametocytes, G) blood stages. Schematic model of telomeric cluster (right) illustrating the physical alignment of heterologous subtelomeric regions, in which increased rates of recombination occur. Figure adapted from ref. (Scherf et al., 2001).

13

Interestingly, analysis of truncated chromosomes demonstrated that subtelomeric sequences, namely the TARE6 element, are crucial to telomere cluster formation but are not involved in nuclear peripheral tethering (Figueiredo et al., 2002; O'Donnell et al., 2002). Furthermore, chromosome-painting studies revealed that a pair of subtelomeric probes distant of 60-80 kb colocalize in the same spot in the nucleus, whereas two internal probes equally distant, are clearly seen as two distinct signals. Thus, indicating that telomere-proximal regions exist in a more condensed form than internal chromosome regions. Consistent with this, electron-dense heterochromatin was observed at the periphery of parasite nuclei in ultrathin sections (Freitas- Junior et al., 2005). Telomeres have the ability to recruit a number of proteins that initiate and/or facilitate heterochromatin formation into the adjacent subtelomeric regions. In yeast, as in higher eukaryotes such humans, genes positioned near to telomeric repeats are silenced by a phenomenon known as telomere position effect (TPE). This form of epigenetic silencing is best characterized in budding yeast, and involves mainly Sir proteins (Sir 2-4), although it can be influenced by many factors. Multiple rounds of histone deacetylation by Sir2 and binding of Sir complex along the nucleosomes lead to heterochromatin spreading and silencing at the proximity of telomeres [reviewed in (Mondoux and Zakian, 2006)]. In fission yeast and metazoans, however, establishment of telomeric heterochromatin requires KMT1 and HP1. KMT1 is involved in methylation of H3K9, which creates binding site for HP1/Swi6. This is followed by HP1-mediated recruitment of more KMT1, which in turn methylates adjacent nucleosomes, with subsequent binding of additional HP1 molecule. HP1 is also able to recruit other chromatin-modifying factors implicated in heterochromatin formation, such as HDACs [reviewed in (Blasco, 2007; Grewal and Jia, 2007)]. The finding that P. falciparum chromosomes are functionally compartmentalized into an actively transcribed central domain and transcriptionally silent subtelomeric domains (Lanzer et al., 1993) is likely the first indication that TPE also operates in this malaria parasite. More recently, the P. falciparum homologue to the yeast Sir2 has been identified and characterized. Like in budding yeast, PfSir2 binds to telomeres, but the silencing it causes spreads over ~50 kb away from the telomere, whereas yeast Sir2-mediated silencing only extends ~3 kb from the telomere (Freitas-Junior et al., 2005). Mutant-Sir2 parasites de-repress some but not all subtelomeric genes (Duraisingh et al., 2005), indicating that TPE in malaria parasites involves additional chromatin factors.

14

var genes: organization and regulation

Much of the success of P. falciparum parasites in establishing persistence infections relies on their ability to evade host immune response, through clonal antigenic variation. This process involves periodically exchanging variants of the major surface antigen P. falciparum erythrocyte membrane protein 1 (PfEMP1). This protein also mediates adherence to host endothelial receptors, resulting in sequestration of parasitized red blood cells in the brain and other tissues, including placenta [reviewed in (Scherf et al., 2008a)]. PfEMP1 is encoded by members of the large var gene family (Baruch et al., 1995; Smith et al., 1995; Su et al., 1995), of which only one is transcribed in each parasite, a process known as monoallelic expression (Scherf et al., 1998).

Figure 4. Organization of the var genes (vertical red bars) in the 14 chromosomes of P. falciparum. The var gene family can be divided in three major groups: A (10 genes), B (22 genes) and C (13 genes). Groups A and B are located on subtelomeric regions, adjacent to subtelomeric repeats (example of a chromosome end organization in the red box). Genes from group A are transcribed towards the telomere, whereas from group B are transcribed towards the centromere. Group C corresponds to genes located on in internal chromosomal regions and are found singly or in groups of 3 to 7 genes (example of an internal var group in the grey box). The var genes have a two-exon structure: the exon1 (4-9 kb) encodes the variable portion of PfEMP1, which mediates cytoadhesion; the exon2 (~1.5 kb) is more conserved and codes for an intracellular region of PfEMP1. Centromeres are shown as white circles. Red bars above the line correspond to sense and under the line to antisense.

15

Distinct P. falciparum strain isolates present a var gene repertoire variable in both sequence and number of genes (3D7, 61 genes; HB3, 54 genes; IT4, 48 genes) (Kraemer et al., 2007). Members of var gene family can be grouped on the basis of chromosomal location and sequence similarity. var genes can be found on either subtelomeric (60%) or chromosomal internal regions (Figure 4). Alignments of var upstream sequences revealed three classes (upsA, upsB and upsC) (Gardner et al., 2002). Subsequent analysis of coding and non-coding sequences demonstrated that 3D7 var genes are organized in three major groups (A, B and C), two intermediate groups (B/A and B/C) and three unique genes (var1, var2 and var3) (Figure 4). Groups A, B and C are flanked by upsA, upsB and upsC sequences, respectively (for more details see Table 2). Group B/A and group B/C appear to represent transitions between the three groups (Kraemer et al., 2007; Lavstsen et al., 2003). The var genes are transcribed during ring asexual stages (Kyes et al., 2000), by an alpha- amanitin-sensitive polymerase, most likely Pol II (Kyes et al., 2007a; Schieck et al., 2007). Specific var gene expression is activated in situ (Scherf et al., 1998) and is controlled at the level of transcription initiation (Kyes et al., 2007a; Schieck et al., 2007) by epigenetic mechanisms. Recent evidence suggests that activation and silencing of var genes involves chromatin modifications (Chookajorn et al., 2007; Duraisingh et al., 2005; Freitas-Junior et al., 2005; Lopez-Rubio et al., 2007a) and perinuclear repositioning (Duraisingh et al., 2005; Mok et al., 2008; Ralph et al., 2005b; Voss et al., 2006). Furthermore, var gene upstream sequences apparently contain elements that nucleate transcriptional silencing and activation, and maintenance of allelic exclusion (Dzikowski et al., 2006; Voss et al., 2006; Voss et al., 2007). Another DNA sequence element possibly implicated in silencing is the intron, which has promoter activity and also appears to play a role in the monoallelic expression mechanism (Deitsch et al., 2001; Frank et al., 2006; Voss et al., 2006). There are also evidences that transcriptional regulation of different groups might occur through distinct processes [reviewed in (Kyes et al., 2007b; Scherf et al., 2008a)]. The distinct genetic and epigenetic elements regulators of each var group are summarized on Table 2. With regard to the organization of var genes in the parasite nucleus, several studies using fluorescence hybridization analysis of single var genes demonstrated that the different var groups reside at the nuclear periphery, regardless of chromosomal location or transcriptional status (Duraisingh et al., 2005; Marty et al., 2006; Mok et al., 2008; Ralph et al., 2005b; Voss et al., 2006). It has also been assumed that the var gene transcription occurs in a euchromatic subdomain in the heterochromatic nuclear periphery. However, there are conflicting conclusions concerning the association of the active var member with the chromosome end clusters (Table 2).

16

Like in yeast, P. falciparum telomeric clusters have been described as silent nuclear compartments, as they are enriched in the histone deacetylase PfSir2 (Freitas-Junior et al., 2005). By using parasite populations expressing a single var gene, it was shown that upon activation the var gene moves away from the silent telomeric cluster to another position at the nuclear periphery (Mok et al., 2008; Ralph et al., 2005b). By contrast, studies using parasites transfected with various episomally maintained telomere-homing constructs or drug resistance genes under control of a var-specific promoter (Duraisingh et al., 2005; Marty et al., 2006; Voss et al., 2006) demonstrated that active var promoters can associate with telomeric structures. Intranuclear repositioning may help to control mutually exclusive transcription and switching of var genes, in a similar manner as it has been demonstrated for the vsg genes in the African trypanosomes (Navarro and Gull, 2001). However, the mechanism and the factors that promote reorganization or the movement of these genes in the nucleus remain elusive.

Table 2. var genes organization and regulation. The var1 group (var1CSA) was not included because its expression appears to fall outside the pattern of mutually exclusive var gene regulation (Kyes et al., 2003). The var 2 group corresponds to the var2CSA gene. The numbers correspond to the following references: 1, (Voss et al., 2007); 2, (Voss et al., 2006); 3, (Duraisingh et al., 2005); 4, (Freitas-Junior et al., 2005); 5, (Lopez-Rubio et al., 2007a); 6, (Chookajorn et al., 2007); 7, (Ralph et al., 2005b); 8, (Mok et al., 2008); 9, (Marty et al., 2006). For the Sir2-mediated repression only upregulated genes with ≥4 fold increase were considered. Ups, upstream; ATS, acidic terminal segment (ATS or exon2). Table adapted from refs. (Lavstsen et al., 2003; Scherf et al., 2008b).

17

rRNA genes: organization and regulation

Ribosomal RNA (rRNA) is the predominant product of transcription, constituting 80-90% of the total mass of cellular RNA in both eukaryotes and prokaryotes. The number of rRNA genes varies from seven in Escherichia coli, 100-200 in lower eukaryotes, to several hundreds in higher eukaryotes. In eukaryotic cells, the multiple identical copies of rRNA form tandem repeated clusters, whereas in bacteria, the rRNA gene pairs are dispersed [reviewed in (Lewin, 2000)]. In P. falciparum parasites the rDNA unit has a typical eukaryotic organization, comprised of 18S, 5.8S and 28S genes (Figure 5), but the family itself is the most unusual among the eukaryotic rDNA families. P. falciparum rRNA genes are not organized in tandem arrays, instead they are spread out on different chromosomes (Langsley et al., 1983; Wellems et al., 1987) (Figure 5).

Figure 5. Right. P. falciparum rDNA unit consists of a ~9kb transcribed region coding for a 45S rRNA precursor. The first nucleotide (+1) of this transcript begins at the 5’ external transcribed spacer (5’ETS) followed by the 18S coding region, the internal transcribed spacer 1 (ITS1), the 5.8S coding region, the internal transcribed spacer 2 (ITS2), the 28S coding region, and finally the 3’ ETS. Left. Organization of rRNA genes in 3D7 in all chromosomes of P. falciparum. Five complete ribosomal units are located on chromosomes 1, 5, 7, 11 and 13, and two incomplete units (lacking 18S gene) on chromosome 8. Because rRNA genes are different in sequence in the schematics are shown in different colours. rDNA units located on chromosomes 5 and 7 are named A1 and A2 genes, respectively. A1 and A2 gene are transcribed in the human host during schizogony. The rDNA unit on is called S1 gene and is transcribed in the human host during gametogony. rDNA copies from chromosomes 11 and 13 correspond to the previously described S2 gene that is preferentially transcribed in the mosquito stages during sporogony (Fang et al., 2004). Representation of the genes is not scale. White circles correspond to centromeres.

18

The number of P. falciparum rDNA copies is very low, it varies from five to eight, depending on the parasite strain [reviewed in (Mercereau-Puijalon et al., 2002)]. In the 3D7 strain, the malaria genome project identified five complete ribosomal units each located on separated chromosomes 1, 5, 7, 11 and 13, and two incomplete units (lacking 18S gene) symmetrically arranged at each end of chromosome 8 (Gardner et al., 2002) (Figure 5). Interestingly, with the exception of the rDNA unit on chromosome 7, all the other rDNAs are positioned in subtelomeric regions. Moreover, P. falciparum individual rDNA units are not identical in sequence (Langsley et al., 1983). The genome sequence (Gardner et al., 2002) revealed an exception on chromosomes 11 and 13. It is possible these two fully identical rDNA units derived from a gene duplication event, given their location on subtelomeric regions, which are regions prone to DNA material exchange. Regarding the other rDNA units, the highest sequence heterogeneity is found in the sequences that either flank or separate the 18S, 5.8S and 28S genes (ETS and ITS sequences, Figure 5); but there are also some particular differences in the 18S and 28S sequences. Another remarkable feature of rRNA genes in Plasmodium spp. is the developmental regulation of stage-specific transcription. This pattern of rRNA gene transcription was initially discovered in the rodent parasite, P. berghei. It was demonstrated that asexual blood stages transcribed one type of genes (termed A-type, from Asexual), whereas in the mosquito stages transcripts from a second type of genes (termed S-type, from Sexual) were the predominant rRNAs (Gunderson et al., 1987). A similar pattern of rRNA transcription was then confirmed for P. falciparum and P. vivax (Li et al., 1994; McCutchan et al., 1988; Waters et al., 1989). In P. falciparum, A-type rRNA genes, include A1 and A2 genes (chromosomes 7 and 5, respectively) and are transcribed during blood stages, in the human host. In gametocyte stages, the dominant transcripts are from the S1 gene (chromosome 1). Lastly, S2 type genes (chromosomes 11 and 13) are transcribed mostly in the mosquito host (oocysts and sporozoites) (Fang and McCutchan, 2002; Fang et al., 2004) (Figure 5). It is not known whether the rRNA genes located on chromosome 8 are transcribed. Hence, only one or two rRNA genes are activated at any one time in this parasite. Transcription of the complete rRNA gene units of eukaryotes occurs in a dedicated nuclear compartment (the nucleolus) and is performed by Pol I. P. falciparum Pol I large subunit has been identified, it has the characteristic domains of a Pol I, but it has also significant amino acid insertions and novel motifs (Fox et al., 1993). The initiation sites and the putative rRNA gene promoter of the different ribosomal units have been recently mapped in P. falciparum. Interestingly, the general structure of the A1 and A2 promoters is very similar to the rRNA gene

19

promoter structure in E. coli (Fang et al., 2004). A GC-rich region of 20 bp was also found just preceding the A1 and A2 start transcription site. This GC-rich DNA is characteristic of Pol I promoters and is conserved from bacteria to human cells [reviewed in (Lewin, 2000)]. No conserved sequences were found upstream the S1 and S2 initiation sites. To date, little is known on the mechanism(s) responsible for transcriptional regulation and switching rRNA genes ON and OFF in the different developmental stages of P. falciparum life cycle. In most organisms rDNA transcription is regulated in response to different stages of the , nutritional state, and altered environmental conditions. Likewise, recent studies have demonstrated that changes in temperature and glucose concentration can influence rRNA transcription pattern in P. falciparum parasites (Fang and McCutchan, 2002; Fang et al., 2004). In fact, low temperature and glucose starvation appear to have synergistic effects on downregulation of A1 and A2 genes and upregulation of S2 genes. These changes are consistent with the environmental conditions found by the parasite in the insect host (temperature 26°C and low glucose concentrations) and mimic the transmission from the human host to the mosquito. In contrast, the rRNA gene transcribed predominantly in gametocytes did not respond to the temperature nor glucose changes, indicating that the distinct rRNA types are differently regulated. The molecular mechanisms that mediate these transcriptional changes however remain to be determined. It is also known, from P. berghei, that A and S rRNA types cannot replace one another. Apparently, parasites lacking A-types are not viable and for the S-types, at least one gene appears to be necessary to complete the cycle in the mosquito (McCutchan et al., 2004; van Spaendonk et al., 2001).

20

Aims of this thesis

The finding of the yeast Sir2 homologous in P. falciparum by Freitas-Junior et al. and Duraisingh et al. (2005) was an important milestone for the understanding of gene regulation in this malaria parasite. PfSir2 histone deacetylase localizes to telomeric clusters and nucleolus of parasite nuclei. Furthermore, PfSir2 associates to upstream sequences of silent subtelomeric virulence genes, and PfSir2 mutants show up-regulation of those genes. These characteristics of PfSir2 raised numerous new questions in the areas of telomere and nuclear biology. It also led to the beginning of investigation of the role of chromatin modifications in regulation of gene expression in P. falciparum, particularly in regulation of genes involved in antigenic variation.

The major objective of this thesis was to understand the contribution of nuclear and telomere biology to gene regulation in P. falciparum. The specific aims were:

- To identify and characterize other telomere-associated proteins involved in heterochromatin formation and TPE. - To investigate the role of histone modifications on the spatial genome organization. - To study nuclear compartmentalization during the blood-stage cycle. - To investigate whether the nuclear architecture and spatial relocation of genes participate in the regulation of differentially transcribed multigene families.

21

22

RESULTS

This section contains the results and their respective discussion. It is organized into seven sub-chapters (indicated in roman numbers) written in article format. Results I and II represent data that have been published and are presented in their published format. Results III, IV and V correspond to results that have been submitted to publication. Finally, the Results VI and VII describe results that are in preparation to be published. The order of the sub-chapters corresponds to the chronological progress of the experimental work, with exception of Results VI and VII.

Results I This sub-chapter contains the first article, which was published in Nucleic Acid Research journal (2005). Here, the gene that encodes the Plasmodial catalytic reverse transcriptase component of telomerase (PfTERT) has been characterized. In the nucleus of the parasite, this molecule localizes in a discrete compartment associated with the nucleolus, hereby defined for the first time in P. falciparum. My contribution to this work was the validation of the PfNop1 antibody as a marker for the P. falciparum nucleolus.

Results II This sub-chapter corresponds to the second article that appeared in Journal of Cell Science (2008). Here, we identified and characterized a novel telomere-associated protein that displays homology with the origin-of-recognition-complex 1 protein, PfOrc1. This protein showed a differential association to telomeric and subtelomeric domains, and it is also differentially localized in the nucleus of the parasite during the blood-stage cycle. In this study, I performed all the nuclear imaging analysis.

23

Results III In this work, we identified and characterized the first member of histone modification- specific recognition proteins in P. falciparum parasites. This protein is an orthologue of the heterochromatin protein 1 (PfHP1) and binds specifically to H3K9me2/3. My experimental contribution was to study the cellular localization of PfHP1 in blood-stage parasites.

Results IV This sub-chapter describes a screening for the presence of histone methylation modifications in P. falciparum, using specific antibodies and determined their spatial nuclear localization. I contributed to this work by investigating the localization of H3K79me3 relative to known nuclear compartments.

Results V Here, we present a high-resolution genome-wide mapping of histone modifications in malaria parasites. H3K9me3 revealed a very unusual pattern, which lead us to perform a quantitative large-scale nuclear localization analysis of the genes enriched in this histone mark. In this work, I performed all the nuclear localization experiments.

Results VI In this sub-chapter, which corresponds to a manuscript in preparation, we used different fluorescent imaging approaches to further characterize the nucleolus in P. falciparum blood- stages. We also explored the factors that determine nucleolar organization and position.

Results VII Here, we present preliminary data showing that rRNA small multi-gene family is regulated epigenetically. We investigated the chromatin modifications pattern associated to active and silent rRNA genes and followed their dynamics in response to temperature changes.

24

RESULTS I

The unusually large Plasmodium telomerase reverse-transcriptase localizes in a discrete compartment associated with the nucleolus

Luisa M. Figueiredo1, Eduardo P. C. Rocha2,3, Liliana Mancio-Silva1, Christine Prevost4, Danièle Hernandez-Verdun5 and Artur Scherf1

1 Institut Pasteur, Biology of Host Parasite Interaction Unit–CNRS URA2581, 25 rue du Dr Roux, 75724 Paris, France 2 Unité GGB URA CNRS 2171 Institut Pasteur 28 rue Dr Roux, 75724 Paris, France 3 Atelier de BioInformatique, Université Pierre et Marie Curie 12 rue Cuvier, 75005 Paris, France 4 Plateforme de Microscopie Electronique, Institut Pasteur 25 rue du Dr Roux, 75724 Paris, France 5 Institut Jacques Monod UMR 7592, 2 place Jussieu, 75251 Paris, France

Published in Nucleic Acids Research. 2005; 33(3): 1111–1122

25

26 Published online February 18, 2005

Nucleic Acids Research, 2005, Vol. 33, No. 3 1111–1122 doi:10.1093/nar/gki260 The unusually large Plasmodium telomerase reverse-transcriptase localizes in a discrete compartment associated with the nucleolus Luisa M. Figueiredo1, Eduardo P. C. Rocha2,3, Liliana Mancio-Silva1, Christine Prevost4, Danie`le Hernandez-Verdun5 and Artur Scherf1,*

1Institut Pasteur, Biology of Host Parasite Interaction Unit–CNRS URA2581, 25 rue du Dr Roux, 75724 Paris, France, 2UniteeGGB, URA CNRS 2171, Institut Pasteur, 28 rue Dr Roux, 75724 Paris, France, 3Atelier de BioInformatique, UniversiteePierre et Marie Curie, 12 rue Cuvier, 75005 Paris, France, 4Plateforme de Microscopie Electronique, Institut Pasteur, 25 rue du Dr Roux, 75724 Paris, France and 5Institut Jacques Monod, UMR 7592, 2 place Jussieu, 75251 Paris, France

Received as resubmission December 31, 2004; Accepted February 1, 2005 DDBJ/EMBL/GenBank accession nos+

ABSTRACT telomeric DNA consists of tandem arrays of short G-rich repeats that are maintained by telomerase, an RNA- Telomerase replicates chromosome ends, a function dependent DNA polymerase. Loss of telomerase activity necessary for maintaining genome integrity. We have leads to a gradual shortening of telomere length, resulting identified the gene that encodes the catalytic reverse in growth arrest and eventually senescence (1–3). The core transcriptase (RT) component of this enzyme in the components of telomerase enzyme consist of a catalytic malaria parasite Plasmodium falciparum (PfTERT)as protein component, called telomerase reverse transcriptase well as the orthologous genes from two rodent and (TERT), and an RNA molecule that serves as an internal one simian malaria species. PfTERT is predicted to template from which repeats are copied to the 30 end of the encode a basic protein that contains the major telomere G-rich strand. Other proteins have been identified sequence motifs previously identified in known that associate with telomerase and, some at least, are required for its action (4,5). telomerase RTs (TERTs). At 2500 amino acids, in vivo TERT was first purified from the ciliate Euplotes aediculatus PfTERT is three times larger than other characterized (1). By , other TERTs were subsequently TERTs. We observed remarkable sequence diversity identified, such as ciliates (6,7) and diplomonads (8), several between TERT proteins of different Plasmodial spe- yeasts (9–11), plants (12) and animals, including humans cies, with conserved domains alternating with hyper- (10,13) and mouse (14). All TERTs contain telomerase- variable regions. Immunofluorescence analysis specific motifs within the N-terminal half of the protein (GQ/ revealed that PfTERT is expressed in asexual blood N, CP, QFP and T) and RT-specific motifs in the C-terminal half stage parasites that have begun DNA synthesis. (1, 2 and A–E; Figure 1) (1,8,15,16). The RT domain of TERT is Surprisingly, rather than at telomere clusters, essential for catalytic activity, whereas the N-terminal half is PfTERT typically localizes into a discrete nuclear com- required for efficient binding of the RNA template, defining the 0 partment. We further demonstrate that this compart- 5 RNA template boundary, multimerization and interactions ment is associated with the nucleolus, hereby defined with associated proteins [reviewed in (17)]. Telomerase is developmentally regulated in higher euka- for the first time in P.falciparum. ryotes (18). In a human being, telomerase activity is undetect- able in most somatic tissues. Interestingly, telomerase activity is detectable in the germline and in highly proliferative cells of INTRODUCTION renewal tissues, such as bone marrow progenitor cells and Telomeres are nucleoprotein complexes that protect chromo- intestinal mucosa (19). Telomerase is necessary for the some ends from degradation and fusion. In most eukaryotes, unlimited proliferation of human somatic cells in vitro (20).

*To whom correspondence should be addressed. Tel: +33 1 4568 8616; Fax: +33 1 4568 8348; Email: [email protected] Present address: Luisa M. Figueiredo, The Rockefeller University, Laboratory of Molecular Parasitology, 1230 York Avenue, New York, NY 10021, USA

+AABL01000132 and AZ525829

ª The Author 2005. Published by Oxford University Press. All rights reserved.

The online version of this article has been published under an open access model. Users are entitled to use, reproduce, disseminate, or display the open access version of this article for non-commercial purposes provided that: the original authorship is properly and fully attributed; the Journal and Oxford University Press are attributed as the original place of publication with the correct citation details given; if an article is subsequently reproduced or disseminated not in its entirety but only in part or as a derivative work this must be clearly indicated. For commercial re-use, please contact [email protected] 1112 Nucleic Acids Research, 2005, Vol. 33, No. 3 Nucleic Acids Research, 2005, Vol. 33, No. 3 1113

These observations along with the prevalence of telomerase loss and eventually death of the parasite. In the present expression in human cancers, makes telomerase an attractive work, we report the identification and characterization of a candidate for cancer therapy [reviewed in (21)]. putative TERT from the human malaria parasite P.falciparum Human malaria is re-emerging as one of the world’s most (PfTERT). We also identified TERT candidate genes from two lethal infections. In tropical regions of the globe, 300 million murine and one simian Plasmodial species. Using immuno- people are infected and of those 1–2 million are killed annually localization techniques, PfTERT and telomeres were shown to (WHO, 1998, Malaria, http://www.who.int:inj-fs/en/fact094. localize in different regions of the nucleus. After identifying html). It is caused by intracellular protozoa of the genus the nucleolus as a sub-nuclear compartment in P.falciparum Plasmodium, of which Plasmodium falciparum is the most for the first time, we were able to show that PfTERT generally virulent species. P.falciparum is transmitted from one person localizes within the nucleolar region. The implications of such to another by the Anopheles mosquito. After a short hepatic spatial segregation are discussed. stage, parasites are released into the bloodstream of the host, where they infect erythrocytes. Massive proliferation of the blood stage parasites causes the clinical symptoms [reviewed in (22)]. Although several anti-malarial drugs exist, resistance MATERIALS AND METHODS is becoming a major problem, and there is an urgent need for Plasmodium falciparum cultures new therapeutics (23). Plasmodium falciparum blood stage parasites from the FCR3 The haploid nuclear genome of P.falciparum is 30 Mb in size and is organized in 14 linear chromosomes [reviewed in strain were cultured as described previously (31). Synchron- (24)]. The ends of the chromosomes consist of a tandem array ization of cultures for time-course immunofluorescence (IF) of a degenerate G-rich heptamer, most frequently GGGTT analysis consisted of two consecutive Plasmagel treatments, at a 48 h interval, followed by Sorbitol treatment. Parasites (T/C)A (25). The mean telomere length is 1.2 kb and is maintained at a constant size during blood-stage proliferation were collected at the beginning of the next cycle (40 h later). (26). Analysis of nuclear architecture reveals that chromosome We estimated that the parasites were synchronized within a ends form clusters of 4–7 telomeres that localize around the window of 6h. nuclear periphery (27). This arrangement was proposed to Genome searches and sequence analysis facilitate gene conversion events between the sub-telomeric regions of heterologous chromosomes. As a result, there is Genes coding for putative TERTs in Plasmodium species were a continuous sequence diversification among gene families obtained using TBLASTN from both NCBI Custom Blast in coding for surface antigens, which promotes the generation Malaria Genetics & Genomics (http://www.ncbi.nlm.nih. of new antigenic and adhesive phenotypes. gov/Malaria/Plasmodiumblcus.html) and PlasmoDB (www. P.falciparum telomeres are partially organized in a non- plasmodb.org). To identify the first Plasmodial TERT gene nucleosomal structure (26), suggesting that specific proteins (PfTERT), motif T and other RT motifs were used as queries. may constitute the telomeric chromatin. In yeast and humans, The PfTERT putative gene has been annotated with the numerous proteins have been shown to interact directly or accession no. chr13_400065.gen2 in PlasmoDB database indirectly with telomeric DNA [for a review see (28)]. In and as GenBank accession no. AX112155. TERT genes from P.falciparum, several putative homologues of known Plasmodium yoelii (GenBank accession no. AABL01000132) telomeric-specific proteins have been identified and are cur- and Plasmodium knowlesi (PlasmoDB accession no. rently being characterized (29). Telomerase activity has been Pk_812g03q1c) were found by blasting the entire PfTERT detected in semi-purified nuclear extracts of P.falciparum gene against the relevant database. For Plasmodium berghei, blood stages (30). Furthermore, it was shown that the best hit was an expressed sequence tag clone (GenBank P.falciparum telomerase can be used as a substrate oligo- accession no. AZ525829). Using oligonucleotides that match nucleotide that mimic not only telomeric sequences, but PfTERT, we amplified a larger fragment containing the also chromosome breaks, suggesting that this enzyme contrib- putative PbTERT gene. The 3942 bp fragment was cloned utes to telomere maintenance and to the de novo telomere and sequenced. Although it lacks the 30 end of the open reading formation necessary to repair broken chromosomes. frame (ORF), it spans motifs GQ/N, QFP, T, 1 and 2. Plasmodial telomerase is likely to be necessary to maintain The number of significant large tandem repeats in a gene telomeres at a constant length during the highly replicative was computed with the mreps software [(32), http://www. stages of the parasite’s life cycle, such as the bloodstream loria.fr/mreps/]. This program identifies all maximal repeats stages. We hypothesize that drugs that block telomerase in a given DNA sequence, compares them with the repeats activity would induce telomere shortening, chromosome obtained in simulations of DNA sequences of the same length

Figure 1. Plasmodium TERT proteins contain telomerase conserved motifs. (A) In-scale diagram showing the difference in size of PfTERT and the S.cerevisiae homologue, Est2p. Grey boxes represent telomerase-specific motifs: GQ/N, CP, QFP and T. Black boxes represent RT motifs. (B) Multiple sequence alignment of the conserved motifs of all TERT proteins described to date and the four Plasmodium TERT proteins identified in this work: PbTERT (P.berghei), PfTERT (P.falciparum), PkTERT (P.knowlesi) and PyTERT (P.yoelii). For a given position within the alignment, identical or similar amino acids to the consensus sequence are shaded in black or grey, respectively. Note: The putative C.elegans TERT gene was excluded because it lacks all TERT-specific motifs. Squares mark telomerase- specific amino acids that are not present in viral RTs; circles indicate amino acids essential for RT activity. The numbers accompanying each protein aligned in (B) represent the number of amino acids that exist between two motifs, or between a terminal motif and the protein end. When an N- or C-terminus is unknown, the amino acid distance is shown as ‘xxx’. Motifs CP and T have not been found in G.lamblia and accordingly are absent from these alignments. At, Arabidopsis thaliana; Ca, Candida albicans; Ea, Euplotes aediculatus; Gl, Giardia lamblia; Hs, Homo sapiens; Mm, Mus musculus; Ot, Oxytricha trifallax; Sc, Saccharomyces cerevisiae; Sp, Schizosaccharomyces pombe; Tt, Tetrahymena thermophila; Xl, Xaenopus laevis; Pb, P.berghei; Pf, P.falciparum; Pk, P.knowlesi; and Py, P.yoelii. 1114 Nucleic Acids Research, 2005, Vol. 33, No. 3 and composition and eventually outputs the ones that are stat- region of PfTERT unique to Plasmodium species. A mouse istically relevant. serum was prepared from the immunization of C129 mice with a glutathione S-transferase (GST)-fusion protein correspond- Construction of motif alignment and phylogenetic tree ing to the 220 amino acids that span motifs 1 and 2. Known alignments of the motifs in other characterized species Human autoimmune serum (S63) directed against fibrillarin were used to identify the motifs in the Plasmodium species by (Nop1 is the Saccharomyces cerevisiae orthologue) was iden- independently using Clustal W and HMMer (33,34). Both tified in the extracts of fibrillarin–green fluorescent protein analyses provided similar results. Once identified and aligned, human cell line using western–blot analysis. These sera motifs were manually checked, in order to remove the posi- were found to cross-react with fibrillarin in many species, tions showing a very large variability (Figure 1). including yeast Nop1 (36). A rabbit anti-PfNop1 serum was For each species (except Giardia lamblia and Caenorhab- obtained by immunizing rabbits with two synthetic peptides ditis elegans), the amino acid sequences of all motifs were coupled with KLH, such as N-GRGNKDRKSFKKDNK-C concatenated in order to produce a single continuous sequence. and N-DLTNMSKKRSNIVPI-C. In P.falciparum, for example, this resulted in a concatenated sequence with 343 amino acids. Phylogenetic trees were con- Immunofluorescence microscopy structed using PROML from the PHYLIP package (35), using Parasitized erythrocytes from a fresh culture of 8–12% para- the JTT model and the Gamma correction with eight discrete sitemia were centrifuged and washed once in RPMI 1640 classes. The 1000 bootstraps values were computed using (Gibco-BRL) and once in phosphate-buffered saline (PBS) SEQBOOT and CONSENSE, also from the PHYLIP package (0.15 M NaCl and 10 mM sodium phosphate, pH 7.2). Cell (Figure 3). pellets were re-suspended in 5 vol of PBS and a monolayer was To compute the average similarity of P.falciparum TERT set onto microscope slides as described previously (37). Cells with the other Plasmodium TERTs, a multiple alignment was were air-dried for 30 min and fixed for 10 min at room tem- constructed using the entire sequences of the four Plasmodium perature in freshly prepared 2% paraformaldehyde in PBS. species. We then used a 20 amino acid sliding window along the Slides were blocked for 10 min in PBS 0.1% BSA. Rabbit multiple alignments to score for alignment quality (Figure 2). anti-PfTERT and human anti-Nop1 sera were diluted 1:200 in PBS 0.1% BSA and incubated with cells for 1 h. After washing Antibodies in PBS 0.1% BSA, cells were incubated for 45 min in a 1:200 Two different sera were produced against distinct regions of dilution of anti-rabbit or anti-human IgG-, fluorescein iso- the PfTERT protein. A rabbit serum was obtained by immun- thiocyanate- or rhodamine-conjugated antibodies (Sigma). izing rabbits with C-PfTERT peptide: a 15 amino acid Slides were washed thoroughly in PBS and mounted in peptide from the PfTERT C-terminus coupled with keyhole VECTASHIELD anti-fading with DAPI (40,6-diamidino-2- limpet hemocyanin (KLH) carrier protein (Sigma–Aldrich), phenylindole) (Vector Labs). Images were captured using a N-CKIKKRLINKYKIGH-C. This peptide corresponds to a Nikon E800 optical microscope.

Figure 2. Plasmodium TERTs are highly divergent outside telomerase and RT motifs. Average similarity of PfTERT to PkTERT (blue), Py TERT (green) and PbTERT (red). Similarity was computed in sliding windows of 20 amino acids over the multiple alignment. A score of 1 means that the 20 amino acids of the window are identical, whereas 0 indicates that the sequences do not share a single identical amino acid for a given position within the alignment (either because of substitutions or insertions/deletions). Above the graphics, the diagram indicates the in-scale position of all motifs in PfTERT protein. The asterisk points to an example of a region of PfTERT that is highly conserved among Plasmodium TERTs, but where no motif has yet been assigned. The lower panel shows the contribution of each Plasmodium TERT to the multiple alignments. Note that PyTERT and PbTERT sequences are not complete. Nucleic Acids Research, 2005, Vol. 33, No. 3 1115

Peptide competition assays were performed to test the spe- 4C were reached. After incubation in LR White resin at 20C, cificity of the anti-PfTERT antibodies. Increasing concentra- samples were cut into ultra-thin sections for immunolabelling. tions of a specific (C-PfTERT) and an unspecific peptide were They were then incubated with anti-PfTERT rabbit serum at added to the serum diluted 1:3 in PBS. The mixture was two different dilutions: 1:50 or 1:25 in TBG (0.1% gelatin and incubated overnight in a rotating wheel at 4C. The following 0.1% BSA in 0.1 M Tris buffer) for 1 h at room temperature. day, antibody/peptide complexes were diluted 1:70 in PBS, in After six 5 min washes in PBG, sections were incubated order to obtain a final dilution of the serum of around 1:200. IF for 1 h at room temperature with protein A conjugated with assay was carried out as described above. Images were cap- 10 nm colloidal gold (Auroprobe TM/EM protein A.G10 tured with constant acquisition parameters. RPN438; Batch 158170) diluted 1:20 in PBG. Controls with either no primary antibody or with pre-immune serum Immunofluorescence combined with in situ were performed. Ultra-thin sections were washed six times for hybridization 5 min in TBG and TBS. After fixation to grids with 1% glut- Cells were prepared as for standard IF except following fixa- araldehyde, thin sections were washed in distilled water and tion cells were permeabilized in 0.1% NP-40 in PBS for 5 min. post-stained for 10 min in uranyl acetate and 3 min in diluted After incubation with primary and secondary antibodies, cells lead citrate (1:2). Observations were made as described for were post-fixed in 2% paraformaldehyde in PBS for 5–10 min. ultrastructure analysis. Fluorescence in situ hybridization (FISH) was performed as described previously (37). Hybridization was performed over- night with 0.5–1 ng/ml of probe in 50% formamide (Roche- RESULTS Boehringer), 10% dextran sulfate (Quantum), 2· SSPE and Identification of P.falciparum TERT gene 250 mg/ml herring sperm DNA. The probe used to visualize the With the perspective of using telomerase as a target for anti- sub-telomeric Rep20 repetitive DNA consists of a 1.4 kb frag- malarials, we have characterized the protein component of ment containing 70 Rep20 units, in which biotin was incor- telomerase in P.falciparum. Two previous observations indic- porated using the Biotin High-Prime kit (Roche). To detect ated the existence of such an enzyme in P.falciparum. First, biotinylated probes (Rep20), slides were incubated for 30 min telomeric DNA is composed of canonical G-rich repeats (25), at room temperature with 50 ng/ml of Rhodamine-conjugated suggesting that telomeres are maintained by telomerase. Avidin (Roche). Second, telomerase enzymatic activity has previously been Ultrastructure analysis detected in semi-purified nuclear extracts (30). Attempts to PCR amplify a PfTERT gene based on consensus motifs gen- Cells were fixed using 0.1 M glutaraldehyde in phosphate erated from other genomes were not successful, and only after buffer containing 0.05 M sucrose for 1 h at 4C. After cent- advances in the sequencing project of malaria genomes we rifugation at 8000 g, cells were washed for 10 min in phos- were able to identify a candidate TERT gene. The P.falciparum phate buffer and twice for 10 min in 0.1 M cacodylate buffer. genome database was searched with the conserved motifs Post-fixation was performed in 1% osmium tetroxide in caco- described previously: the T motif and the seven RT motifs dylate buffer for 1 h, followed by a 30 min washing step in (10). Two overlapping contigs displayed a high score. Their cacodylate buffer. Cells were washed in 30% methanol for alignment resulted in a sequence of 10 kb, containing a large 10 min and stained in 2% uranyl acetate in 30% methanol for ORF of 7554 bp, which was predicted to encode a protein of 1 h at room temperature. After rinsing in 30% methanol, cells 2518 amino acids. We have named this gene PfTERT and its were dehydrated in a series of increasing concentrations of corresponding protein PfTERT. methanol (30–100%). Dehydrated cells were washed twice in propyl oxide and embedded in Epon 812 (Polysciences). Ultra- Features of P.falciparum TERT protein thin sections were cut with a Leica UltraCut UCT. Cuts were PfTERT contains all of the canonical motifs of RTs (motifs 1, post-stained with uranyl acetate and Reynold’s lead citrate. 2 and A–E; Figure 1A), as well as those conserved amino acids Observations were made on a Jeol 1200 EX II at 80 kV accel- known to be critical for the RT activity (1,38) (marked with erating voltage electron microscope. circles in Figure 1B). In the N-terminal half of PfTERT, we found three previously described motifs specific to TERTs: Immuno-electron microscopy GQ/N, QFP and T (16), but we failed to identify the CP motif. Cells were fixed for 1 h at 4C in 4% paraformaldehyde in Amino acids known to be specific for were found phosphate buffer. After washing in phosphate buffer twice for in PfTERT (marked with squares in Figure 1B): Arg in motif 1, 5 min, cells were incubated for 30 min at room temperature in an aromatic residue (Phe or Tyr) following the two critical Asp 0.25% NH4Cl in phosphate buffer in order to inactivate resid- residues in motif C and a motif similar to the Trp-X-Gly-X- ual aldehyde groups. Cells were washed twice in phosphate Ser/Leu in motif E. buffer and finally embedded in 10% gelatin buffer on ice. The PfTERT predicted molecular weight is 280 kDa, Specimens were next infused in 1.7 M sucrose + 15% which is almost three times the size of other TERTs described polyvinylpirrolidone for at least 4 h at 4C. Small blocks previously (e.g. S.cerevisiae Est2p is 103 kDa). Among most (1 mm3) were cryofixed by immersion in liquid nitrogen at of the TERTs, slight differences in length are chiefly due to 196C and by adding substitution medium in an automate size variations in the N-terminal half of the protein and freeze substitution (Leica). Substitution was performed for between motifs A and B0 (6). In contrast, the difference in 16 h at 90C in anhydrous methanol containing 0.5% uranyl size of the Plasmodial TERT results from increased distance acetate, followed by temperature increments of 5C/h, until between nearly all motifs. The in-scale diagram of Figure 1A 1116 Nucleic Acids Research, 2005, Vol. 33, No. 3

Table 1. Comparative analysis of the DNA coding sequence of some polymerases in P.falciparum and S.cerevisiae

Gene Function Length of Pf/Yeast %(A+T) of No. of tandem repeats in Pf gene (nt) ratio Pf gene Pf gene (no. single nt repeats)

PFE0465c RNA pol I (LSU) 8745 1.75 76.6 20 (2) PF11_0264 DNA-directed RNA pol, mitochondrial 4596 1.13 76.3 16 (7) PF13_0150 DNA-directed RNA pol 3 (LSU) 7071 1.61 76.4 27 (4) PFD0590c DNA pol a 5739 1.30 77.7 19 (4) PFC0805w DNA-directed RNA pol II 7374 1.42 72.3 22 (0) PFC0340w DNA pol d (SSU) 1497 1.02 76.3 10 (1) PF10_0165 DNA pol d 3285 1.00 74.8 2 (0) PfTERT TERT 7554 2.85 80.8 28 (11)

Pf/Yeast ratio is the length of the P.falciparum gene divided by the length of the yeast orthologue (as defined in PlasmoDB). The far right column indicates the number of tandem repeats in each Pf gene (‘mreps’ software). The number of nucleotide repeats is shown in parentheses. Abbreviations: LSU, larger subunit; and SSU, smaller subunit. shows that the spacing between the motifs is much greater in of this genome project. Nevertheless, PyTERT and PfTERT PfTERT than in ScEst2p. Only two blocks of motifs show share 43% identical amino acids. PbTERT gene, although relatively conserved spacing: motifs 1+2 and motifs B0+ incomplete, spans motifs GQ/N, QFP, T, 1 and 2. The amino C+D (Figure 1B). As observed in other P.falciparum proteins, acid identity between PfTERT and this truncated PbTERT PfTERT contains insertions of basic amino acids that are candidate is 54%. The two species that infect rodents, PyTERT absent from its orthologues in other genomes (39). For and PbTERT are highly homologous (76% identity), suggest- example, one of these segments (between motifs T and 1) ing closer ancestry. consists of a stretch of 30 asparagines, interspaced by a few The overall percentage of amino acid identity between lysines, isoleucines and tyrosines. PfTERT and the other Plasmodial TERTs is surprisingly Since Plasmodium proteins are systematically larger than low for a protein that has a conserved function in eukaryotes their orthologues in yeast or humans (40), we tested whether (e.g. hTERT and msTERT share 63% amino acid identity). that difference is the same for PfTERT and other DNA and We carried out a multiple alignment of the four Plasmodial RNA polymerases (Table 1). As expected, most polymerases TERT proteins and plotted the frequency of identical amino genes are larger in P.falciparum (Pf/Yeast ratio between 1.00 acids along the alignment (Figure 2). As expected, the regions and 1.75), but surprisingly the difference between TERT containing the telomerase motifs are most highly conserved. sequences is much bigger (Pf/Yeast ratio = 2.85). To test Notably, we can also see regions of high similarity, where whether this is related to the number of stretches of basic motifs have not yet been described (asterisk in Figure 2). amino acids easily detectable in the PfTERT protein sequence, Nevertheless, large domains share very little similarity we computed the number of significant large tandem repeats in between Plasmodium candidate TERT proteins. These typic- these genes (Table 1). We observed a larger number of repeats ally correspond to the regions showing repetitive sequences in PfTERT, 28, than in other Plasmodium polymerases of and characterized by many small insertions and deletions, as similar size. This difference is mainly due to the higher num- well as an unusual sequence composition, as discussed above. ber of nucleotide repeats, 11, such as tandem A or T nucle- otides. This is consistent with the higher A+T content of this Phylogenetic analysis of TERT proteins gene (80.8%) relative to the others. In conclusion, we assume To test the extent of conservation of the four Plasmodium that the large size of PfTERT gene is due partially to a general TERTs among eukaryotes, we undertook a phylogenetic ana- trend of Plasmodium genes being larger than the yeast ortho- lysis of all eukaryotic TERTs described to date. By concat- logues, and also due to the presence of more A/T repeats. enating all the conserved motifs within each TERT, we built a phylogenetic tree using the maximum-likelihood algorithm. Identification of other Plasmodial TERT The statistical significance of branches was tested by bootstrap candidate genes analysis (35). We eliminated G.lamblia and C.elegans from the analysis, as both lack several of the motifs present in the For future tests of anti-telomerase drugs in experimental mal- other TERTs. Besides, the low conservation of the G.lamblia aria models (mice and monkeys), it will be useful to know the and C.elegans TERT has been pointed out previously (8), and TERT sequences of the Plasmodium species that cause malaria evidence has not yet been obtained of telomerase activity in in these animals. To search for TERT homologues in non- these species. The final tree (Figure 3), despite displaying a human malaria species, we searched the genome databases few branches with low bootstrap values, is compatible with the of rodent and simian Plasmodium species with the PfTERT phylogeny based on the small subunit ribosomal RNA for protein sequence. We found three candidate TERT proteins: these species. Moreover, it lends support to the hypothesis PkTERT, in the simian P.knowlesi species and PyTERT and that telomerase was present in early eukaryotes. PbTERT, in the rodent species P.yoelii and P.berghei, respect- ively. PkTERT shares 42% amino acid identity with PfTERT Sub-nuclear localization of PfTERT in blood stages and contains all motifs present in PfTERT. PyTERT also contains all PfTERT motifs, but it appears to lack a complete Given the role of telomerase in DNA replication, a nuclear N-terminus, probably as a consequence of the ongoing status localization is expected for this enzyme. It is known that the Nucleic Acids Research, 2005, Vol. 33, No. 3 1117

Figure 3. The phylogenetic tree based on sequences of concatenated telomerase motifs is fully compatible with the known phylogeny for these species. The tree was built as described in Materials and Methods. A total of 1000 bootstraps experiments were performed, and bootstraps >50% are shown. Abbreviations of the species are as in Figure 1. The unit of branch length is the expected fraction of changed amino acids. nuclear localization signals (NLS) are conserved among euka- ryotes. Based on an algorithm that predicts sorting signals and localization sites encoded in protein sequences (PSORTII) (http://psort.ims.u-tokyo.ac.jp/; Kenta Nakai, Center, Tokyo, Japan), we found that PfTERT contains several NLS-like motifs. In fact, the k-nearest neighbours classifier (k-NN) algorithm predicts that PfTERT primary sequence refers to a protein that has 78.3% probability to be nuclear (for Tetrahymena TERT, for example, the probability is 52.2%). In order to test this prediction, we performed IF assays on Figure 4. PfTERT localizes in a discrete nuclear region in certain blood stream blood stage parasites. The blood stage cycle takes 48 h to be stages. (A) Diagram indicating the source of two different antibodies used to completed. It begins with the invasion of non-infected eryth- characterize PfTERT: a mouse serum raised against a GST–PfTERT fusion rocytes by merozoites present in the bloodstream. Once inside protein and a rabbit serum raised against a 15 amino acid peptide at the PfTERT C-terminus (C-PfTERT peptide). (B) The 48 h time course in blood stages, the erythrocyte, the parasite will undergo three stages: ring during which parasites develop inside an erythrocyte. PfTERT (green) was (0–18 h), trophozoite (18–38 h) and schizont (38–48 h) detected by IF using the rabbit anti-peptide antibody; nuclear DNA was stained stages. S phase begins in the trophozoite stage, 28–31 h after with DAPI (blue). (C) Peptide competition assay. The specificity of PfTERT merozoite invasion. Nuclear division occurs throughout serum was determined by IF after pre-incubating the antibodies with increasing concentrations of a specific (C-PfTERT) or an unspecific peptide. Detection of schizogony, leading to the production of up to 32 individual anti-PfTERT antibodies and staining of DNA was performed as in (B). merozoites. At the end of the 48 h cycle, erythrocytes burst releasing merozoites and the cycle starts anew. Previous stud- ies have shown that telomerase activity is detectable by TRAP pattern was observed: a single dot within a region of the only in trophozoite and schizont stages (41). To follow the nucleus. During schizogony, the number of PfTERT foci expression and localization of PfTERT during the blood stage increased proportionally to the number of nuclei. When schi- cycle by IF, we collected infected erythrocytes from a syn- zonts become fully mature (48 h) and merozoites are released chronized culture at different time points: 0, 12, 24, 36, 40 and into the blood stream, PfTERT was no longer detectable. 48 h. Studies using rabbit a-PfTERT-peptide (Figure 4), Two lines of evidence confirm the specificity of the PfTERT revealed no staining in the first 12 h (ring stages). When signal. First, when the above IF assay was undertaken in the parasite matures into trophozoite stage, however, a clear the presence of competing epitopes, the PfTERT signal 1118 Nucleic Acids Research, 2005, Vol. 33, No. 3

PfTERT localizes in the nucleolus, we searched for nucleolar markers. By IF, we tested different sera against highly con- served nucleolar proteins. Strikingly, human anti-hNop1 serum revealed a hat-like structure polarized towards one side of the nucleus (Figure 6A), very similar to that described for S.cerevisiae. Rabbit antibodies raised against a recombin- ant PfNop1 protein showed an identical staining pattern (Figure 6A). Immuno-electron microscopy (EM) observations further support that the anti-hNop1 serum is a nucleolar mar- ker, as it preferentially stains a sub-nuclear region (data not shown). Taken together, our data provide evidence for the Figure 5. PfTERT is not detectable at telomeres. Indirect-IF of PfTERT (green) combined with FISH of Rep20 (red), a sub-telomeric repetitive motif. DNA was existence of a nucleolus in P.falciparum. stained with DAPI (blue). Both cells are blood stage trophozoites. In order to test whether PfTERT localizes to the nucleolus, we performed IF in which cells were double-labelled using the rabbit anti-PfTERT and the human anti-hNop1 sera. Strik- disappeared with increasing concentrations of the C-terminal ingly, PfTERT co-localized with the Nop1-defined region, PfTERT peptide (C-PfTERT); in contrast, when an unspecifc indicating it localizes to the nucleolus. Remarkably, PfTERT peptide was used as competitor, the PfTERT signal intensity is not found all over the nucleolus, but is restricted to a sub- remained unchanged (Figure 4C). Second, the mouse sera compartment of this organelle (Figure 6B). This contrasts with produced against a recombinant PfTERT protein generated the situation in other eukaryotes (42–44), in which ectopically a similar localization pattern (data not shown). expressed TERT accumulates uniformly throughout the nuc- Thus, our results show that PfTERT is expressed at detect- leolus. Taken together, these data strongly suggest that able levels in trophozoites and schizonts, the same stages that PfTERT localizes in a sub-compartment of the nucleolus. were previously shown to have telomerase activity and where This finding is further supported by immuno-EM studies. DNA replication occurs. Moreover, our studies also revealed The anti-PfTERT serum revealed a single cluster of gold that PfTERT is not detectable throughout the entire nucleo- beads close to the nuclear membrane (Figure 6C). plasm, but rather in a discrete sub-nuclear compartment.

PfTERT: localization distinct from telomere clusters DISCUSSION The sub-nuclear localization of PfTERT is surprising consid- In the present work, we identified and characterized a gene in ering that telomeres, the substrate, are organized in 4–7 clus- P.falciparum that has the characteristics of the protein compon- ters around the nuclear periphery (27). We hypothesized that ent of telomerase, named here PfTERT. It contains structural the plasmodial telomerase complex may localize to a single features that are common with the telomerases known in other cluster and then ‘hop’ to another. To test this hypothesis, we species: RT motifs, telomerase-specific motifs and pI > 10. We combined IF with FISH, using a probe (Rep20) that recognizes also identified homologous genes in other Plasmodial species: a sub-telomeric repeat (telomeric DNA and Rep20 FISH sig- the simian P.knowlesi, PkTERT, and two murine species, nals co-localize; Scherf et al., unpublished data) and examined P.yoelii and P.berghei (PyTERT and PbTERT genes). if PfTERT co-localizes with a telomere cluster (Figure 5). A comparison of PfTERT and other eukaryote TERTs shows Optimization of the IF/FISH technique allowed to have very different degrees of conservation along the sequence. This >50% of the cells double-labelled. Strikingly, we concluded is in agreement with experimental data in S.cerevisiae, where that in the majority of the cells, telomeres and PfTERT did not hypermutable segments separate conserved and vital regions co-localize. We observed occasional colocalization of of the molecule (15). Within the conserved regions, we were PfTERT with telomere clusters, which may be due to errors able to identify almost all of the TERT-specific and all of the prone to the 2D analysis of the images. We conclude that, RT-specific motifs. The CP motif, which was not found in any within the detection limits of IF, PfTERT is located in a sub- Plasmodia, is the only exception. This motif is also absent from nuclear compartment that is distinct from telomeric clusters. G.lamblia TERT, and given its very low-sequence conserva- tion, doubts have been put forward for its significance outside PfTERT: localization in a sub-compartment the ciliates group (6,15). It is thus unclear if this motif has of the nucleolus diverged beyond recognition or if it has appeared after the The largest sub-nuclear compartment of eukaryotic cells is the separation between Plasmodia and Giardia and the other nucleolus, the main function of which is the transcription of eukaryotes analysed. rRNA genes and the assembly of ribosomes. Some ribonuc- PfTERT contains several stretches of 10–20 basic amino leoproteins are also assembled in this compartment. In recent acids, such as asparagines, which are encoded by A-rich studies, human and yeast TERTs were shown to be enriched at codons. Interestingly, among P.falciparum polymerases, the the nucleolus [reviewed in (4)]. Given PfTERT’s localization TERT gene is the one that contains more A+T nucleotides, to a discrete domain of the nucleus away from the telomeres, more repeats and a higher increase in size relative to the yeast we hypothesized that it localizes to the nucleolus. orthologue (Table 1), suggesting that PfTERT sequence is less In P.falciparum, a nucleolus has never been identified and functionally constrained than that of the other polymerases. its presence has even been questioned given the low number of These repeats are therefore more likely to reflect the propen- ribosomal genes in the genome (40). Thus, in order to test if sity of the P.falciparum genome by polymerase slippage Nucleic Acids Research, 2005, Vol. 33, No. 3 1119

Figure 6. PfTERT localizes in a sub-compartment of the nucleolus. (A) IF with an anti P.falciparum Nop1 serum (red) shows the nucleolus as a hat-like structure in the nucleus of a trophozoite stage. An anti-human Nop1 serum (green) recognizes exactly the same sub-nuclear region. (B) Double-labelling reveals that PfTERT (red) co-localizes with part of the nucleolus (green); yellow spot in the merge image. (C) Immuno-EM shows that PfTERT (white arrow head) localizes to a peripheral region of the nucleus. In (A and B), DNA was stained with DAPI (blue). events due to the presence of stretches of A and/or T (45,46), A phylogenetic tree built from sequences consisting of a instead of selection process that leads to increased gene length. concatenation of the conserved motifs, ignoring the hypervari- If this is so, the extreme A+T richness of Plasmodia genomes able stretches, respects the known phylogenetic relationships, (80% A+T for P.falciparum) may favour the increase in gene both among Plasmodia species and among eukaryotes in gen- size and led to the accumulation of the repetitive regions in eral. Our results confirm previous analysis made with a smaller non-essential gene areas. dataset (6), and reinforce the hypothesis that telomerase was Given the common biological role of telomerase in euka- present in early eukaryotes. The phylogenetic analysis is com- ryotes, we expected its catalytic component to be much con- plicated by a few branches lacking robustness (low-bootstrap served among species of the same genus. Surprisingly, we values) and by the lack of motifs in the sequences of the found very little sequence conservation between the TERT putative TERT of C.elegans and G.lamblia. Addition of sequences of the different Plasmodium species: PfTERT newly sequenced eukaryotic genomes will allow to further shares 42% identical amino acids with PkTERT, 43% with resolve these questions. PyTERT and 54% with PbTERT. The regions between the It was previously shown that, in P.falciparum cell extracts, conserved motifs show lower amino acid identity (Figure 2). telomerase can be efficiently inhibited by RT type of drugs, The existence of such hypervariable regions is indicative of a such as nucleoside analogues (30). These same drugs are cur- continuous process of genome evolution of malaria species, as rently being tested on in vitro cultures and preliminary data it was previously shown by a phylogenetic analysis based on show killing of P.falciparum parasites after 3–5 blood stage comparison of the small-subunit ribosomal RNA gene cycles at micromolar concentrations (data not shown). More- sequences (47). These differences among Plasmodial TERT over, we were unable to generate a knock-out of the PfTERT proteins may play a role in the specific mean telomere length gene, supporting the idea that telomerase activity is needed for observed for the different species (2.0 kb for P.yoelii, 1.2 kb blood stage parasite proliferation. Thus, a prophylactic therapy for P.falciparum and 6.7 kb for P.vivax) (26). based on plasmodial telomerase inhibitors might be possible. 1120 Nucleic Acids Research, 2005, Vol. 33, No. 3

In this work, the comparison of the TERT sequences from human, rodent and simian Plasmodium species provides valu- able information for the design of specific anti-telomerase drugs. Obviously, drugs should be targeted to conserved regions and functional domains, in order to avoid rapid emer- gence of drug resistance. The cell-cycle progression in P.falciparum is not yet com- pletely understood (48). DNA synthesis begins in the relat- ively small trophozoites (intra-erythrocytic form), but nuclear subdivision, which leads to the formation of multinucleate cells, occurs only during schizogony. Whether or not any gap phases (G) exist between each round of DNA synthesis and is unknown. Eventually, a schizont composed of 8–32 nuclei undergoes segmentation, which culminates with the formation of individual merozoites that burst from the erythrocyte into the blood stream. Our observations indicate that PfTERT is only detectable in trophozoites and schizonts, the stages where multiple rounds of DNA synthesis occur and Figure 7. Model for the mechanism of action of telomerase in P.falciparum. where telomerase activity can be detected (41). Thus, our We speculate that a sub-compartment of the nucleolus (green) is the location data reveal a correlation between the expression of PfTERT where telomerase complex (red) is assembled and stored. When telomeres protein and the pattern of telomerase activity in the parasite. (yellow) replicate, two scenarios can be envisioned. A, A subfraction (unde- As is known to occur in human cells, we speculate that tectable by IF) of the assembled/active telomerase is released from the sub- the regulation of PfTERT expression may be one of the nucleolar compartment and goes to the telomeres, where replication takes place. B, Telomeres form a bouquet-like structure during mitosis and move to the mechanisms that P.falciparum uses to control telomerase proximity of the nucleolus. Telomerase replicates the chromosome ends, while activity (49). remaining anchored in the nucleolus. Studies in S.cerevisiae and human cells have shown that ectopically expressed TERTs are partially enriched in the nucleolus (42–44). In humans, the telomerase RNA compon- ent also seems to be enriched at the nucleolus (50,51) or Cajal Surprisingly, a combination of immunolabelling with anti- bodies (52,53). These observations led to the speculation that PfTERT sera and FISH, did not reveal a co-localization of the nucleolus may play a dual role: providing an environment PfTERT with the telomeric clusters. The general view for the biogenesis of telomerase [reviewed in (4)] and regu- point in the telomerase field is that only a few telomerase lating the release of telomerase to its telomeric substrates (54). molecules are necessary to lengthen telomeres. This may In this work, we described the localization of the endogenous explain the perceptible absence of telomerase at telomeres telomerase catalytic component. We observed that PfTERT is using immunolocalization assays. By chromatin immuno- not evenly distributed in the nucleoplasm, but accumulates in precipitation, however, it was shown that anti-telomerase anti- a discrete peripheral site of the nucleus, which we identified as a bodies were capable of precipitating telomere repeats, sub-compartment of the nucleolus. In the P.falciparum related providing evidence that telomerase can be associated with apicomplexan parasite Toxoplasma, the nucleolus appears as telomeres in vivo (58). a dot in the middle of the nucleus (as detected by anti-human We propose two models that could explain our observations fibrillarin or a YFP-tagged nucleolar protein) (55). In (Figure 7). In the first model, we predict that a fraction of the P.falciparum, we observe a significantly larger structure activated telomerase is released from the nucleolus to the with a hat-like pattern mostly at the periphery of the nucleus. nucleoplasm, to allow complete telomeric DNA replication, This pattern was observed with human serum against hNop1 as observed with ectopically expressed hTERT (54). However, and confirmed with antibodies that we raised against in P.falciparum, levels of telomerase released to the nucleo- P.falciparum Nop1. Thus, our experimental data strongly sug- plasm may be insufficient to be detected by IF. After the gest that the nucleolus of P.falciparum has a more expanded addition of telomeric repeats onto the chromosome ends, structure than in Toxoplasma. telomerase is either degraded or shuttled back to the nucleolus. It is possible that, as in S.cerevisiae, the sub-nucleolar The second model speculates that telomerase is not released to PfTERT-containing structure is functionally equivalent to the nucleoplasm. Instead, it remains in a sub-nucleolar com- the mammalian Cajal bodies (56). The RNA component of partment where telomeres are elongated. In this scenario, chro- P.falciparum telomerase has not yet been identified and mosome ends would move to the nucleolus (or its proximity) thus we cannot test where the assembled telomerase com- when telomere replication takes place. In support of this plex localizes. We speculate that the presence of PfTERT second model, preliminary studies in our laboratory show a in the nucleolus may be necessary for the assembly of the bouquet-like structure of telomeres proximal to the nucleolus telomerase complex, given that in some eukaryotes this during DNA replication (data not shown). In both models, the sub-nuclear compartment is involved in the assembly of ribo- telomerase sub-nucleolar compartment may fulfil a role in the nucleoproteins (57). assembly of the telomerase complex and/or a role in the regu- Given the role of telomerase at telomeres, one might lation of telomerase activity by determining when telomerase predict a release of the telomerase activity from the nucleolus should have access to its substrate. Further studies are required to the nucleoplasm during the replication of telomeric DNA. to test these hypotheses. Nucleic Acids Research, 2005, Vol. 33, No. 3 1121

ACKNOWLEDGEMENTS 18. Shay,J.W. and Wright,W.E. (1996) Telomerase activity in human cancer. Curr. Opin. Oncol., 8, 66–71. We thank Christine Scheidig Benatar for technical assistance, 19. Marciniak,R. and Guarente,L. (2001) Human genetics. Testing Joanna Lowell for the revision of the manuscript and Stuart telomerase. Nature, 413, 370–371, 373. Ralph for help with electron microscopy. This work was sup- 20. Bodnar,A.G., Ouellette,M., Frolkis,M., Holt,S.E., Chiu,C.P., ported by a grant from the European Community (Network of Morin,G.B., Harley,C.B., Shay,J.W., Lichtsteiner,S. and Wright,W.E. (1998) Extension of life-span by introduction of telomerase into normal excellence BioMalPar). Funding to pay the open access pub- human cells. Science, 279, 349–352. lication charges for this article was provided by the same grant. 21. Mergny,J.L., Riou,J.F., Mailliet,P., Teulade-Fichou,M.P. and Gilson,E. L.M.F and L.M.S. were supported by fellowships (Ref. (2002) Natural and pharmacological regulation of telomerase. Nucleic PRAXIS-XXI/16020/98 and SFRH/BD/11756/2003, respec- Acids Res., 30, 839–865. 22. Knell,A.J. (1991) Malaria. Trustees of the Welcome Trust, Oxford tively) from the Fundac¸~aao para a Ci^eencia e Tecnologia, University Press, Oxford. Portugal. 23. Ridley,R.G. (2002) Medical need, scientific opportunity and the drive for antimalarial drugs. Nature, 415, 686–693. 24. Hoffman,S.L., Subramanian,G.M., Collins,F.H. and Venter,J.C. (2002) Plasmodium, human and Anopheles genomics and malaria. Nature, REFERENCES 415, 702–709. 25. Vernick,K.D. and McCutchan,T.F. (1988) Sequence and structure of a 1. Lingner,J., Hughes,T.R., Shevchenko,A., Mann,M., Lundblad,V. and Plasmodium falciparum telomere. Mol. Biochem. Parasitol., 28, 85–94. Cech,T.R. (1997) Reverse transcriptase motifs in the catalytic subunit of 26. Figueiredo,L.M., Pirrit,L.A. and Scherf,A. (2000) Genomic organisation telomerase. Science, 276, 561–567. and chromatin structure of Plasmodium falciparum chromosome ends. 2. Nakamura,T.M., Cooper,J.P. and Cech,T.R. (1998) Two modes of Mol. Biochem. Parasitol., 106, 169–174. survival of fission yeast without telomerase. Science, 282, 493–496. 27. Freitas-Junior,L.H., Bottius,E., Pirrit,L.A., Deitsch,K.W., Scheidig,C., 3. Blasco,M.A., Lee,H.W., Hande,M.P., Samper,E., Lansdorp,P.M., Guinet,F., Nehrbass,U., Wellems,T.E. and Scherf,A. (2000) Frequent DePinho,R.A. and Greider,C.W. (1997) Telomere shortening and tumor ectopic recombination of virulence factor genes in telomeric formation by mouse cells lacking telomerase RNA. Cell, 91, 25–34. chromosome clusters of P. falciparum. Nature, 407, 1018–1022. 4. Vega,L.R., Mateyak,M.K. and Zakian,V.A. (2003) Getting to the end: 28. McEachern,M.J., Krauskopf,A. and Blackburn,E.H. (2000) Telomeres telomerase access in yeast and humans. Nature Rev. Mol. Cell. Biol., and their control. Annu. Rev. Genet., 34, 331–358. 4, 948–959. 29. Scherf,A., Figueiredo,L.M. and Freitas-Junior,L.H. (2001) Plasmodium 5. Evans,S.K. and Lundblad,V. (2000) Positive and negative regulation of telomeres: a pathogen’s perspective. Curr. Opin. Microbiol., 4, 409–414. telomerase access to the telomere. J. Cell. Sci., 113, 3357–3364. 30. Bottius,E., Bakhsis,N. and Scherf,A. (1998) Plasmodium falciparum 6. Bryan,T.M., Sperger,J.M., Chapman,K.B. and Cech,T.R. (1998) telomerase: de novo telomere addition to telomeric and nontelomeric Telomerase reverse transcriptase genes identified in Tetrahymena sequences and role in chromosome healing. Mol. Cell. Biol., 18, 919–925. thermophila and Oxytricha trifallax. Proc. Natl Acad. Sci. USA, 31. Scherf,A., Hernandez-Rivas,R., Buffet,P., Bottius,E., Benatar,C., 95, 8479–8484. Pouvelle,B., Gysin,J. and Lanzer,M. (1998) Antigenic variation in 7. Collins,K. and Gandhi,L. (1998) The reverse transcriptase component of malaria: in situ switching, relaxed and mutually exclusive transcription of the Tetrahymena telomerase ribonucleoprotein complex. Proc. Natl var genes during intra-erythrocytic development in Plasmodium Acad. Sci. USA, 95, 8485–8490. falciparum. EMBO J., 17, 5418–5426. 8. Malik,H.S., Burke,W.D. and Eickbush,T.H. (2000) Putative telomerase 32. Kolpakov,R., Bana,G. and Kucherov,G. (2003) mreps: efficient and catalytic subunits from Giardia lamblia and Caenorhabditis elegans. flexible detection of tandem repeats in DNA. Nucleic Acids Res., Gene, 251, 101–108. 31, 3672–3678. 9. Counter,C.M., Meyerson,M., Eaton,E.N. and Weinberg,R.A. (1997) 33. Eddy,S.R. (1998) Profile hidden Markov models. Bioinformatics, The catalytic subunit of yeast telomerase. Proc. Natl Acad. Sci. USA, 14, 755–763. 94, 9202–9207. 34. Thomson,J.D., Higgins,D.G. and Gibson,T.J. (1994) Clustal W: 10. Nakamura,T.M., Morin,G.B., Chapman,K.B., Weinrich,S.L., improving the sensitivity of progressive multiple sequence alignment Andrews,W.H., Lingner,J., Harley,C.B. and Cech,T.R. (1997) through sequence weighting, positions-specific gap penalties and weight Telomerase catalytic subunit homologs from fission yeast and human. matrix choice. Nucleic Acids Res., 22, 4673–4680. Science, 277, 955–959. 35. Felsenstein,J. (1993) PHYLIP (Phylogeny Inference Package) version 11. Metz,A.M., Love,R.A., Strobel,G.A. and Long,D.M. (2001) Two 3.6a. Department of Genetics, University of Washington, Seattle, WA. telomerase reverse transcriptases (TERTs) expressed in Candida 36. Gautier,T., Fomproix,N., Masson,C., Azum-Gelade,M.C., Gas,N. and albicans. Biotechnol. Appl. Biochem., 34, 47–54. Hernandez-Verdun,D. (1994) Fate of specific nucleolar 12. Oguchi,K., Liu,H., Tamura,K. and Takahashi,H. (1999) Molecular perichromosomal proteins during mitosis: cellular distribution and cloning and characterization of AtTERT, a telomerase reverse association with U3 snoRNA. Biol. Cell., 82, 81–93. transcriptase homolog in Arabidopsis thaliana. FEBS Lett., 37. Freitas-Junior,L.H. and Scherf,A. (2000) Methods in Malaria Research. 457, 465–469. ATCCC Books, Waldorf, MD. 13. Meyerson,M., Counter,C.M., Eaton,E.N., Ellisen,L.W., Steiner,P., 38. Weinrich,S.L., Pruzan,R., Ma,L., Ouellette,M., Tesmer,V.M., Holt,S.E., Caddle,S.D., Ziaugra,L., Beijersbergen,R.L., Davidoff,M.J., Liu,Q. et al. Bodnar,A.G., Lichtsteiner,S., Kim,N.W., Trager,J.B. et al. (1997) (1997) hEST2, the putative human telomerase catalytic subunit gene, Reconstitution of human telomerase with the template RNA component is up-regulated in tumor cells and during immortalization. Cell, hTR and the catalytic protein subunit hTRT. Nature Genet., 17, 90, 785–795. 498–502. 14. Greenberg,R.A., Allsopp,R.C., Chin,L., Morin,G.B. and DePinho,R.A. 39. Aravind,L., Iyer,L.M., Wellems,T.E. and Miller,L.H. (2003) (1998) Expression of mouse telomerase reverse transcriptase during Plasmodium biology: genomic gleanings. Cell, 115, 771–785. development, differentiation and proliferation. Oncogene, 40. Gardner,M.J., Hall,N., Fung,E., White,O., Berriman,M., Hyman,R.W., 16, 1723–1730. Carlton,J.M., Pain,A., Nelson,K.E., Bowman,S. et al. (2002) Genome 15. Friedman,K.L. and Cech,T.R. (1999) Essential functions of amino- sequence of the human malaria parasite Plasmodium falciparum. Nature, terminal domains in the yeast telomerase catalytic subunit revealed by 419, 498–511. selection for viable mutants. Genes Dev., 13, 2863–2874. 41. Sriwilaijareon,N., Petmitr,S., Mutirangura,A., Ponglikitmongkol,M. and 16. Xia,J., Peng,Y., Mian,I.S. and Lue,N.F. (2000) Identification of Wilairat,P. (2002) Stage specificity of Plasmodium falciparum functionally important domains in the N-terminal region of telomerase telomerase and its inhibition by berberine. Parasitol. Int., 51, 99–103. reverse transcriptase. Mol. Cell. Biol., 20, 5196–5207. 42. Etheridge,K.T., Banik,S.S., Armbruster,B.N., Zhu,Y., Terns,R.M., 17. Liu,J.-P. (2002) Molecular mechanisms regulating telomerase activity. Terns,M.P. and Counter,C.M. (2002) The nucleolar localization domain In Telomerase, Aging and Disease Mattson,M.P. and Pandita,T.K. (eds), of the catalytic subunit of human telomerase. J. Biol. Chem., 277, Elsevier Science, The Netherlands, pp. 33–59. 24764–24770. 1122 Nucleic Acids Research, 2005, Vol. 33, No. 3

43. Yang,Y., Chen,Y., Zhang,C., Huang,H. and Weissman,S.M. (2002) functions to localize the RNA within the nucleus. RNA, 7, Nucleolar localization of hTERT protein is associated with telomerase 1833–1844. function. Exp. Cell. Res., 277, 201–209. 52. Zhu,Y., Tomlinson,R.L., Lukowiak,A.A., Terns,R.M. and Terns,M.P. 44. Teixeira,M.T., Forstemann,K., Gasser,S.M. and Lingner,J. (2002) (2004) Telomerase RNA accumulates in Cajal bodies in human cancer Intracellular trafficking of yeast telomerase components. EMBO Rep., cells. Mol. Biol. Cell, 15, 81–90. 3, 652–659. 53. Jady,B.E., Bertrand,E. and Kiss,T. (2004) Human telomerase RNA and 45. Achaz,G., Netter,P. and Coissac,E. (2001) Study of intrachromosomal box H/ACA scaRNAs share a common Cajal body-specific localization duplications among the eukaryote genomes. Mol. Biol. Evol., 18, signal. J. Cell. Biol., 164, 647–652. 2280–2288. 54. Wong,J.M., Kusdra,L. and Collins,K. (2002) Subnuclear shuttling of 46. Levinson,G. and Gutman,G.A. (1987) Slipped-strand mispairing: a major human telomerase induced by transformation and DNA damage. Nature mechanism for DNA sequence evolution. Mol. Biol. Evol., Cell. Biol., 4, 731–736. 4, 203–221. 55. Gubbels,M.J., Wieffer,M. and Striepen,B. (2004) Fluorescent protein 47. Waters,A.P., Higgins,D.G. and McCutchan,T.F. (1991) Plasmodium tagging in Toxoplasma gondii: identification of a novel inner membrane falciparum appears to have arisen as a result of lateral transfer between complex component conserved among Apicomplexa. Mol. Biochem. avian and human hosts. Proc. Natl Acad. Sci. USA, 88, 3140–3144. Parasitol., 137, 99–110. 48. Arnot,D.E. and Gull,K. (1998) The Plasmodium cell-cycle: facts and 56. Verheggen,C., Lafontaine,D.L., Samarsky,D., Mouaikel,J., questions. Ann. Trop. Med. Parasitol., 92, 361–365. Blanchard,J.M., Bordonne,R. and Bertrand,E. (2002) Mammalian and 49. Cong,Y.S., Wright,W.E. and Shay,J.W. (2002) Human telomerase and its yeast U3 snoRNPs are matured in specific and related nuclear regulation. Microbiol. Mol. Biol. Rev., 66, 407–425Table of contents. compartments. EMBO J., 21, 2736–2745. 50. Mitchell,J.R., Cheng,J. and Collins,K. (1999) A box H/ACA small 57. Pederson,T. (1998) The plurifunctional nucleolus. Nucleic Acids Res., nucleolar RNA-like domain at the human telomerase RNA 30 end. 26, 3871–3876. Mol. Cell. Biol., 19, 567–576. 58. Taggart,A.K., Teng,S.C. and Zakian,V.A. (2002) Est1p as a cell 51. Lukowiak,A.A., Narayanan,A., Li,Z.H., Terns,R.M. and Terns,M.P. cycle-regulated activator of telomere-bound telomerase. Science, 297, (2001) The snoRNA domain of vertebrate telomerase RNA 1023–1026.

RESULTS II

Differential association of Orc1 and Sir2 proteins to telomeric domains in Plasmodium falciparum

Liliana Mancio-Silva1, Ana Paola Rojas-Meza1,2, Miguel Vargas2, Artur Scherf1 and Rosaura Hernandez-Rivas2

1 Unité de Biologie des Interactions Hôte-Parasite, CNRS URA2581, Institut Pasteur, 25 Rue du Dr Roux, 75724 Paris, France 2 Departamento de Biomedicina Molecular, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional (IPN), Apartado postal 14-740, 07360 México, D. F., México

Published in Journal of Cell Science. 2008 Jun 15;121(Pt 12):2046-53.

39

40 2046 Research Article Differential association of Orc1 and Sir2 proteins to telomeric domains in Plasmodium falciparum

Liliana Mancio-Silva1, Ana Paola Rojas-Meza1,2, Miguel Vargas2, Artur Scherf1,* and Rosaura Hernandez-Rivas2,* 1Unité de Biologie des Interactions Hôte-Parasite, CNRS URA 2581, Institut Pasteur, 25, Rue du Dr Roux, 75724 Paris, France 2Departamento de Biomedicina Molecular, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional (IPN), Apartado postal 14-740, 07360 México, D. F., México *Authors for correspondence (e-mails: [email protected]; [email protected])

Accepted 31 March 2008 Journal of Cell Science 121, 2046-2053 Published by The Company of Biologists 2008 doi:10.1242/jcs.026427

Summary Telomeres have the capacity to recruit proteins that facilitate Relocation of Orc1 and Sir2 was also linked to the partial the spreading of heterochromatin into subtelomeric DNA dissociation of telomere clusters. Super gel-shift and chromatin- regions. In the human protozoan pathogen Plasmodium immunoprecipitation experiments showed the physical falciparum, the telomere-associated protein Sir2 has been shown association of Orc1 with telomere repeats but revealed a to control the silencing of members of virulence genes at some, differential association with adjacent non-coding repeat DNA but not all, chromosome-end loci, indicating that additional elements. Our data suggest that Plasmodium telomeres might proteins are involved in telomere position effect. Here, we fold back and that Orc1 cooperates with Sir2 in telomeric identified, in P. falciparum, a novel telomere-associated protein silencing. that displays homology with the origin-of-recognition-complex 1 protein Orc1. Antibodies raised against this P. falciparum protein localized to telomeric clusters in the nuclear periphery Supplementary material available online at and the nucleolus. It was found that, prior to DNA replication, http://jcs.biologists.org/cgi/content/full/121/12/2046/DC1 P. falciparum Orc1 and Sir2 undergo drastic subcellular reorganization, such as dissociation from the telomere cluster Key words: Telomeres, Subtelomeric repeats, Nucleolus, Sir2, Orc1, and spreading into the nucleus and parasite cytoplasm. Blood-stage cycle, Telomere folding, Plasmodium falciparum

Journal of Cell Science Introduction to seven heterologous chromosome ends at the nuclear periphery Telomeres are essential protein-DNA structures that protect all (Freitas-Junior et al., 2000). Telomere repeats are followed by a chromosome ends from degradation and fusion, and are vital for mosaic of six different telomere-associated repetitive elements complete replication of chromosomes (Blackburn, 2006). In several (TAREs 1-6), which are always found in the same order but vary in organisms, telomeres are anchored to the nuclear periphery and recruit size (Figueiredo et al., 2000; Gardner et al., 2002). This subtelomeric a number of proteins that initiate and/or facilitate heterochromatin region is important to maintain chromosome ends clustered in the formation into the adjacent subtelomeric regions (Hediger and nuclear periphery and to regulate telomere length (Figueiredo et al., Gasser, 2002). In yeast, as in higher eukaryotes such as humans, genes 2002). The proteins that crosslink telomeres, however, remain positioned near to telomeric repeats are silenced by a phenomenon unknown (Marty et al., 2006). Members of the var multigene family called telomere position effect (TPE) (Baur et al., 2001; Gottschling are located adjacent to the non-coding TAREs. var genes encode a et al., 1990). This form of epigenetic silencing is best characterized key virulence factor expressed at the surface of infected red blood in budding yeast, and involves mainly Rap1 and Sir (silent information cells, which is strongly linked to malaria pathogenesis in humans regulator) proteins, which assemble in a step-wise manner. Rap1 binds (Kyes et al., 2001). The sequential expression of different members to telomeric repeats and recruits Sir2 via interaction with Sir4. Sir2 of the var gene family in a mono-allelic fashion (antigenic variation) deacetylates nucleosomes that favour the binding of Sir3 and Sir4. leads to immune escape and chronic infection of the parasite in its Multiple rounds of deacetylation and binding of Sir complex along human host (Craig and Scherf, 2001). the nucleosomes lead to heterochromatin propagation and silencing We previously identified orthologues to several yeast telomere- at the proximity of telomeres. This spreading extends for ~3 kb from associated proteins in the P. falciparum genome (Scherf et al., 2001) the telomere (Mondoux and Zakian, 2006). A rather unusual TPE and characterized in detail a protein homologous to the yeast Sir2 was observed recently in the human malaria parasite Plasmodium (Duraisingh et al., 2005; Freitas-Junior et al., 2005). P. falciparum falciparum. In this unicellular protozoan pathogen, heterochromatin Sir2 is a telomeric protein that reversibly associates with the spreads from telomeres and adjacent non-coding subtelomeric repeats promoter regions of silent but not active subtelomeric var genes over 20-30 kb into coding regions, thereby establishing gene (Freitas-Junior et al., 2005). Mutant-Sir2 parasites derepress var repression on different chromosome ends (Duraisingh et al., 2005; genes; however, transcription of only a subgroup of this gene family Freitas-Junior et al., 2005). is upregulated (Duraisingh et al., 2005), indicating that TPE is a P. falciparum telomeres, which are composed of degenerate G- rather complex process involving additional chromatin factors in rich heptamer repeats (5Ј-GGGTT[T/C]A-3Ј), form clusters of four malaria parasites. Relocation of telomeric proteins 2047

In this work, we identified a protein associated with telomeric and subtelomeric heterochromatin in P. falciparum. This molecule has been suggested to be the origin-of-recognition-complex 1 molecule (Orc1) of P. falciparum parasites (Gupta et al., 2006; Li and Cox, 2003; Mehra et al., 2005). Here, we show that P. falciparum Orc1 localizes specifically with distinct perinuclear sub- compartments: the telomeres and the nucleolus. In addition, we show that telomeric clusters disassemble prior to DNA replication, which is linked to P. falciparum Sir2 and Orc1 relocation within the nucleus and to the cytoplasm. Our findings indicate that nucleo-cytoplasmic shuttling of telomere-associated proteins might reveal novel functions of these molecules.

Results Sir2 and Orc1 colocalize to telomeric and nucleolar regions In order to search for putative factors involved in TPE in P. falciparum, we previously identified a sequence homologous to the yeast telomere-associated silencing factor Sir3 (Scherf et al., 2001). We generated antibodies against the putative P. falciparum Sir3 recombinant GST fusion protein. These antibodies recognized a band of 130 kDa in P. falciparum nuclear blood-stage extracts (supplementary material Fig. S1). Subsequent analysis revealed that the open reading frame identified is now annotated in the P. falciparum database (www.plasmodb.org) as being homologous to Fig. 1. Orc1 localizes to telomeric foci and the nucleolus. (A) Double IF using anti-rabbit Orc1 (red) and anti-rat Sir2 (green). Orc1 is preferentially localized the origin recognition complex 1 (Orc1) protein (PFL01050w). In at the periphery of the parasite nucleus and colocalizes with Sir2 signals. fact, Sir3 and Orc1 amino acid sequences are very similar in budding (B) Dual-colour IF by using anti-rabbit Orc1 (red) and anti-rat Nop1 (green). yeast because SIR3 resulted from duplication of the ORC1 gene A fraction of Orc1 signals colocalize with Nop1, suggesting that Orc1 also (Kellis et al., 2004). Orc1 has been found in yeast subtelomeric localizes in the nucleolus of the parasite. (A,B) Parasites are in ring stage and nuclear DNA was stained with DAPI (blue). Scale bars: 1 μm. regions (Pryde and Louis, 1999; Wyrick et al., 2001) and in other transcriptionally repressed domains, such as the mating-type loci in yeast (Bell et al., 1993) and the pericentric heterochromatin in Drosophila (Pak et al., 1997), where it participates in shift assay, the probes for telomere, TARE3 and TARE6 formed heterochromatin assembly. This led us to hypothesize that Orc1 is retarded complexes, indicating that telomeres and subtelomeric involved in telomeric silencing in P. falciparum. To test whether P. repeats contain elements specifically recognized by nuclear proteins. falciparum Journal of Cell Science Orc1 localizes to telomeric heterochromatic regions, we Telomere showed a more complex pattern, suggesting that distinct performed immunofluorescence assays in which parasites were protein complexes can bind to these sequences. The specificity of double-labelled using anti-Orc1 and -Sir2 antibodies. We observed these complexes was confirmed using 50-fold molar excess of either that more than 80% (n=104) of the Orc1 foci colocalized with the homologous or heterologous competitors (Fig. 2B). Sir2 signals (Fig. 1A), indicating that Orc1 localizes to telomeres. Anti-Sir2 antibodies produced supershifted complexes with However, Sir2 is not an exclusive marker for telomeric clusters but telomere, TARE3 and TARE6 probes. Anti-Orc1 antibodies formed also stains the nucleolus (Freitas-Junior et al., 2005). We used anti- a supershifted complex only with telomere and TARE3 (Fig. 2C). Nop1 antibodies as a nucleolar marker to test whether Orc1 also These results demonstrate that Sir2 and Orc1 bind specifically to localizes in the nucleolus. A fraction of Orc1 signals indeed telomeric and subtelomeric repeats. Moreover, the data suggest that colocalized with the region defined by Nop1 (Fig. 1B). We Orc1 spreads from the telomere to TARE3, whereas Sir2 expands confirmed the association of Orc1 to the telomeres and the nucleoli from telomeres to TARE6. by immunofluorescence/fluorescence in situ hybridization (IF- We used chromatin immunoprecipitation assays (ChIP) to further FISH) using telomeric (TARE6) (supplementary material Fig. S2) investigate the in vivo association of Sir2 and Orc1 with telomeric and nucleolar (rRNA) probes (data not shown), respectively. and subtelomeric chromatin. For this mapping study, we used DNA Our results show that, at ring stage, Orc1 localizes mainly at the probes for telomere repeats, TARE1, TARE2, a sequence between nuclear periphery rather than within the entire nucleoplasm. We TARE2 and TARE3, TARE 3, and TARE6 (Fig. 2A). We found also show that it localizes to similar compartments as Sir2, that Orc1 and Sir2 proteins are present at the telomere, TARE1-3 supporting the idea that Orc1 might participate in heterochromatin and TARE6 (Fig. 3A,B). This confirmed most of the EMSA data, formation in P. falciparum. with the difference that the ChIP experiments show that Orc1 was also enriched in TARE6. The observed difference in Orc1 location Sir2 and Orc1 proteins spread differentially into subtelomeric might be explained by the fact that EMSA does not take into repeats consideration possible particular telomere architecture that could To study the binding of Sir2 and Orc1 to telomeric and subtelomeric bring Orc1 from the telomere into the TARE6 region by a fold- domains in P. falciparum, we performed mobility-shift and back structure. Telomeres are known to form fold-back structures supershift assays (EMSA). These assays were carried out using in yeast (de Bruin et al., 2000; de Bruin et al., 2001; Strahl-Bolsinger nuclear extracts and specific radioactively labelled probes directed et al., 1997; Zaman et al., 2002) and the same was proposed recently against telomere, TARE3 and TARE6 (Fig. 2A). In the mobility for P. falciparum (Figueiredo and Scherf, 2005). 2048 Journal of Cell Science 121 (12)

Fig. 2. Orc1 and Sir2 specifically recognize elements on the telomeric and subtelomeric repeats. (A) Schematic representation of telomere and subtelomeric regions of P. falciparum chromosome ends. The probes used in the EMSA and ChIP assays are indicated. (B) Binding of nuclear proteins to telomeric and subtelomeric regions. 32P-labelled telomere (left panel), TARE3 (middle panel) and TARE6 (right panel) probes were incubated with nuclear extracts. Competition shift assays were done using either 50-fold molar excess of unlabelled homologous (third lane) or heterologous (KAHRP and Sp1, fourth and fifth lanes) competitor. A free probe was run in the first lane in all cases. The telomere probe formed three protein-DNA complexes, whereas TARE3 and TARE6 formed only a single retarded complex. Arrowheads indicate DNA-protein complexes. (C) Binding of Orc1 and Sir2 to telomeric and subtelomeric repeats. A total of 30 μg of antiserum against Sir2 and Orc1, and respective non- immune sera, were pre-incubated for 15 minutes in the binding reaction, followed by the addition of the labelled telomere, TARE3 and TARE6 probes. Telomere and TARE3 probes produced supershifted complexes with both anti-Sir2 and -Orc1 antibodies (left and middle panels), whereas the TARE6 probe only formed a supershifted complex in the presence of Sir2 (right panel). NE, nuclear extracts; C, complex; SC, supershifted complex; KAHRP, upstream region of kahrp gene; Sp1, consensus Sp1-binding-site factor; PI, pre-immune serum.

Sir2 and Orc1 relocate during the blood-stage cycle During the ~48-hour blood-stage cycle, the parasite matures through the ring, trophozoite and schizont stages, and undergoes multiple rounds of asynchronous nuclear division (Leete and Rubin, 1996). Because Orc1 was assumed to be involved in DNA replication (Li and Cox, 2003; Mehra et al., 2005), we sought to analyze whether Orc1 localizes through S-M phase (trophozoite and schizont stages). We collected parasites from a synchronized culture at three different time points: rings, trophozoites and schizonts (Fig. 4A). Double- labelling IF studies revealed that, as the parasite

Journal of Cell Science differentiates, there is an increase in Sir2 and Orc1 protein levels together with their relocalization to a non- nuclear region (Fig. 4B-D). Strikingly, the IF pattern changes from well-defined perinuclear spots in ring stages to rather (Inselburg and Banyal, 1984). To understand the redistribution diffuse small dots in trophozoite stages. Moreover the of Sir2 and Orc1 throughout the cycle, we cultivated the parasites colocalization of Orc1 and Sir2 (>80% in ring stage) decreases in medium containing aphidicolin for about 24 hours. This culture, as the parasite matures. composed of arrested late trophozoites (Fig. 4A), was collected DNA synthesis in P. falciparum can be reversibly blocked by and subjected to IF analysis. Orc1 localized inside the nucleus aphidicolin, an inhibitor of DNA-polymerase-α in eukaryotes in non-replicative trophozoites, which is consistent with its

Fig. 3. Orc1 and Sir2 distribution at telomeric and subtelomeric chromatin. (A,B) ChIP analysis of ring- stage parasites using antibodies against Orc1 (A) and Sir2 (B). Immunoprecipitated DNA was analyzed by dot-blots hybridized with probes specific to telomeric sequences, subtelomeric sequences (TARE1, TARE2, TARE2-3, TARE3 and TARE6) and HRP. A representative dot-blot is shown for each antibody (left panels). The right panels show a quantitative presentation of the data shown in the left panels. In all cases, ChIP values represent a percentage of the total input DNA after subtraction of the background signal value (i.e. the material immunoprecipitated by the pre- immune sera or IgG). Association of Orc1 with telomeric and subtelomeric repeats is similar to that of Sir2. Relocation of telomeric proteins 2049

putative replicative function. In addition, we observed a punctate parasites. Sir2 and Orc1 are thus partially released to the and diffuse pattern outside of the nucleus. Surprisingly, Sir2 cytoplasm in mature stages. displayed a similar IF staining pattern (Fig. 4E). We concluded In order to exclude possible IF fixation artefacts, we performed that reorganization of Orc1 and Sir2 occurred prior to DNA double-IF assays in which we used anti-Sir2 and -histone-H3 replication. To identify the subcellular localization of Sir2 and antibodies. Clearly, histone H3 remained associated with nuclear Orc1 in mature stages, an anti-Hsp70 antibody was used as a DNA, whereas Sir2 relocated to the cytoplasm during parasite cytoplasmic marker (Mattei et al., 1989). Sir2 (Fig. 4F) and Orc1 maturation (Fig. 4G). We also tested an alternative fixation (data not shown) colocalize with Hsp70 in aphidicolin-treated protocol described previously (Tonkin et al., 2004), and the same Journal of Cell Science

Fig. 4. Sir2 and Orc1 relocalize during the developmental cycle of P. falciparum. (A) DNA synthesis during the P. falciparum cell cycle. During the 48-hour blood- stage cycle, the parasite differentiates through ring [between 0- and 18-hours post-invasion (p.i)], trophozoite (18- to 36-hours p.i.) and schizont (36- to 48-hours p.i.) stages. DNA replication in P. falciparum takes place in the trophozoite stage; it peaks at 30 hours and can be blocked by adding aphidicolin to the culture medium. Nuclear division occurs by schizogony, leading to the production of 16-32 merozoites; these are released in the bloodstream and can initiate a new cycle by invasion of a new red blood cell. (B-D) IF analysis of Sir2 (green) and Orc1 (red) during the blood-stage cycle. (B) Ring stages display a punctate pattern at the nuclear periphery. (C) In the trophozoite stages, anti-Sir2 and -Orc1 antibodies reveal an apparent increase of both protein levels, and an additional punctate and diffuse pattern inside and outside of the nucleus. (D) The schizont stage, Sir2 and Orc1 seem to relocalize at the nuclear periphery. (E-G) Double-labelling IF using parasites cultivated with aphidicolin for ~24 hours. (E) Aphidicolin-treated parasites demonstrate that the redistribution of Sir2 (green) and Orc1 (red) occurs prior to initiation of DNA replication. (F) Labelling with an anti-mouse Hsp70 (red) revealed that Sir2 (green) is displaced into the cytoplasm of the parasite before S- phase. (G) Histone H3 (dimethyl K4, 2mK4H3; red) remained associated with the nucleus. Nuclei were detected by DAPI staining (blue) in all the figures. Scale bars: 1 μm. 2050 Journal of Cell Science 121 (12)

Orc1 and Sir2 redistribution pattern was observed (data not of virulence genes involved in antigenic variation and pathogenesis shown). (Ralph and Scherf, 2005). Although the finding of Sir2 has shed some light on epigenetic control of antigenic variation in this Telomeric clusters break up before DNA replication pathogen (Duraisingh et al., 2005; Freitas-Junior et al., 2005), it Chromosome ends of P. falciparum form four to seven telomeric also taught us that a single telomeric protein is not sufficient to clusters at the nuclear periphery (Freitas-Junior et al., 2000). In control repression of all subtelomeric var genes. In this study, we this previous work, only pre-replicative-stage parasites were report the Orc1 protein as another telomere-associated protein in analyzed. Considering the relocation of the telomere-binding P. falciparum. EMSA and ChIP studies clearly demonstrate that proteins to the cytoplasm, we studied whether the telomeric Orc1 can interact specifically with telomere and with various clusters were rearranged in the nucleus of the parasite through its subtelomeric repeats in a similar way as Sir2 (Figs 2 and 3), pointing maturation. To address this issue, we analyzed rings, trophozoites to Orc1 as a strong candidate molecule that contributes to telomeric and schizonts by FISH. Telomeres were visualized by using a silencing in Plasmodium. Remarkably, Orc1 appears to have fluorescein-labelled DNA probe corresponding to TARE6 (or retained the replication and silencing functions observed in other rep20). We observed an increase from 4.3±0.3 signals per nucleus eukaryotes (Chesnokov, 2007). (n=64) in rings (Fig. 5A) to 12.5±1.2 (n=59) in trophozoite stages In addition, our EMSA and ChIP results suggest that Plasmodium (Fig. 5B). Schizont stages displayed more than 30 FISH signals telomeres fold back, allowing the telomeric chromatin to interact per parasite. It was difficult to determine the exact number of FISH with the subtelomeric domains, as shown schematically in Fig. 6A. dots because of overlapping signals of densely packed 16 to 32 This fold-back structure might account for the stabilization of nuclei (Fig. 5C). We also assessed the number of telomeric clusters telomeric and subtelomeric chromatin at the nuclear periphery in in the non-replicative aphidicolin-arrested parasites. Unexpectedly, Plasmodium, as previously suggested in yeast (de Bruin et al., 2000; we scored 10.0±1.0 FISH signals (n=56), indicating that the Strahl-Bolsinger et al., 1997). telomeric clusters are partially disrupted before DNA replication Interestingly, the Orc1 N-terminal region has a leucine zipper (Fig. 5D). It is interesting to note that these nuclei show both weak motif and many charged amino acid residues potentially involved and strong signals, and also that some of the telomeres seem to be in DNA-protein and protein-protein interactions (Li and Cox, 2003; dislocated from the nuclear periphery. Mehra et al., 2005). In our model (Fig. 6A), we hypothesize that the putative N-terminal DNA-binding motif allows the direct Discussion binding of Orc1 to the telomeric DNA. By contrast, Sir2 association P. falciparum telomeres and their adjacent subtelomeric regions have to telomeric and subtelomeric chromatin probably involves other been described as being important elements in the nuclear molecules, because Sir2 has no evident DNA-binding motifs. positioning of chromosome ends and in the control of expression Although no interaction between Orc1 and Sir2 was reported in the Journal of Cell Science

Fig. 5. Visualization of the telomeric clusters during the blood-stage cycle. (A-D) FISH analysis of nuclei stained with DAPI (blue) and hybridized with TARE6 (green) probe to visualize the chromosome ends on rings (A), trophozoites (B), schizonts (C) and aphidicolin-treated parasites (D). Scale bars: 1 μm. (E) Quantification of TARE6 FISH signals in rings, trophozoites and aphidicolin- treated parasites. n refers to the chromosome number per nucleus; in trophozoite stage, n>1, depending on the number of nuclear divisions that occurred on each individual parasite. In the case of aphidicolin-treated parasites, all signals (weak and strong) were counted. Error bars are 95% confidence intervals (±1.96 s.e.m.). The standard deviations for rings, aphidicolin-treated parasites and trophozoites are 1.20, 3.93 and 4.71, respectively. Relocation of telomeric proteins 2051

Fig. 6. Model for dynamic telomeric heterochromatin assembly and relocation during blood-stage development. (A) Hypothetical model of the differential spreading of P. falciparum telomeric proteins. Orc1 (green ovals) might bind directly to telomere and to TARE1-3 repeats via its N-terminal DNA-binding domain. The interaction of Orc1 with other unknown TARE6-binding molecules (adaptor proteins) might lead to telomere bending. Sir2 (red circles) might be recruited via interaction with an unknown telomere-associated protein (black curve) or alternatively with Orc1. Sir2 might then deacetylate telomeric histone tails (Merrick and Duraisingh, 2007), promoting heterochromatin formation and spreading towards the coding region. (B) Model for the dynamics of telomere chromatin factors during the P. falciparum blood-stage cycle. During ring stage (left), parasite telomeres form clusters at the nuclear periphery and associate with Sir2 (red circles) and Orc1 (green circles) (i). This period (G phase) is followed by multiple rounds of DNA synthesis and nuclear mitosis (trophozoite stage, middle), which produce a multinucleate schizont (right). Our data show that Orc1, Sir2 and telomeric clusters disassemble prior to DNA replication (ii). Telomeric components assemble in the newly formed nuclei (iii) and will be maintained for the next cycle. We speculate that the relocation events are driven by specific post-translational Journal of Cell Science modifications. Alternatively, cytoplasmic Orc1 and Sir2 might correspond to newly synthesized proteins necessary to accommodate the demands of the rapid nuclear divisions occurring during S-M phase.

yeast two-hybrid screening assay (LaCount et al., 2005), Orc1 might not been determined so far. Sir2-mutant parasites apparently do not work as the initiator of heterochromatin assembly and be involved change the transcription pattern of plasmodial rRNA genes (L.M.- in Sir2 recruitment into the telomere tract. Apart from Sir2, the S. and A.S., unpublished). Regarding Orc1, to our knowledge, no other proteins of the Sir complex (Sir1, Sir3 and Sir4) are not present localization studies in other organisms have shown this protein to in this parasite. It has been suggested that the lack of the Sir complex be present in the nucleolus. It is striking, though, that three proteins might be compensated for by the presence of the heterochromatin localizing in the nucleolus and having functions elsewhere in the protein 1 (HP1) (Moazed, 2001). Strikingly, the orthologue to HP1 nucleus have already been observed: TERT (Figueiredo et al., 2005), has been recently identified in P. falciparum (K. Toledo and R.H.- Sir2 (Freitas-Junior et al., 2005) and Orc1 (this work). An attractive R., unpublished data). ORC-HP1 interactions have been described hypothesis is that the nucleolus might serve as a reservoir for in Drosophila (Pak et al., 1997), Xenopus (Pak et al., 1997) and telomere-associated proteins, as has been proposed for yeast Sir mammals (Auth et al., 2006; Prasanth et al., 2004) in centromeric proteins (Gotta et al., 1997). heterochromatin, implicating HP1 as another candidate protein that One of the most remarkable features of Orc1 and Sir2 is that interacts with Orc1 in Plasmodium telomeres. Taking into account they undergo dynamic relocations throughout the 48-hour blood- all this data, it is becoming evident that the molecular components stage cycle in P. falciparum. They localize to distinct perinuclear contributing to TPE in P. falciparum resemble more the silencing compartments after merozoite invasion (ring stage) all the way complexes in metazoans and fission yeast, rather the budding-yeast through early trophozoite stage (20-hours post-invasion), and complexes. reposition to the entire nucleus and parasite cytoplasm during the Unexpectedly, Orc1 colocalized with Sir2 not only in the late trophozoite stage, the part of the cycle in which schizogony telomeric clusters but also in the nucleolus (Fig. 1). In yeast, Sir2 takes place to generate between 16 and 32 nuclei (Fig. 4 and also localizes in the nucleolus (Gotta et al., 1997) and represses schematically shown in Fig. 6B). Nuclear mitosis in Plasmodium both recombination and polymerase-II-dependent transcription is asynchronous and occurs without cytoplasmic division (Leete within the ribosomal (r)DNA repeats (Gottlieb and Esposito, 1989; and Rubin, 1996). DNA-replication machinery might cause the Smith and Boeke, 1997). The function of Sir2 in the nucleolus has disruption of telomeric clusters, unfolding of telomeres, and 2052 Journal of Cell Science 121 (12)

consequent dispersion of Orc1 and Sir2. Arresting the parasites by Immunofluorescence microscopy Synchronized cultures of P. falciparum were washed in phosphate-buffered saline aphidicolin allowed us to take a picture of the nuclear organization (PBS), lysed in saponine (0.015%) and fixed in suspension with 4% paraformaldehyde of the parasite prior to S-M phase. This demonstrated, however, solution for 15 minutes. Parasites were then incubated with the primary antibodies that Orc1 and Sir2 relocations occur before DNA replication and diluted in 1% bovine serum albumin (BSA) at 37°C for 30 minutes. After washing, might be linked with the break-up of telomeric clusters (Fig. 5 and parasites were incubated at 37°C for 30 minutes with the secondary antibodies conjugated with fluorochromes. After final washes, parasites were deposited on schematically shown in Fig. 6B). Cell-cycle-dependent microscope slides and mounted in Vectashield anti-fading with 4-6-diamidino-2- relocalization and remodelling of silencing factors have also been phenylindole (DAPI). Images were captured using a Nikon Eclipse 80i optical reported for yeast (Laroche et al., 2000; Smith et al., 2003). microscope. Anti-Orc1 and anti-Sir2 antibodies were pre-absorbed with lysed non-infected red However, Rap1 and Sir proteins partially disperse from the telomeric blood cells before incubation with the fixed parasites. The final antibody dilutions clusters only in G2 phase and mitosis (Laroche et al., 2000). This were: rabbit anti-Orc1 1:50-100, rabbit anti-Sir2 1:100, rat anti-Sir2 1:50, rat anti- discrepancy between Plasmodium and yeast might be due to the Nop1 1:50, monoclonal mouse anti-Hsp70 1:1200, rabbit anti-H3 2mK4 1:200, Alexa- unusual nuclear mitosis that occurs in this parasite. Fluor-488-conjugated goat anti-rabbit highly cross-absorbed 1:500, Alexa-Fluor-568- conjugated goat anti-rabbit highly cross-absorbed 1:500, Alexa-Fluor-568-conjugated Differential post-translation modifications might explain goat anti-mouse highly cross-absorbed 1:500 and Fluorescein-conjugated goat anti- differential localizations, and might account for other functions of rat 1:500. Orc1 and Sir2 at different time points of the Plasmodium blood- Fluorescence in situ hybridization stage cycle. In fact, a recent study in our laboratory indicates that FISH was performed with the same lysed and fixed parasites used for the IF to allow the nuclear form of Sir2 is modified by SUMO (N. Issar, E. Roux, a close correlation between IF and FISH imaging. Briefly, the fixed parasites were D. Mattei and A.S., unpublished). The amino acid sequence of Orc1 deposited on microscope slides, permeabilized in 0.1% Triton X-100 and hybridized has several SUMO consensus motifs, implying that Orc1 is also with heat-denaturized TARE6 probe (see ChIP probes) at 80°C for 30 minutes and at 37°C overnight. After hybridization, slides were washed as described previously sumoylated. It is tempting to speculate that reversible sumoylation (Freitas-Junior et al., 2000). might cause the relocation of Sir2 and Orc1. Sumoylation has been For immunofluorescence combined with in situ hybridization (IF-FISH) parasites described as regulating subcellular localization (Heun, 2007); it has were prepared as described above for IF; incubation with secondary antibody, parasites were post-fixed in suspension with 4% paraformaldehyde solution for 15 minutes, also been shown that telomere-associated proteins are often deposited on microscope slides and followed for FISH, in the same conditions sumoylated (Potts and Yu, 2007; Xhemalce et al., 2007), pointing described above. to this as a feature conserved in evolution. Orc1 also contains phosphorylation sites in the N-terminus (Mehra et al., 2005), Nuclear extract preparation, gel-shift and super-shift assays Preparation of nuclear extracts, gel shift and super-gel shift were performed as indicating that phosphorylation state might also dictate localization described previously (Freitas-Junior et al., 2005; Ruvalcaba-Salazar et al., 2005). and activity of this protein. Telomere probe of 175 bp was amplified from a plasmid obtained in a previous work In summary, this work reveals dynamic changes that occur in (Figueiredo et al., 2000) using the primers listed in supplementary material Table S1. A 550-bp DNA fragment of TARE3 was amplified from pCR2.1TOPO-TARE3 (see the perinuclear chromatin organization and telomere architecture ChIP probes); it was then digested with EcoRI and AvaII enzymes to get a TARE3 during P. falciparum blood-stage differentiation and proliferation. probe of 180 bp. TARE6 probe contains three Rep20 units and was obtained by Our study of Orc1 reveals another dimension of plasmodial hybridization of a 79-bp oligonucleotide with the respective anti-sense chromosome ends, namely that, during blood-stage development, oligonucleotides (supplementary material Table S1). Kharp and Sp1 probes were obtained as described previously (Ruvalcaba-Salazar et al., 2005). telomeres share the same molecule with other DNA elements, such

Journal of Cell Science as the origin-of-replication sequences and possibly rDNA, within ChIP assay and dot-blots the nucleolus. Future work should address directly the role of Orc1 The ChIP assay was performed as described previously (Freitas-Junior et al., 2005; in the nucleolus, cytoplasm and TPE in P. falciparum. TPE is of Lopez-Rubio et al., 2007). Chromatin fragments were incubated overnight at 4°C with the following antibodies: 15 μl of rabbit anti-Sir2, 4 μl of rabbit anti-IgG, 10 particular interest owing to its central role in the regulation of μl of rabbit anti-Orc1 and the respective preimmune sera. The immunoprecipitated virulence-factor genes and antigenic variation in this parasite. DNA was radioactively labelled and hybridized with a Hybond N+ membrane dot- Because ORC genes are essential for cell survival (Chesnokov, blotted with 35 ng of telomeric, TARE1, TARE2, TARE2-3, TARE3 or TARE6 probes. TARE1, TARE 2-3, TARE3 and HRP were PCR amplified from genomic 2007) and because of the absence of an applicable inducible gene- DNA using the primers listed in supplementary material Table S1 and cloned in knockdown system in P. falciparum, only mutations touching pCR2.1-TOPO (Invitrogen). The telomeric probe was PCR amplified using the dispensable functions such as telomere silencing can be studied in telomere primers (supplementary material Table S1) from the plasmid previously obtained (Figueiredo et al., 2000). TARE2 (2 kb) and TARE6 (1.5 kb) probes were this parasite; namely, mutations in the N-terminal region, because amplified using the primers M13 from plasmids obtained in a previous work the function in DNA replication seems to reside exclusively in the (Figueiredo et al., 2000). C-terminus (Mehra et al., 2005). Further studies should clarify For total DNA samples (input), an aliquot of lysate used in the immunoprecipitation whether nuclear and cytoplasmic forms show different post- was processed along with the rest of the samples. Signals were quantified using the ImageQuant software. Calculation of the amount of immunoprecipitated DNA in each translational modifications and whether or not they are linked with ChIP was based on the relative signal to the corresponding total DNA signal. the potential multiple functions of Orc1. We would like to thank the members of the Scherf laboratory for Materials and Methods helpful discussions; Cosmin Saveanu for critically reading the Parasites manuscript; and PlasmoDB for the invaluable malaria-database support. P. falciparum parasites were cultivated according to standard conditions (Trager and Jensen, 1976). Aphidicolin (Sigma) was added to a synchronized ring-stage culture We are also grateful to Anabela Cordeiro da Silva and Faculdade de as described previously (Inselburg and Banyal, 1984). Farmácia da Universidade do Porto, Portugal for their support. This work was supported by the European Commission (BioMalPar, contract Antibodies no.: LSPH-CT-2004-503578). L.M.-S. has financial support from the A rabbit anti-Orc1 antibody was prepared by immunizing rabbits with a GST fusion GABBA PhD program and Fundação para a Ciência e Tecnologia, protein corresponding to 150 amino acids on the C-terminus of the Orc1 protein Portugal. A.P.R.-M. was funded by Institut Pasteur, France and (aa794-944). Rat sera against Sir2 and against Nop1 were obtained by immunizing CONACYT, Mexico. A.S. is supported by a grant from the ‘Fonds rats with two synthetic peptides coupled to KLH for each protein. The sequence of the peptides is described elsewhere (Figueiredo et al., 2005; Freitas-Junior et al., dédié: Combattre les Maladies Parasitaires’ Sanofi Aventis-Ministère 2005). Rabbit polyclonal antibody to histone H3 (dimethyl K4) was purchased from de la Recherche and an ANR grant (ANR-06-MIME-026-01). R.H.-R. Abcam. has been supported by a grant from CONACYT (45687/A-1), Mexico. Relocation of telomeric proteins 2053

References Laroche, T., Martin, S. G., Tsai-Pflugfelder, M. and Gasser, S. M. (2000). The dynamics Auth, T., Kunkel, E. and Grummt, F. (2006). Interaction between HP1alpha and replication of yeast telomeres and silencing proteins through the cell cycle. J. Struct. Biol. 129, proteins in mammalian cells. Exp. Cell Res. 312, 3349-3359. 159-174. Baur, J. A., Zou, Y., Shay, J. W. and Wright, W. E. (2001). Telomere position effect in Leete, T. H. and Rubin, H. (1996). Malaria and the cell cycle. Parasitol. Today 12, 442- human cells. Science 292, 2075-2077. 444. Bell, S. P., Kobayashi, R. and Stillman, B. (1993). Yeast origin recognition complex Li, J. L. and Cox, L. S. (2003). Characterisation of a sexual stage-specific gene encoding functions in transcription silencing and DNA replication. Science 262, 1844-1849. ORC1 homologue in the human malaria parasite Plasmodium falciparum. Parasitol. Int. Blackburn, E. (2006). A history of telomere biology. In Telomeres. 2nd edn (ed. T. de 52, 41-52. Lange, V. Lundblad and E. Blackburn), pp. 1-19. Cold Spring Harbor: Cold Spring Harbor Lopez-Rubio, J. J., Gontijo, A. M., Nunes, M. C., Issar, N., Hernandez Rivas, R. and Laboratory Press. Scherf, A. (2007). 5Ј flanking region of var genes nucleate histone modification patterns Chesnokov, I. N. (2007). Multiple functions of the origin recognition complex. Int. Rev. linked to phenotypic inheritance of virulence traits in malaria parasites. Mol. Microbiol. Cytol. 256, 69-109. 66, 1296-1305. Craig, A. and Scherf, A. (2001). Molecules on the surface of the Plasmodium falciparum Marty, A. J., Thompson, J. K., Duffy, M. F., Voss, T. S., Cowman, A. F. and Crabb, infected erythrocyte and their role in malaria pathogenesis and immune evasion. Mol. B. S. (2006). Evidence that Plasmodium falciparum chromosome end clusters are cross- Biochem. Parasitol. 115, 129-143. linked by protein and are the sites of both virulence gene silencing and activation. Mol. de Bruin, D., Kantrow, S. M., Liberatore, R. A. and Zakian, V. A. (2000). Telomere Microbiol. 62, 72-83. folding is required for the stable maintenance of telomere position effects in yeast. Mol. Mattei, D., Scherf, A., Bensaude, O. and da Silva, L. P. (1989). A heat shock-like protein Cell. Biol. 20, 7991-8000. from the human malaria parasite Plasmodium falciparum induces autoantibodies. Eur. de Bruin, D., Zaman, Z., Liberatore, R. A. and Ptashne, M. (2001). Telomere looping J. Immunol. 19, 1823-1828. permits gene activation by a downstream UAS in yeast. Nature 409, 109-113. Mehra, P., Biswas, A. K., Gupta, A., Gourinath, S., Chitnis, C. E. and Dhar, S. K. Duraisingh, M. T., Voss, T. S., Marty, A. J., Duffy, M. F., Good, R. T., Thompson, J. (2005). Expression and characterization of human malaria parasite Plasmodium K., Freitas-Junior, L. H., Scherf, A., Crabb, B. S. and Cowman, A. F. (2005). falciparum origin recognition complex subunit 1. Biochem. Biophys. Res. Commun. 337, Heterochromatin silencing and locus repositioning linked to regulation of virulence genes 955-966. in Plasmodium falciparum. Cell 121, 13-24. Merrick, C. J. and Duraisingh, M. T. (2007). Plasmodium falciparum Sir2: an unusual Figueiredo, L. and Scherf, A. (2005). Plasmodium telomeres and telomerase: the usual sirtuin with dual histone deacetylase and adp-ribosyltransferase activity. Eukaryot. Cell actors in an unusual scenario. Chromosome Res. 13, 517-524. 6, 2081-2091. Figueiredo, L. M., Pirrit, L. A. and Scherf, A. (2000). Genomic organisation and chromatin Moazed, D. (2001). Common themes in mechanisms of gene silencing. Mol. Cell 8, 489- structure of Plasmodium falciparum chromosome ends. Mol. Biochem. Parasitol. 106, 498. 169-174. Mondoux, M. A. and Zakian, V. A. (2006). Telomere position effect: silencing near the Figueiredo, L. M., Freitas-Junior, L. H., Bottius, E., Olivo-Marin, J. C. and Scherf, end. In Telomeres. 2nd edn (ed. T. de Lange, V. Lundblad and E. Blackburn), pp. 261- A. (2002). A central role for Plasmodium falciparum subtelomeric regions in spatial 317. Cold Spring Harbor: Cold Spring Harbor Laboratory Press. positioning and telomere length regulation. EMBO J. 21, 815-824. Pak, D. T., Pflumm, M., Chesnokov, I., Huang, D. W., Kellum, R., Marr, J., Figueiredo, L. M., Rocha, E. P., Mancio-Silva, L., Prevost, C., Hernandez-Verdun, D. Romanowski, P. and Botchan, M. R. (1997). Association of the origin recognition and Scherf, A. (2005). The unusually large Plasmodium telomerase reverse-transcriptase complex with heterochromatin and HP1 in higher eukaryotes. Cell 91, 311-323. localizes in a discrete compartment associated with the nucleolus. Nucleic Acids Res. Potts, P. R. and Yu, H. (2007). The SMC5/6 complex maintains telomere length in ALT 33, 1111-1122. cancer cells through SUMOylation of telomere-binding proteins. Nat. Struct. Mol. Biol. Freitas-Junior, L. H., Bottius, E., Pirrit, L. A., Deitsch, K. W., Scheidig, C., Guinet, 14, 581-590. F., Nehrbass, U., Wellems, T. E. and Scherf, A. (2000). Frequent ectopic recombination Prasanth, S. G., Prasanth, K. V., Siddiqui, K., Spector, D. L. and Stillman, B. (2004). of virulence factor genes in telomeric chromosome clusters of P. falciparum. Nature Human Orc2 localizes to centrosomes, centromeres and heterochromatin during 407, 1018-1022. chromosome inheritance. EMBO J. 23, 2651-2663. Freitas-Junior, L. H., Hernandez-Rivas, R., Ralph, S. A., Montiel-Condado, D., Pryde, F. E. and Louis, E. J. (1999). Limitations of silencing at native yeast telomeres. Ruvalcaba-Salazar, O. K., Rojas-Meza, A. P., Mancio-Silva, L., Leal-Silvestre, R. EMBO J. 18, 2538-2550. J., Gontijo, A. M., Shorte, S. et al. (2005). Telomeric heterochromatin propagation Ralph, S. A. and Scherf, A. (2005). The epigenetic control of antigenic variation in and histone acetylation control mutually exclusive expression of antigenic variation genes Plasmodium falciparum. Curr. Opin. Microbiol. 8, 434-440. in malaria parasites. Cell 121, 25-36. Ruvalcaba-Salazar, O. K., del Carmen Ramirez-Estudillo, M., Montiel-Condado, D., Gardner, M. J., Hall, N., Fung, E., White, O., Berriman, M., Hyman, R. W., Carlton, Recillas-Targa, F., Vargas, M. and Hernandez-Rivas, R. (2005). Recombinant and J. M., Pain, A., Nelson, K. E., Bowman, S. et al. (2002). Genome sequence of the native Plasmodium falciparum TATA-binding-protein binds to a specific TATA box

Journal of Cell Science human malaria parasite Plasmodium falciparum. Nature 419, 498-511. element in promoter regions. Mol. Biochem. Parasitol. 140, 183-196. Gotta, M., Strahl-Bolsinger, S., Renauld, H., Laroche, T., Kennedy, B. K., Grunstein, Scherf, A., Figueiredo, L. M. and Freitas-Junior, L. H. (2001). Plasmodium telomeres: M. and Gasser, S. M. (1997). Localization of Sir2p: the nucleolus as a compartment a pathogen’s perspective. Curr. Opin. Microbiol. 4, 409-414. for silent information regulators. EMBO J. 16, 3243-3255. Smith, C. D., Smith, D. L., DeRisi, J. L. and Blackburn, E. H. (2003). Telomeric protein Gottlieb, S. and Esposito, R. E. (1989). A new role for a yeast transcriptional silencer distributions and remodeling through the cell cycle in Saccharomyces cerevisiae. Mol. gene, SIR2, in regulation of recombination in ribosomal DNA. Cell 56, 771-776. Biol. Cell 14, 556-570. Gottschling, D. E., Aparicio, O. M., Billington, B. L. and Zakian, V. A. (1990). Position Smith, J. S. and Boeke, J. D. (1997). An unusual form of transcriptional silencing in yeast effect at S. cerevisiae telomeres: reversible repression of Pol II transcription. Cell 63, ribosomal DNA. Genes Dev. 11, 241-254. 751-762. Strahl-Bolsinger, S., Hecht, A., Luo, K. and Grunstein, M. (1997). SIR2 and SIR4 Gupta, A., Mehra, P., Nitharwal, R., Sharma, A., Biswas, A. K. and Dhar, S. K. (2006). interactions differ in core and extended telomeric heterochromatin in yeast. Genes Dev. Analogous expression pattern of Plasmodium falciparum replication initiation proteins 11, 83-93. PfMCM4 and PfORC1 during the asexual and sexual stages of intraerythrocytic Tonkin, C. J., van Dooren, G. G., Spurck, T. P., Struck, N. S., Good, R. T., Handman, developmental cycle. FEMS Microbiol. Lett. 261, 12-18. E., Cowman, A. F. and McFadden, G. I. (2004). Localization of organellar proteins Hediger, F. and Gasser, S. M. (2002). Nuclear organization and silencing: putting things in Plasmodium falciparum using a novel set of transfection vectors and a new in their place. Nat. Cell Biol. 4, E53-E55. immunofluorescence fixation method. Mol. Biochem. Parasitol. 137, 13-21. Heun, P. (2007). SUMOrganization of the nucleus. Curr. Opin. Cell Biol. 19, 350-355. Trager, W. and Jensen, J. B. (1976). Human malaria parasites in continuous culture. Science Inselburg, J. and Banyal, H. S. (1984). Plasmodium falciparum: synchronization of asexual 193, 673-675. development with aphidicolin, a DNA synthesis inhibitor. Exp. Parasitol. 57, 48-54. Wyrick, J. J., Aparicio, J. G., Chen, T., Barnett, J. D., Jennings, E. G., Young, R. A., Kellis, M., Birren, B. W. and Lander, E. S. (2004). Proof and evolutionary analysis of Bell, S. P. and Aparicio, O. M. (2001). Genome-wide distribution of ORC and MCM ancient genome duplication in the yeast Saccharomyces cerevisiae. Nature 428, 617-624. proteins in S. cerevisiae: high-resolution mapping of replication origins. Science 294, Kyes, S., Horrocks, P. and Newbold, C. (2001). Antigenic variation at the infected red 2357-2360. cell surface in malaria. Annu. Rev. Microbiol. 55, 673-707. Xhemalce, B., Riising, E. M., Baumann, P., Dejean, A., Arcangioli, B. and Seeler, J. LaCount, D. J., Vignali, M., Chettier, R., Phansalkar, A., Bell, R., Hesselberth, J. R., S. (2007). Role of SUMO in the dynamics of telomere maintenance in fission yeast. Schoenfeld, L. W., Ota, I., Sahasrabudhe, S., Kurschner, C. et al. (2005). A protein Proc. Natl. Acad. Sci. USA 104, 893-898. interaction network of the malaria parasite Plasmodium falciparum. Nature 438, 103- Zaman, Z., Heid, C. and Ptashne, M. (2002). Telomere looping permits repression “at a 107. distance” in yeast. Curr. Biol. 12, 930-933. RESULTS III

Plasmodium falciparum heterochromatin protein 1 binds to trimethylated histone H3 lysine 9 and is a new component of the subtelomeric chromatin

Karla Perez-Toledo¤1, Ana Paola Rojas-Meza¤1,2, Liliana Mancio-Silva2, Nora Adriana Hernandez-Cuevas1, Dulce Maria Delgadillo1, Miguel Vargas1, Santiago Martinez-Calvillo3, Artur Scherf2 and Rosaura Hernandez-Rivas1*

1 Departamento de Biomedicina Molecular, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional (IPN), Apartado postal 14-740, 07360 México, D. F., México 2 Unité de Biologie des Interactions Hôte-Parasite, CNRS URA2581, Institut Pasteur, 25 Rue du Dr Roux, 75724 Paris, France 3 Unidad de Biomedicina, Facultad de Estudios Superiores Iztacala, Universidad Nacional Autonoma de México. Av. de los Barrios 1, Col. Los Reyes Iztacala, Tlalnepantla, Edo. de México, CP 54090, México

¤These authors contributed equally to this work * Corresponding author

Submitted.

49 50 Summary Increasing experimental evidence shows the prominent role of histone modifications in the coordinated control of gene expression in the human malaria parasite Plasmodium falciparum. The search for histone mark reading machinery that translates histone modifications into biological processes, such as formation of heterochromatin and antigenic variation is of foremost importance. In this work, we identified the first member of a histone methyl mark specific recognition protein, an ortholog of heterochromatin protein 1 (PfHP1). Analysis of the PfHP1 amino acid sequence revealed the presence of the two characteristic HP1 domains: a chromo domain (CD) and a chromo shadow domain (CSD). Recombinant CD binds to di- and tri-methylated lysine 9 from histone H3, but not to unmodified or methylated histone H3 in lysine 4. The CSD domain is able to interact with itself to form dimers underlying its potential role in aggregating nucleosomes to form heterochromatin. Antibodies raised against PfHP1 detect this molecule in foci at the perinuclear region. ChIP assays demonstrated that PfHP1 associates with subtelomeric non-coding repeat regions but is not detectable at telomere repeats. Our data strongly suggest that PfHP1 works in concert with the histone deacetylase PfSir2 to form silent chromatin regions.

Introduction Plasmodium falciparum, the causal protozoan agent of the most severe form of human malaria, has a complex life cycle with two different hosts, the Anopheles mosquito and humans. In order to complete its life cycle, P. falciparum invades different types of cells and self-propagates in very distinct environments in the mosquito (gut, hemolymph and salivary glands) as well as in the human host (skin, liver and erythrocytes). Each of these distinct environments exerts selective pressure related to morphological changes that force P. falciparum to exhibit differential gene expression during its life cycle (Bozdech et al., 2003, Le Roch et al., 2003). A surprising finding was the identification of a relatively low number of genes encoding transcription factors in the parasite, although the basal core transcriptional machinery and the protein-coding genes involved in nucleosome assembly and regulation of chromatin structure were found to be conserved (Aravind et al., 2003, Templeton et al., 2004, Callebaut et al., 2005). To date, no specific DNA binding proteins have been identified other than PfMyb1 (Gissot et al., 2005) and the ApiAp2 gene family (De Silva et al., 2008). This suggested that epigenetic mechanisms play a significant role in controlling gene expression in Plasmodium (Aravind et al., 2003, Hakimi et al., 2007, Ralph et al., 2005, Freitas-Junior et al., 2005). The importance of reversible chromatin modifications involving histone modifications and the chromatin-associated protein PfSir2 was first demonstrated for a

51 subtelomeric gene family (var gene family) involved in the control of antigenic variation in P. falciparum (Freitas-Junior et al., 2005). Knock out of PfSir2 resulted in the de-repression of a fraction of the members of the var gene family (Duraisingh et al., 2005). A screen for histone modifications revealed that H3K9-me3 plays a particular role in var gene repression (Lopez- Rubio et al., 2007, Chookajorn et al., 2007) and H3K4-me2/3 in transcriptional activity and epigenetic memory (Lopez-Rubio et al., 2007). These dynamic histone modifications occur mainly in the 5’UTR regions of var genes. It was shown that histone modifications do also play a major role in the regulation of non-var genes (Cui et al., 2007, Cui et al., 2008, Lopez- Rubio et al., 2007, Miao et al., 2006). The identification of factors able to ‘read’ specific histone marks and translate them into changes in gene activity remains elusive in Plasmodium. In this work, we identified the first major factor that displays characteristic features of the histone reading machinery that converts particular histone methyl modifications into heterochromatin. The identified protein is an ortholog to heterochromatin protein 1 (PfHP1) and shows a specific location to telomeric clusters in the nuclear periphery. We propose that PfHP1 works in concert with histone deacetylase such as PfSir2 in the formation and maintenance of subtelomeric heterochromatin in P. falciparum.

Results Identification and modelling of the HP1 ortholog in P. falciparum The HP1 protein is highly preserved through evolution, from S. pombe to human, and participates in the formation of highly condensed chromatin by recognizing the di- or tri- methylated H3 histone lysine 9 (H3K9-me2 or H3K9-me3) (Hiragami et al., 2005). In order to determine the existence of a protein similar to HP1 and its possible participation in heterochromatin formation in P. falciparum, we performed a search in the P. falciparum databank (www.plasmodb.cis.upenn.edu). This analysis allowed us to identify the PFL1005c gene, which encodes a 30 kDa protein. A detailed analysis of its amino acid sequence with the SMART program revealed the presence of the two characteristic HP1 domains defined to date: a chromo domain (CD) located between amino acids 6 to 60, and a low-homology chromo shadow domain (CSD) that spans from amino acid 197 to 267. The P. falciparum CD's highest identity (66%) is with the mouse CD. While the CSD's highest identity is with S. pombe CSD (37%) (Figure 1A). Similar to what has been reported in other eukaryotes, the PfHP1 CD has three of the most conserved aromatic cage residues (F9, W30 and Y33); in mice, these residues recognize the H3K9-me3, and abrogate silencing when they are mutated (Jacobs et al., 2002). Sequence alignment analysis allowed us to establish that the secondary structure of the PfHP1 CD is very similar to that described in mouse, since it presents the three ß chains, as well as

52 the α2 helix structures (Figure 1A). These findings suggest that the PfHP1 protein might be able to recognize the H3K9-me2 or H3K9-me3, similarly to the HP1 protein in higher eukaryotes. With respect to the CSD, PfHP1 consists of three beta chains and two α helixes (α1 and α2), as described in S. pombe and other eukaryotes, and also presents some of the residues reported to be important for dimerization because HP1 must form a dimer to participate in heterochromatin formation (Figure 1A).

Figure 1. Structural modelling of the chromo domain (CD) and chromoshadowdomain (CSD) of PfHP1. A) Schematic diagram showing the complete ORF and location of CD and CSD of PfHP1 separated by the hinge. Numbers correspond to residue positions for each domain. The primary sequence alignment of CD from mouse HP1ß (19-73 a.a.) and PfHP1 (6-60 a.a.) was used for the modelling of the CD domain. The boxes indicate the aromatic cage residues potentially involved in the recognition of the H3K9-me peptide. Conserved residues that form a complementary surface responsible for H3 peptide recognition in the CD from Drosophila are indicated in red (Jacobs et al.,2002). The primary sequence alignment of CSD from S. pombe SpSwi6 (261-320 a.a.) and from PfHP1 (197-267 a.a.) was used to perform the model of the CSD domains. Residues implicated in the formation of the dimer interface are enclosed in boxes. Residues that have shown to be important for dimer formation in S. pombe (Brasher et al., 2000) are indicated with pink letters. Both alignments were performed with ClustalW and used for modelling with the program MOE (www.chemcomp.com). The secondary structure elements are shown above the alignment: bars represent α-helices (α-1 and α-2) and arrows represent ß-strands (ß1 to ß3). B) Tertiary structure of CD and CSD domains of PfHP1. The structure is conserved between mouse and P. falciparum CD domains, and between S. pombe and P. falciparum CSD domains. C) Cartoon representation of the CSD dimer. The two monomers are indicated in red. The molecular surface of the putative PfHP1-CSD dimer is shown. C1) One putative monomer is shown by cartoon and the other one by molecular surface. Surface convex zones are shown in red (protuberances); concave zones that correspond to binding or recognition sites are shown in green (hydrophobic regions) and the polar surface is denoted in blue. C2) The hydrophobic region and the crucial L residues that contact the other subunit in the formation of the dimer are indicated. D) Superposition of the three- dimensional models of the CSD domains homodimers from SpSwi6 (magenta-yellow) and PfHP1 homodimers (green-red). As observed, the CSD architecture is strongly conserved between S. pombe and P. falciparum.

53 In order to determine whether the CD and CSD domains of the PfHP1 protein have a tertiary structure similar to that described in S. pombe and other eukaryotes, we made homology models, taking as a reference the mouse HP1 CD domain and Swi6 protein CSD of S. pombe. The tertiary structure obtained showed that the PfHP1 CD and CSD domains have a folding similar to that of the mouse CD and S. pombe CSD (Figure 1B). Dimer formation in S. pombe is centred on helix 2, which interacts symmetrically with the helix 2 of the other monomer, and the primary contacts involve amino acids L161, Y164 and L168. In PfHP1, L161 and L168 are present, but Y164 has been replaced by a L. Nevertheless, in mouse Y164 has also been changed and that does not affect dimer formation (Brasher et al., 2000). Therefore, amino acids L161, L164 and L168 could play a role in PfHP1 dimer formation (Fig. 1C-C2). In the case of S. pombe, it has been found that the bonding of the two monomers through the alpha 2 helix generates a hydrophobic groove that favours protein- protein interactions (Brasher et al., 2000). As seen in figure 1C, the PfHP1 CSD folding generates a hydrophobic groove through which the PfHP1 monomers can interact with themselves. Finally, the overlap of the S. pombe CSD homodimers (magenta-yellow) and CSD of PfHP1 (red-green) shows a similar structural organization (Figure 1D). In conclusion, the predicted tertiary structure suggests that PfHP1 could interact with the trimethylated H3 histone lysine 9, as well as participate in the formation of homodimers.

Figure 2. Nuclear expression of PfHP1. Antibodies against GST-PfHP1 were used to probe a Western blot on nuclear (NE) and cytoplasmic (Cyt) extracts from P. falciparum. Preimmune serum (PI) did not react with nuclear extracts. Anti-acetyl histone 4 antibodies were used to probe a Western blot on NE.

54 PfHP1 is enriched in the nuclear periphery at the telomere foci To investigate the functions of PfHP1, we cloned the entire PfHP1 gene (801 bp) in the pGEX4T1 expression vector. The GST-PfHP1 recombinant protein identified with an anti- GST antibody showed the expected molecular weight of about 55 kDa (data not shown). The soluble GST-PfHP1 protein was then purified and used to obtain polyclonal antibodies. Immunoblot analysis with this anti-PfHP1 antibody detected the 30 kDa PfHP1 protein in nuclear extracts. We observed a similar sized band in the cytoplasmic fraction, however, only after longer exposition. The specificity of the nuclear extracts was verified using an anti-acetyl histone H4 antibody, which recognizes both histones H2B and H4 (Figure 2). No signal was detected in the cytoplasmic fraction (data not shown). Preimmune serum gave no signal with the parasite’s nuclear extracts. These data show that PfHP1 is a nuclear protein in P. falciparum. To determine the cellular localization of PfHP1 we performed immunofluorescence assays (IFA), which demonstrated that this protein localizes predominantly in the periphery of the nucleus. Next, to study whether PfHP1 is associated to telomeric regions, attempts to do immunofluorescence combined with fluorescent in situ hybridization for telomeric repeats failed, therefore we performeda dual-labelling IFA using antibodies against PfSir2 protein, a marker for telomeres and nucleolus in P. falciparum (Freitas-Junior et al., 2005). PfHP1 protein forms foci that colocalize with most of PfSir2 signals. However, some PfHP1 IFA signals were not associate with PfSir2 (Figure 3A). In contrast to PfSir2 and PfOrc1, PfHP1 did not or only marginally colocalize with a nucleolar marker Nop1 (Figure 3B).

Figure 3. Cellular localization of PfHP1. A) Antibodies directed against PfHP1 (red) and PfSir2 (green) stain subnuclear compartments. In ring stage parasites, PfHP1 and PfSir2 display a punctuated pattern at nuclear periphery. There is a partial co-localization between both antibodies (yellow spots). B) Co-localization with a nucleolar marker PfNop1 shows minimal overlap of PfHP1 with this compartment. Scale bars represent 1 µm.

55

PfHP1 binds to subtelomeric TAREs The IF assays suggested that the PfHP1 protein is located on the telomeric and/or subtelomeric regions of P. falciparum. To determine the in vivo distribution of PfHP1 over the chromosome ends of the parasite, we performed ChIP assays. DNA pulled down with anti-PfHP1 was hybridized to different repetitive TARE elements and telomere probes in a dot blot assay (Figure 4A). The ChIP analysis showed that the subtelomeric region is enriched in PfHP1 (Figure 4B). In contrast, PfHP1 is almost absent from telomeric repeats (Figure 4B). As expected, two single copy blood-stage expressed genes (HRP1 and GBP) did not immunoprecipate with this antibody. Thus, our ChIP results show that PfHP1 is associated to chromatin across subtelomeric repeat elements, most likely mediating chromatin compaction. We assume that PfHP1 contributes directly to the previously demonstrated genetic and structural properties of TAREs such as gene silencing (Duraisingh et al., 2005) and the high chromatin condensation at chromosome ends of P. falciparum (Freitas-Junior et al., 2005).

Figure 4. PfHP1 participates in the formation of subtelomeric heterochromatin of P. falciparum. A) Schematic representation of subtelomeric region of P. falciparum chromosome. ChIP assays using anti-PfHP1 antibody were carried out on parasites at ring stage. The dot blot based ChIP analysis of PfHP1 abundance at telomeric and subtelomeric region is shown. The immunoprecipitated DNA was radioactively labeled and hybridized with membranes dot-blotted with telomeric and subtelomeric sequences TARE 1,2,3,6 and a sequence between TARE 2-3. 5´UTR regions of GBP-130 and HRP genes were used as control. B) Quantification of PfHP1 levels at telomeres and subtelomeres, relative to the histidine rich protein 1 (HRP1). Identical results were obtained using glycophorin binding protein (GBP) to normalize the data. Results are the average of two independent experiments and the standard error of the mean is indicated on each bar.

56 In vitro binding of the PfHP1 chromodomain to H3K9me2 and H3K9me3 peptides The CD in eukaryotes is responsible for recognizing H3K9-me2 or H3K9-me3 (Lachner et al., 2001, Bannister et al., 2001). To determine whether the PfHP1 protein is able to perform the same function in malaria parasites, soluble GST fusion protein corresponding to the CD of the PfHP1 protein was purified (Figure 5A) and immunoblotted with anti-GST antibodies. A band of the expected size (33 kDa) was observed together with minor bands which are probably degradation products as often seen for GST fusion proteins. Then, the GST-PfHP1- CD protein was used to investigate its binding specificity to different histone methyl marks. The GST-HP1-CD protein was dotted on nylon membranes and subsequently incubated with biotinylated histone H3 peptides, which were H3K9-me3, H3K9-me2, H3K9-ac, H3K4-me3, H3K4-me2 and unmodified histone H3. Binding of the peptides to the CD of PfHP1 was detected using streptavidin conjugated to HRP. Signal was observed only with H3K9-me2 and H3K9-me3 (Figure 5B). To show that these modified histones did not bind the recombinant GST-chromo protein through GST, the GST protein was used as a negative control. These results demonstrated that the CD from the PfHP1 protein is able to interact to H3K9-me2 or H3K9-me3, but not to H3K9-ac, H3K4-me2, H3K4-me3 or to unmodified histone H3 (Figure 5B).

Figure 5. PfHP1 binds to a histone H3 peptide methylated at lysine 9. A) Total proteins from DH5α strain expressing the CD of PfHP1 before (NI) and after IPTG induction (I), as well as the purified GST-PfHP1-CD fusion protein (P), were separated by SDS-PAGE and stained with Coomassie blue. An anti-GST antibody was used to detect the presence of the fusion protein. B) Purified protein GST-PfHP1-CD was assayed for binding to histone H3 peptides. Two hundred nanograms of the GST-chromo domain protein (GST-PfHP1-CD) were dotted to a nitrocellulose membrane, which was incubated with two micrograms of biotinylated peptides that were: H3K9- me3, H3K9-me2, H3K9-ac, H3K4-me3, H3K4-me2 and unmodified H3. C) 200 ng of GST-PfHP1-CD fusion protein were dotted to nitrocellulose and incubated with differents amounts of biotinylated peptides. Quantification of the GST-PfHP1-CD affinity to different biotinylated peptides. The average of two independent experiments and the standard error is indicated on each bar.

57 To determine the affinity of the GST-PfHP1-CD protein for both H3K9-me3 and H3K9-me2 peptides, dot blot binding-assays were performed using a fixed amount of GST-PfHP1-CD fusion protein with different concentrations of peptides (Figure 3C). Densitometry analysis showed that GST-PfHP1-CD binds 3 times more efficiently to H3K9-me3 than to H3K9-me2 (Figure 5C) in vitro.

Figure 6. In vitro dimerization of the PfHP1 protein. A).The His-PfHP1 protein was purified using nickel- nitriloacetic acid resin (TALON). Western blot analysis of this sample with anti-histidine antibodies identified the monomer of PfHP1. B) Interaction between His-PfHP1 and GST-PfHP1 fusion proteins. Purified refolded His- PfHP1 was mixed with total bacterial lysate that over-express the GST-PfHP1 fusion protein (left panels) or only GST (right panels). Protein complexes were recovered by TALON, separated by SDS-PAGE and detected by Western blot with anti-histidine and anti-GST antibodies.

Recombinant PfHP1 forms homodimers Cowieson et al. demonstrated that recombinant CSD in its unmodified form is able to dimerize in solution through the alpha helixes, forming a non-polar cavity where proteins that have the PxVxL pentapeptide can bind (Cowieson et al., 2000). As indicated above, both the alignment of the PfHP1 CSD domain and its tertiary structure strongly suggest that the PfHP1 protein is able to form homodimers (Figure 1). To demonstrate it, we generated the complete PfHP1 protein fused to six histidines (His-PfHP1). Inclusion bodies containing His-PfHP1 were incubated in the presence of urea, renatured and incubated with Nickel-beads. The beads were then, extensively washed and finally the bound protein was eluted and analysed in a SDS-PAGE gel, blotted and tested with the anti-histidine antibody. As expected, this antibody recognized specifically a highly enriched 31 kDa protein, which corresponds to the monomeric form of PfHP1 (Figure 6A).

58 To further analyze the ability of PfHP1 to interact with itself, we incubated lysates from bacterial cells over-expressing the GST-PfHP1 fusion protein (55 kDa) with the renatured His- PfHP1 fusion protein bound to the nickel beads. Bound proteins were eluted, separated by SDS-PAGE and immunoblotted with anti-histidine and anti-GST antibodies (Figure 6B). The Western blot analysis confirmed that His-PfHP1 interacts specifically to GST-PfHP1, because GST alone (26 kDa) did not bind to His-PfHP1 under the same conditions (Figure 6B). In summary, our in vitro experiments suggest the ability of PfHP1 to form dimers, which apparently is important for HP1-mediated heterochromatin formation (for model see Figure 7).

Discussion In this work, we identified and characterized PfHP1, the first protein from malaria that binds specifically to a histone mark, thus demonstrating the existence of a histone mark reading machinery in this pathogen. PfHP1 specifically recognizes H3K9-me2 and H3K9-me3 peptides by its chromodomain. The recombinant PfHP1 show protein-protein interaction, which strongly suggests that PfHP1 is involved in the formation of compact chromatin at chromosome ends in P. falciparum (Freitas-Junior et al., 2005, Duraisingh et al., 2005). The affinity of this protein seems to be greater with trimethylated than with dimethylated peptide (Figure 2C), suggesting that the degree of methylation at K9 may lead to the formation of more or less tight chromatin. Nuclear staining with anti-PfHP1 suggested that this protein is located at the perinuclear region visible as foci similar to PfSir2 and PfOrc1, two proteins known to be components of subtelomeric heterochromatin in P. falciparum (Mancio-Silva et al., 2008, Freitas-Junior et al., 2005). Our co-localization experiments indicated that, in contrast to PfSir2 and PfOrc1, that PfHP1 does not stain, or only marginally, the parasite nucleolus and may therefore not participate in the silencing of rDNA genes in blood stage parasites. However, the perinuclear signal of PfHP1 shows frequently signals not linked to telomeres, indicating that other chromosome regions may be associated to PfHP1. In fact, genome wide analysis of histone marks H3K9me3 in P. falciparum links silent subtelomeric and chromosome central virulence genes (Lopez-Rubio et al., submitted). This raises the possibility that PfHP1 may bind to subtelomeric and chromosome central virulence gene families. The notion that PfHP1 is a component of telomeric heterochromatin was further demonstrated by showing the binding of PfHP1 to subtelomeric chromatin (Figure 4). ChIP studies clearly showed that the anti-PfHP1 antibody immunoprecipitated mainly the subtelomeric regions, and reacted poorly with telomeric repeats. As in yeast, the telomeres of P. falciparum are mostly not packed within nucleosomes; whereas the subtelomeric chromatin contains

59 nucleosomes (Figueiredo et al., 2000). The association of PfHP1 with the subtelomeric chromatin correlates well with the nucleosomal nature of HP1 heterochromatinization function. Thus, our results suggests that in P. falciparum HP1, similarly to its orthologs in S. pombe, Drosophila and human, participates in the formation and maintenance of the heterochromatin (Lomberk et al., 2006). In mammals, the HP1 family is composed of three variants (alpha, beta and gamma), which are coded by distinct genes (Jones et al., 2001) and each of these proteins is involved in different events, HP1 alpha and beta in heterochromatin formation and HP1 gamma in euchromatin formation (Lomberk et al., 2006). As in S. pombe, in P. falciparum we have found only a single HP1 gene in the P. falciparum gene database. This point raises the question of whether a single PfHP1 protein performs different functions. Given the presence of H3K9-me3 present in the 5’flanking region of the var genes adjacent to subtelomeric TARES, it is tempting to speculate that PfHP1 may play a role in mediating more compact chromatin structure in plasmodial genes. Genome wide analysis by using ChIP-on-chip may reveal if PfHP1 spreads into subtelomeric coding regions and contribute to virulence factor gene control and centromere formation. The hinge region of the HP1 protein is highly amenable to post-translational modifications in other eukaryotes, especially phosphorylation (Lomberk et al., 2006). In addition, mutation in this region has been shown to affect the location, interactions and functions of HP1 (Zhao et al., 2001). For example, the specific phosphorylation of Ser83 of HP1 gamma allows the localization of this protein only in the euchromatin region. In silico analysis of the PfHP1 protein suggests that a number of post-translational changes such as: sumoylation, ubiquitination and phosphorylation may occur mainly in the hinge region (Perez-Toledo K and Hernandez-Rivas R unpublished data). Therefore, it is possible that different post-translational modifications may define distinct functions of PfHP1. Another component of the subtelomeric chromatin, PfSir2, has recently been shown to be sumoylated (Issar et al., 2008). Thus, it will be interesting to investigate whether mutating the PfHP1 hinge region will affect its function and nuclear location. In line with our data, we propose a model in which PfHP1 mediated heterochromatin formation depends on the presence of a histone deacetylase and a histone methyltransferase. In our model PfSir2, a histone deacetylase (Freitas-Junior et al., 2005, Merrick et al., 2007), is first recruited to telomeric chromatin and deacetylates H3K9 residues. In a subsequent step, we hypothesize that the H3K9 methyltransferase ortholog PfKMT1 (Cui et al., 2008) methylates the H3K9 residues. As consequence, PfHP1 binds to H3K9-me3, dimerizes and promotes chromatin condensation. This scenario is illustrated in figure 7A. What recruits PfHP1 to subtelomeres remains guesswork (the same is true for PfSir2 and PfORC1). In mammals targeting of HP1 to chromatin requires not only K9 methylation but also a direct

60 protein-protein interaction between SUV39 (an ortholog of PfKMT1) and HP1 (Blasco, 2007). HP1 can also interact with Orc1 as has been demonstrated in Drosophila and Xenopus (Pak et al., 1997, Huang et al., 1998). Future studies are necessary to determine potential protein- protein interaction between plasmodial PfHP1 and chromatin components by TAP-tag or pull down assays. The spreading of PfHP1 along the different TAREs is most likely controlled by complex machinery including known components such as PfSir2 and PfSet3. One of the key questions is whether PfHP1 does contribute to the reversible silencing of virulence factor gene families immediately adjacent of TAREs (Figure 7B). For example, silent subtelomeric var gene members are enriched for H3K9-me3 (Chookajorn et al., 2007, Lopez-Rubio et al., 2007). Thus, investigating the role of PfHP1 in facultative heterochromatin formation is of foremost interest in understanding the molecular mechanism of antigenic variation (Scherf et al., 2008).

Figure 7. Hypothetical model for heterochromatin formation at P. falciparum chromosome ends. A) PfHP1- mediated heterochromatin model depends on the action of the histone deacetylase PfSir2 and histone methyltransferase PfKMT1 (Freitas-Junior et al., 2005; Merrick and Duraising, 2007; Cui et al., 2008). B) General view of known chromatin components at P. falciparum subtelomeres. The spreading of PfHP1 along the different TAREs is orchestrated by PfSir2 and PfKMT1. The role of PfOrc1 in this process remains unknown.

61 In conclusion, this work identifies a key component of the epigenetic machinery involved in heterochromatin formation. We demonstrated that PfHP1 is closely related to its counterpart in higher eukaryotes, namely homodimer formation and recombinant CD binds methylated histone H3K9-me2 and -me3. PfHP1 is the first plasmodial protein identified that recognizes specific histone marks linked to silent chromatin. It appears that PfHP1 is an essential protein since we failed to produce a parasite line with a disruption in the PfHP1 gene. Together with the previously identified PfSir2, PfHP1 might form the core of the subtelomericheterochromatin machinery in P. falciparum. Our anti-PfHP1 antibodies are a new tool to investigate facultative and constitutive heterochromatin formation in malaria parasites. Telomere-linked control of phenotypic variation is an important issue in malaria pathogenesis and a better knowledge of the epigenetic factors controlling those is vital for understanding the disease and to develop new anti-parasite control strategies.

Material and Methods Parasites P. falciparum FCR3 strain was cultivated according to standard culture condition (Trager et al., 1976).

Recombinant proteins A 801-base pair (bp) DNA fragment of the PfHP1 gene was obtained by RT-PCR. The sequences of direct and reverse oligonucleotides were: PfHP1EcoRI 5’- CCGCTCGAGTTAAGCTGTACGGTATCTTAG-3´ and PfHP1Xho 5’- CGGAATTCATGACGGGTCAGATGAAGAA-3’, respectively. The PCR fragment was digested, retrieved and inserted into the EcoRI-XhoI-digested pGEX4T1 vector (Amersham) fused to a GST-tag. The resulting construct was sequenced and called GST-PfHP1. Primers PfHP1-CDEcoRI 5´-CGGAATTCATGACAGGGTCAGATGAAGAA-3’ and PfHP1- CDXhoI 5’-CCGCTCGAG CTCATTAGCTTTCGATAAAAAATT-3’ were used to amplify a DNA fragment of 204 bp that contains PfHP1 chromo domain (PfHP1-CD). Integrity of all recombinants clones was confirmed by sequencing with 5’ and 3’ primers to the pGEX cloning site (Pharmacia) and they were transformed into Escherichia coli DH5α strain. Expression of GST fusion protein was induced with 0.5 mM IPTG at 30ºC for 4 hr. GST fusion proteins were purified with glutathione sepharose (Pharmacia) by standard methods. The PfHP1 DNA fragment of 801 bp was cloned into BamH1 (5’- CCGCTCGAGTTAAGCTGTACGGTATCTTAG-3’) and NotI sites (5’- CGGAATTCATGACGGGTCAGATGAAGAA-3’) of the pROEXTM HTb vector

(Invitrogen). 6HisPfHP1 protein was induced in BL21 bacterial cells by addition of 0.6 mM

62 IPTG. Recombinant proteins were purified using nickel-nitriloacetic acid resin (TALON) according to manufacturer protocol. Purified recombinant proteins were verified by Western blot with 6His-Tag –specific antibodies.

Production of PfHP1 antibodies The GST-PfHP1 recombinant protein was purified as recommended by the manufacturers (see above). For polyclonal antibodies against PfHP1, two rabbits were inoculated i.c. with 100 µg of GST-PfH1 protein fusion, for the first dose, protein was emulsified with complete adjuvant (Sigma). Following the inoculation series, animals were sacrificed and serum was collected. Immunoglobulins were purified from serum by protein A-sepharose (Pharmacia) chromatography, following standard procedures.

Nuclear and cytoplasmic extracts preparation Nuclear and cytoplasmic extracts were done as described previoiusly (Freitas-Junior et al 2005 ; Ruvalcaba-Salazar et al., 2005) with some modifications. Briefly, 5X109 parasites of an asynchronous culture of FCR3 P. falciparum strain were isolated from infected erytrhrocytes by saponine lysis, resuspended in 1 ml of lysis buffer (10 mM Hepes, pH7.9, 10 mM KCl, 0.1 mM EDTA, 0.1 mM EGTA, 1 mM DTT, 0.65% NP-40) and incubated 30 min a 4ºC. Then parasites were lysed by 200 strokes on a prechilled douncer homogenizer. The nuclei were collected by centrifugation. The supernatant that contains the cytoplasmic fraction was recovered, aliquot and keept at -80ºC. The nuclei were resuspended in 1 ml of lysis buffer and placed in a cushion of 0.34 M sucrose. Nuclei were collected by centrifugation and nuclear protein was extracted using 100 µl of extraction buffer (20 mM HEPES, pH 7.9, 0.4 M NaCl, 1 mM EDTA, 1 mM EGTA and 1 mM DTT). Following 15 min of vigorous shaking on ice, the extract was centrifuged and the supernatant containing nuclear proteins was collected. All buffers used in this protocol contained a mixture of protease inhibitors (Complete).

Immunofluorescence Microscopy Immunofluorescence assays were performed as described previously (Mancio-Silva et al., 2008). For immunofluorescence the parasites were fixed in suspension with 4% paraformaldehyde solution for 15 minutes on ice. Next, the fixed parasites were incubated with the primary antibody for 60 min at room temperature followed by incubation for 30 min with a secondary antibody conjugated with fluorochrome and deposited on microscope slides. The final antibody dilutions were rabbit anti-PfHP1 1:500, rat anti-PfSir2 1:50, rat anti- PfNop1 1:50, Alexa Fluor 568 goat anti-rabbit highly cross-absorbed 1:500 and Fluorescein- conjugated goat anti-rat 1:500. The samples were analysed in a Nikon microscope.

63 Chromatin Immunoprecipitation and dot-blots The ChIP assay was performed as described previously (Freitas-Junior et al., 2005) with minor modifications. The parasites were treated with saponin and cross-linked in 1% formaldeyde for 10 min at 37°C. Chromatin fragments were incubated overnight at 4°C with 2.5 µl of rabbit polyclonal antibody anti-PfHP1 or preimmune serum. The immunoprecipitated DNA was radioactively labeled and hybridized with a Hybond N+ membrane dot-blotted with 35 ng of Telomeric, TARE1, TARE2, TARE2-3, TARE3 and TARE6 repeats as previously published (Mancio-Silva et al., 2008). For total DNA samples, an aliquot of lysate used in the immunoprecipitation was processed along with the rest of the samples at the crosslink reversal step. Quantification of the signal was done with the ImageQuant software (Molecular Dynamics, Sunnyvale). The amount of telomeric and subtelomeric DNA immunoprecipitated in each ChIP was calculated based on the signal relative to the corresponding total input signal. Additionally, relative enrichment was calculated by normalization of HRP or GBP signal.

In vitro binding assays The His-PfHP1 protein was denatured and renatured by dialysis against 8M urea, followed by 6 M , 4 M, 2 M, 1 M and 0.5 M urea and finally with nickel binding buffer with no urea. Then, 40 µg of purified His-PfHP1 fusion protein refolded was incubated with Nickel- nitriloacetic acid resin (TALON) (50 µl) according to manufacturer protocol. By this way the His-PfHP1 fusion protein was immobilized on talon resin, mixed with 2mg/ml total bacterial lysate that over-expressed GST-PfHP1 protein fusion and incubated for 2 hr at 4ºC with 1 ml binding buffer (20 mM Tri-HCl pH 8.0, 100 mM NaCl). After, the beads were washed three times with 1 ml washing buffer (20 mM Tri-HCl pH 8.0, 100 mM NaCl, 15 mM Imidazol pH 8.0). Bound proteins were eluted with sample buffer and heating before loading on a 10% SDS-PAGE gel. Then, electrophoresis, proteins were transferred to Hybond ECL nitrocellulose (Amersham) as described (Towbin et al., 1979) and the blot was probed with anti-GST and anti- 6His-Tag –specific antibodies.

Dot-blot Overlay Assays The dot-blot assays were performed as described by Sambrook et al., 1989 (Sambrook et al., 2001). Two hundred nanograms of the GST-chromo domain protein (GST-PfHP1-CD) were dotted to Hybond ECL nitrocellulose (Amersham) and blocked using 4% fat–free dried milk in TBS-Tween (10 mM Tris-HCl, 150 mM NaCl, 1% Tween, pH 7.5). Membranes were incubated overnight at 4ºC in BC100 buffer (25 mM HEPES, pH7.6; 100 mM NaCl, 1 mM

MgCl2, 0.5 mM EGTA, 0.1 mM EDTA, 10% glycerol, 1 mM DTT and 0.2 mM PMSF) with 2

64 µg or different amounts of biotynilated peptides H3K9-me3 (Upstate), H3K9-me2 (Washington Biotechnology), H3K3-ac (Upstate), H3K4-me3 (Upstate), H3K4-me2 (Washington Biotechnology) and non-modified H3 (Washington Biotechnology). The membranes were washed twice (10 min each) with BC-100 plus 0.05% NP40 and twice with

BC200 (25 mM HEPES, pH7.6; 200 mM NaCl, 1 mM MgCl2, 0.5 mM EGTA, 0.1 mM EDTA, 10% glycerol, 1 mM DTT and 0.2 mM PMSF) plus 0.05% NP-40. Membranes were incubated with streptavidin, coupled to HRP (1:10,000 dilutions in TBS plus 1% BSA) 30 min at 37ºC. Finally the membranes were washed seven times with TBS plus 1% Tween-100 and bound peptides were detected by chemiluminescence using an ECL kit. Quantification of the signal was done with the ImageQuant software (Molecular Dynamics, Sunnyvale).

Molecular modeling of PfHP1 protein Homology modeling of the PfCD and PfSCD complexes was performed with MOE package (molecular operating environment, http://www.chemcomp.com) using the crystallographic structure of CD from mouse HP1β (19-73) and CSD from S. pombe SpSwi6 (261-320) as template (PDB ID 1APO and 1E0B, respectively). Search for templates and sequence alignments were done with BLAST, with minor adjustments in the alignment based on template structure inspection to facilitate indel modeling. Sequence identity between CD from mouse and CD from P. falciparum and CSD from S. pombe and CSD from P. falciparum was 66% and 37%, respectively. For the search of the conformation of replaced side chains we used a library of rotamers included in MOE; for indel modeling, a loop library of this package was used. By random combination of these libraries we constructed 1000 alternative models and minimized them to reach a gradient of 0.1 kcal/ (mol A°). The model with the best packing was subjected to more extensive minimization, until an RMS gradient lower than 0.05 kcal/ (mol A°) was obtained, and used thereafter for the structural analysis. Minimizations used the CHARMM22 force field with its partial-charge assignments for protein atoms. The stereochemical quality of the model was verified with MOE and with the Mol-Probity server (http://kinemage.biochem.duke.edu/molprobity)

65 Acknowledgements We would like to thank J.J. Lopez-Rubio, L. Riviere and F. Iracheta Tovar for critical comments. This work was supported by a grant from CONACYT (45687/A-1), Mexico to RHR. A.P.R.M. was funded by Institut Pasteur, France EGIDE and CONACYT, Mexico. A.S. is supported by a grant from the European Commission (BioMalPar) (contract No: LSPH-CT- 2004-503578), the "Fonds dédié: Combattre les Maladies Parasitaires" Sanofi Aventis - Ministère de la Recherche and an ANR grant (ANR-06-MIME-026-01). L.M.S. has financial support from the GABBA PhD program and Fundação para a Ciência e Tecnologia, Portugal.

References Aravind, L., Iyer, L.M., Wellems, T.E. and Miller, L.H. (2003). Plasmodium biology: genomic gleanings. Cell 115, 771-785. Bannister, A.J., Zegerman, P., Partridge, J.F., Miska, E.A., Thomas, J.O., Allshire, R.C. and Kouzarides, T. (2001). Selective recognition of methylated lysine 9 on histone H3 by the HP1 chromo domain. Nature 410, 120-124. Blasco, M.A. (2007). The epigenetic regulation of mammalian telomeres. Nat Rev Genet 8, 299-309. Bozdech, Z., Llinas, M., Pulliam, B.L., Wong, E.D., Zhu, J. and DeRisi, J.L. (2003). The transcriptome of the intraerythrocytic developmental cycle of Plasmodium falciparum. PLoS Biol 1, E5. Brasher, S.V., Smith, B.O., Fogh, R.H., Nietlispach, D., Thiru, A., Nielsen, P.R., et al. (2000). The structure of mouse HP1 suggests a unique mode of single peptide recognition by the shadow chromo domain dimer. EMBO J 19, 1587-1597. Callebaut, I., Prat, K., Meurice, E., Mornon, J.P. and Tomavo, S. (2005). Prediction of the general transcription factors associated with RNA polymerase II in Plasmodium falciparum: conserved features and differences relative to other eukaryotes. BMC Genomics 6, 100. Cowieson, N.P., Partridge, J.F., Allshire, R.C. and McLaughlin, P.J. (2000). Dimerisation of a chromo shadow domain and distinctions from the chromodomain as revealed by structural analysis. Curr Biol 10, 517-525. Cui, L., Fan, Q. and Miao, J. (2008). Histone lysine methyltransferases and demethylases in Plasmodium falciparum. Int J Parasitol 38, 1083-1097. Cui, L., Miao, J., Furuya, T., Li, X. and Su, X.Z. (2007). PfGCN5-mediated histone H3 acetylation plays a key role in gene expression in Plasmodium falciparum. Eukaryot Cell 6, 1219-1227. Chookajorn, T., Dzikowski, R., Frank, M., Li, F., Jiwani, A.Z., Hartl, D.L. and Deitsch, K.W. (2007). Epigenetic memory at malaria virulence genes. Proc Natl Acad Sci U S A 104, 899-902. De Silva, E.K., Gehrke, A.R., Olszewski, K., Leon, I., Chahal, J.S., Bulyk, M.L. and Llinas, M. (2008). Specific DNA-binding by apicomplexan AP2 transcription factors. Proc Natl Acad Sci U S A 105, 8393-8398. Duraisingh, M.T., Voss, T.S., Marty, A.J., Duffy, M.F., Good, R.T., Thompson, J.K., et al. (2005). Heterochromatin silencing and locus repositioning linked to regulation of virulence genes in Plasmodium falciparum. Cell 121, 13-24. Figueiredo, L.M., Pirritt, L.A. and Scherf, A. (2000). Genomic organisation and chromatin structure of Plasmodium falciparum chromosome ends. Molecular and Biochemical Parasitology 106, 169-174. Freitas-Junior, L.H., Bottius, E., Pirrit, L.A., Deitsch, K.W., Scheidig, C., Guinet, F., et al. (2000). Frequent ectopic recombination of virulence factor genes in telomeric chromosome clusters of P. falciparum. Nature 407, 1018-1022. Freitas-Junior, L.H., Hernandez-Rivas, R., Ralph, S.A., Montiel-Condado, D., Ruvalcaba-Salazar, O.K., Rojas-Meza, A.P., et al. (2005). Telomeric Heterochromatin Propagation and Histone Acetylation Control Mutually Exclusive Expression of Antigenic Variation Genes in Malaria Parasites. Cell 121, 25-36. Gissot, M., Briquet, S., Refour, P., Boschet, C. and Vaquero, C. (2005). PfMyb1, a Plasmodium falciparum transcription factor, is required for intra-erythrocytic growth and controls key genes for cell cycle regulation. J Mol Biol 346, 29-42.

66 Hakimi, M.-A. and Deitsch, K.W. (2007). Epigenetics in Apicomplexa: control of gene expression during cell cycle progression, differentiation and antigenic variation. Current Opinion in Microbiology 10, 357-362. Hiragami, K. and Festenstein, R. (2005). Heterochromatin protein 1: a pervasive controlling influence. Cell Mol Life Sci 62, 2711-2726. Huang, D.W., Fanti, L., Pak, D.T., Botchan, M.R., Pimpinelli, S. and Kellum, R. (1998). Distinct cytoplasmic and nuclear fractions of Drosophila heterochromatin protein 1: their phosphorylation levels and associations with origin recognition complex proteins. J Cell Biol 142, 307-318. Issar, N., Roux, E., Mattei, D. and Scherf, A. (2008). Identification of a novel post-translational modification in Plasmodium falciparum: Protein SUMOylation in different cellular compartments. Cell Microbiol. Jacobs, S.A. and Khorasanizadeh, S. (2002). Structure of HP1 chromodomain bound to a lysine 9- methylated histone H3 tail. Science 295, 2080-2083. Jones, D.O., Mattei, M.G., Horsley, D., Cowell, I.G. and Singh, P.B. (2001). The gene and pseudogenes of Cbx3/mHP1 gamma. DNA Seq 12, 147-160. Lachner, M., O'Carroll, D., Rea, S., Mechtler, K. and Jenuwein, T. (2001). Methylation of histone H3 lysine 9 creates a binding site for HP1 proteins. Nature 410, 116-120. Le Roch, K.G., Zhou, Y., Blair, P.L., Grainger, M., Moch, J.K., Haynes, J.D., et al. (2003). Discovery of Gene Function by Expression Profiling of the Malaria Parasite Life Cycle. Science 301, 1503- 1508. Lomberk, G., Wallrath, L. and Urrutia, R. (2006). The Heterochromatin Protein 1 family. Genome Biology 7, 228. Lopez-Rubio, J.J., Gontijo, A.M., Nunes, M.C., Issar, N., Hernandez Rivas, R. and Scherf, A. (2007). 5' flanking region of var genes nucleate histone modification patterns linked to phenotypic inheritance of virulence traits in malaria parasites. Mol Microbiol 66, 1296-1305. Mancio-Silva, L., Rojas-Meza, A.P., Vargas, M., Scherf, A. and Hernandez-Rivas, R. (2008). Differential association of Orc1 and Sir2 proteins to telomeric domains in Plasmodium falciparum. J Cell Sci 121, 2046-2053. Merrick, C.J. and Duraisingh, M.T. (2007). Plasmodium falciparum Sir2: an unusual sirtuin with dual histone deacetylase and ADP-ribosyltransferase activity. Eukaryot Cell 6, 2081-2091. Miao, J., Fan, Q., Cui, L. and Li, J. (2006). The malaria parasite Plasmodium falciparum histones: organization, expression, and acetylation. Gene 369, 53-65. Pak, D.T., Pflumm, M., Chesnokov, I., Huang, D.W., Kellum, R., Marr, J., et al. (1997). Association of the origin recognition complex with heterochromatin and HP1 in higher eukaryotes. Cell 91, 311-323. Ralph, S.A. and Scherf, A. (2005). The epigenetic control of antigenic variation in Plasmodium falciparum. Current Opinion in Microbiology 8, 434-440. Sambrook, J. and Rusell, D.W. (2001) Molecular Cloning a Laboratory Manual Cold Spring Harbor, New York., Cold Spring Harbor Laboratory Press. Scherf, A., Riviere, L. and Lopez-Rubio, J.J. (2008). SnapShot: var gene expression in the malaria parasite. Cell 134, 190. Templeton, T.J., Iyer, L.M., Anantharaman, V., Enomoto, S., Abrahante, J.E., Subramanian, G.M., et al. (2004). Comparative analysis of apicomplexa and genomic diversity in eukaryotes. Genome Res 14, 1686-1695. Towbin, H., Staehelin, T. and Gordon, J. (1979). Electrophoretic transfer of proteins from polyacrylamide gels to nitrocellulose sheets: procedure and some applications. Proc Natl Acad Sci U S A 76, 4350-4354. Trager, W. and Jensen, J.B. (1976). Human malaria parasites in continuous culture. Science 193, 673- 675. Zhao, T., Heyduk, T. and Eissenberg, J.C. (2001). Phosphorylation site mutations in heterochromatin protein 1 (HP1) reduce or eliminate silencing activity. J Biol Chem 276, 9512-9518.

67

68 RESULTS IV

Differential sub-nuclear localization of repressive and active histone methyl modifications in P. falciparum

Neha Issar*, Stuart Ralph*, Liliana Mancio-Silva, Catherine Keeling and Artur Scherf

Institut Pasteur, Unité de Biologie des Interactions Hôte-Parasite, CNRS URA2581, 25 Rue du Dr Roux, 75724 Paris, France * These authors contributed equally to this work

Correspondence: Artur Scherf, tel: + 33 (01) 4568-8616, fax: + 33 (01) 4568-8348, e-mail: [email protected]

Submitted.

69 70 Summary Epigenetic factors play a crucial role in the co-ordinated expression of Plasmodium falciparum genes, especially those implicated in antigenic variation and pathogenesis. Post- translational histone modifications and sub-nuclear organization have the potential to epigenetically influence gene regulation. Here we screened for the presence of histone methylation modifications using specific antibodies and determined, for the first time, their spatial nuclear localisation. We observed that several marks were differentially enriched in sub-nuclear compartments, suggesting high local concentrations of particular methyl marks such as histone H3 lysine 79. This mark does not co-localize with already defined nuclear compartments, namely the perinuclear var gene expression site, telomere cluster or nucleolus. Thus, antibodies specific for H3K79me apparently define a nuclear compartment not observed previously in other eukaryotes. Our data show the presence of discrete sub nuclear zones enriched for histone methyl marks, pointing to the existence of specialized transcription factories and repressive regions in malaria parasites.

Introduction Post-translational modifications (PTMs) of histones present a remarkable array of chemical variations that can control gene expression in eukaryotes [1]. These provide an additional layer of regulation on top of the DNA code, called epigenetic regulation, with implications for transcriptional activation, silencing and cellular memory [2, 3]. Covalent histone modifications on the protruding N-terminal tails of the H3 and H4 subunits contribute to such epigenetic control. Several studies have found important roles for both histone acetylation [4] and methylation[5, 6] in the regulation of Plasmodium genes, indicating that they are involved in defining active and repressive transcriptional states as well as the inheritance of these states. This tight regulation applies, not only to global transcriptional states of P. falciparum genes, but characteristically to var genes; a 60 member antigenically-variant gene family responsible for the polymorphism of a major virulence factor: P. falciparum erythrocyte membrane protein 1 (PfEMP1). The mechanism of monoallelic expression of a single var gene from a repertoire of many is still not fully resolved. Investigations to decipher this led us to investigate the organization of P. falciparum nucleus, revealing distinct sub-nuclear domains, such as telomere clusters, heterochromatin-rich nuclear periphery and the perinuclear expression site of a var gene [7]. Mounting evidence hints towards existence of decided zones: non-permissive or `silenced zones’ and transcriptionally permissive ‘expression zones’. However, the malaria community still lacks markers to distinguish between these, especially markers recognising the active

71 expression sites. Having already observed connections between certain histone marks and active and repressive chromatin states in P. falciparum[4, 8], we wondered if histone modifications may also be markers that define the spatial compartmentalisation of active and silenced genes. To this end we conducted a survey of major histone methyl modifications in blood stage P. falciparum parasites, in anticipation that knowledge about nuclear distribution of known and novel histone modifications may provide insight into their function. Here we show that particular marks are highly enriched using immunofluorescence analysis and reveal distinct sub-nuclear areas that point to specialized transcription factories.

Results Sub-nuclear enrichment of particular histone methyl marks A distinct separation of purified P. falciparum histone proteins was obtained and examined for the presence of specific histone modifications by Western Blot (Fig. 1. A). Strong specific bands corresponding to the predicted molecular weight of histone subunit H3 and H4 were observed using commercial antibodies against H3-Core, H3-K4me2, H3-K4me3, H3- K79me3, H4-Core and H4-K20me3. Antibodies against H3-K4me1, H3-K9me1 and H3- K9me3 also recognised bands at the predicted motility of histone H3, although at a slightly weaker intensity. Antibodies against H3-K9me2 and H3-K27me3 did not recognise any parasite specific bands, even at higher concentrations, suggesting that these modifications are either very low in abundance or absent in P. falciparum. H3-K36me3 did not recognise any band either, which may be due to the non-utilisation of this mark or the divergence of the residues flanking the PfH3-36me3. However, in a recently published study, another antibody (different manufacturer) against this mark has been shown to react with PfH3 [11] We further examined the sub-nuclear enrichment of histone methyl marks. For our analysis, we focussed on the ring stage, as the major virulence factor genes, the var genes are transcribed at this time. To this end, IFA was performed on P. falciparum-infected ring stage erythrocytes to localise specific histone modifications (Fig. 1 B). Anti-H3-K4me1 labelling was moderate, and diffused equally throughout the nucleus (panel a). Strong nuclear labelling was observed for antibodies against H3-K4me2 and H3-K4me3K4, two marks generally associated with active loci. In both these cases, the labelling patterns from these antibodies changed slightly with the developmental stage. In ring stages (presented here), H3-K4me2 was distributed in a punctate pattern (panel b) but became more diffuse in the later stages (data not shown). In very early ring stages the H3-K4me3 labelling was less intense and often restricted to punctate localisation of 4-6 dots around the edge of nucleus, while in later ring stage cells, these dots coalesced into a strong uniform labelling that extended most of the way

72

Fig.1. A. Immunoblot analysis for histone methyl modifications. ~8 µg of partially purified P. falciparum histones from mid trophozoites (~30 hrs post invasion) were separated by 15% SDS-PAGE and stained with coomassie blue (right most lane). Methyl modifications present on histone H3/H4 were determined by immunoblotting with respective site-specific antibodies in equal dilutions. B. Immunofluorescence assays with histone methyl modification antibodies. IFAs were carried out with antibodies against specific histone modifications (panel a.-g) to reveal differential sub nuclear compartmentalisation of histone marks. DAPI stained nucleus in blue. Comparison of transcriptional states defined by specific histone marks in model systems [21] and P. falciparum [6] respectively. + denotes activating, - denotes repressive. Scale bar depicts 1 micron.

73 around the periphery of the nucleus (panel c), leaving one contiguous area largely devoid of labelling at the nuclear core and at one end of the nuclear periphery resulting in a horseshoe- shaped pattern. Neither the gap in this horseshoe pattern, nor the H3-K4me3-positive region corresponded with the previously described nucleolar marker PfNop1 [4] partially co- localising with both regions (data not shown). Antibodies directed against H3-K9me3, a repressive mark for var genes[6], apparently concentrate in 2-3 discrete foci polarising at the periphery of the nucleus (panel d), which is compatible with our previously observed peripheral heterochromatin–like region[4]. No detectable labelling was seen for H3-K9me1 or H3-K9me2, consistent with the lack of reactivity by western blotting. H3-K36 showed full nuclear staining (panel e), though it was undetected by western analysis. Two other modifications with a marked sub-nuclear distribution were H3-K79me3 and H4-K20me3, generally associated with activation and repression respectively. We observed a pronounced peripheral labelling with the repressive modification H4-K20me3, comparable to H3-K4me3 but slightly more concentrated at the outer extremity of the nuclear periphery (panel g). A recent publication also localised Plasmodium H3-K4me3 to foci at the periphery of the schizont nuclei [12]. The methylase mediating this mark has been characterised in Toxoplasma gondi, a related apicomplexan [12], although its reported nuclear localisation does not mirror the compartmentalisation of its apparent modified histone product. H3-K79me3, generally linked to gene activation, gave the most striking pattern of all marks analyzed. This modification forms 3 to 5 distinctive dots polarised to one end in the nucleus (panel f). From all of the methylations analysed by IFA 4 out of 7, namely H3-K4me2, H3-K4me3, H3- K4me 79 and H4-K20me3, indicated differential sub-nuclear enrichments, H3-K79me3 being the most intriguing. This is an interesting observation since some marks have recently been shown to be antagonistically related at a specific var gene loci, e.g. H3-K4me3 and H3- K9me3 that define activation and repression respectively [6].

H3-K79me3 defines a unique sub-nuclear compartment Our observation led us to question if any of these modifications are linked to the postulated P. falciparum transcription sites. We were particularly intrigued by the polarised area stained by H3-K79me3, a modification hereby reported in Plasmodium (Fig 1 B, panel f). The fact that H3-K79me3 is generally known to specifically mark active genes in other eukaryotes and that it showed a sub-compartmentalization within the P. falciparum nucleus, led us to question whether this specific histone modification marks the potential var ‘transcription-competent’ area. We aimed to test whether H3-K79me3 mark is linked to any of the previously

74

Fig.2. Co-localisation of H3-K79me3 with known cellular compartments. A. Nucleolus: Double labelling IFAs were carried out with antibodies against H3-K79me3 and nucleolus marker, PfNop1. B. Telomere clusters: IF- FISH analysis to investigate relative nuclear association between H3-K79me3 (red) and Rep20 (green). C. Active var gene: IF-FISH analysis between H3-K79me3 (red) and active var gene (green). Two representative examples show no association between the two. DAPI stained nucleus in blue. Font colour represents corresponding fluorescein labelled antibody. Scale bar depicts 1 micron D. Schematic of the dynamic compartmentalisation of the P. falciparum nucleus. The distinct sub-nuclear distribution of histone methylation modification, taking example of H3-K79me3 as observed in this study is presented as a diagram. Known compartments compatible with active transcription are indicated. H3-K79 occupies a distinct zone.

75 characterised P.falciparum nuclear compartments, namely, the telomere clusters[13], the nucleolar expression sites for nuclear rRNA genes (Mancio-Silva and Scherf, unpublished data) or the perinuclear expression site for var genes [10]. To examine the possibility that this area is linked to transcription of rDNA in the nucleolus, we conducted double labelling with an antibody against previously described nucleolar marker PfNop1[4](Fig 2 A). Double labelling showed that the two did not co-localise, and no consistent spatial relationship was observed between them. We have shown the presence of distinct telomeric clusters marking the periphery of the parasite nucleus [13]. We also demonstrated previously that the var genes localize at the nuclear periphery irrespective of their transcriptional state (Ralph et al., 2005 and unpublished data). Moreover, we described by electron microscopy the existence of a gap in the perinuclear heterochromatin, suggestive of a transcriptionally competent zone [10]. We thus examined the localization of H3-K79me3 in relation to telomere clusters and an active var gene. Synchronized FCR3-CSA parasites [9] (actively transcribing the var2CSA gene) were simultaneously labelled with H3-K79me3 antibody (IF) and a specific probe for the var2CSA gene or Rep20 respectively (FISH). In each case we scored the co-localization of H3-K79me3 with Rep20 signals (Fig. 2 B) and H3-K79me3 with var2CSA (Fig. 2 C.), where only a total overlap between the fluorescent foci was considered positive. The majority of the telomere clusters were distinct from and did not co-localise with the H3-K79me3 signals (Fig. 2 B). The var2CSA locus and the H3-K79me3zone co-localised only in 16% of the nuclei, arguing against an association of the var-transcription zone with H3-K79me3 (Fig. 2 C.). From our results, H3-K79me3 defines a novel substructure in the P. falciparum nucleus. It is possibly linked to a subgroup of genes other than var genes for transcription.

Discussion Here we have documented the presence of methylation specific histone modifications in Plasmodium with the aim of visualizing the functionally differentiated sub-nuclear compartments. Half of the screened methyl modifications indicated sub-nuclear regions differentially enriched for methyl marks (Fig.1 A), indicating that this may reflect functionally distinct states of chromatin. It has been demonstrated in other eukaryotes that genes can be dynamically recruited to sites of active transcription or ‘transcription factories’ [14] and that similar active genes cluster in specialized transcription factories. Moreover, reports show that epigenetic markers such as histone modifications can cause a region of chromatin to undergo nuclear compartmentalisation in a non random fashion [15]. It is increasingly becoming evident that Plasmodium, like many other eukaryotes, possesses a

76 higher order nuclear structure and it’s reasonable to envisage that similar genes and / or gene families may utilise this organization in a co-ordinated manner. So far only few studies have addressed this issue in Plasmodium. Our lab and others have shown that silenced var genes associate with telomeric clusters at the perinuclear heterochromatin and activation requires relocation in a ‘transcriptionally competent’ area, still at the periphery [10]. P. falciparum nucleolus is also an active transcription site at least for rRNA genes as in other organisms (Mancio Silva and Scherf, unpublished data). In Plasmodium, epigenetic marks have been studied as factors regulating antigenic variation. Recently it was reported that a singularly active var gene is defined by a combination of active (H3-K4me2/3) [5, 6]and repressive (H3-K9me3) marks at the 5’ coding region [5, 6]. In addition, lately, putative lysine methyltransferases and demethylases have been reported in Plasmodium, elucidating existence of a dynamic histone modification machinery [11]. Reports suggest that in higher eukaryotes, certain factories contain high concentrations of particular factors [16], including histone marks [17], which are hypothesised to ‘mark’ certain genes for targeting to specific factories[18]. Although the above-mentioned studies reveal the presence of many histone marks in Plasmodium, specifically methyl marks [6, 11], there has been none that systematically looked into the nuclear spatial distribution of these marks in the plasmodial nuclei. We show that H3-K79me3 is one of the most confined marks in P. falciparum resulting in a unique sub-nuclear pattern. Although no co-localisation was observed with the known Plasmodium compartments linked either to polymerase I or II transcription (Fig 2 A-C), it may be a mark for a novel putative transcription site (Fig 2 D) specific for similarly transcribed gene families other than the var genes. Interestingly however, in yeast it has an indirect role in maintaining heterochromatin by limiting the spread of Sir2/3 proteins [19]. Furthermore, recently this mark was shown to modulate antigenic variation in T. brucei. Deletion of DOT1B, the methyl transferase responsible for H3-K76me3 (corresponding to H3-K79 in yeast and mammals), compromises silencing of telomeric VSG (Variant Surface Glycoprotein) genes as well as monoallelic expression of VSGs and switching [20]. Given that both this mark and similar mechanism of monoallelic expression exist in Plasmodium, we need to envisage active and repressive activities with respect to the role of this modification. Moreover, as unlike in other eukaryotes, active and repressive marks co-exist in the nuclear periphery of P. falciparum. H3-K4me3 (activation) and H3-K9me3 (repression) being examples elucidating this scenario [6]. It will thus require an in depth analysis to reveal the gene(s) and / or families enriched for these interesting marks and their spatial location. Techniques like ChIP-on-chip or ChIP-sequencing that allow global profiling of histone PTMs will need to be undertaken to decipher this and our data offers a list of exciting

77 candidates to subject to such global analysis. Furthermore tools like Br-UTP labelling of transcription sites may be combined with our IFA study. Although feasible, this is technically challenging in a model system like Plasmodium (Freitas-Junior et al, personal communication). Some of these marks may be involved in promoting specific inter- chromosomal loop interactions, which could favour particular higher order nuclear substructures in malaria parasites. Such interactions can be detected using the chromosome conformation capture (3C) methodology. Now that different functional elements are being identified in the nucleus of P. falciparum, a major challenge is to unravel functional relationships between them. All in all, a more detailed genome wide investigation is now required to reveal the link of distinct subnuclear chromatin domains on the nuclear architecture. Understanding the dynamics of these histone PTMs will help build a ‘high-resolution’ picture of the P.falciparum nucleus to better link sub nuclear architecture with gene expression and possibly the identification of distinct transcription factories [20].

Materials and methods Parasite cultures and panning P. falciparum blood stage parasites from the 3D7 and FCR3 strain were used. Culture and panning assays for selection of FCR3 parasites transcribing var genes associated with CSA binding (referred to as FCR3-CSA) were performed as described [9].

Antibodies Commercially available anti-sera against respective histone modifications were used at following dilutions: Westernblot/IFA: H3-Core abcam1791, 1:1000/1:100, H3-K4me1 abcam8895, 1:2000/1:100, H3-K4me2 ab7766, 1:1000/ 1:100, H3-K4me3 ab8580, 1: 2000/1:100, H3-K9me1 ab9045, 1:2000/ 1:50, H3-K9me2 ab32521, 1:500/ 1:50, H3-K9me3 07-442(upstate), 1:200/1:100, H3-K27me3 ab6147, 1:500/1:50, H3-K36me3 07-549 (upstate), 1:500/ 1:100, H3-K79me3 ab2621, 1:500/ 1:100, H4-Core ab16483, 1:500/1:100, H4-K20me3 07-463 (upstate) 1:500/1:100. Rat anti-PfNop1 sera were generated as described [4] and used at a dilution 1:50.

Purification of histones and Western blot analysis Approx. 2.5 x 109 infected red blood cells were harvested 30-32 hours post invasion and parasite histones purified as described [6]. Precipitated histones were pelleted, air dried and analysed by 15% SDS-PAGE and transferred onto nitrocellulose membranes for Immunoblotting. Membranes were probed with antibodies against core histone H3, H4 and

78 respective site-specific methyl modifications (see Antibodies). Secondary antibodies conjugated to horseradish peroxidase (Pierce) were developed with Super-Signal West Pico Chemiluminescent Substrate (Pierce).

Immunofluorescence Immunofluorescence (IFA) was conducted both on mixed stage cultures (Rings, Trophozoites and Schizont stages), as well as synchronised ring-stage cultures. Parasite-infected erythrocytes were washed, saponin lysed and fixed in suspension with 4% paraformaldehyde. Parasites were incubated with the primary antibodies diluted in 1% bovine serum albumin at 37°C for 30 minutes, followed by incubation at 37°C for 30 minutes with fluorochrome conjugated secondary antibodies. Final secondary antibody dilutions were 1:500 Alexa Fluor 568 / 488 goat anti-rabbit. After final washes, parasites were deposited on microscope slides and mounted with Vectashield anti-fading containing 4-6-diamidino-2-phenylindole (DAPI) (Vector Labs). Images were captured using a Nikon Elipse 80i optical microscope. Double labelling IFAs were performed with final antibody dilutions of 1:200 for rabbit anti- H3-K79me3 and 1:50 for anti-rat PfNop1. Secondary detection was performed with 1:500 Alexa Fluor 488 / 568 goat anti-rabbit and 1:500 fluorescein-conjugated goat anti-rat.

Combined IFA - FISH IFA was performed on FCR3-CSA panned parasites as above. Final antibody dilutions were: 1:200 anti-H3-K79me3, 1:500 Alexa Fluor 568 goat anti-rabbit. Parasites were post-fixed in suspension with 4% paraformaldehyde, and followed for FISH [4]. Fluorescein labelled DNA var2CSA[10] and REP20 were generated using Fluorescein-High Prime (Roche) as per manufacturers instructions, and used as probes for active var gene and telomere clusters respectively.

Acknlowledgements This work was supported by the European Commission (BioMalPar) contract No; LSPH-CT- 2004-503578. N.I. has financial support from the BioMalPar PhD program. L.M.-S. has financial support from the Portuguese National Foundation for Science and Tecnology. A.S. is supported by the ANR grant (ANR-06-MIME-026–01).

79 References [1] T. Jenuwein, C.D. Allis, Translating the histone code, Science 293 (2001) 1074-1080. [2] W. Fischle, Y. Wang, C.D. Allis, Histone and chromatin cross-talk, Curr Opin Cell Biol 15 (2003) 172-183. [3] B.M. Turner, Cellular memory and the histone code, Cell 111 (2002) 285-291. [4] L.H. Freitas-Junior, R. Hernandez-Rivas, S.A. Ralph, D. Montiel-Condado, O.K. Ruvalcaba- Salazar, A.P. Rojas-Meza, L. Mancio-Silva, R.J. Leal-Silvestre, A.M. Gontijo, S. Shorte, A. Scherf, Telomeric heterochromatin propagation and histone acetylation control mutually exclusive expression of antigenic variation genes in malaria parasites, Cell 121 (2005) 25-36. [5] T. Chookajorn, R. Dzikowski, M. Frank, F. Li, A.Z. Jiwani, D.L. Hartl, K.W. Deitsch, Epigenetic memory at malaria virulence genes, Proc Natl Acad Sci U S A 104 (2007) 899-902. [6] J.J. Lopez-Rubio, A.M. Gontijo, M.C. Nunes, N. Issar, R. Hernandez Rivas, A. Scherf, 5' flanking region of var genes nucleate histone modification patterns linked to phenotypic inheritance of virulence traits in malaria parasites, Mol Microbiol 66 (2007) 1296-1305. [7] S.A. Ralph, A. Scherf, The epigenetic control of antigenic variation in Plasmodium falciparum, Curr Opin Microbiol 8 (2005) 434-440. [8] L. Cui, J. Miao, T. Furuya, X. Li, X.Z. Su, L. Cui, PfGCN5-mediated histone H3 acetylation plays a key role in gene expression in Plasmodium falciparum, Eukaryot Cell 6 (2007) 1219-1227. [9] A. Scherf, R. Hernandez-Rivas, P. Buffet, E. Bottius, C. Benatar, B. Pouvelle, J. Gysin, M. Lanzer, Antigenic variation in malaria: in situ switching, relaxed and mutually exclusive transcription of var genes during intra-erythrocytic development in Plasmodium falciparum, Embo J 17 (1998) 5418-5426. [10] S.A. Ralph, C. Scheidig-Benatar, A. Scherf, Antigenic variation in Plasmodium falciparum is associated with movement of var loci between subnuclear locations, Proc Natl Acad Sci U S A 102 (2005) 5414-5419. [11] L. Cui, Q. Fan, L. Cui, J. Miao, Histone lysine methyltransferases and demethylases in Plasmodium falciparum, International journal for parasitology 38 (2008) 1083-1097. [12] C.F. Sautel, D. Cannella, O. Bastien, S. Kieffer, D. Aldebert, J. Garin, I. Tardieux, H. Belrhali, M.A. Hakimi, SET8-mediated methylations of histone H4 lysine 20 mark silent heterochromatic domains in apicomplexan genomes, Mol Cell Biol 27 (2007) 5711-5724. [13] L.H. Freitas-Junior, E. Bottius, L.A. Pirrit, K.W. Deitsch, C. Scheidig, F. Guinet, U. Nehrbass, T.E. Wellems, A. Scherf, Frequent ectopic recombination of virulence factor genes in telomeric chromosome clusters of P. falciparum, Nature 407 (2000) 1018-1022. [14] C.S. Osborne, L. Chakalova, K.E. Brown, D. Carter, A. Horton, E. Debrand, B. Goyenechea, J.A. Mitchell, S. Lopes, W. Reik, P. Fraser, Active genes dynamically colocalize to shared sites of ongoing transcription, Nature genetics 36 (2004) 1065-1071. [15] M.R. Branco, A. Pombo, Chromosome organization: new facts, new models, Trends in cell biology 17 (2007) 127-134. [16] J. Bartlett, J. Blagojevic, D. Carter, C. Eskiw, M. Fromaget, C. Job, M. Shamsher, I.F. Trindade, M. Xu, P.R. Cook, Specialized transcription factories, Biochem Soc Symp (2006) 67-75.

80 [17] W.G. Muller, D. Rieder, T.S. Karpova, S. John, Z. Trajanoski, J.G. McNally, Organization of chromatin and histone modifications at a transcription site, J Cell Biol 177 (2007) 957-967. [18] M. Xu, P.R. Cook, Similar active genes cluster in specialized transcription factories, J Cell Biol 181 (2008) 615-623. [19] F. van Leeuwen, P.R. Gafken, D.E. Gottschling, Dot1p modulates silencing in yeast by methylation of the nucleosome core, Cell 109 (2002) 745-756. [20] L.M. Figueiredo, C.J. Janzen, G.A. Cross, A histone methyltransferase modulates antigenic variation in African trypanosomes, PLoS biology 6 (2008) e161. [21] B. Li, M. Carey, J.L. Workman, The role of chromatin during transcription, Cell 128 (2007) 707-719.

81

82

RESULTS V

Genome-wide localization of heterochromatin couples clonally variant gene regulation to perinuclear epigenetic factories in malaria parasites

Jose-Juan Lopez-Rubio*, Liliana Mancio-Silva* and Artur Scherf

Institut Pasteur, Unité de Biologie des Interactions Hôte-Parasite, CNRS URA2581, 25 Rue du Dr Roux, 75724 Paris, France * These authors contributed equally to this work

Correspondence: Artur Scherf, tel: + 33 (01) 4568-8616, fax: + 33 (01) 4568-8348, e-mail: [email protected]

Submitted.

83

84

Summary

Clonally variant gene families provide the blueprint for phenotypic plasticity in Plasmodium falciparum, a process indispensable for survival of the pathogen in its human host. Using high-resolution ChIP-chip analysis, genome wide mapping of heterochromatin mark H3K9me3 reveals, for the first time, a mechanism of coordinated gene repression in this parasite. This is virtually limited to differentially transcribed virulence gene families. We show that high level of H3K9me3 links mutually exclusive gene expression, spatial gene positioning and perinuclear epigenetic repression factories (PERF). In PfSir2 knock out parasites, we observe that the change of H3K9me3 to H3K9ac at 5’UTRs uncouples var genes from monoallelic transcription. This discontinuous manner of the histone mark changes along chromosomes points to a complex spatial chromatin formation via multiple loops of var 5’UTRs to PERF. Our data show that P. falciparum has evolved a spatially confined epigenetic factory to control genes involved in phenotypic variation and pathogenesis.

Introduction

The human protozoan malaria parasite Plasmodium falciparum kills more than 2 million people (mainly children) per year (Snow et al., 2005). This parasite is well adapted to survive in changing host environments such as the Anopheles mosquito mid gut and salivary glands, human liver hepatocytes and infected red blood cells (IRBC). Malaria pathogenesis is linked to the parasite’s capacity to develop different phenotypic states in genetically clonal blood stage parasites (Fried and Duffy, 1996; Jensen et al., 2004). However, most processes causing pathogenesis are not yet understood at the molecular level and may involve expression of particular combinations of clonally variant molecules. In malaria parasites, phenotypic diversity is commonly achieved by the expression of clonally variant surface molecules either at the erythrocyte membrane or merozoite surface (Cortes et al., 2007; Scherf et al., 1998; Stubbs et al., 2005). Switching of expression to another variant molecule prolongs the period of infection and generates alternative evasion pathways. Little is known about factors that govern phenotypic plasticity in malaria parasites. The best-studied example of variant surface proteins is the P. falciparum Erythrocyte Membrane Protein 1 (PfEMP1) (Leech et al., 1984). PfEMP1 is a major virulence factor involved in antigenic variation and adherence of IRBC to host receptors, resulting in sequestration of infected erythrocytes to capillaries in the brain and other critical organs (Kyes et al., 2007). PfEMP1 is encoded by genes of the 60-member var family, which are located on subtelomeric and

85

internal chromosome loci. In a single parasite, only one var gene is expressed in a mutually exclusive manner under the control of epigenetic factors (Scherf et al., 1998). A number of recent publications point to reversible chromatin changes, such as histone methylation and acetylation, as important control elements of P. falciparum gene silencing and mono-allelic activation (Scherf et al., 2008a). Enzymes that add or remove acetyl and methyl marks in P. falciparum have also been identified (Cui et al., 2008; Merrick and Duraisingh, 2007; Miao et al., 2006). It has been demonstrated that trimethylation of histone H3 lysine 9 (H3K9me3) is associated to transcriptionally silent var genes (Chookajorn et al., 2007; Lopez-Rubio et al., 2007). Upon var gene activation, at the 5’flanking region (5’UTR), methylation is replaced by acetylation at lysine 9 of histone H3, and histone H3 lysine 4 is di- and trimethylated (Lopez-Rubio et al., 2007). Furthermore, PfSir2 histone deacetylase has been shown to spread from telomeres into the subtelomeric coding regions, leading to stable but reversible repression of subtelomeric var genes (Duraisingh et al., 2005; Freitas-Junior et al., 2005; Mancio-Silva et al., 2008). Genetic elements such as the 5’UTR (Voss et al., 2006) contain essential information able to interact with the epigenetic control machinery. Experimental data suggest that also the var intron-linked promoter activity contribute to mono-allelic expression (Dzikowski et al., 2006). Sequencing of the P. falciparum genome has led to the discovery of multiple novel gene families mainly clustered at subtelomeres adjacent to var genes (Gardner et al., 2002). A few of these have been studied, such as rif (Fernandez et al., 1999; Kyes et al., 1999), clag (Cortes et al., 2007), stevor and PfMC-2TM (Blythe et al., 2008; Lavazec et al., 2007) and, similarly to var genes, are clonally variant and are possibly involved in immune evasion (Scherf et al., 2008a). Telomeres are spatially restricted to nuclear periphery, where they form clusters of 3- 7 heterologous chromosome ends (Freitas-Junior et al., 2000). This compartment is believed to enhance continual genetic exchange between members of gene families and may explain the vast var repertoire diversity observed between field isolates (Barry et al., 2007). Investigating the role of clusters in gene expression has resulted in inconsistent data. Experimental evidence, using parasites transfected with various constructs, reported relocation of an active var gene into a cluster (Duraisingh et al., 2005; Marty et al., 2006; Voss et al., 2006) whereas other studies, using parasite populations expressing a single var gene, observed the relocation outside of cluster (Ralph et al., 2005). Up to now, it remained unknown if phenotypic variation shares common epigenetic control mechanisms in any pathogenic parasite. Here we report that epigenetic factories at the perinuclear space control virulence gene family silencing in P. falciparum. Genome-wide analysis of histone marks revealed that heterochromatin mark H3K9me3 is enriched in clonally variant gene families and demonstrate that this mark is linked to the perinuclear

86

location of central and subtelomeric regions enriched for H3K9me3. This histone modification is controlled by the histone deacetylase PfSir2, however only for a subset of variant gene family members, revealing an unusual mechanism of PfSir2 repression distinct from the postulated telomere position effect. In addition, this study allowed us to identify novel subtelomeric gene families, which are apparently differentially transcribed and represent potential new virulence factors participating in phenotypic plasticity of malaria parasites.

Results H3K9me3 shows a highly localized atypical distribution in P. falciparum chromosomes We performed ChIP-chip analysis to determine the localization of H3K9me3 across the P. falciparum 14 chromosomes. Because most of the dynamic histone modifications changes controlling gene regulation in P. falciparum appear to occur at gene flanking non-coding regions (Cui et al., 2007; Lopez-Rubio et al., 2007), and taking advantage of the small size of the parasite genome, we developed a custom-made high-resolution tilling array covering coding and intergenic regions, and also non-coding subtelomeric regions (for details see Supplemental experimental procedures). We observed several prominent H3K9me3 enriched domains restricted to all subtelomeric regions as well as to central regions of chromosomes 4, 6, 7, 8 and 12 (Figure 1A). ChIP-chip analysis of another repressive mark (H4K20me3) and activation marks (H3K4me3 and H3K9ac) showed a broad distribution across the genome (as exemplified for chromosome 12 in Figure 2), contrasting with the pattern of H3K9me3. H3K4me3 and H3K9ac were basically absent at the major heterochromatin loci coated with H3K9me3. Our results demonstrate the almost universal utilization of H4K20me3, H3K4me3 and H3K9ac, whereas H3K9me3 is highly restricted to chromosome ends and some central regions. P. falciparum chromosome ends display a conserved higher order structure being composed of telomeric DNA followed by six repetitive non-coding polymorphic DNA blocks (TAREs 1-6), with an average total length of approximately 30 kb (Figueiredo et al., 2000). Since these regions are composed of repetitive DNA, our custom tilling array included probes only to the extremities of one chromosome (chr.12) (see Supplemental experimental procedures). We observed a strong enrichment in H3K9me3 at the entire TAREs 1-6 (Figure 1B right, Figure 2 both flanks and Figure S3). On the contrary, telomeric repeat DNA presented only a modest H3K9me3 level (Figure S3 right). H3K4me3 and H3K9ac were completely absent from TAREs and telomere repeats, and H4K20me3 was detectable only at low levels (Figure 2 both flanks).

87

Figure 1. ChIP-chip genome map of H3K9me3. (A) Paired tracings of raw data (top tracing for each chromosome) and the distribution of peaks (lower tracing for each chromosome) for H3K9me3 for P. falciparum 14 chromosomes (non-coding chromosomes ends are excluded). Regions represented in Figure 1B and 1C are boxed. Regions analysed in Figure 1D are marked with i, ii and iii. (B) H3K9me3 distribution at telomeric and subtelomeric regions of chromosome 12. TAREs and coding regions are separated with a dotted line. (C) H3K9me3 distribution for a central H3K9me3 enriched domain in chromosome 12. Paired tracing for (B) and (C) as in (A). (D) Table presenting H3K9me3 enriched genes that are neither var genes nor genes associated with var gene loci. Transcriptional data were collected from PlasmoDB. Gene marked with asterisk indicates that the enrichment covers partially the locus. In all cases, raw data are presented as the log2 ratio of the hybridization signal given by DNA immunoprecipitated using specific antibody against H3K9me3 compared with the signal given by the input DNA sample. Peaks data are represented as described in Supplementary experimental procedures. The position on the genome sequence is indicated at the bottom for (A) and at the top for (B) and (C). The position and annotation of the predicted genes are presented at the bottom (CDS; blue bars, sense; green bars, antisense).

88

Figure 2. Chromosome 12 distribution profiles of different histone modifications. The raw data (top tracing) and the distribution of peaks (lower tracing) identified for H3K9me3, H4K20me3, H3K4me3 and H3K9ac are presented. Non-coding chromosome ends (TAREs and telomeres) are separated with a dotted line. Schematic representation of chromosome 12 showing main features and the position on the genome is indicated at the top. Raw and peaks data are presented as in Figure 1.

Eukaryotic centromeres and pericentromeric regions are embedded in heterochromatin, and exhibit unique chromatin architecture necessary for their chromosome segregation function (Grewal and Jia, 2007). P. falciparum centromeric DNA differs, in terms of size and organization, from centromeres of other organisms (Kelly et al., 2006). Our tilled array also covered pericentromeric DNA. In contrast to other organisms where this region is enriched in H3K9me3, our ChIP-chip analysis showed that pericentromeric chromatin in P. falciparum apparently was not enriched for H3K9me3 (example for pericentromere of chromosome 12 in Figure 2), which is consistent with the unusual nature of centromeric DNA in this parasite.

H3K9me3 is linked to differentially transcribed virulence gene families

Detailed examination of the major heterochromatin regions revealed that the entire var gene family was enriched in H3K9me3, which is consistent with previous predictions (Chookajorn et al., 2007; Lopez-Rubio et al., 2007). In addition, we observed that many distinct gene families adjacent to subtelomeric var genes were also enriched in H3K9me3 at the 5’UTR and coding region (Figure 1B). Chromosome central H3K9me3 enriched domains contained generally only var, rif and stevor genes (Figure 1C). We found H3K9me3 enrichments in seven previously identified subtelomeric gene families: var, rif, stevor, Pfmc-2TM, clag,

89

PfAcs and PHIST (Table 1). For almost all of these gene families there are experimental evidences demonstrating a differential transcription mode of individual members (Scherf et al., 2008a). Another common feature is that many of these gene families code for variant surface proteins participating in the vast phenotypic plasticity of blood stage parasites and immune evasion. Additionally, based on our H3K9me3 findings together with OrthoMCL database (Chen et al., 2006), we identified five new putative subtelomeric gene families. In the absence of any biological data, we called these families PFD0075w-like, PF14_0742-like, PF14_0758-like, PF14_0743-like and MAL8P1.335-like (Table 1). Published expression profile data of these genes (Le Roch et al., 2003) analysed using PlasmoDB (http://plasmodb.org/plasmo/) strongly suggest a differential transcription profile of these novel gene families, which is a hallmark for phenotypic variation. Silent members presented high levels of H3K9me3, whereas transcribed members had low levels of this mark in the 5’UTR (Table 1). Therefore, our results indicate that H3K9me3 at 5’UTR may predict genes that are controlled by a similar epigenetic mechanism.

90

However, our analysis also showed that some subtelomeric gene families were not enriched for H3K9me3, such as surfin, FIKK and ETRAMPS (Table S1). These genes were interspersed into H3K9me3 enriched regions, resulting in a mosaic pattern of this histone modification. The question arises of whether this non-H3K9m3 enriched gene families are differentially transcribed or not. For the 20-member FIKK family, quantitative PCR data show that some members are transcribed at very low levels during blood stage development (Nunes et al., 2007). For the surfin and ETRAMPS families, expression profile data (PlasmoDB) show that some members are not expressed during blood stages. Further quantitative expression data of cloned parasites are necessary to determine whether these families undergo clonal variation. We also detected three H3K9me3 short-ranged peaks outside the large var-defined heterochromatic domains on chromosomes 10, 11 and 12 (i, ii and iii, respectively in Figure 1A). Genes encoded in these heterochromatic ‘islands’ are listed in Figure 1D. Apart from PFL1085w, which encodes a member of the ApiAP2 family of transcription factors (De Silva et al., 2008), very little information is available for the other genes. Further analysis is needed to determine whether these genes participate in phenotypic diversity of parasites or are epigenetically co-regulated.

PfKMT1 and H3K9me3 enriched genes are localized to the nuclear periphery Given the unusual highly localized H3K9me3 pattern in P. falciparum genome, we were interested in investigating the nuclear organization of H3K9me3 domains. H3K9 methylation is performed by a histone lysine 9 methyltransferase KMT1 (Allis et al., 2007), of which an homologue (PF08_0012) has been found in P. falciparum genome (Cui et al., 2008). Antibodies raised against the P. falciparum KMT1 were used to study the cellular location of this molecule. Surprisingly, we observed that PfKMT1 fluorescent signals located largely at the periphery of P. falciparum nuclei (Figure 3B). Likewise, antibodies specific for H3K9me3 showed a predominant enrichment in the perinuclear space (Issar et. al, submitted) strongly suggesting that H3K9me3 enriched genes might locate in this nuclear compartment. To test this hypothesis, we performed a quantitative large-scale localization study of genes enriched in H3K9me3 using fluorescent in situ hybridization (FISH). We chose to examine the entire var gene family and two non-var genes, which form distinct H3K9me3 peaks at internal (PFL1085w) and subtelomeric chromosomal position (PF11_0479) (iii and ii respectively in Figure 1A and 1D). For analysis of the var 60-members, we used exon2 probes that cross-hybridize to var genes of each major group: group A (10 subtelomeric genes), group B (22 subtelomeric genes) and group C (13 internal genes grouped in 7 central

91

Figure 3. PfKMT1 and genes enriched in H3K9me3 are positioned at the nuclear periphery. (A) Schematic of the var loci organization. Each var gene consists of a large variable exon1 (4-9 kb) and a conserved exon2 (~1.5 kb). The 60-member var gene family can be divided in three major groups (A, B and C) based on chromosome location and sequence similarity. Group A contains 10 telomeric genes transcribed towards the telomere, 22 telomeric genes belong to group B and are tanscribed towards the centromere, and group C includes 13 internal genes (Lavstsen et al., 2003). Among the var arrays there are members of other multi-gene families, such as rifin and stevor. The location of FISH probes is indicated with horizontal bars. (B) Immunofluorescence staining of P. falciparum putative H3K9 methyltransferase (PfKMT1). 3D7 parasites are in late schizont stage and nuclear DNA was detected with Dapi (red). (C) DNA FISH analysis of 3D7 ring stage parasites nuclei. First row: subtelomeric (var exon2A, var exon2B and PF11_0479) and internal H3K9me3 enriched genes (var exon2C, var PFF0845c and PFL1085w). Second row: internal gene non-enriched in H3K9me3 (PFB0540w). FISH signals are in green and the nuclear DNA was stained with Dapi (red). Bar graph: subnuclear position of FISH signals with respect to the three concentric zones of equal surface (zone 1, 2 and 3). The number of nuclei analyzed and the 95% confidence values (p) for the χ2 test between random and test distributions are: var exon2A n=143, p=1.50x10-8; var exon2B n=166, p=3.02x10-16; var exon2C n=194, p=6.92x10-15; var PFF0845c n=180, p=1.90x10-2; PF11_0479 n=226, p=7.59x10-3; PFL1085w n=159, p=5.30x10-2; PFB0540w n=151, p=1.10x10-2. (D) Two color DNA FISH of var exon2C probe (green) and telomeric clusters (TARE6 probe, red), on 3D7 ring-stage parasites. Nuclear DNA was detected with Dapi (not shown). Bar graph: colocalization was 48.06% (n=105) and the 95% confidence values (p) for the χ2 test between random and test distributions was p=0.615. (E) RNA/DNA FISH analysis of var transcripts and telomeric clusters. First row: RNA FISH signals are shown in green and DNA FISH signals are shown in red (TARE6 probe). Nuclear DNA was visualized with Dapi (not shown). var2CSA transcripts were detected in ring-stage FCR3-CSA selected parasites (first panel). Exon2 probes of each var group were hybridized on a bulk cultured 3D7 ring-stage parasites. Second row: Quantification of colocalization of var transcripts with telomeric clusters (var2CSA 77.94% of non-colocalization, n=68; var exon2A 83.64%, n=55; var exon2B 81.48%, n=54; var exon2C 78.95%, n=38). In all cases: representative examples are shown as merged images, FISH signals were scored using nuclei from multiple experiments, and scale bars represent 1µm. For quantification of colocalization only nuclei containing at least three TARE6 signals were considered as positive and colocalization was defined as any overlap between two fluorescent signals. (F) Proposed model for nuclear organization of H3K9me3 enriched domains. The chromosome 12 was used as an example (see text for details).

92

regions, see Figure 1A and 3A) (Lavstsen et al., 2003). We observed 2 to 5 foci per nucleus for each var exon2 probe, indicating clustering of var genes from group A, B and C (Figure 3C and S1). The other FISH probes tested were gene specific probes (single internal var PFF0845c, PF11_0479 and PFL1085w) and produced single foci per nuclei (Figure 3C). In order to precisely define the nuclear boundaries, we developed an antibody that recognizes a P. falciparum protein encoded by a gene orthologous to a yeast nuclear pore protein (PF14_0706) (Figure S2). The area of nuclear section obtained in immunofluorescence staining was calculated and then divided into equal thirds. FISH signals were scored as localizing to zone 1, 2 and 3 (Hediger et al., 2002), as illustrated in Figures 3C and S2. Quantitative analysis of nuclear position revealed that all var and non-var genes enriched for H3K9me3 analyzed were not randomly distributed, but rather located predominantly at the nuclear periphery (zone 1, Figure 3C). By contrast, an internal gene non-enriched in H3K9me3, PFB0540w (Figueiredo et al., 2002), was found mostly in internal nuclear regions (zone 3, Figure 3C). Thus, virulence gene families and single copy genes enriched in H3K9me3 are localized at the perinuclear space, irrespective of their chromosomal position. This led us to hypothesize that PfKMT1 and H3K9me3 might determine the nuclear organization of genes. P. falciparum chromosome-ends are tethered to the nuclear membrane forming clusters (Freitas-Junior et al., 2000). To exclude the possibility of internal H3K9me3 enriched genes (var group C, var PFF0845c and PFL1085w) being at the nuclear periphery as a consequence of being associated to telomeric clusters, we analyzed their relative position to these clusters. Simultaneous FISH analysis of var exon2C with a telomeric cluster marker TARE6, demonstrated that var group C clusters localized randomly relative to chromosome-end clusters (Figure 3D). Similar random distribution was obtained for co-localization of TARE6 with 5’UTR from var group C and the single internal var PFF0845c (data not shown), thus suggesting that perinuclear tethering of chromosome internal genes is independent of telomeric clustering. Taken together, our results indicate that heterochromatic peripheral region is composed of telomeric and non-telomeric clusters. Moreover, it implies that certain Plasmodium chromosomes (the case of chromosome 12, which has several domains of H3K9me3 enrichment, Figure 1A) should form several large loops into the nuclear membrane, as exemplified in Figure 3F.

93

RNA FISH locates var transcripts to the nuclear periphery dissociated from telomeric clusters Previous work has demonstrated that upon activation of a var gene (var2CSA), H3K9me3 at the 5’UTR is lost but persists at the 3’coding region (Lopez-Rubio et al., 2007). It has also been shown that active var2CSA gene locates at the nuclear periphery. However, there are conflicting conclusions with regard to its association with the telomeric clusters (Scherf et al., 2008b). To overcome potential problems linked to DNA FISH such as possible relocation of the var2CSA gene to telomeres once transcription ceases, we located var2CSA transcripts using RNA FISH. Telomeric clusters were then visualized performing a second hybridization using the TARE6 probe by DNA FISH. We observed that var2CSA transcripts were clearly dissociated from telomeric clusters (Figure 3E). To rule out the possibility that this observation was specific for the var2CSA gene, we localized transcripts of the entire var gene repertoire. We performed RNA/DNA FISH using the exon2 var group-specific probes. We observed that mRNA transcripts from the different var groups did not associate with the telomeric clusters (Figure 3E), implying that activation of any var gene involves nuclear reposition into a peripheral region compatible with transcription. With regard to var group A and B, which are linked to telomeric repeats, activation involving acquisition of active chromatin marks, such as H3K9ac and H3K4me2/3, presumably leads to chromatin decondensation and formation of a single-gene loop, which separate out the transcribed gene from more compact and silent telomeric cluster (schematically shown in Figure 3F). Although anchoring at the nuclear periphery must require appropriate proteins, the fact that the H3K9me3 subsists in the active gene (at the 3’coding region), may be important for maintenance of the gene at the nuclear periphery.

PfSir2 controls H3K9me3 in a discontinuous manner In order to determine whether the putative PfKMT1 was responsible of H3K9 methylation and its role in the localization of the heterochromatin domains to the nuclear periphery, we attempted to delete the unique PfKMT1 locus. Using the currently available knock-out methods for P. falciparum, we were incapable to generate viable parasites with PfKMT1 gene disruption. H3K9 trimethylation requires the action of a histone deacetylase to remove the acetyl group before the methylation step (Nakayama et al., 2001). We used a parasite knock- out strain for a histone deacetylase PfSir2 (Duraisingh et al., 2005) to analyse the molecular events leading to de-repression of a subset of subtelomeric genes and its role in the formation of heterochromatin and nuclear organization. ChIP-Chip analysis of PfSir2 disrupted 3D7 parasites showed that H3K9me3 levels were reduced and H3K9ac levels increased,

94

comparing to wild-type parasites, at the 5’UTR of the previously published upregulated genes (Figure 4 and Table 2). Competition between acetylation and methylation at H3K9 in the 5’UTR has been previously shown to influence transcription activity of var genes (Lopez- Rubio et al., 2007). In the absence of PfSir2, apparently this balance is disturbed in favour of H3K9ac. Unexpectedly, this effect is seen only in two out of 12 gene families enriched in H3K9me3. Histone modification changes associated with up-regulated var and rif genes in the absence of PfSir2 are illustrated in Figure 4A-D. Our results suggest that PfSir2-mediated H3K9 deacetylation is required for establishing repressive H3K9 methylation for a subset of var and rif genes.

Figure 4. ChIP-chip analysis of PfSir2 knock-out (sir2 KO) parasites. Raw data for H3K9me3 (black) and H3K9ac (grey), in 3D7 wild-type (wt) and sir2 KO parasites. Schematic representation of the analyzed regions showing main features and the position on the genome is indicated at the top. The position, annotation and whether upregulated or not in sir2 KO parasites of the predicted genes are presented at the bottom. (A-D) Local effects: decrease of H3K9me3 and increase of H3K9ac at 5’UTR of var and rif genes. (E and F) Regional effects: decrease of H3K9me. Grey squares highlight the regions with histone modifications changes. Transcription data have been obtained from PlasmoDB. Raw data are presented as in Figure 1. (G) Proposed model for a common epigenetic default silencing mechanism at the nuclear periphery for genes families linked to virulence. Intergenic DNA regions may interact with PERF via small loop formation. PERF is composed of PfSir2, PfOrc1, PfHP1 and presumably PfKMT1 and other histone deacetylases.

95

Surprisingly, the lack of PfSir2 resulted in very local effects, in a discontinuous manner (Figure 4A-D). The postulated telomere position effect of PfSir2 (Freitas-Junior et al., 2005) would have predicted a continuous H3K9me3 depletion from the telomere into the coding region. Together, these data suggest that different separated DNA regions may interact with subtelomeric PfSir2 via small loop formation (schematically shown in Figure 4G). Consistent with this confined effect of PfSir2 in H3K9me3, analysis of the nuclear position of PfKMT1 and H3K9me3 enriched genes in ΔPfSir2 parasites did not produce major changes in nuclear localization (data not shown). A number of additional puzzling features of PfSir2 emerged from our study. First, a general decrease of H3K9me3 with no increase of H3K9ac at basically all intergenic regions of the heterochromatic domains was observed, even in genes whose transcriptional state did not change (Figure 4E). Second, nearly the entire subtelomeric domain in the left arm of chromosome 7 (comprising 13 members of gene families) was depleted of H3K9me3 and H3K9ac levels remained low (Figure 4F). Unfortunately, there are no available transcription data for these loci in PfSir2 disrupted parasites to see whether the absence of H3K9 methylation without induction of acetylation affects repression. Third, we observed only a very weak reduction of H3K9me3 at subtelomeric repeats (Figure S3) associated to a slight increase in H3K9ac, suggesting that other histone deacetylases may cooperate with the process of H3K9me.

96

Discussion

In this study we demonstrate for the first time a link between malaria parasite virulence genes enriched for H3K9me3, their spatial location at the nuclear periphery and their epigenetic silencing via a perinuclear repression factory. This unique association is almost entirely restricted to clonally variant multi-gene families involved in phenotypic plasticity. Here, we established the first high-resolution genome-wide H3K9me3 heterochromatin map for a malaria parasite. As in other organisms (Grewal and Jia, 2007), P. falciparum chromosome-end regions (TAREs 1-6) are enriched in H3K9me3 and may represent constitutive heterochromatin. In contrast to other organisms, however, no enrichment was found on pericentromeric DNA and high enrichment was found in all subtelomeric coding regions. A few small blocks of H3K9m3 were also detected in chromosome internal regions (Figure 1). Other histone modifications showed a broad distribution along chromosomes (Figure 2) and their genome wide link to gene expression has been initiated in collaboration with PlasmoDB. The fact that we used a high-resolution tilling array including non-coding regions may explain why in a previous ChIP-chip study, using only coding regions with a very low coverage, the authors did not observe the restricted localization of H3K9me3 (Cui et al., 2007). Importantly, we found five new subtelomeric gene families not yet characterized, which are enriched in H3K9me3, differentially transcribed, and that may participate in the phenotypic makeup of genetically clonal blood stage parasites. Insight into these potential virulence factors is very important for our understanding of malaria pathogenesis. It is noteworthy that other variant gene families such as surfin, FIKK and ETRAMP (Table S1), which are not enriched in H3K9me3, are interspersed within the subtelomeric region, demonstrating that two types of gene families, apparently controlled by distinct mechanisms, co-exist in P. falciparum subtelomeres. The data also reveal that epigenetic boundaries do exist in the heterochromatin domains to prevent the spreading of the repressive mark into neighbouring regions. We provide, for the first time, genome wide precise physical limitations for H3K9me3 in P. falciparum. It has been suggested that RNA Polymerase III (Pol III) transcribed genes act as barriers against self-propagating heterochromatin in single-celled eukaryotes (Lunyak, 2008). Indeed, some tRNA genes localize in the proximity of heterochromatin domains borders of P. falciparum (for example in chromosome 4 left end and chromosome 13 right end). Assays need to be developed to elucidate whether Pol III transcribed genes or other specific DNA elements are involved in blocking the spreading of H3K9me3. The present work highlights the idea that P. falciparum genomic DNA is non-randomly distributed in the nucleus and that gene expression and spatial organization are strongly linked

97

(see the model in Figure 3F). Nuclear position analysis of actively transcribed var genes, in their native chromosomal context, confirmed their dissociation from telomeric clusters. We speculate that chromatin-loop formation may be associated to activation and relocation of subtelomeric var genes. Relocalization of a subtelomeric gene from a telomeric focus associated to transcriptional activation has also been described in S. cereviseae (Taddei et al., 2006). Notably, induction of the same gene through an alternative artificial pathway produces relocalization to a different region in the yeast nucleus. This underlines the importance of the natural chromosomal context to decipher epigenetic mechanisms of gene activation. One of the most striking findings is that perinuclear large loop formation of central chromosome coding regions is associated to H3K9me3, and this is independent of either telomeric clustering or a GC-rich element (Figure 3). This ~200bp GC-rich element is strictly associated with internal var arrays, and is not found in subtelomeric var genes, nor near the single internal var gene on chromosomes 6 and 12, and apparently produces non-coding RNAs (Hall et al., 2002; Mourier et al., 2008). Together, our observations point to PfKMT1 and H3K9me3 as determinants of spatial organization of chromosomes in P. falciparum nuclei. This is supported by data obtained in S. pombe showing SpKMT1 to be crucial for correct subnuclear localization of the mating-type region (Alfredsson-Timmins et al., 2007). In P. falciparum, however, deletion of PfKMT1 produces a lethal effect on parasite proliferation. Improved inducible gene-knockdown systems will be necessary to gain insight into the PfKMT1 function in spatial organization of P. falciparum nuclei. Our results are compatible with the existence of ‘Perinuclear-linked Epigenetic Repressive Factories’ (PERF) (see model in Figure 4G) composed of PfSir2 and other potential silencing factors such as PfOrc1 (Mancio-Silva et al., 2008), PfHP1 (Perez-Toledo et al., submitted) and probably PfKMT1. Although experiments are needed to clarify whether these telomeric associated proteins form complexes, evidences from studies in S. pombe (Bannister et al., 2001; Nakayama et al., 2001; Yamada et al., 2005) indicate that H3K9me provides the binding site for HP1 to recruit other proteins, such as histone deacetylases and methylases, involved in heterochromatin stabilization and spreading. DNA binding proteins and/or transcribed repetitive subtelomeric non-coding RNAs (Mourier et al., 2008) are possibly involved in the initial targeting of histone modifying enzymes to the PERF. It will be also interesting to investigate whether telomeric and internal heterochromatic regions share the same PERF components. Having established high-resolution ChIP-chip assays for P. falciparum, we were in a position to analyze chromatin changes in PfSir2 mutant parasites, which display transcriptional de- repression of a subset of var and rif genes (Duraisingh et al., 2005). We show that PfSir2 acts precisely on H3K9 methylation at the 5’UTR of genes, which have been shown to be

98

upregulated in ΔPfsir2 parasites. The discontinuous pattern of H3K9me3 removal along subtelomeric regions and replacement by acetylation was unexpected given the current model of telomere position effect, which predicts a gradient along the telomere into the subtelomeric coding regions (Freitas-Junior et al., 2005; Mancio-Silva et al., 2008). The results strongly suggest a distinct mechanism of PfSir2-mediated silencing. We hypothesize that multiple small loop structures of chromosome regions harbouring variant multi-gene families interact directly with PERF (see model in Figure 4G). In the case of PfSir2, this interaction apparently is focused on the 5’UTR of some var and rif genes, underlining the important role of this DNA region in var gene control. Several elements suggest that subtelomeric chromatin is made-up of other histone deacetylases. First, PfSir2 does only control repression of a small subset of the H3K9me3- enriched genes (5%), and second, PfSir2 inactivation has no major change in the H3K9me3 levels in the non-coding subtelomeric region. This is in contrast to S. pombe where loss of Sir2 resulted in a general decrease in H3K9me3 and increase in H3K9ac at telomeres and telomeric-associated sequences (Shankaranarayana et al., 2003). We had searched the genome for other Sir-like genes and found an open reading frame on chromosome 14 that is homologous to human Sirt6, which acts in mammals at subtelomeric repeats (Michishita et al., 2008). Antibodies raised against this protein show a similar location at the nuclear periphery as PfSir2 (L. Mancio-Silva and A. Scherf unpublished data), strongly suggesting that a second telomere-associated histone deacetylase contributes to epigenetic silencing of virulence gene families. In conclusion, our study is the first demonstration that a pathogen has developed a specific mechanism to restrict repressive H3K9me3 almost exclusively to clonally variant gene families. Moreover, our data show a complex spatial distribution of plasmodial chromosomes in the nucleus, which apparently creates functional distinct compartments important for epigenetic gene expression of a subgroup of genes. We postulate that distinct DNA loops may exist: large chromosomal loops, which bring specific chromosome regions to the nuclear periphery; small DNA loops that allow interaction with different components of PERF; and chromatin loops associated to relocation from telomeric clusters of active genes into a perinuclear expression site. In addition, we provide, for the first time, a detailed insight into the underlying chromatin modifications that are necessary to overcome the precise monoallelic counting mechanism of var genes. Our PfSir2 mutant analysis demonstrates that the removal of the H3K9me3 and replacement by acetylation at the 5’UTR is a key step towards unlocking the default silent state of var genes. It will be interesting to investigate if other protozoan pathogens that rely on subtelomeric gene families for immune escape (Borst, 2003), may use this type of epigenetic silencing mechanism. Insight into the complexity of

99

control of gene families involved in malaria pathogenesis is vital for understanding the disease and will provide a major step towards controlling it.

Materials and Methods Parasites

P. falciparum blood stage parasites were cultivated in the conditions previously described (Nunes et al., 2007). 3D7ΔSir2 parasites were cultivated in the presence of Pyrimethamine 2 µM. Selection of FCR3 parasites for CSA binding was performed according to (Scherf et al., 1998).

Antibodies

Antibodies against histone modifications were purchased from Upstate for H3K9me3, H3K9ac and H4K20me3 (07-442, 07-352 and 07-463, respectively) and from Abcam for H3K4me3 (ab8580). To obtain polyclonal antibody serum against PfKMT1 we used GenScript Corporation standard protocols for immunizing a rabbit with a synthetic peptide (N-CASRDIQPNEPLKYH-C).

Custom P. falciparum high-resolution tilled array design

The design of the array was done with the assistance of NimbleGen Systems Inc. We downloaded the P. falciparum genome sequence from The Sanger Center website and subjected it to several iterative rounds of probe selection (see Supplemental experimental procedures).

ChIP-chip analysis

ChIP assays were carried out as previously described (Lopez-Rubio et al., 2007) using ring- stage 3D7 parasites. These parasites express several var genes, however, at low percentage in ‘unselected’ bulk cultured 3D7.

We subjected precipitated DNA and DNA from parasite extracts (input) recovered after reverse cross-linking to amplification using Sigma GenomePlex WGA kit. After amplification, the immunoprecipitated DNA was tested for enrichment of control loci by qPCR (Realplex4 EpgradientS thermalcycler, Eppendorf) and co-hybridized to the tiling array

100

with input DNA. DNA was labelled, NimbleGen arrays were hybridized and the data were extracted according to standard operating procedures by NimbleGen Systems Inc. At least two biological replicates were performed for each condition and two technical replicates for each biological replicate. Detailed procedures and quantitative real-time PCR validation of ChIP-chip data are described in Supplemental experimental procedures.

Immunofluorescence and FISH

Immunofluorescence for PfKMT1 was conducted on air-dried IRBCs fixed with 4% paraformaldehyde (PFA), as described in (Nunes et al., 2007). PfKMT1 antibodies (1:500 dilution) were incubated for 1 hour and secondary antibodies (Alexa-Fluor-488-conjugated anti-rabbit highly cross-absorbed, 1:500 dilution) were incubated for 30 minutes at room temperature.

For FISH, IRBCs were lysed with saponine and the released parasites fixed in suspension with 4% PFA. Parasites were then deposited on microscope slides and subjected to DNA FISH in the conditions previously described (Mancio-Silva et al., 2008). For combined RNA/DNA FISH, parasites were deposited on slides, pretreated with 0,1% Triton X100 for 5 minutes and hybridized first with var probes at 50°C for 16 hours. The slides were washed three times in 2xSSC at 50°C, denaturated at 80°C and hybridized with the TARE6 probe at 37°C for16 hours. Finally, the slides were washed as for DNA FISH.

TARE6 and var2CSA probes were obtained as described before (Ralph et al., 2005). All other FISH probes were PCR amplified from genomic DNA using the primers listed on Table S2. Images were taken using a Nikon Eclipse 80i microscope with a CoolSnap HQ2 camera (Photometrics). NIS Elements 3.0 software (Nikon) was used for acquisition and ImageJ (http://rsbweb.nih.gov/ij/) for composition. Chromatic aberration was corrected in all images by aligning the red, the green and the blue channel signals from 0.2µm TetraSpeck microspheres.

101

Acknowledgments

We would like to thank to A. Cowman for providing the 3D7ΔSir2 strain, M. Nunes for the FCR3-CSA selected parasites, D. Roos for help with annotation files for genome wide analysis, E. Bischoff for assistance with microarray design, and PlasmoDB for the invaluable malaria data base support. We are also grateful to B. Arcangioli, P. Navarro and L. Riviere for critical reading of the manuscript, and to A. Cordeiro da Silva and F.F.U.P., Portugal for their support. This work was finance by BioMAlPar (contract No: LSPH-CT-2004-503578). J.J.L.R. has financial support from the Human Frontier Science Program and L.M.S. from the Portuguese Foundation for Science and Technology. A.S. is supported by a grant ANR Microbiologie (ANR-06-MIME-026-01).

References

Alfredsson-Timmins, J., Henningson, F., and Bjerling, P. (2007). The Clr4 methyltransferase determines the subnuclear localization of the mating-type region in fission yeast. Journal of cell science 120, 1935-1943.

Allis, C.D., Berger, S.L., Cote, J., Dent, S., Jenuwien, T., Kouzarides, T., Pillus, L., Reinberg, D., Shi, Y., Shiekhattar, R., et al. (2007). New nomenclature for chromatin-modifying enzymes. Cell 131, 633- 636.

Bannister, A.J., Zegerman, P., Partridge, J.F., Miska, E.A., Thomas, J.O., Allshire, R.C., and Kouzarides, T. (2001). Selective recognition of methylated lysine 9 on histone H3 by the HP1 chromo domain. Nature 410, 120-124.

Barry, A.E., Leliwa-Sytek, A., Tavul, L., Imrie, H., Migot-Nabias, F., Brown, S.M., McVean, G.A., and Day, K.P. (2007). Population genomics of the immune evasion (var) genes of Plasmodium falciparum. PLoS pathogens 3, e34.

Blythe, J.E., Yam, X.Y., Kuss, C., Bozdech, Z., Holder, A.A., Marsh, K., Langhorne, J., and Preiser, P.R. (2008). Plasmodium falciparum STEVOR proteins are highly expressed in patient isolates and located in the surface membranes of infected red blood cells and the apical tips of merozoites. Infection and immunity 76, 3329-3336.

Borst, P. (2003). Mechanisms of antigenic variation: an overview. In Antigenic variation, A. Craig, and A. Scherf, eds. (London: Academic Press), pp. 1-16.

Chen, F., Mackey, A.J., Stoeckert, C.J., Jr., and Roos, D.S. (2006). OrthoMCL-DB: querying a comprehensive multi-species collection of ortholog groups. Nucleic Acids Res 34, D363-368.

Chookajorn, T., Dzikowski, R., Frank, M., Li, F., Jiwani, A.Z., Hartl, D.L., and Deitsch, K.W. (2007). Epigenetic memory at malaria virulence genes. Proceedings of the National Academy of Sciences of the United States of America 104, 899-902.

Cortes, A., Carret, C., Kaneko, O., Lim, B.Y., Ivens, A., and Holder, A.A. (2007). Epigenetic Silencing of Plasmodium falciparum Genes Linked to Erythrocyte Invasion. PLoS pathogens 3, e107.

Cui, L., Fan, Q., Cui, L., and Miao, J. (2008). Histone lysine methyltransferases and demethylases in Plasmodium falciparum. International journal for parasitology 38, 1083-1097.

Cui, L., Miao, J., Furuya, T., Li, X., Su, X.Z., and Cui, L. (2007). PfGCN5-mediated histone H3 acetylation plays a key role in gene expression in Plasmodium falciparum. Eukaryot Cell 6, 1219-1227.

102

De Silva, E.K., Gehrke, A.R., Olszewski, K., Leon, I., Chahal, J.S., Bulyk, M.L., and Llinas, M. (2008). Specific DNA-binding by apicomplexan AP2 transcription factors. Proceedings of the National Academy of Sciences of the United States of America 105, 8393-8398.

Duraisingh, M.T., Voss, T.S., Marty, A.J., Duffy, M.F., Good, R.T., Thompson, J.K., Freitas-Junior, L.H., Scherf, A., Crabb, B.S., and Cowman, A.F. (2005). Heterochromatin silencing and locus repositioning linked to regulation of virulence genes in Plasmodium falciparum. Cell 121, 13-24.

Dzikowski, R., Frank, M., and Deitsch, K. (2006). Mutually exclusive expression of virulence genes by malaria parasites is regulated independently of antigen production. PLoS pathogens 2, e22.

Fernandez, V., Hommel, M., Chen, Q.J., Hagblom, P., and Wahlgren, M. (1999). Small, clonally variant antigens expressed on the surface of the Plasmodium falciparum-infected erythrocyte are encoded by the rif gene family and are the target of human immune responses. Journal of Experimental Medicine 190, 1393-1403.

Figueiredo, L.M., Freitas-Junior, L.H., Bottius, E., Olivo-Marin, J.C., and Scherf, A. (2002). A central role for Plasmodium falciparum subtelomeric regions in spatial positioning and telomere length regulation. The EMBO journal 21, 815-824.

Figueiredo, L.M., Pirrit, L.A., and Scherf, A. (2000). Genomic organisation and chromatin structure of Plasmodium falciparum chromosome ends. Molecular & Biochemical Parasitology 106, 169-174.

Freitas-Junior, L.H., Bottius, E., Pirrit, L.A., Deitsch, K.W., Scheidig, C., Guinet, F., Nehrbass, U., Wellems, T.E., and Scherf, A. (2000). Frequent ectopic recombination of virulence factor genes in telomeric chromosome clusters of P. falciparum. Nature 407, 1018-1022.

Freitas-Junior, L.H., Hernandez-Rivas, R., Ralph, S.A., Montiel-Condado, D., Ruvalcaba-Salazar, O.K., Rojas-Meza, A.P., Mancio-Silva, L., Leal-Silvestre, R.J., Gontijo, A.M., Shorte, S., and Scherf, A. (2005). Telomeric heterochromatin propagation and histone acetylation control mutually exclusive expression of antigenic variation genes in malaria parasites. Cell 121, 25-36.

Fried, M., and Duffy, P. (1996). Adherence of Plasmodium falciparum to chondroitin sulfate A in the human placenta. Science (New York, N.Y 272, 1502-1504.

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. (2002). Genome sequence of the human malaria parasite Plasmodium falciparum. Nature 419, 498-511.

Grewal, S.I., and Jia, S. (2007). Heterochromatin revisited. Nature reviews 8, 35-46.

Hall, N., Pain, A., Berriman, M., Churcher, C., Harris, B., Harris, D., Mungall, K., Bowman, S., Atkin, R., Baker, S., et al. (2002). Sequence of Plasmodium falciparum chromosomes 1, 3-9 and 13. Nature 419, 527-531.

Hediger, F., Neumann, F.R., Van Houwe, G., Dubrana, K., and Gasser, S.M. (2002). Live imaging of telomeres: yKu and Sir proteins define redundant telomere-anchoring pathways in yeast. Curr Biol 12, 2076-2089.

Jensen, A.T., Magistrado, P., Sharp, S., Joergensen, L., Lavstsen, T., Chiucchiuini, A., Salanti, A., Vestergaard, L.S., Lusingu, J.P., Hermsen, R., et al. (2004). Plasmodium falciparum associated with severe childhood malaria preferentially expresses PfEMP1 encoded by group A var genes. The Journal of experimental medicine 199, 1179-1190.

Kelly, J.M., McRobert, L., and Baker, D.A. (2006). Evidence on the chromosomal location of centromeric DNA in Plasmodium falciparum from etoposide-mediated -II cleavage. Proceedings of the National Academy of Sciences of the United States of America 103, 6706-6711.

Kyes, S.A., Kraemer, S.M., and Smith, J.D. (2007). Antigenic variation in Plasmodium falciparum: gene organization and regulation of the var multigene family. Eukaryot Cell 6, 1511-1520.

Kyes, S.A., Rowe, J.A., Kriek, N., and Newbold, C.I. (1999). Rifins: A second family of clonally variant proteins expressed on the surface of red cells infected with Plasmodium falciparum. Proceedings of the National Academy of Sciences of the United States of America 96, 9333-9338.

103

Lavazec, C., Sanyal, S., and Templeton, T.J. (2007). Expression switching in the stevor and Pfmc-2TM superfamilies in Plasmodium falciparum. Mol Microbiol 64, 1621-1634.

Lavstsen, T., Salanti, A., Jensen, A.T., Arnot, D.E., and Theander, T.G. (2003). Sub-grouping of Plasmodium falciparum 3D7 var genes based on sequence analysis of coding and non-coding regions. Malaria journal 2, 27.

Le Roch, K.G., Zhou, Y., Blair, P.L., Grainger, M., Moch, J.K., Haynes, J.D., De La Vega, P., Holder, A.A., Batalov, S., Carucci, D.J., and Winzeler, E.A. (2003). Discovery of gene function by expression profiling of the malaria parasite life cycle. Science (New York, N.Y 301, 1503-1508.

Leech, J.H., Barnwell, J.W., Miller, L.H., and Howard, R.J. (1984). Identification of a strain-specific malarial antigen exposed on the surface of Plasmodium falciparum-infected erythrocytes. The Journal of experimental medicine 159, 1567-1575.

Lopez-Rubio, J.J., Gontijo, A.M., Nunes, M.C., Issar, N., Hernandez Rivas, R., and Scherf, A. (2007). 5' flanking region of var genes nucleate histone modification patterns linked to phenotypic inheritance of virulence traits in malaria parasites. Mol Microbiol 66, 1296-1305.

Lunyak, V.V. (2008). Boundaries. Boundaries...Boundaries??? Current opinion in cell biology 20, 281- 287.

Mancio-Silva, L., Rojas-Meza, A.P., Vargas, M., Scherf, A., and Hernandez-Rivas, R. (2008). Differential association of Orc1 and Sir2 proteins to telomeric domains in Plasmodium falciparum. Journal of cell science 121, 2046-2053.

Marty, A.J., Thompson, J.K., Duffy, M.F., Voss, T.S., Cowman, A.F., and Crabb, B.S. (2006). Evidence that Plasmodium falciparum chromosome end clusters are cross-linked by protein and are the sites of both virulence gene silencing and activation. Mol Microbiol 62, 72-83.

Merrick, C.J., and Duraisingh, M.T. (2007). Plasmodium Falciparum Sir2: an Unusual Sirtuin with Dual Histone Deacetylase and Adp-Ribosyltransferase Activity. Eukaryot Cell.

Miao, J., Fan, Q., Cui, L., and Li, J. (2006). The malaria parasite Plasmodium falciparum histones: Organization, expression, and acetylation. Gene 369, 53-65.

Michishita, E., McCord, R.A., Berber, E., Kioi, M., Padilla-Nash, H., Damian, M., Cheung, P., Kusumoto, R., Kawahara, T.L., Barrett, J.C., et al. (2008). SIRT6 is a histone H3 lysine 9 deacetylase that modulates telomeric chromatin. Nature 452, 492-496.

Mourier, T., Carret, C., Kyes, S., Christodoulou, Z., Gardner, P.P., Jeffares, D.C., Pinches, R., Barrell, B., Berriman, M., Griffiths-Jones, S., et al. (2008). Genome-wide discovery and verification of novel structured RNAs in Plasmodium falciparum. Genome research 18, 281-292.

Nakayama, J., Rice, J.C., Strahl, B.D., Allis, C.D., and Grewal, S.I. (2001). Role of histone H3 lysine 9 methylation in epigenetic control of heterochromatin assembly. Science (New York, N.Y 292, 110-113.

Nunes, M.C., Goldring, J.P., Doerig, C., and Scherf, A. (2007). A novel protein kinase family in Plasmodium falciparum is differentially transcribed and secreted to various cellular compartments of the host cell. Mol Microbiol 63, 391-403.

Ralph, S.A., Scheidig-Benatar, C., and Scherf, A. (2005). Antigenic variation in Plasmodium falciparum is associated with movement of var loci between subnuclear locations. Proceedings of the National Academy of Sciences of the United States of America 102, 5414-5419.

Scherf, A., Hernandez-Rivas, R., Buffet, P., Bottius, E., Benatar, C., Pouvelle, B., Gysin, J., and Lanzer, M. (1998). Antigenic variation in malaria: in situ switching, relaxed and mutually exclusive transcription of var genes during intra-erythrocytic development in Plasmodium falciparum. The EMBO journal 17, 5418-5426.

Scherf, A., Lopez-Rubio, J.J., and Riviere, L. (2008a). Antigenic Variation in Plasmodium falciparum. Annu Rev Microbiol 62, 445-470.

Scherf, A., Riviere, L., and Lopez-Rubio, J.J. (2008b). SnapShot: var gene expression in the malaria parasite. Cell 134, 190.

104

Shankaranarayana, G.D., Motamedi, M.R., Moazed, D., and Grewal, S.I. (2003). Sir2 regulates histone H3 lysine 9 methylation and heterochromatin assembly in fission yeast. Curr Biol 13, 1240-1246.

Snow, R.W., Guerra, C.A., Noor, A.M., Myint, H.Y., and Hay, S.I. (2005). The global distribution of clinical episodes of Plasmodium falciparum malaria. Nature 434, 214-217.

Stubbs, J., Simpson, K.M., Triglia, T., Plouffe, D., Tonkin, C.J., Duraisingh, M.T., Maier, A.G., Winzeler, E.A., and Cowman, A.F. (2005). Molecular mechanism for switching of P. falciparum invasion pathways into human erythrocytes. Science (New York, N.Y 309, 1384-1387.

Taddei, A., Van Houwe, G., Hediger, F., Kalck, V., Cubizolles, F., Schober, H., and Gasser, S.M. (2006). Nuclear pore association confers optimal expression levels for an inducible yeast gene. Nature 441, 774-778.

Voss, T.S., Healer, J., Marty, A.J., Duffy, M.F., Thompson, J.K., Beeson, J.G., Reeder, J.C., Crabb, B.S., and Cowman, A.F. (2006). A var gene promoter controls allelic exclusion of virulence genes in Plasmodium falciparum malaria. Nature 439, 1004-1008.

Yamada, T., Fischle, W., Sugiyama, T., Allis, C.D., and Grewal, S.I. (2005). The nucleation and maintenance of heterochromatin by a histone deacetylase in fission yeast. Molecular cell 20, 173-185.

105

Supplemental materials and methods

Custom P. falciparum high resolution tilled array design. Probes selection: for intergenic and coding regions an initial pool of probes was selected according to the following criteria: approximately isothermal about 76ºC and probe length between 45-85 bases. The pool was then reduced in size by computing 13-mer frequencies for each probe and filtering out probes that had frequencies higher than 100. The number of genome matches was determined for each of the remaining probes and the final probes were selected from this reduced pool with the criterion that the final probes be unique within the genome. A final interval of about 100 bases in alternating strands for coding regions up to 1.5 Mb; about 300 bases for coding regions above 1.5 Mb; and about 50 bases for intergenic regions. For the telomeric regions of chromosome 12, tiled at interval of 50 bases in alternating strands and no masking or frequency filtering neither uniqueness tests were performed. Taking advantage of the small size of the P. falciparum genome, each chip included two genome blocks, which generated two technical replicates.

ChIP-chip analysis. DNA was labelled using random primers coupled to a fluorochrome and hybridized according to NimbleGen Systems procedures. At least two biological replicates were performed for each condition and two technical replicates for each biological replicate. Each feature on the array has a corresponding scaled log2-ratio. This is the ratio of the input signals for the experimental and test samples that were co-hybridized to the array. The log2- ratio is computed and scaled to center the ratio data around zero. Scaling is performed by subtracting the bi-weight mean for the log2-ratio values for all features on the array from each log2-ratio value. Peak data are generated from the scaled log2-ratio data. NimbleScan detects peaks by searching for 4 or more probes whose signals are above the specified cutoff values, ranging from 90% to 15%, using a 500 bp sliding window. The cutoff values are a percentage of a hypothetical maximum, which is the mean + 6 (standard deviation). The ratio data is randomized 20 times to evaluate the probability of “false positives.” Each peak is then assigned a false discovery rate (FDR) score based on the randomization. Peaks with FDR score between 0 and 0.2 and present in the four replicates are shown.

Quantitative real-time PCR validation of ChIP-chip data. The ChIP-chip results were validated using qRT-PCR. Chromatin was isolated and immunoprecipitated as described above using the different antibodies. ChIP DNA and corresponding input DNA were assayed by qRT-PCR (Realplex4 EpgradientS thermal cycler from Eppendorf) as described previously (Lopez-Rubio et al., 2007) for eight single-locus. We found enrichment of the modified histones in regions identified as peaks in the ChIP-chip experiments and not significant enrichment in regions in which ChIP-chip assays did not find any peak.

106

Supplemental figures

107

Figure S1. Clustering of var group A, B and C. Distribution of the number of FISH signals per nuclei using the exon2 probes to each var group. (A) The 10 subtelomeric var genes group A form on average 3.14 ± 0.08 clusters (n= 288). (B) The 22 subtelomeric genes of group B form on average 3.85 ± 0.08 clusters (n= 336). (C) The 13 central genes of group C form on average 3.22 ± 0.09 clusters (n= 208). The errors are indicated for 95% confidence intervals. The standard deviations for groups A, B and C are 1.28, 1.40 and 1.33, respectively. Images from multiple FISH experiments were combined and counted.

108

Figure S2. Nuclear pore and quantitative analysis of nuclear position. (A) The nuclear periphery was defined using an antibody to a putative nuclear pore protein in P. falciparum (PF14_0706). This hypothetical protein shows homology with the Nucleoporin NUP116 from Saccharomyces cerevisiae (34% identity and 50% similarity). Antibodies raised against synthetic peptides for this molecule react with parasite total extracts and in immunofluorescence assays show a punctuated ring pattern to some extent outside of the Dapi staining. (B) The diameter of the nucleus combined with the nuclear pore immunofluorescence signals was measured. (C) The area of the nuclear section was calculated and divided into three equal surfaces as illustrated. The FISH signals were then scored as localizing to zone 1, 2 and 3. For an average nucleus, the zone1 corresponds to a diameter of 2.06 – 1.67 µm and was defined as the peripheral zone; zone 2 corresponds to a diameter of 1.67 – 1.19 µm and zone 3 the internal zone, corresponds to a diameter of 1.19 µm. Measurements were performed using the NIS-Elements 3.0 imaging software (Nikon). The scoring was performed by direct optical observation and registered using the NIS-Elements 3.0 software. For each FISH probe images from two to three independent cultures were combined and analyzed. In the schematics, the Dapi staining is collared in grey and the three nuclear zones are indicated as punctuated lines. To allow a better correlation between immunofluorescence and FISH imaging we used the same sample of paraformaldeyde fixed parasites in suspension. Bars: 1µm.

109

Figure S3. ChIP-chip analysis of H3K9me3 and H3K9ac at non-coding chromosome 12 right arm in PfSir2 knock-out parasites. Data are presented as in Figure 4.

110

RESULTS VI

Nucleolar organization of dispersed rDNA in Plasmodium falciparum asexual stages

Liliana Mancio-Silva, Christine Scheidig-Benatar and Artur Scherf

Institut Pasteur, Unité de Biologie des Interactions Hôte-Parasite, CNRS URA2581, 25 Rue du Dr Roux, 75724 Paris, France

Correspondence: Artur Scherf, tel: + 33 (01) 4568-8616, fax: + 33 (01) 4568-8348, e-mail: [email protected]

Manuscript in preparation.

111

112

Introduction In eukaryotic cells nucleus many functions are organized into sub-compartments. The most dominant function-based nuclear compartment is the nucleolus, the site of rDNA transcription by RNA Polymerase I (Pol I), rRNA maturation and assembly of ribosomes. Nucleoli vary greatly in appearance depending on type and growth rate of the cell. In yeast, the nucleolus appears as a single crescent-shaped structure juxtaposed to the nuclear envelope, occupying up to one third of the nucleus. In multicellular organisms, cells contain multiple nucleoli, adjacent to heterochromatin (Carmo-Fonseca et al., 2000; Taddei et al., 2004). In most eukaryotes, the ribosomal units are composed of 18S-5.8S-28S genes, identical in sequence and tandem arrayed in large numbers. In P. falciparum, the parasite responsible for human malaria, rDNA unit has a typical eukaryote organization but the family itself is the most unusual among the eukaryotic rDNA families. P. falciparum parasites contain few rDNA units, slightly different in sequence and distributed in different chromosomes (Langsley et al., 1983; Wellems et al., 1987). The malaria genome project identified five complete ribosomal units each located on separated chromosomes 1, 5, 7, 11 and 13, and two incomplete units (lacking 18S gene) on chromosome 8 (Gardner et al., 2002) (see Fig. S1). Moreover, the rRNA genes are differentially expressed over the course of the parasite developmental cycle (Waters et al., 1989). P. falciparum has a complex life cycle that involves the human and mosquito hosts and three major developmental stages: schizogony, gametogony and sporogony. Schizogony occurs in the human red blood cells and the dominant rRNA transcripts are from the units located on chromosomes 5 and 7, named A1 and A2 rRNA genes, respectively. In gametocyte stages, the predominant transcripts are from the unit on chromosome 1, called S1 rRNA gene. Sporogony occurrs in the mosquito and the dominant forms are from units located on chromosomes 11 and 13, named S2 type gene (Fang and McCutchan, 2002; Fang et al., 2004). It is not known whether the genes located on chromosome 8 are functional. Here, we used the P. falciparum rDNA multi-gene family as a model for understanding the importance of nuclear architecture in gene regulation in this parasite. By using different fluorescent imaging approaches we explored the determinants of nucleolar organization and position.

113

Results

P. falciparum nucleolus localizes at the nuclear periphery In previous work, we have identified for the first time the nucleolus in P. falciparum parasites (Figueiredo et al., 2005). IFA using anti-PfNop1 serum revealed a hat-like structure polarized towards one side of the nucleus, very similar to that described in yeast. In this study, we aimed to further characterize the nucleolus compartment in Plasmodium. In order to preserve the nucleolar integrity we improved our fixation protocol for IFA and FISH, by performing the fixation step(s) in free parasites in suspension. This protocol produced smaller nucleolus comparing to the classical protocol of air-dried infected red blood cells. Furthermore, we generated new antibodies against distinct nucleolar proteins: the highly conserved RNA Polymerase I (PFE0465c) and also to a putative nucleolar protein Nop5 (PF10_0085). These anti-peptide antibodies were validated on western blot analysis (not shown) and/or peptide competition assays (Fig. S2). On IFA, anti- PfRNA Polymerase I (PfPol I) and anti-PfNop5 antibodies recognize the same nuclear region previously identified by PfNop1 (Fig. 1A and 1B). As a complementary study for the localization of the nucleolus, we used IFA of PfPol I combined with RNA FISH of 18S rRNA. 18S probe recognizes a conserved region among the different P. falciparum rDNA units, we therefore considered it as a universal probe. rRNA transcripts showed a diffuse pattern in a crescent shape at the nuclear periphery that indeed colocalizes with the PfPol I staining (Fig. 1C). Together, the data from Figure 1 clearly confirmed that the region previously defined as the P. falciparum nucleolus corresponds to the site of rDNA transcription by PfPol I.

P. falciparum nucleolus is a dynamic structure In yeast, the nucleolus remains intact during mitosis, but in multicellular organisms, the nucleolus disassembles in metaphase and reassembles in late telophase (Carmo-Fonseca et al., 2000; Heun et al., 2001). In the P. falciparum 48-hour schizogonic cycle, DNA replication and mitotic division initiate at 30 hours (see Fig. 2A). To follow the P. falciparum nucleolus throughout the cycle, we collected parasites from a synchronized culture at three time points: ~12h (ring stage), ~30h (trophozoite stage) and ~40h (schizont stage). We examined by IFA the different nucleolar markers: PfPol I (Fig. 2B), PfNop1 (not shown) and PfNop5 (not shown). Early parasites stages (Fig. 2B left panel) form a condensed perinuclear nucleolus (as shown above). In trophozoite stage (Fig. 2B middle panel), the nucleolus become dispersed and disorganized throughout the

114

Fig. 1. Nucleolus of P. falciparum localizes at the nuclear periphery A and B. Double labelling IFA using rabbit anti-PfPol I (red) and rat anti-PfNop1 or anti-PfNop5 (green) antibodies, showing the localization of the nucleolus at the periphery of the nucleus, forming a hat-like structure. C. Combination of RNA FISH (using 18S universal probe that hybridizes to all rRNAs produced in the parasite, green) with IFA (using anti-PfPolI antibody, red) demonstrating that the region defined by the antibodies corresponds indeed to the nucleolus of the parasite. Parasites are on ring stage and the nuclear DNA is stained with Dapi (blue). Bars: 1µm. D. Structural organization of a Plasmodium rDNA unit. It consists of a ~9kb transcribed region coding for a 45S rRNA precursor. The first nucleotide (+1) of this transcript begins at the 5’ external transcribed spacer (5’ETS) followed by the 18S coding region, the internal transcribed spacer 1 (ITS1), the 5.8S coding region, the internal transcribed spacer 2 (ITS2), the 28S coding region, and finally the 3’ ETS. This precursor is processed to mature 18S, 5.8S and 28S rRNAs for ribosome biogenesis. The coding regions display a modest sequence similarity whereas the non-coding regions are considerable distinct among the members of rDNA multi-gene family. The position of the probes used in FISH and the primers used in real time PCR are indicated.

nucleoplasm and cytoplasm (not shown). In later multinucleated schizont stages (Fig. 2B right panel), we observed some reposition of the nucleolar components at the periphery of the newly formed nuclei. We obtained the same results using PfNop1 and PfNop5 antibodies. To investigate whether these structural changes in the nucleolus affect rRNA synthesis or vice- versa, we monitored rRNA gene transcription levels by quantitative real-time reverse transcriptase PCR analysis using specific pairs of primers for the ETS of four rRNA genes (A1,

115

A2, S1 and S2) (Fig. 1D). The ETS sequence is known to be rapidly removed from the pre-rRNA transcripts during rRNA processing and therefore can be used to measure on going rRNA transcription. Parasite total RNA was harvested in the same three time points as for IFA. We compared the transcription levels of each rRNA gene from ring and trophozoite stages to the transcription levels of schizont stage. We observed that genes known to be active during schizogony, A1 and A2 types, are preferentially transcribed on ring stages. In throphozoites and schizonts stages rRNA synthesis is suppressed (Fig. 2C). Thus, the expression profile of rRNA genes correlates with nucleolar structural changes during the cycle.

Fig. 2. A. Cell cycle for P. falciparum schizogonic stage parasites (adapted from (Leete and Rubin, 1996)). B. The nucleolus is a dynamic compartment. PfPol I IFA (green) throughout the blood stage cycle. The nucleolus looses the moon-shape (rings) and disassembles during the mitotic divisions (trophozoite and schizont stages). Nuclei were staining by Dapi in blue. C. Expression profile of the rRNA genes on rings, trophozoites and schizonts. The y-axis represents the fold change of rings and trophozoites relative to schizonts, for each ribosomal gene (A1, A2, S1 and S2).

116

P. falciparum nucleolus hosts active and silent rRNA genes It is generally assumed that nuclear position is dependent on the transcriptional state of the gene. We hypothesized that the same principal applies to ribosomal genes in P. falciparum. We studied the nuclear location of the active (A1 and A2) and the silent (S1 and S2) ribosomal genes in Plasmodium at ring stage. To discriminate between active and silent rRNA genes, we generated specific DNA FISH probes that recognize the upstream regions of each rRNA type (Fig. 1D). These probes were tested in combination with the anti-PfPol I serum on IFA-DNA FISH. We observed that active genes (Fig. 3A) and silent rRNA genes (Fig. 3A) locate within the region defined as the nucleolus. We concluded that the nucleolar position is important but not sufficient for transcriptional activation of members of this multi-gene family.

Fig. 3. A. Active and Silent rRNA genes localize in the nucleolus. IFA-DNA FISH analysis using anti-PfPol I antibody as nucleolar marker (red) and specific FISH probes for rRNA genes (green). Dapi, blue. B. There are two rRNA expression sites in the nucleolus. Two-colour DNA FISH of the two active rRNA genes (A1, red and A2, green). Dapi, blue.

P. falciparum nucleolus contains two expression sites We next investigated the location of the two active genes in schizogonic stages (A1 and A2). We used the upstream-specific FISH probes in two-colour DNA FISH analysis. On ring stage nuclei, >90% of nuclei the two specific probes do not colocalize (Fig. 3B). This result suggests the presence of two independent transcriptional factories within the nucleolus.

117

P. falciparum nucleolus excludes the telomeres and is cross-linked by proteins Chromosome ends of P. falciparum form four to seven telomeric clusters at the nuclear periphery. These clusters are sites enriched in silencing factors and are also important for parasite survival as it promotes gene conversion between members of subtelomeric virulence factor genes (Freitas- Junior et al., 2000; Freitas-Junior et al., 2005). Previous observations have shown that the telomeric clusters are tethered to one of nucleus (our unpublished data and (Marty et al., 2006)). We hypothesized that the nucleolus might locate in the other nuclear pole. To test this hypothesis, we carried out a two colour DNA FISH using the subtelomeric TARE 6 (or Rep20) probe for the telomeric clusters and the 18S universal probe for the rDNA genes. The majority but not all telomeric clusters foci do not colocalize with rDNA. Thus telomeric clusters and nucleolus appear to form discrete subcompartments in P. falciparum nuclei (Fig. 4A). To further examine the nature of the nucleolar compartment we treated unfixed parasites with Proteinase K prior to DNA FISH analysis. Untreated nuclei displayed a clustering at the nuclear periphery (Fig. 4A) whereas the Proteinase K treated showed dispersed rDNA foci (Fig. 2B). The telomeric clusters were also disrupted by the Proteinase K treatment, consistent with previous report (Marty et al., 2006). This result demonstrates that protein-protein and protein-DNA interactions are important for the stabilization of the P. falciparum nucleolus.

Fig. 4. Two-colour DNA FISH of telomeric clusters (TARE6 probe, green) and the 18S rDNA universal probe (red). A. In ring stage parasites, telomeric clusters are excluded from the nucleolus. B. The nucleolus is disrupted by Proteinase K. In the Proteinase K treated nuclei the telomeric clusters are dissociated (Marty et al. 2006) and also the individual units of rDNA are dispersed to multiple sites. Nuclei were detected by Dapi staining (blue). Bars: 1µm.

118

rDNA sequence is sufficient for nucleolar position The finding that active and silenced rRNA genes locate in the nucleolus raised the question of what determines the localization of the P. falciparum rRNA genes in the nucleolus. To address this question, we divided an rDNA unit in four fragments (A, B, C and D) (Fig. 5), cloned them in a plasmid and transfected into P. falciparum parasites. As negative control we transfected the parasites using plasmids with a non-related rDNA sequence (Histone H3 gene). Transfectant parasites were screened by PCR to confirm that the sequence of interest was not altered (not shown). We only succeed to obtain parasites carrying plasmids with rDNA fragment D, which corresponds to the most downstream fragment. We examined the nuclear position of the plasmids relative to the nucleolus by RNA/DNA FISH. RNA FISH of the ribosomal RNA with 28S probe was used to detect the nucleolus, after washing a second hybridization using a plasmid-specific probe to detect the episomes by DNA FISH. The parasites carrying control plasmids did not show association with the nucleolus whereas the parasites carrying plasmids containing rDNA D fragment showed colocalization in the nucleolus in most of the nuclei analyzed (Fig. 5). Taken together, these results show that P. falciparum rDNA sequence contains positional information that determines nucleolar localization.

Fig. 5. Position analysis of plasmids (green) relative to the nucleolus (red) on transfected parasites by RNA/DNA FISH. In the negative control, the plasmids with Histone H3 sequence (first row) do not associate to the nucleolus, whereas the plasmids containing a rDNA sequence (second row) colocalize with the nucleolus. The nuclei were detected with Dapi (blue).

119

Discussion The results presented here strongly support the idea of a functional compartmentalization of the P. falciparum nucleus. We have focused our work on the nucleolar compartment: composition, dynamics, relative localization to other sub-compartments, determinants for nucleolar position and also transcription gene regulation. Based on the presented data we propose a model for nuclear architecture organization of the ring stage P. falciparum nucleus (Fig. 6). In our model, the nucleolus consists of a single major compartment, similar to the yeast, it appears to occupy roughly one third of nucleus (Fig. 1). Nucleolus and telomeric clusters define distinct sub- compartments in the nucleus (Fig. 2); the telomeric clusters specialized on gene silencing (Freitas-Junior et al., 2005; Ralph et al., 2005) and the nucleolus dedicated to gene silencing and activation (Fig. 5). The different chromosomes containing the members of the unusual P. falciparum ribosomal multi-gene family must loop into the nucleolus, in the nuclear periphery. This unusual physical link must have significant implications on three-dimensional organization of the parasite genome. Our study provides the first evidence that the P. falciparum nucleolus is a dynamic structure. As in mammalian cells (Raska et al., 2006), in Plasmodium there is an apparent correlation between the cell cycle, the nucleolar structure and the rRNA synthesis (Fig. 3). Maximum levels of ongoing rRNA transcription were detected on proliferating ring stage, when the nucleolus appears a crescent shape at the nuclear periphery. As the parasite progress to S/M phases (trophozoite and schizont stages) the nucleolus is disassembled and the rRNA synthesis is decreased.

Fig. 6. Model for the nuclear organization of rDNA in P. falciparum. The nucleolus localizes at the nuclear periphery and is represented in a hat-like shape (dark grey) where are located all rDNA units existing in the P. falciparum genome (A1 and A2, green; S1 and S2, red; incomplete rDNA units, yellow). Activation and silencing co-exists in this the major sub-compartment. The clusters of chromosome ends (black ovals) also locate at the nuclear periphery (Freitas-Junior et al., 2000), opposing the nucleolus and form a silent compartment.

120

In mammalian cells, transcription of ribosomal genes is completely blocked in mitosis; this involves phosphorylation by cyclins that leads to inactivation of Pol I and consequently nucleolar disassemble; at the end of mitosis, rRNA synthesis re-starts and nucleoli reassemble (Boisvert et al., 2007). It is likely that similar mechanisms of cell-cycle regulation may dictate morphology and transcription in the P. falciparum nucleolus. One conserved feature of the rDNA organization in eukaryotes is the localization in heterochromatic regions in the genome (centromeric or subtelomeric positions) and in the nucleus. Interestingly, in P. falciparum all the dispersed rDNA units (except one) are in subtelomeric regions (Fig. S1) and the parasite nucleolus localizes at the nuclear periphery (Fig. 1), a heterochromatin-like region (Freitas-Junior et al., 2005). Heterochromatin seems to function as a general strategy to suppress recombination on rRNA genes (Pikaard and Pontes, 2007). In flies, mutants of proteins involved in heterochromatin formation, such histone H3K9 methyltransferase, heterochromatin protein 1 (HP1) and proteins of the RNAi pathway display disrupted nucleolus, dispersed rDNA and production of extrachromosomal circles of rDNA (Peng and Karpen, 2007). In budding yeast, Sir2 is also known to suppress recombination among rRNA genes (Gottlieb and Esposito, 1989; Kobayashi et al., 2004). In fact, some heterochromatin related proteins have been already localized in the Plasmodium nucleolus: PfSir2 (Freitas-Junior et al., 2005) and PfOrc1 (Mancio-Silva et al., 2008). It is noteworthy that, given the small number of rDNA copies per haploid genome of Plasmodium, avoid homologous recombination must be extremely important. This raises the question of the minimal rDNA genes required for viability. Knock out experiments conducted in P. berghei suggest that there is indeed a minimal number of two A-types and one S-type genes for complete development in both the vertebrate and mosquito hosts (McCutchan et al., 2004; van Spaendonk et al., 2001). Surprisingly, we observed that in P. falciparum rRNA active and silenced members co-exist in this heterochromatic-like region (Fig. 3). The association of rDNA units with the nucleolus irrespective of their transcriptional status has also been observed in mouse>human cell hybrids. In these hybrid cells, the human rRNA genes are silenced by a mechanism of nucleolar dominance. Despite of being silent, the human rRNA genes are found in the nucleoli, together with the active mouse rRNA genes (Sullivan et al., 2001). Together, this underscores the necessity of rDNA locate in a heterochromatic environment. Unlike most eukaryotes, P. falciparum ribosomal units are not tandem arrayed in large numbers, the seven members are dispersed in the genome. Nevertheless, we observed that rDNA units are clustered forming a single nucleolus (Fig. 1). Our results contradict previous studies performed on yeast cells with deleted chromosomal rDNA repeats and carrying ribosomal genes on plasmids

121

that showed multiple small mini-nuceloli localized in the perinuclear region (Oakes et al., 1998; Oakes et al., 2006). Thus, P. falciparum is the natural example that proves that tandem arrangement is not an absolute requirement for nucleolar clustering of rDNA. Although the rRNA genes are clustered in the same compartment, the two rRNA genes transcriptionally active in schizogony localized in discrete foci, suggesting the existence of two expression sites within the ribosomal factory (Fig. 4). These observations are to some extent in agree with α-amanitin experiments, in which Pol II and Pol III were inhibited and two sites transcription are visualized at the nuclear periphery in 40% of the nuclei (Moraes and Freitas- Junior, poster MPM2006). Moreover, we did not observed much colocolization between the silent rRNA genes, nor between silent and active rRNA genes (not shown), indicating that there is no apparent sub-compartments within the nucleolus and that the nucleolar structure is reasonably open. Importantly, we found that subnuclear localization of a multi-gene family in P. falciparum is sequence-dependent. To understand the nucleolar clustering of rRNA genes we have cloned different portions of rDNA and transfected into parasites. Initially, we have tested two approaches: 1) introduce plasmids containing rDNA portions, and 2) integrate the rDNA fragments into a chromosomal region that has not a ribosomal unit. To obtain the chromosomal integration we used the bacterial integrase mediated transfection system as described elsewhere (Nkrumah et al., 2006). Unfortunately, we failed to obtain integration of the rDNA fragments. The fact that the fragments were too large (~3Kb) and/or interference with the original chromosomal organization may justify this unsuccess. Regarding the episomal strategy (Fig. 5), it should be noted that the plasmids with rDNA sequence did not associate with the nucleolus in all the nuclei. This may be explained by a possible competition with Pol II transcription of the selectable drug that preventes localization into the nucleolus. Further analysis of nuclear position of the plasmids in the absence of drug will be necessary. These observations raise the question of what drives plasmids into the nucleolus. The presence of the rDNA itself may attract nucleolar proteins responsible for binding and targeting into the nucleolus. An alternative, although not mutually exclusive, may involve epigenetic events. Our recent experiments (Lopez-Rubio, Mancio-Silva and Scherf, unpublished) showed that position of a gene at the nuclear periphery is linked with the presence of a specific histone modification (H3K9 methylation). Thus, it is possible that a specific histone modification, present in the rDNA determines its location. In summary, our data favour a model in which nucleolar position is dependent on sequence composition and independent of transcriptional status. Further studies will be necessary to elucidate the factors that determine the differential transcription of rRNA multi-gene family in

122

this protozoan pathogen. The methodology of transfecting portions of rDNA sequence as we performed in this study may be used to identify the proteins involved in structural integrity, in recognition and targeting to the nucleolus. It may also be used to further define the minimal region required for targeting into the nucleolus or any other subcompartment in the parasite nucleus.

Materials and Methods Parasites P. falciparum blood stage parasites from the 3D7 strain were cultured as detailed described on (Nunes et al., 2007). Parasites were synchronized by consecutive Plasmagel and Sorbitol treatments, with an estimated window of 6 hours. Ring stages were collected approximately 12 hours post-invasion (p.i.), trophozoites stages 30 hours p.i. and schizonts stages 40 hours p.i. within the same 48 hour cycle.

Antibodies A rabbit anti-PfPol I antibody was prepared by Sigma-Genosys. Rabbits were immunized with a synthetic peptide coupled with KLH/MBS carrier, N-CYEEGRGYDIDEQND-C. Antibodies against PfNop1 and PfNop5 were obtained from Eurogentec, by immunizing rats with two synthetic peptides coupled to KLH/MBS carrier. The sequences are as follows: PfNop1, N- CKPKEKLTLEPFHRHH-C and N-TDSFRGGSGNFKRNSC-C; PfNop5, N- CFNKRKSDFRFKREAD-C and N-KFQNAFDETNKLMESC-C.

Immunofluorescence and FISH Immunofluorescence assays (IFA) were performed on synchronized parasites fixed in suspension as described previously (Mancio-Silva et al., 2008). Up to fixation with paraformaldeyde 4%, all steps were performed at 37°C. Primary antibodies were absorbed with lysed non-infected red blood cells. The final antibody dilutions were: rabbit anti-PfPol I 1:500, rat anti-PfNop1 1:50, anti-PfNop5 1:50, Alexa-Fluor-488-conjugated goat anti-rabbit highly cross-absorbed 1:500, Alexa-Fluor-568-conjugated goat anti-rabbit highly cross-absorbed 1:500 and Fluorescein- conjugated goat anti-rat 1:500. For peptide competition experiments increasing concentrations of specific and non-specific peptides were incubated with specific and non-specific sera overnight at 4°C and examined by

123

IFA. IFA was conducted on air-dried infected red blood cells fixed with 4% paraformaldeyde. Primary and secondary antibodies were incubated for 30 minutes at room temperature in PBS- BSA 3%. Intermediate and final washes were made in PBS three times for 5 minutes. DNA FISH and combined IFA-DNA FISH on parasites fixed in suspension were performed as described previously (Mancio-Silva et al., 2008). For IFA-RNA FISH, parasites were subjected to IFA, post-fixed in 4% paraformaldeyde, deposited on slides, hybridize at 37°C with denatured probes and washed three times with 2xSSC 10 minutes. For RNA/DNA FISH, parasites were deposited on slides, pretreated with 0,1% Triton X100 (5 minutes) and hybridized first with 28S rRNA probe at 37°C for 16 hours. The slides were washed three times in 2xSSC at 50°C, denaturated at 80°C and hybridized with the plasmid probe at 37°C for 16 hours, and washed as for DNA FISH. Proteinase K treatment was performed prior to fixation of the parasites using 0.5 µg/µl Proteinase K (Eurobio) for 1 minute at room temperature. TARE6 probe was obtained as described elsewhere (Figueiredo et al., 2000). All other FISH probes were PCR amplified from 3D7 genomic DNA using the primers listed on Table S1. Fluorescent samples were analyzed using a Nikon Eclipse 80i microscope. NIS Elements 3.0 software (Nikon) was used for images acquisition and ImageJ (http://rsbweb.nih.gov/ij/) for composition.

RNA extraction and real time PCR analysis Total RNA was isolated from synchronized cultures using the Trizol method as described in (Kyes et al., 2000). The temperature was maintained at 37°C in all the operations. Extracted RNA was treated with DNA-free kit (Ambion) and reverse transcribed with Super-scriptII Rnase H Reverse Transcriptase (Invitrogen) using random hexameres (Invitrogen). Quantitative real-time PCR was done in a Realplex4 EpgradientS thermal cycler (Eppendorf) using SYBR Green PCR master mix (Applied Biosystems) and standard settings (Eppendorf). PCR was performed in duplicates. Water, no-reverse transcriptase controls and standard curves with serial dilutions of genomic DNA were included in each run. Specificity of amplification was examined by performing a melting-curve analysis at the end of each PCR program and also by showing the presence of a single band on ethidium bromide-stained gel. The pairs of primers used are listed in Table S1. The fold change for each sample was calculated using the 2-∆∆Ct method (Livak and Schmittgen, 2001). Each rRNA gene on each sample was normalized to the housekeeping gene calmodulin (PF14_0323).

124

Plasmid construct and parasite transfection rDNA ~3.2 Kb fragment A, B, C and D were amplified from 3D7 genomic DNA using Expand High Fidelity PCR System (Roche) with primers containing restriction sites for the enzymes BamHI and ApaI and cloned into the backbone plasmid vector pLN-ENR-GFP (Nkrumah et al., 2006). 3D7 ring stage parasites were transfected by electroporation according to the Functional Genomics of Malaria Parasites Workshop Manual, 2007. Parasites were selected using 2.5 µg/ml of Blasticidin (Invitrogen). The presence of the plasmid and the integrity of the fragment were confirmed by PCR on transfected parasites DNA, using different combinations of primers for the vector and the cloned fragment. All the primers referred are listed on Table S1.

Acknowledgments We would like to thank to J.J. Lopez-Rubio for helpful comments and discussions, to M. Nunes for help with the real time reverse transcriptase PCR analysis and parasite transfection, to L. Riviere for providing the pLN-H3 plasmid, to A. Cordeiro-da-Silva and Faculdade de Farmácia da Universidade do Porto, Portugal for their support. This work was supported by the European Commission (BioMalPar) with contract number: LSPH-CT-2004−503578. L.M.S. has financial support from the GABBA PhD program and Fundação para a Ciência e Tecnologia, Portugal. A.S. is supported by a grant from the "Fonds dédié: Combattre les Maladies Parasitaires" Sanofi Aventis - Ministère de la Recherche and an ANR grant (ANR-06-MIME-026-01).

References Boisvert, F.M., van Koningsbruggen, S., Navascues, J., and Lamond, A.I. (2007). The multifunctional nucleolus. Nat Rev Mol Cell Biol 8, 574-585.

Carmo-Fonseca, M., Mendes-Soares, L., and Campos, I. (2000). To be or not to be in the nucleolus. Nature cell biology 2, E107-112.

Fang, J., and McCutchan, T.F. (2002). Thermoregulation in a parasite's life cycle. Nature 418, 742.

Fang, J., Sullivan, M., and McCutchan, T.F. (2004). The effects of glucose concentration on the reciprocal regulation of rRNA promoters in Plasmodium falciparum. The Journal of biological chemistry 279, 720- 725.

Figueiredo, L.M., Pirrit, L.A., and Scherf, A. (2000). Genomic organisation and chromatin structure of Plasmodium falciparum chromosome ends. Molecular & Biochemical Parasitology 106, 169-174.

125

Figueiredo, L.M., Rocha, E.P., Mancio-Silva, L., Prevost, C., Hernandez-Verdun, D., and Scherf, A. (2005). The unusually large Plasmodium telomerase reverse-transcriptase localizes in a discrete compartment associated with the nucleolus. Nucleic acids research 33, 1111-1122.

Freitas-Junior, L.H., Bottius, E., Pirrit, L.A., Deitsch, K.W., Scheidig, C., Guinet, F., Nehrbass, U., Wellems, T.E., and Scherf, A. (2000). Frequent ectopic recombination of virulence factor genes in telomeric chromosome clusters of P. falciparum. Nature 407, 1018-1022.

Freitas-Junior, L.H., Hernandez-Rivas, R., Ralph, S.A., Montiel-Condado, D., Ruvalcaba-Salazar, O.K., Rojas-Meza, A.P., Mancio-Silva, L., Leal-Silvestre, R.J., Gontijo, A.M., Shorte, S., et al. (2005). Telomeric heterochromatin propagation and histone acetylation control mutually exclusive expression of antigenic variation genes in malaria parasites. Cell 121, 25-36.

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. (2002). Genome sequence of the human malaria parasite Plasmodium falciparum. Nature 419, 498-511.

Gottlieb, S., and Esposito, R.E. (1989). A new role for a yeast transcriptional silencer gene, SIR2, in regulation of recombination in ribosomal DNA. Cell 56, 771-776.

Heun, P., Taddei, A., and Gasser, S.M. (2001). From snapshots to moving pictures: new perspectives on nuclear organization. Trends in cell biology 11, 519-525.

Kobayashi, T., Horiuchi, T., Tongaonkar, P., Vu, L., and Nomura, M. (2004). SIR2 regulates recombination between different rDNA repeats, but not recombination within individual rRNA genes in yeast. Cell 117, 441-453.

Kyes, S., Pinches, R., and Newbold, C. (2000). A simple RNA analysis method shows var and rif multigene family expression patterns in Plasmodium falciparum. Molecular & Biochemical Parasitology 105, 311-315.

Langsley, G., Hyde, J.E., Goman, M., and Scaife, J.G. (1983). Cloning and characterisation of the rRNA genes from the human malaria parasite Plasmodium falciparum. Nucleic acids research 11, 8703-8717.

Leete, T.H., and Rubin, H. (1996). Malaria and the cell cycle. Parasitology today (Personal ed 12, 442-444.

Livak, K.J., and Schmittgen, T.D. (2001). Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods (San Diego, Calif 25, 402-408.

Mancio-Silva, L., Rojas-Meza, A.P., Vargas, M., Scherf, A., and Hernandez-Rivas, R. (2008). Differential association of Orc1 and Sir2 proteins to telomeric domains in Plasmodium falciparum. Journal of cell science 121, 2046-2053.

Marty, A.J., Thompson, J.K., Duffy, M.F., Voss, T.S., Cowman, A.F., and Crabb, B.S. (2006). Evidence that Plasmodium falciparum chromosome end clusters are cross-linked by protein and are the sites of both virulence gene silencing and activation. Mol Microbiol 62, 72-83.

McCutchan, T.F., van Spaendonk, R.M., Janse, C.J., Fang, J., and Waters, A.P. (2004). Regulation of rRNA transcription and processing during the Plasmodium life cycle. Malaria Parasites: Genomes and Molecular Biology, ed AP Waters and CJ Janse, 229-247.

Nkrumah, L.J., Muhle, R.A., Moura, P.A., Ghosh, P., Hatfull, G.F., Jacobs, W.R., Jr., and Fidock, D.A. (2006). Efficient site-specific integration in Plasmodium falciparum chromosomes mediated by mycobacteriophage Bxb1 integrase. Nature methods 3, 615-621.

126

Nunes, M.C., Goldring, J.P., Doerig, C., and Scherf, A. (2007). A novel protein kinase family in Plasmodium falciparum is differentially transcribed and secreted to various cellular compartments of the host cell. Mol Microbiol 63, 391-403.

Oakes, M., Aris, J.P., Brockenbrough, J.S., Wai, H., Vu, L., and Nomura, M. (1998). Mutational analysis of the structure and localization of the nucleolus in the yeast Saccharomyces cerevisiae. The Journal of cell biology 143, 23-34.

Oakes, M.L., Johzuka, K., Vu, L., Eliason, K., and Nomura, M. (2006). Expression of rRNA genes and nucleolus formation at ectopic chromosomal sites in the yeast Saccharomyces cerevisiae. Molecular and cellular biology 26, 6223-6238.

Peng, J.C., and Karpen, G.H. (2007). H3K9 methylation and RNA interference regulate nucleolar organization and repeated DNA stability. Nature cell biology 9, 25-35.

Pikaard, C., and Pontes, O. (2007). Heterochromatin: condense or excise. Nature cell biology 9, 19-20.

Ralph, S.A., Scheidig-Benatar, C., and Scherf, A. (2005). Antigenic variation in Plasmodium falciparum is associated with movement of var loci between subnuclear locations. Proceedings of the National Academy of Sciences of the United States of America 102, 5414-5419.

Raska, I., Shaw, P.J., and Cmarko, D. (2006). Structure and function of the nucleolus in the spotlight. Current opinion in cell biology 18, 325-334.

Sullivan, G.J., Bridger, J.M., Cuthbert, A.P., Newbold, R.F., Bickmore, W.A., and McStay, B. (2001). Human acrocentric chromosomes with transcriptionally silent nucleolar organizer regions associate with nucleoli. The EMBO journal 20, 2867-2874.

Taddei, A., Hediger, F., Neumann, F.R., and Gasser, S.M. (2004). The function of nuclear architecture: a genetic approach. Annual review of genetics 38, 305-345. van Spaendonk, R.M., Ramesar, J., van Wigcheren, A., Eling, W., Beetsma, A.L., van Gemert, G.J., Hooghof, J., Janse, C.J., and Waters, A.P. (2001). Functional equivalence of structurally distinct ribosomes in the malaria parasite, Plasmodium berghei. The Journal of biological chemistry 276, 22638-22647.

Waters, A.P., Syin, C., and McCutchan, T.F. (1989). Developmental regulation of stage-specific ribosome populations in Plasmodium. Nature 342, 438-440.

Wellems, T.E., Walliker, D., Smith, C.L., do Rosario, V.E., Maloy, W.L., Howard, R.J., Carter, R., and McCutchan, T.F. (1987). A histidine-rich protein gene marks a linkage group favored strongly in a genetic cross of Plasmodium falciparum. Cell 49, 633-642.

127

Supplementary figures

Fig. S1. rRNA gene positions in 3D7 P. falciparum linear chromosome map. The malaria genome project identified five complete ribosomal units each located on chromosomes 1, 5, 7, 11 and 13, and two incomplete units (lacking 18S gene) on chromosome 8. rDNA units located on chromosomes 5 and 7 are named A1 and A2 genes, respectively. A1 and A2 gene are transcribed in the human host during schizogony. The rDNA unit on chromosome 1 is called S1 gene and is transcribed in the human host during gametogony. rDNA copies from chromosomes 11 and 13 correspond to the previously described S2 gene that is preferentially transcribed in the mosquito stages during sporogony (Fang et al., 2004; Gardner et al., 2002). We confirmed the duplication of the S2 gene by bioinformatic analysis and also by pulse- field gradient gel electrophoresis (not shown).

128

Fig. S2. Peptide competition assay. The specificity of PfPol I serum was determined by IFA after pre-incubating the antibodies with increasing concentrations of a specific (PfPol I) or a non-specific peptide. The PfPol I signal intensity decreased with increasing concentrations of the specific peptides but not with the non-specific peptides. Anti-PfSir2 antibody was also pre-incubated with the increasing concentrations of PfPol I peptides and no change on fluorescence intensity was observed. The dilutions of antibodies used and the different concentrations of the peptides are indicated.

129

Table S1. Primers used in the study. Name Primers Product size (bp) ATTTCTGTTTTGTGCGTG A1 FISH 1678 TCCTTTGCCTTTTCATTC TCGTCATAATCATCTGAATC A2 FISH 1777 GAGCATACACTATTTTTTAGC AAATGGACGACTAATCGG S1 FISH 1989 GTCTTTTTTTTCTCACAGG ATTTATGTAATCTATTCTTTCTG S2 FISH 1781 AAAAAAACACACTCTCCTC CAGTAGTCATATGCTTGTC 18S FISH 2132 TAATGATCCTTCCGCAGG CCTGTCTCACGACGGTCTA 28S FISH 2182 CGCTTGGATAAGGTGCCTAA TGTTTTCTTTTTTCTAAGTTT A1 qPCR 152 TCACCTCATTTGAAGCAA TGTTTTCTTTTTTCTAAGTTT A2 qPCR 152 TCACCTCATTTGAAGCAA TTAAAGTATCACAATATGAAAA S1 qPCR 199 ATAAACGACAACGCCAAT TTTTTTTTTATTGTGTTGTC S2 qPCR 105 GCTTCACTTTTGTTGTATGT CGAGAGGATCCATAGAAAAAAGTAGGTAAGGG Fragment D 3204 GTAGTCGGGCCCAAATTACTTATTCATGTTTA TTCCGTGTCGCCCTTATTCC pLN FISH 1286 CCTGACGAGCATCACAAAAATCG

130 RESULTS VII

Dynamic chromatin structure of rDNA in Plasmodium falciparum depends on the temperature

Liliana Mancio-Silva and Artur Scherf

Unité de Biologie des Interactions Hôte-Parasite, CNRS URA2581, Institut Pasteur, 25, Rue du Dr Roux, 75724 Paris, France

Correspondence: Artur Scherf, tel: + 33 (01) 4568-8616, fax: + 33 (01) 4568-8348, e-mail: [email protected]

Preliminary data.

131 132 Introduction Post-transcriptional histone modifications play important roles in the regulation of chromatin structure and gene expression. The tails and globular domains of core histones (H2A, H2B, H3 and H4) are subject to number of modifications including acetylation, methylation, phosphorilation, ubiquitination and SUMOylation. There are modifications that associate with active transcription, such H3 and H4 acetylation or H3K4 di- and trimethylation, whereas H3K9, H3K27 and H4K20 trimethylation marks inactive genes or regions [reviewed in (Li et al., 2007)]. These modifications distinguish silent heterochromatin from permissive euchromatin and correlate with the activity status of rDNA repeats. In mammals and plants, hypomethylated, active rRNA genes are associated with acetylated histones H4 and H3 as well as H3K4 trimethylation. In contrast, the promoter of hypermethylated silent genes is associated with methylated H3K9, H3K27 and H4K20 [reviewed in (McStay and Grummt, 2008)]. In P. falciparum, the parasite responsible for human malaria, the rDNA unit has a typical eukaryote organization (see Fig. 1A), but the family itself is the most unusual among the eukaryotic rDNA families. The P. falciparum individual rDNA units are substantial different in sequence and are spread on different chromosomes (Langsley et al., 1983; Wellems et al., 1987). The number of copies is very low, it varies from five to eight, depending on the strain (Mercereau-Puijalon et al., 2002). In the 3D7 strain, the malaria genome project identified five complete ribosomal units each located on separated chromosomes 1, 5, 7, 11 and 13, and two incomplete units (lacking 18S gene) on chromosome 8 (Gardner et al., 2002). Furthermore, P. falciparum rRNA genes are differentially transcribed over the course of the parasite developmental cycle (Waters et al., 1989). rDNA units located on chromosomes 5 and 7 (named A1 and A2 genes) are transcribed in blood stages, in the human host. In gametocyte stages, the dominant transcripts are from the unit on chromosome 1 (S1 gene). Finally, rDNA copies from chromosomes 11 and 13 (S2 type gene) are transcribed only in the mosquito host (Fang and McCutchan, 2002; Fang et al., 2004). It is not known whether the rRNA genes located on chromosome 8 are functional. Thus, only one or two rRNA genes are activated at any one time. To date, little is known on the mechanism(s) responsible for switching rRNA genes ON and OFF in the different developmental stages of the cycle. Recent studies have shown that S2 genes become activated by changing the temperature from 37°C to 26°C and also by glucose starvation (Fang and McCutchan, 2002; Fang et al., 2004). However, the mechanistic basis for the temperature-mediated switch remains elusive. In this study, we have investigated whether chromatin modifications are involved in silencing and activation of P. falciparum rRNA gene transcription by RNA Polymerase I (Pol I).

133 Results and Discussion

Active rRNA promoters are highly enriched in PfPol I and display low nucleosome density To investigate the chromatin structure and organization at the P. falciparum rDNA locus, we used chromatin immunoprecipitation assays (ChIP) coupled to quantification by real time PCR. For this study, we set up specific primers pairs to the upstream (-500 bp) and downstream (+500 bp) of the initiation site (+1) for each A1, A2, S1 and S2 rRNA genes. Examination of individual P. falciparum rRNA genes was possible because the non-coding transcribed sequences are sufficient distinct among the four types of rRNA complete units (Fig. 1A). This allowed us to directly analyse rDNA chromatin in its native chromosomal context, with no need of specific enzyme treatments or transfection of rRNA minigenes, as it has been done in plants and mammalians cells (Lawrence et al., 2004; Santoro et al., 2002). To first check the specificity of our analysis, we used antibodies to the large sub-unit of PfPol I (PFE0465c). As expected in blood stages, PfPol I was present at the start transcription site (+1) of the two active genes (A1 and A2) and absent in the silent genes (S1 and S2) (Fig. 1B). As negative controls, we tested the pre-immune serum and we also analyzed the occupancy of PfPol I on genes transcribed by PfPol II (the calmodulin gene). No enrichment in PfPol I was found in this gene, showing that the rabbit anti-PfPol I antibody recognizes specifically active ribosomal genes (Fig. 1B). We obtained similar results when the immunoprecipitated DNA was detected by Dot-blot (data not shown). In addition, preliminary genome-wide ChIP-chip analysis shows PfPol I enrichment only at the two active rRNA genes, indicating that PfPol I transcribes exclusively rRNA genes (data not shown). We next monitored the occupancy of PfPol I by ChIP during the 48-hour cycle: ~12h (ring stage), ~30h (trophozoite stage) and ~40h (schizont stage), at the non-coding transcribed position +500 bp. ChIP analysis of Pol I in transcribed regions is considered as a measure of on-going gene transcription and gives similar results to those of methods such nuclear run-on (Sandoval et al., 2004). As shown in Fig. 1C, we observed that PfPol I associates to rDNA only on ring stages. These results are consistent with the rRNA transcription profile during the blood stage cycle, which shows that maximum levels of rDNA transcription occur in ring stage parasites (Results VI of this thesis).

134

Fig. 1. Chromatin analysis by ChIP and real time PCR of PfPol I and nucleosome density. A. Schematic representation of one rDNA transcription unit in P. falciparum. It consists of a ~9kb transcribed region coding for a 45S rRNA precursor. The first nucleotide (+1) of this transcript begins at the 5’ external transcribed spacer (5’ETS) followed by the 18S coding region, the internal transcribed spacer 1 (ITS1), the 5.8S coding region, the internal transcribed spacer 2 (ITS2), the 28S coding region, and finally the 3’ ETS. This precursor is processed to mature 18S, 5.8S and 28S rRNAs for ribosome biogenesis. The coding regions display a modest sequence similarity whereas the non-coding regions are considerable distinct among the members of rDNA multi-gene family. The arrow indicates the transcription start site and the direction of transcription. The position of primers used in real time PCR analysis is indicated (small horizontal bars). B. Levels of PfPol I (black bars) and respective pre-immune serum (white bars) at the transcription start site of individual ribosomal genes (A1, A2, S1 and S2) and at Pol II dependent gene, the calmodulin gene (CAL). FCR3 synchronized parasites on ring stage. The transcriptional state of each gene is indicated as ON or OFF. C. Levels of PfPol I at the ETS (+500bp) in three time points of the blood stage cycle: rings (black bars), trophozoites (grey bars) and schizonts (white bars). The transcriptional state of each gene is indicated as ON or OFF. D. Distribution of PfPol I along the ribosomal unit for the active (A1, red and A2, green) and the silent (S1, blue and S2, violet) rRNA genes. 3D7 synchronized parasites in ring stage. E. Distribution of histone H3 core along the ribosomal unit for the active (A1, red and A2, green) and the silent (S1, blue and S2, violet) rRNA genes. 3D7 synchronized parasites in ring stage. The results shown are representative examples.

135 Mapping of PfPol I distribution throughout the ribosomal units revealed that PfPol I bound all the transcribed region of the active rRNA genes, with particular enrichment at the start transcription site (+1) (Fig. 1D). Transcription of genes involves remodelling of nucleosomes, such as histone loss or replacement, which provides access for the transcription apparatus along the transcription unit [reviewed in (Williams and Tyler, 2007)]. To assess the nucleosome density throughout the active and inactive P. falciparum rRNA genes, we used an antibody against the histone H3 core. Remarkably, we found that histone H3 levels are on average approxily ten times higher in silent genes than in active rRNA genes (Fig. 1E). Together, the data from Fig. 1D and 1E show that active rRNA promoters are highly enriched in PfPol I and have low nucleosome density, whereas the silent promoters do not associate with PfPol I and show elevated nucleosome density. Clearly, in P. falciparum we can distinguish two types of rDNA chromatin states: an open (transcriptionally active) and closed conformation (transcriptionally inactive). It should be noted that in Plasmodium parasites such discrepancy on nucleosome density does not occur on Pol II dependent genes. We did not find major differences between active genes (calmodulin and GBP130 genes) and silent genes (PfGam and var2CSA genes) (Fig. 2A). Previous work on the var multi-gene family, which are also highly transcribed genes, did not reveal differences between active, poised and inactive states (Lopez-Rubio et al., 2007).

Histone methylation operates differently in Pol I and Pol II transcribed genes To ascertain whether distinct histone modifications may be implicated in the maintenance of these two chromatin states of rRNA genes in malaria parasites, we performed ChIP assays using antibodies specific to histone methylation. In P. falciparum, H3K4 methylation correlates with transcriptionally active genes, whereas H3K9 methylation marks silent chromatin (Chookajorn et al., 2007; Cui et al., 2007; Lopez-Rubio et al., 2007). Surprisingly, we observed that P. falciparum Pol I associated genes (A1 and A2 rRNA genes) are enriched in both H3K4 and H3K9 methylation at the non-coding transcribed position (+500bp) (Fig. 2B, C and D). As control of our experiments, we analysed Pol II transcribed genes in distinct ON and OFF states. We observed that actives genes (GBP130 and calmodulin) are enriched in H3K4me2/3, while silent genes (inactive var) are enriched on H3K9me3, consistent with published reports (Lopez-Rubio et al., 2007). Also intriguingly, silent Pol I-dependent genes show similar levels of H3K4me2/3 to the active Pol II-dependent genes (Fig. 2B, C and D). Together, our results suggest that in P. falciparum parasites histone methylation apparently operate differently in Pol I and Pol II transcribed genes.

136 With regard to H3K9me3, although several studies have reported the presence of H3K9me3 in the coding region of actively transcribed genes, its function remains unclear (Barski et al., 2007; Vakoc et al., 2005). More recently, H3K9me3 has also been shown associated to actively transcribed mammalian rRNA genes (Yuan et al., 2007). It has been proposed that it might suppress aberrant transcription initiation in the coding regions during Pol I elongating, similarly to the model of Pol II elongation and deposition of H3K36me [reviewed in (Li et al., 2007)].

Fig. 2. Study of the histone methylation patterns of Pol I transcribed genes (A1, red bars; A2, green bars; S1, blue bars; S2, violet bars) and Pol II transcribed genes (calmodulin, CAL and GBP130, GBP – white bars; PfGam, Gam and var2CSA, var – black bars), in 3D7-synchronized parasites in ring stage. The transcriptional state of each gene is indicated as ON or OFF. For the rRNA genes the presented levels correspond to the ETS position (+500bp). Levels of: histone H3 core (A), dimethylated H3K4 (B), trimethylated H3K4 (C) and trimethylated H3K9 (D). For (B), (C) and (D) the values were normalized with the levels of histone H3 core (A). The results shown are representative examples.

Dynamic of position and composition of nucleosomes in rDNA is dependent on the temperature The P. falciparum S2 rRNA genes, which are silent on the asexual blood stage parasites at 37°C, can be transcriptionally activated in vitro by decreasing the temperature to 26°C (Fang and McCutchan, 2002). Synchronized ring-stage parasites were incubated at 26°C for about 7 hours and total RNA collected before and after incubation. RT-PCR analysis, using pairs of primers for each rRNA type, confirmed the increase on S2-type rRNA transcripts relative to 37°C. S1 gene

137 also appeared to increase its transcription though (Fig. 3A). Next, we examined ribosomal chromatin by ChIP for PfPol I, nucleosome density and histones modifications. Although we did not observed a significant shift on PfPol I from A-type to S-type genes at the non-coding transcribed position (+500bp) (Fig. 3B), the levels of the non-modified histone H3 (not shown) and non-modified histone H4 (Fig. 3C) were strikingly changed at 26°C. The transcription of S- type genes was accompanied by a decrease on nucleosome density whereas in the A-types we observed an increase of the H4 levels. We therefore concluded that ribosomal chromatin states are reversible and dependent on the temperature. Moreover, we also detected dynamic changes in the histone modifications (H3K4me2) that correlate with the decrease on transcription on the A-type genes (Fig. 3D). This preliminary data suggests an important role of histone methylation in regulating the epigenetic switch rRNA genes in P. falciparum.

Fig. 3. Dynamic chromatin structure in rRNA genes is dependent on the temperature. A. Transcription of individual rRNA genes at 26°C of 3D7 ring stage parasites. RNA levels were detected by real-time PCR. The amount of transcription at 37°C was taken as unity. B. ChIP analysis of PfPol I levels at 26°C and 37°C for the individual rRNA genes. The levels of PfPol I at 37°C were taken as unity. C. ChIP analysis of histone H4 core levels at 26°C and 37°C for the individual rRNA genes. The values were normalized with the calmodulin gene values. D. Levels of dimethylated H3K4 at 26°C for the rRNA genes and the calmodulin gene. The values were normalized with the levels of histone H4 core.

138 In summary, this work shed some light on the molecular mechanisms of rRNA gene regulation in malaria parasites. Here, we show that P. falciparum active and silent rRNA genes exist in distinct epigenetic states. P. falciparum active rRNA genes display euchromatic features, such open conformation, association with Pol I and methylated H3K4 and H3K9. By contrast, silent rRNA genes exhibit a condensed heterochromatic structure (Fig. 1 and 2). In addition, we demonstrate that these chromatin conformations can be modulated by temperature changes (Fig. 3). The presence of H3K9me3 on active rRNA genes and the absence of PfPol I in the cold-activated S2 genes need further investigation. Regarding the PfPol I, we do not exclude the possibility that this enzyme does not work at 26°C and that S2 rRNA genes might be transcribed at 26°C by another RNA Polymerase, namely Pol II. Some evidences support our hypothesis: first, the upstream region of S2 genes (and also S1 gene) does not contain the conserved features of a Pol I promoter; second, P. falciparum genome appears to have a single copy gene for Pol I; and third, several studies performed on yeast, have shown that rDNA can be transcribed by Pol II. It will be interesting to analyse transcription of S2 genes at 26°C in the presence of alpha-amanitin, a toxin inhibitor of Pol II and Pol III. Future work should also investigate whether there are in P. falciparum enzymes specifically located in the nucleolus that perform histone modifications (histone methyltransferases) and/or also that remove modifications (histone demethylases). Those enzymes would be interesting targets for anti-malarial drug design.

Materials and Methods Parasites P. falciparum blood stage parasites from the 3D7 strain were cultured as detailed described on (Nunes et al., 2007). Parasites were synchronized by consecutive Plasmagel and Sorbitol treatments. Ring stages were collected approximately 12 hours post-invasion (p.i.), trophozoites stages 30 hours p.i. and schizonts stages 40 hours p.i. within the same 48 hour cycle.

ChIP The ChIP assay was performed using the protocol described earlier (malaria methods) with some modifications. Parasites cultures were first treated with saponin and cross-linked with 1% formaldeyde for 10 minutes. ~10µg of chromatin was incubated at 4°C with the following antibodies: 7.5 µl of anti-PfPol I and respective pre-immune serum, 2 µl anti-H3core (ab1791,

139 abcam), 7.5 µl anti-H4core (ab31827, abcam), 1µl anti-H3K9me3 (07-442, Upstate), 2 µl anti- H3K4me3 (ab8580, abcam) and 2 µl anti-H3K4me2 (ab7766, abcam). Immunoprecipitated DNA was analyzed by real-time PCR using SYBR Green PCR master mix (Applied Biosystems) on a Realplex4 EpgradientS thermal cycler (Eppendorf). The primer pairs specific to rRNA genes are listed on Table 1. The primer sequence used as controls (calmodulin, GBP130, PfGam and var2CSA) are described on (Lopez-Rubio et al., 2007). PCR was performed in duplicates and serial dilutions of purified input DNA were measured together with the immunoprecipitated DNA samples. The % of input calculated corresponds to the amount of target sequence in immunoprecipitated chromatin relative to the amount of target sequence in input DNA.

RT-PCR Total RNA from synchronized cultures was isolated according to (Kyes et al., 2000). For RNA at 37°C, all the operations were performed at 37°C. For RNA at 26°C, parasites were maintained at 26°C for 7 hours and operations performed at room temperature. Extracted RNA was treated with DNA-free kit (Ambion) and reverse transcribed with Super- scriptII Rnase H Reverse Transcriptase (Invitrogen) using random hexameres (Invitrogen). Quantification was performed by real-time PCR. Water, no-reverse transcriptase controls and standard curves with serial dilutions of genomic DNA were included in each run. The pairs of primers used are specific to the 5’ETS (+500); this sequence is rapidly processed and degraded and is commonly used to measure on going rDNA transcription. The values obtained for each rRNA gene were normalized to the housekeeping gene calmodulin (PF14_0323). Using the 2-∆∆Ct method (Livak and Schmittgen, 2001) we calculated the amount of RNA of each gene at 26°C relative to the amount of the gene at 37°C.

Acknowledgments We would like to thank to Jose Juan Lopez-Rubio for thoughtful discussions and to Anabela Cordeiro-da-Silva and Faculdade de Farmácia da Universidade do Porto, Portugal for help and support. This work was supported by the European Commission (BioMalPar) with contract number: LSPH-CT-2004−503578. L.M.S. has financial support from the GABBA PhD program and Fundação para a Ciência e Tecnologia, Portugal. A.S. is supported by a grant from the "Fonds dédié: Combattre les Maladies Parasitaires" Sanofi Aventis - Ministère de la Recherche and an ANR grant (ANR-06-MIME-026-01).

140 References Barski, A., Cuddapah, S., Cui, K., Roh, T.Y., Schones, D.E., Wang, Z., Wei, G., Chepelev, I., and Zhao, K. (2007). High-resolution profiling of histone methylations in the human genome. Cell 129, 823-837.

Chookajorn, T., Dzikowski, R., Frank, M., Li, F., Jiwani, A.Z., Hartl, D.L., and Deitsch, K.W. (2007). Epigenetic memory at malaria virulence genes. Proceedings of the National Academy of Sciences of the United States of America 104, 899-902.

Cui, L., Miao, J., Furuya, T., Li, X., Su, X.Z., and Cui, L. (2007). PfGCN5-mediated histone H3 acetylation plays a key role in gene expression in Plasmodium falciparum. Eukaryot Cell 6, 1219-1227.

Fang, J., and McCutchan, T.F. (2002). Thermoregulation in a parasite's life cycle. Nature 418, 742.

Fang, J., Sullivan, M., and McCutchan, T.F. (2004). The effects of glucose concentration on the reciprocal regulation of rRNA promoters in Plasmodium falciparum. The Journal of biological chemistry 279, 720- 725.

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. (2002). Genome sequence of the human malaria parasite Plasmodium falciparum. Nature 419, 498-511.

Kyes, S., Pinches, R., and Newbold, C. (2000). A simple RNA analysis method shows var and rif multigene family expression patterns in Plasmodium falciparum. Molecular & Biochemical Parasitology 105, 311-315.

Langsley, G., Hyde, J.E., Goman, M., and Scaife, J.G. (1983). Cloning and characterisation of the rRNA genes from the human malaria parasite Plasmodium falciparum. Nucleic acids research 11, 8703-8717.

Lawrence, R.J., Earley, K., Pontes, O., Silva, M., Chen, Z.J., Neves, N., Viegas, W., and Pikaard, C.S. (2004). A concerted DNA methylation/histone methylation switch regulates rRNA gene dosage control and nucleolar dominance. Molecular cell 13, 599-609.

Li, B., Carey, M., and Workman, J.L. (2007). The role of chromatin during transcription. Cell 128, 707- 719.

Livak, K.J., and Schmittgen, T.D. (2001). Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods (San Diego, Calif 25, 402-408.

Lopez-Rubio, J.J., Gontijo, A.M., Nunes, M.C., Issar, N., Hernandez Rivas, R., and Scherf, A. (2007). 5' flanking region of var genes nucleate histone modification patterns linked to phenotypic inheritance of virulence traits in malaria parasites. Mol Microbiol.

McStay, B., and Grummt, I. (2008). The Epigenetics of rRNA Genes: From Molecular to Chromosome Biology. Annual review of cell and developmental biology.

Mercereau-Puijalon, O., Barale, J.C., and Bischoff, E. (2002). Three multigene families in Plasmodium parasites: facts and questions. International journal for parasitology 32, 1323-1344.

Nunes, M.C., Goldring, J.P., Doerig, C., and Scherf, A. (2007). A novel protein kinase family in Plasmodium falciparum is differentially transcribed and secreted to various cellular compartments of the host cell. Mol Microbiol 63, 391-403.

Sandoval, J., Rodriguez, J.L., Tur, G., Serviddio, G., Pereda, J., Boukaba, A., Sastre, J., Torres, L., Franco, L., and Lopez-Rodas, G. (2004). RNAPol-ChIP: a novel application of chromatin immunoprecipitation to the analysis of real-time gene transcription. Nucleic acids research 32, e88.

141 Santoro, R., Li, J., and Grummt, I. (2002). The nucleolar remodeling complex NoRC mediates heterochromatin formation and silencing of ribosomal gene transcription. Nature genetics 32, 393-396.

Vakoc, C.R., Mandat, S.A., Olenchock, B.A., and Blobel, G.A. (2005). Histone H3 lysine 9 methylation and HP1gamma are associated with transcription elongation through mammalian chromatin. Molecular cell 19, 381-391.

Waters, A.P., Syin, C., and McCutchan, T.F. (1989). Developmental regulation of stage-specific ribosome populations in Plasmodium. Nature 342, 438-440.

Wellems, T.E., Walliker, D., Smith, C.L., do Rosario, V.E., Maloy, W.L., Howard, R.J., Carter, R., and McCutchan, T.F. (1987). A histidine-rich protein gene marks a linkage group favored strongly in a genetic cross of Plasmodium falciparum. Cell 49, 633-642.

Williams, S.K., and Tyler, J.K. (2007). Transcriptional regulation by chromatin disassembly and reassembly. Current opinion in genetics & development 17, 88-93.

Yuan, X., Feng, W., Imhof, A., Grummt, I., and Zhou, Y. (2007). Activation of RNA polymerase I transcription by cockayne syndrome group B protein and histone methyltransferase G9a. Molecular cell 27, 585-595.

142 Table 1. List of primers used in the real time PCR analysis. The primer pair name indicates the position relative to the start transcription start site.

Primer pair Forward primer Reverse primer Product size (bp) A1 -500 TATATAGTATATGAATAAGACC ATTAATTAATAATTGCATTAC 111 A1 +1 CCATTTCTAAATAAATAAACGC TTTTATTTCTTTTTTTCTTTACTA 82 A1 +500 TTTTTTCTAATGTTCCAAT AATATAAAACTTTCACCTCAT 201 A2 -500 AAAGATTTTACAACCTAT AATACATACCCTACCTTA 91 A2 +1 TCTGAAAGAAAAATCTCACT CCTCCACCAGTCAATCAT 92 A2 +500 TGTTTTCTTTTTTCTAAGTTT TCACCTCATTTGAAGCAA 152 S1 -500 ATAAGACAACTGGCATTT TACACTTTCATTAGGGAA 111 S1 +1 TTCCTTTTTCCTTTTATCTA TCCTCTCCTCCCTTTTCA 189 S1 +500 GAAATATAAAACATCAATTGGCGT ATAAATATAAATATATCTACATGTTGA 135 S2 -500 AAGACAAATAAGTAATAGTAT CTTCAATAAATCCAACCT 85 S2 +1 TGAAAAATCAAAAGGCAA CCTAAAGAATAAAACCTAACA 138 S2 +500 TTTTTTTTTATTGTGTTGTC GCTTCACTTTTGTTGTATGT 105 A +2000 ATCCTTTGATTTTTATCTTTGG ATAAAATAAATTTATTACGTG 108 S +2000 GTATCATTAAGGTATGTATTTG CACATATTTTATGTGTTTCTTCT 86

143

CONCLUDING REMARKS

This final section summarizes the major conclusions of the presented results. We attempt to place together our observations, the questions that arise and discuss the future research directions.

P. falciparum heterochromatin and transcriptional silencing In most eukaryotes, heterochromatin is concentrated in telomeric and pericentromeric regions and is enriched for repetitive sequences. Heterochromatin is characterized by specific histone modifications and associated proteins that maintain these regions in highly condensed nucleosomal arrays inaccessible for transcription [reviewed in (Grewal and Jia, 2007)]. Plasmodium telomeres are thought to have characteristics of heterochromatin, as suggested by the fact that virulence subtelomeric genes are transcriptionally silenced. Like in yeasts, P. falciparum telomeric and subtelomeric repeats are hypoacetylated and enriched for a deacetylase protein, PfSir2 (Freitas-Junior et al., 2005; Merrick and Duraisingh, 2007; Shankaranarayana et al., 2003). In this work, using immunofluorescence, ChIP and EMSA studies, we demonstrated that other proteins can associate to P. falciparum telomeric domains: PfOrc1 (Results II), PfHP1 (Results III) and possibly PfKMT1 (Results V). Using ChIP coupled to high-resolution tilled arrays, we demonstrated that P. falciparum telomeric and subtelomeric chromatin is highly enriched in H3K9m3. H4K20me3 was also found in these regions, but not H3K9ac or H3K4m3 (Results V). This histone modifications pattern is commonly found on heterochromatin in metazoan organisms [reviewed in (Grewal and Jia, 2007)]. A part from the telomeric domains, H3K9me3 was also detected highly concentrated in internal chromosomal regions but not in all chromosomes. Surprisingly, these internal enriched regions do not correspond to P. falciparum centromeres. Instead, they are composed mostly of var, rif and stevor genes, like the subtelomeric coding regions. The fact that H3K9me3 appears to be restricted to clonally variant virulence gene families is astonishing and implies that this

145

specific modification has an important role in the epigenetic regulation of phenotypic variation in this parasite. This highly restricted localization of H3K9me3 in P. falciparum genome (Results V) points to the existence of regulatory mechanisms controlling silent heterochromatin spreading in this parasite. Histone variants (Meneghini et al., 2003), DNA elements that recruit active chromatin factors (Oki et al., 2004) and tRNAs associated to Pol III machineries (Oki and Kamakaka, 2005) are examples of heterochromatin boundaries described in other eukaryotes [reviewed in (Gaszner and Felsenfeld, 2006; Grewal and Jia, 2007)]. It will be interesting to investigate whether these or distinct boundary elements operate in P. falciparum. It will be also interesting to perform ChIP-chip analysis of PfSir2, PfOrc1 and PfHP1 to elucidate whether internal H3K9me3 enriched regions share the same molecules with telomeres. PfOrc1 genome-wide binding map may additionally reveal the DNA replicating sequences, which have not been established in P. falciparum. Although further investigation is needed to gain mechanistic insight on whether these telomeric proteins interact, evidences from studies in other organisms (Bannister et al., 2001; Nakayama et al., 2001; Rea et al., 2000) led us to propose the following model for heterochromatin assembly in P. falciparum: H3K9me provides the binding site for PfHP1 to recruit other proteins, such as histone deacetylases (possibly PfSir2) and methyltransferases (possibly PfKMT1) that modify adjacent nucleosomes. Multiple rounds of deacetylation and methylation of H3K9, with subsequent binding of PfHP1 and recruitment of more modifying enzymes may lead to heterochromatin propagation. Importantly, PfHP1 and PfSir2 appear to have ability of multimerization [Results III; (Chakrabarty et al., 2008)], which may facilitate their spreading across the subtelomeric repeats. The mechanism for initiation of P. falciparum heterochromatin assembly also remains to be resolved. In analogy to Drosophila pericentromeric heterochromatin (Pak et al., 1997; Shareef et al., 2003), PfOrc1 may function as a DNA-binding factor that recruits PfHP1 to nucleation sites. Alternatively, but not mutually exclusive, recently identified subtelomeric noncoding RNAs in P. falciparum (Mourier et al., 2008) may also cooperate in targeting heterochromatin. Although the conventional RNAi machinery is absent in this malaria parasite, it is possible that P. falciparum possess distinct silencing machinery that uses these non-coding RNAs in a similar manner as the RNAi mechanism (Hall et al., 2002). Experiments using RNAse treatments of parasite nuclei will possibly reveal if nucleation of heterochromatin in P. falciparum is a DNA- and/or RNA-based mechanism. Further studies should also clarify whether the PfOrc1 putative leucine zipper motif (Mehra et al., 2005) has ability to bind DNA.

146

A striking result was that PfSir2 is not essential for telomeric heterochromatin formation and propagation. In PfSir2 depleted parasites the H3K9me3 pattern remained barely unchanged (Results V). The effect of PfSir2 appears to be restricted to the 5’upstream regions of specific subtelomeric genes, where H3K9ac replaced H3K9m3. This local rather than a regional effect of H3K9me3 depletion was unexpected given the postulated model of TPE (Freitas-Junior et al., 2005), which predicted a gradient of PfSir2 from the telomere into the subtelomeric coding region. These observations suggest that other histone deacetylases are implicated in heterochromatin formation and maintenance. In fact, P. falciparum has a second Sir2 homologue, PF14_0489. This protein shows similar nuclear localization as PfSir2 and deletion of this molecule produces a more significant decrease (but not entire depletion) of H3K9me3 in subtelomeric regions (L. Mancio-Silva and J.J. Lopez-Rubio, unpublished). It is possible that both Sir2 proteins work together and that one can compensate for lack of the other. A double knock-out would be necessary to ascertain the function of these histone deacetylases. The finding that Plasmodium telomeres and subtelomeric regions are enriched in epigenetic marks indicates that chromatin modifications may be important regulators of telomeres. In mammalian cells, for example, deletion of KMT1A and KMT1B (SUV39H1 and SUV39H2) results in decreased levels of H3K9me at telomeres and aberrant telomere elongation (Garcia-Cao et al., 2004). Our preliminary work indicates that PfSir2 is a negative regulator of telomere length, as deletion of this protein results in elongated telomeres (L. Mancio-Silva and A. Scherf, unpublished).

P. falciparum rRNA gene regulation in response to environmental changes rRNA synthesis is tightly regulated in response to metabolic and environmental changes [reviewed in (Grummt and Ladurner, 2008)]. In P. falciparum, it is also known that changes in temperature and glucose concentration affect the rRNA transcription pattern (Fang and McCutchan, 2002; Fang et al., 2004). In this study, we demonstrated that P. falciparum rDNA transcription, similarly to eukaryotic rRNA genes, is regulated epigenetically. Active rRNA genes display high levels of PfPol I, low nucleosome density and active histone modifications, whereas silent genes show high nucleosome density and absence of PfPol I and active epigenetic marks. In addition, we demonstrated that the two epigenetic states are reversible and modulated by temperature changes (Results VII). Plasmodium ribosomes and rRNA genes were always seen as interesting targets for antimalarial drugs. The finding that rRNA genes are epigenetically

147

regulated revealed novel enzymes, such as histone methyltransferases involved in rRNA gene activation, to be explored in the future as drug targets. In humans, epigenetic control of rDNA loci in response to glucose starvation, involves a protein complex composed of SIRT1 (Sir2 homologue), KMT1A and nucleomethylin (NML) (Murayama et al., 2008). Under low glucose concentrations, the NAD-dependent SIRT1 is activated, leading to H3K9 deacetylation and methylation by KMT1A, thus establishing silencing. NML binds to H3K9me2 in the nucleolus and recruits SIRT1 and KMT1A. Several evidences indicate that a mechanism akin is behind the downregulation of A-type genes under glucose starvation (Fang et al., 2004) in malaria parasites: first, PfSir2 is highly enriched in the nucleolus (Freitas-Junior et al., 2005); second, PfKMT1 also appears to localize in the nucleolus (J.J. Lopez-Rubio, unpublished); and third, P. falciparum genome encodes a NML homologue (PFI1235w). Moreover, preliminary ChIP-chip analysis of H3K9me3 in PfSir2 knock-out parasites shows a decrease in H3K9me3 and increase in H3K9ac at the rDNA locus (L. Mancio- Silva and J.J. Lopez-Rubio, unpublished). This result indicates that H3K9 methylation in P. falciparum rDNA requires PfSir2-dependent H3K9 deacetylation, like in humans. It will be interesting to submit these PfSir2 knock-out parasites to low glucose concentrations and evaluate whether the A-type genes cannot be downregulated in the absence of PfSir2. Given that glucose starvation and low temperature have synergistic effects on rRNA gene regulation (Fang et al., 2004), future work should examine parasites in both conditions. Analysis of the nucleolar structure under glucose and low temperature should also be performed.

P. falciparum gene regulation through subnuclear compartmentalization An important feature of eukaryotic nuclei is the existence of distinct structural and functional compartments. These compartments are characterized by the presence or enrichment of specific proteins, sequences or chromosomal elements. Increasing evidences suggest that repositioning of genomic regions in the nuclear space participate in regulation of gene expression [reviewed in (Sexton et al., 2007)]. The first evidence that P. falciparum nuclei are compartmentalized comes from the observation that telomeres are clustered at the nuclear periphery (Freitas-Junior et al., 2000). Later, it was shown that they create silencing compartments (Freitas-Junior et al., 2005). In the present study, we identified and characterized additional subcompartments in P. falciparum nucleus (schematically represented in Figure 1), and we attempted to understand whether they are important for gene regulation in this parasite.

148

Figure 1. Model for nuclear architecture organization of ring-stage P. falciparum parasites. The nucleolus is represented in a half-moon shape at the nuclear periphery, opposing the telomeric clusters. Our results show that chromosomal domains enriched in H3K9me3 (red flags) localize at perinuclear regions irrespective of their chromosomal location, implying formation of large chromosomal loops. Chromosomal loops must also ensure the clustering of the dispersed rDNA in the nucleolus. Active (green rectangles) and silent (black rectangles) rRNA genes co-exist in the nucleolus. The two active rRNA genes are marked with histone modifications (H3K4me2/3, blue flags and H3K9me3, red flags) and are associated to two Pol I expression sites. Regarding to the var multigene family (white rectangles), we observed that all members localize in the nuclear periphery regardless of their chromosomal location or transcriptional status. Upon activation, colocalization of var genes (for all groups) with the telomeric clusters is reduced, suggesting that to be transcribed var genes moves out from the telomeric cluster into a putative var transcription zone. The figure illustrates a hypothetical relocation of a subtelomeric var gene. We speculate that var gene activation, including acquisition of histone marks such as H3K9ac (green flags) and H3K4me2/3 (blue flags) (Lopez-Rubio et al., 2007a), may lead to chromatin de-condensation and formation of a chromatin loop, which separate out the transcribed var gene (green rectangle) from more compact and silent telomeric cluster.

149

The rRNA transcription zone P. falciparum nucleolus is the most prominent subcompartment, it localizes in the periphery occupying roughly one third of the ring-stage parasite nucleus (Figure 1). The nucleolus is known as a compartment dedicated to Pol I transcription of rRNA genes. In P. falciparum, rDNA units are dispersed in distinct chromosomes and not tandem arrayed. Our initial hypothesis was that the parasite would take advantage of this unusual distribution in the genome to regulate differential transcription. We expected that actively transcribed rDNA units would locate in the nucleolus, whereas silent rDNA would be elsewhere in the nucleus. Using IFA combined with FISH of individual rRNA genes, we demonstrated that active and silent members localized in the same compartment, indicating that activation and silencing can co-exist in the nucleolus. Thus, in this differentially transcribed multigene family, nuclear position is important, but not sufficient for transcriptional activation. We further investigated factors that determine nucleolar localization and found that apparently rDNA sequence itself contains positional information necessary for spatial location (Results VI). Future experiments should try to find the minimal region of rDNA required for nucleolar targeting and identify possible proteins involved in recognition. Nucleolus and telomeric clusters define distinct subcompartments in P. falciparum nucleus (Results VI) (Figure 1); however, they appear to share a number of proteins, such as PfSir2 (Freitas-Junior et al., 2005), PfTERT (Results I) and PfOrc1 (Results II). The presence of telomeric components in nucleoli has also been observed in yeast and mammalian cells. In yeast nucleolus, Sir2 is known to repress both recombination and Pol II transcription within the rDNA locus (Gottlieb and Esposito, 1989; Kobayashi et al., 2004; Smith and Boeke, 1997). Thus, a possible function for PfSir2 in the nucleolus is to suppress recombination among rDNA repeats and to guarantee the minimal number of rRNA genes required for parasite viability. Future work should address the rDNA stability in the PfSir2 knock-out parasite strain. Regarding the telomerase, the mammalian nucleolus appears to regulate telomerase activity by controlling its distribution in subnuclear compartments. Translocation of telomerase from the nucleolus to the nucleoplasm coincides with the time of telomere synthesis. Sequestering of telomerase in the nucleolus has been suggested to prevent its potential interaction with substrates other than the telomeres [reviewed in (Olson, 2004)]. It is expected that similar regulation occur for PfTERT in malaria parasites. Likewise, the nucleolus may sequester PfOrc1 until the time that this molecule is necessary for initiation of replication, and thus protect the DNA from inappropriate replication.

150

The var transcription zone Apart from the nucleolus other compartments exist in the nuclear periphery where active transcription can occur. This is the case for another multigene family, the var genes that are expressed in a monoallelic manner. We observed that all var genes, regardless of their activity or chromosomal location, localize predominately in perinuclear regions (Results V) (Figure 1). In their silent state, for subtelomeric var genes (groups A and B) colocalization with telomeric clusters is elevated, while for internal var genes (group C) colocalization appears to be random. Upon activation, colocalization of var transcripts (of all groups) with the telomeric clusters is reduced, suggestive of relocation of the var gene to a putative transcription zone outside of the silent telomeric cluster (Results V). This result, obtained using RNA/DNA FISH and parasites in their native chromosomal context, clarified a discrepancy of results from previous studies that used parasites transfected with various constructs. It is noteworthy that in yeast, activation using a natural and artificial pathway leads to repositioning of a gene into distinct nuclear regions (Taddei et al., 2006). This exemplifies the importance of the native chromosomal context to study mechanisms of gene activation. Regulation of monoallelic expression by relocation of genes to distinct subnuclear compartments has also been described in mammals and other parasites. For example, during early mammalian female development, inactivation of one of the two X chromosomes involves relocation to a more internal location, within the Xist noncoding RNA compartment devoid of Pol II transcription machinery (Chaumeil et al., 2006). In African trypanosomes, the single active VSG (variant surface glycoprotein) is transcribed by Pol I in an extranucleolar expression site (Navarro and Gull, 2001). Upon developmental silencing, the active VSG is subject to repositioning to the nuclear periphery (Landeira and Navarro, 2007). In analogy to the extranucleolar VSG expression site, we speculate that upon activation the single var gene relocates into a privileged var-specific expression site (Figure 1). This perinuclear subcompartment predicts exclusion of the silent var genes and could also explain how activity could be restricted to one gene at the time. Experiments that forced the activation of two var promoters in episomes showed elevated colocalization, supporting the idea of existence of a var- specific expression site, however this site can accommodates more than one active var gene (Dzikowski et al., 2007). It will be interesting to re-analyze this question using var genes in their native chromosomal context. This can possibly be done using PfSir2 knock-out parasites, in which the mutually exclusive transcription is overruled, i.e. that several var genes appear to be transcribed in a single parasite.

151

The understanding of the var counting/choosing mechanism stills puzzling. Apparently it is regulated at a level other than the subnuclear compartmentalization. Lopez-Rubio et al. (2007b) have proposed an ‘enhancer-mediated mechanism’, in which var genes would compete for physical interaction with a regulatory element (such as an enhancer) that localizes in the var- specific expression site. If this var-specific enhancer exists, it can possibly be detected using the chromosome conformation capture (3C) methodology. The nature of the ‘var-specific expression site’ in the nuclear periphery also remains to be determined. Interestingly, a number of recent studies have reported the association of actively transcribed genes with the nuclear pore complex [reviewed in (Akhtar and Gasser, 2007)]. In yeast, for example, relocation of a subtelomeric gene from a telomeric focus and association with the nuclear pore has been shown to confer optimal gene expression levels (Taddei et al., 2006). It is very likely that, similar to yeast, var genes associate to nuclear pores upon activation. Nuclear pores appear to serve as docking sites for chromatin remodeling factors, transcription machinery and also mRNA processing factors. In this scenario is possible that the previously suggested var enhancer (Lopez-Rubio et al., 2007b) may also be anchored at the pore. Association of the active var gene to the nuclear pore, with the formation of a single-gene loop (Figure 1), would satisfy an additional issue: the repression of neighboring var genes. The nuclear pores can act as barriers boundaries that insulate active domains from flanking repressed regions (Ishii et al., 2002). The hypothesis of var genes being transcribed at the pores is quite straightforward to test, as an antibody for the nuclear pore has now been validated (Results V). We can combine immunofluorescence of nuclear pores with RNA FISH of var genes and/or perform ChIP analysis of the nuclear pore protein.

Post-translational histone modifications Previous studies have shown that post-translation histone modifications have important roles in gene regulation in P. falciparum [reviewed in (Coleman and Duraisingh, 2008)]. In this study, we demonstrated that they might also be important for nuclear organization of genomic DNA. Quantitative analysis of subnuclear position of genes enriched in H3K9me3 revealed a predominant localization at the nuclear periphery, irrespective of their chromosomal location. This implies that large chromosomal loops must be formed in order to bring the internal H3K9me3 enriched regions to the nuclear periphery (Figure 1). PfKMT1 was also detected in the perinuclear region, and led us to propose that together with PfSir2 and other potential silencing factors, such as PfOrc1 and PfHP1 may create ‘Perinuclear Epigenetic Repressive Factories’

152

(PERF), a specialized compartment for gene silencing (Results V). Co-immunoprecipitation experiments should elucidate whether these proteins form complexes in the nuclear periphery. Consistent with the existence of active transcription and silencing in the nuclear periphery (visualized in this study for the var and rRNA genes), we found that histones modifications such as H3K4me (active) and H4K20me (repressive) co-exist in this region (Results IV). Interestingly, some histone modifications appear to form distinct subnuclear compartments. This is the case of H3K79me3, which in higher eukaryotes associates to active genes (Barski et al., 2007). In P. falciparum, H3K79me3 appeared to occupy unique regions in the nucleus that did not overlap with the nucleolus, the telomeric clusters nor the putative var expression site. We speculate that H3K79me3 may mark Pol II transcription factories. ChIP-chip analysis of H3K79me will possibly reveal the genes transcribed in these factories.

Dynamic changes during the blood stage cycle A remarkable feature of P. falciparum nuclear architecture is that it undergoes dynamic changes throughout the 48-hour blood stage cycle. The nucleolus disassembles in S/M phase (~30hours); this is apparently correlated with the decrease of rRNA synthesis (Results VI). Nucleolar proteins such as PfPol I, PfNop1, PfNop5 (Results VI), and also PfSir2 and PfOrc1 (Results II) re-distribute in the nucleoplasm and cytoplasm prior to DNA replication. Also, the telomeric clusters partially disassemble by the same time (Results II). It remains to be shown whether these relocations are linked to differential functions at different time points of the cycle, and whether they are controlled by differential post-translational modifications, such as phosphorylation or SUMOylation (Issar et al., 2008).

Future directions on nuclear architecture Although the observations presented here provided considerable advances in the understanding the nuclear architecture and its role on regulation of gene expression in P. falciparum parasites, further studies are needed to identify more nuclear landmarks, and they should definitely include three-dimensional analysis. Future studies should address the protein complexes that tether telomeres and internal chromosome regions to the nuclear periphery and that lead to repression of perinuclear virulence gene families. Chromosome looping formation and whether it is compatible with the organization of the chromosomes in nuclear territories also needs closer investigation, possibly using FISH chromosome mapping and/or 3C methodology. The localization of Pol II transcription factories, namely the one involved in var gene

153

transcription at the perinuclear space also ought to have further examination to better understand the molecular mechanism(s) that control monoallelic expression. Improved technologies such as P. falciparum transfection inducible systems, ideally a GFP-lac repressor/lac operator system, currently used in yeast for imaging of a single DNA locus in living cells should give us the opportunity of working with live parasites and overcome the problems associated to fixation and harsh treatments of FISH techniques. Protein tagging and live imaging must provide a more accurate view of the spatial and temporal aspects of the parasite nuclear dynamics across the 48-hour cycle. In particular, real-time fluorescence microscopy tracking of individual loci should shed more light on the chromatin repositioning movements associated to transcriptional gene activation.

154

ACKNOWLEDGMENTS

I would like to express my gratitude towards Prof. Artur Scherf for his supervision, guidance and encouragement during these years. Also for proposing me challenging questions and for giving me the scientific freedom to conduct the experimental work.

I would like to extend my appreciation to Prof. Anabela Cordeiro da Silva for her co-supervision, for helpful suggestions, guidance and for dealing with the administrative part of this thesis.

I also would like to thank to all past and present members in the Biology of Host Parasite Interactions at Institut Pasteur. In particular, I would like to thank: Lucio Freitas-Junior and Luisa M. Figueiredo for introducing me to research and for sharing with me their experience and knowledge. Jose Juan Lopez-Rubio for his interest, advices, thoughtful discussions, friendship and also for careful reading of all my manuscripts (including this thesis). Marta Nunes for always being supportive, for immense technical advices and discussions. Christine Scheidig-Benatar and Yvon Sterkers for their helpful technical french assistance. Loïc Riviere for always being accessible and for proofreading this thesis. Ana Paola Rojas-Meza and Neha Issar, my PhD colleagues, for their companionship and support.

It is a great pleasure to thank the GABBA Program, all the scientific coordinators and my colleagues of class 2003 for their friendship and scientific networking.

I would like to end by thanking to my parents, my sister, all my family and Gonçalo for their endless support and patience.

The studies described in this thesis were performed at the Institut Pasteur (Paris, France), in the Parasitology and Micology Department. I was a student of Graduate Program in Areas of Basic and Applied Biology (GABBA) and financially supported by a Doctoral Fellowship (SFRH/BD/11756/2003) from Fundação para a Ciência e Tecnologia, Portugal.

155

156

REFERENCES

Afonso, A., Hunt, P., Cheesman, S., Alves, A.C., Cunha, C.V., do Rosario, V., and Cravo, P. (2006). Malaria parasites can develop stable resistance to artemisinin but lack mutations in candidate genes atp6 (encoding the sarcoplasmic and endoplasmic reticulum Ca2+ ATPase), tctp, mdr1, and cg10. Antimicrobial agents and chemotherapy 50, 480-489.

Akhtar, A., and Gasser, S.M. (2007). The nuclear envelope and transcriptional control. Nature reviews 8, 507-517.

Allis, C.D., Berger, S.L., Cote, J., Dent, S., Jenuwien, T., Kouzarides, T., Pillus, L., Reinberg, D., Shi, Y., Shiekhattar, R., et al. (2007). New nomenclature for chromatin-modifying enzymes. Cell 131, 633-636.

Alonso, P.L., Sacarlal, J., Aponte, J.J., Leach, A., Macete, E., Milman, J., Mandomando, I., Spiessens, B., Guinovart, C., Espasa, M., et al. (2004). Efficacy of the RTS,S/AS02A vaccine against Plasmodium falciparum infection and disease in young African children: randomised controlled trial. Lancet 364, 1411- 1420.

Andrews, K.T., Tran, T.N., Lucke, A.J., Kahnberg, P., Le, G.T., Boyle, G.M., Gardiner, D.L., Skinner- Adams, T.S., and Fairlie, D.P. (2008). Potent antimalarial activity of histone deacetylase inhibitor analogues. Antimicrobial agents and chemotherapy 52, 1454-1461.

Balaji, S., Babu, M.M., Iyer, L.M., and Aravind, L. (2005). Discovery of the principal specific transcription factors of Apicomplexa and their implication for the evolution of the AP2-integrase DNA binding domains. Nucleic acids research 33, 3994-4006.

Bannister, A.J., Zegerman, P., Partridge, J.F., Miska, E.A., Thomas, J.O., Allshire, R.C., and Kouzarides, T. (2001). Selective recognition of methylated lysine 9 on histone H3 by the HP1 chromo domain. Nature 410, 120-124.

Barski, A., Cuddapah, S., Cui, K., Roh, T.Y., Schones, D.E., Wang, Z., Wei, G., Chepelev, I., and Zhao, K. (2007). High-resolution profiling of histone methylations in the human genome. Cell 129, 823-837.

Baruch, D.I., Pasloske, B.L., Singh, H.B., Bi, X., Ma, X.C., Feldman, M., Taraschi, T.F., and Howard, R.J. (1995). Cloning of P. falciparum gene encoding PfEMP1, a malarial variant antigen and adherence receptor on the surface of parasitized human erythrocytes. Cell 82, 77-87.

Ben Mamoun, C., Gluzman, I.Y., Hott, C., MacMillan, S.K., Amarakone, A.S., Anderson, D.L., Carlton, J.M., Dame, J.B., Chakrabarti, D., Martin, R.K., et al. (2001). Co-ordinated programme of gene expression during asexual intraerythrocytic development of the human malaria parasite Plasmodium falciparum revealed by microarray analysis. Mol Microbiol 39, 26-36.

Blasco, M.A. (2007). The epigenetic regulation of mammalian telomeres. Nature reviews 8, 299-309.

Bojang, K.A., Milligan, P.J., Pinder, M., Vigneron, L., Alloueche, A., Kester, K.E., Ballou, W.R., Conway, D.J., Reece, W.H., Gothard, P., et al. (2001). Efficacy of RTS,S/AS02 malaria vaccine against Plasmodium falciparum infection in semi-immune adult men in The Gambia: a randomised trial. Lancet 358, 1927-1934.

Bowman, S., Lawson, D., Basham, D., Brown, D., Chillingworth, T., Churcher, C.M., Craig, A., Davies, R.M., Devlin, K., Feltwell, T., et al. (1999). The complete nucleotide sequence of chromosome 3 of Plasmodium falciparum. Nature 400, 532-538.

157

Bozdech, Z., Llinas, M., Pulliam, B.L., Wong, E.D., Zhu, J., and DeRisi, J.L. (2003). The transcriptome of the intraerythrocytic developmental cycle of Plasmodium falciparum. PLoS biology 1, E5.

Branco, M.R., and Pombo, A. (2007). Chromosome organization: new facts, new models. Trends in cell biology 17, 127-134.

Briquet, S., Boschet, C., Gissot, M., Tissandie, E., Sevilla, E., Franetich, J.F., Thiery, I., Hamid, Z., Bourgouin, C., and Vaquero, C. (2006). High-mobility-group box nuclear factors of Plasmodium falciparum. Eukaryot Cell 5, 672-682.

Callebaut, I., Prat, K., Meurice, E., Mornon, J.P., and Tomavo, S. (2005). Prediction of the general transcription factors associated with RNA polymerase II in Plasmodium falciparum: conserved features and differences relative to other eukaryotes. BMC genomics 6, 100.

Cary, C., Lamont, D., Dalton, J.P., and Doerig, C. (1994). Plasmodium falciparum chromatin: nucleosomal organisation and histone- like proteins. Parasitol Res 80, 255-258.

Chakrabarty, S.P., Saikumari, Y.K., Bopanna, M.P., and Balaram, H. (2008). Biochemical characterization of Plasmodium falciparum Sir2, a NAD+-dependent deacetylase. Molecular and biochemical parasitology 158, 139-151.

Chandra, B.R., Olivieri, A., Silvestrini, F., Alano, P., and Sharma, A. (2005). Biochemical characterization of the two nucleosome assembly proteins from Plasmodium falciparum. Molecular and biochemical parasitology 142, 237-247.

Chaumeil, J., Le Baccon, P., Wutz, A., and Heard, E. (2006). A novel role for Xist RNA in the formation of a repressive nuclear compartment into which genes are recruited when silenced. Genes & development 20, 2223-2237.

Choi, S.W., Keyes, M.K., and Horrocks, P. (2006). LC/ESI-MS demonstrates the absence of 5-methyl-2'- deoxycytosine in Plasmodium falciparum genomic DNA. Molecular and biochemical parasitology 150, 350-352.

Chookajorn, T., Dzikowski, R., Frank, M., Li, F., Jiwani, A.Z., Hartl, D.L., and Deitsch, K.W. (2007). Epigenetic memory at malaria virulence genes. Proceedings of the National Academy of Sciences of the United States of America 104, 899-902.

Christophides, G.K. (2005). Transgenic mosquitoes and malaria transmission. Cell Microbiol 7, 325-333.

Coleman, B.I., and Duraisingh, M.T. (2008). Transcriptional control and gene silencing in Plasmodium falciparum. Cell Microbiol.

Coulson, R.M., Hall, N., and Ouzounis, C.A. (2004). Comparative genomics of transcriptional control in the human malaria parasite Plasmodium falciparum. Genome research 14, 1548-1554.

Cui, L., Fan, Q., Cui, L., and Miao, J. (2008a). Histone lysine methyltransferases and demethylases in Plasmodium falciparum. International journal for parasitology 38, 1083-1097.

Cui, L., Miao, J., Furuya, T., Fan, Q., Li, X., Rathod, P.K., Su, X.Z., and Cui, L. (2008b). Histone acetyltransferase inhibitor anacardic acid causes changes in global gene expression during in vitro Plasmodium falciparum development. Eukaryot Cell 7, 1200-1210.

Cui, L., Miao, J., Furuya, T., Li, X., Su, X.Z., and Cui, L. (2007). PfGCN5-mediated histone H3 acetylation plays a key role in gene expression in Plasmodium falciparum. Eukaryot Cell 6, 1219-1227.

158

De Silva, E.K., Gehrke, A.R., Olszewski, K., Leon, I., Chahal, J.S., Bulyk, M.L., and Llinas, M. (2008). Specific DNA-binding by apicomplexan AP2 transcription factors. Proceedings of the National Academy of Sciences of the United States of America 105, 8393-8398.

Deitsch, K.W., Calderwood, M.S., and Wellems, T.E. (2001). Malaria. Cooperative silencing elements in var genes. Nature 412, 875-876.

Dillon, N. (2006). Gene regulation and large-scale chromatin organization in the nucleus. Chromosome Res 14, 117-126.

Duraisingh, M.T., Voss, T.S., Marty, A.J., Duffy, M.F., Good, R.T., Thompson, J.K., Freitas-Junior, L.H., Scherf, A., Crabb, B.S., and Cowman, A.F. (2005). Heterochromatin silencing and locus repositioning linked to regulation of virulence genes in Plasmodium falciparum. Cell 121, 13-24.

Dzikowski, R., Frank, M., and Deitsch, K. (2006). Mutually exclusive expression of virulence genes by malaria parasites is regulated independently of antigen production. PLoS pathogens 2, e22.

Dzikowski, R., Li, F., Amulic, B., Eisberg, A., Frank, M., Patel, S., Wellems, T.E., and Deitsch, K.W. (2007). Mechanisms underlying mutually exclusive expression of virulence genes by malaria parasites. EMBO Rep 8, 959-965.

Ekland, E.H., and Fidock, D.A. (2007). Advances in understanding the genetic basis of antimalarial drug resistance. Current opinion in microbiology 10, 363-370.

Fan, Q., An, L., and Cui, L. (2004). Plasmodium falciparum Histone Acetyltransferase, a Yeast GCN5 Homologue Involved in Chromatin Remodeling. Eukaryot Cell 3, 264-276.

Fang, J., and McCutchan, T.F. (2002). Thermoregulation in a parasite's life cycle. Nature 418, 742.

Fang, J., Sullivan, M., and McCutchan, T.F. (2004). The effects of glucose concentration on the reciprocal regulation of rRNA promoters in Plasmodium falciparum. The Journal of biological chemistry 279, 720- 725.

Figueiredo, L.M., Freitas-Junior, L.H., Bottius, E., Olivo-Marin, J.C., and Scherf, A. (2002). A central role for Plasmodium falciparum subtelomeric regions in spatial positioning and telomere length regulation. The EMBO journal 21, 815-824.

Figueiredo, L.M., Pirrit, L.A., and Scherf, A. (2000). Genomic organisation and chromatin structure of Plasmodium falciparum chromosome ends. Molecular & Biochemical Parasitology 106, 169-174.

Fox, B.A., Li, W.B., Tanaka, M., Inselburg, J., and Bzik, D.J. (1993). Molecular characterization of the largest subunit of Plasmodium falciparum RNA polymerase I. Molecular and biochemical parasitology 61, 37-48.

Frank, M., Dzikowski, R., Costantini, D., Amulic, B., Berdougo, E., and Deitsch, K. (2006). Strict pairing of var promoters and introns is required for var gene silencing in the malaria parasite Plasmodium falciparum. The Journal of biological chemistry 281, 9942-9952.

Freitas-Junior, L.H., Bottius, E., Pirrit, L.A., Deitsch, K.W., Scheidig, C., Guinet, F., Nehrbass, U., Wellems, T.E., and Scherf, A. (2000). Frequent ectopic recombination of virulence factor genes in telomeric chromosome clusters of P. falciparum. Nature 407, 1018-1022.

Freitas-Junior, L.H., Hernandez-Rivas, R., Ralph, S.A., Montiel-Condado, D., Ruvalcaba-Salazar, O.K., Rojas-Meza, A.P., Mancio-Silva, L., Leal-Silvestre, R.J., Gontijo, A.M., Shorte, S., et al. (2005). Telomeric

159

heterochromatin propagation and histone acetylation control mutually exclusive expression of antigenic variation genes in malaria parasites. Cell 121, 25-36.

Fujioka, H., and Aikawa, M. (1999). The malaria parasite and its life cycle. Malaria - Molecular and Clinical Aspects, ed Wahlgren M, Perlmann P, 19-55.

Garcia-Cao, M., O'Sullivan, R., Peters, A.H., Jenuwein, T., and Blasco, M.A. (2004). Epigenetic regulation of telomere length in mammalian cells by the Suv39h1 and Suv39h2 histone methyltransferases. Nature genetics 36, 94-99.

Gardiner, D.L., Holt, D.C., Thomas, E.A., Kemp, D.J., and Trenholme, K.R. (2000). Inhibition of Plasmodium falciparum clag9 gene function by antisense RNA. Molecular & Biochemical Parasitology 110, 33-41.

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. (2002). Genome sequence of the human malaria parasite Plasmodium falciparum. Nature 419, 498-511.

Gaszner, M., and Felsenfeld, G. (2006). Insulators: exploiting transcriptional and epigenetic mechanisms. Nature reviews 7, 703-713.

Gottlieb, S., and Esposito, R.E. (1989). A new role for a yeast transcriptional silencer gene, SIR2, in regulation of recombination in ribosomal DNA. Cell 56, 771-776.

Greenwood, B.M., Fidock, D.A., Kyle, D.E., Kappe, S.H., Alonso, P.L., Collins, F.H., and Duffy, P.E. (2008). Malaria: progress, perils, and prospects for eradication. The Journal of clinical investigation 118, 1266-1276.

Grewal, S.I., and Jia, S. (2007). Heterochromatin revisited. Nature reviews 8, 35-46.

Grummt, I., and Ladurner, A.G. (2008). A metabolic throttle regulates the epigenetic state of rDNA. Cell 133, 577-580.

Gunasekera, A.M., Patankar, S., Schug, J., Eisen, G., Kissinger, J., Roos, D., and Wirth, D.F. (2004). Widespread distribution of antisense transcripts in the Plasmodium falciparum genome. Molecular and biochemical parasitology 136, 35-42.

Gunderson, J.H., Sogin, M.L., Wollett, G., Hollingdale, M., de la Cruz, V.F., Waters, A.P., and McCutchan, T.F. (1987). Structurally distinct, stage-specific ribosomes occur in Plasmodium. Science (New York, NY 238, 933-937.

Hakimi, M.A., and Deitsch, K.W. (2007). Epigenetics in Apicomplexa: control of gene expression during cell cycle progression, differentiation and antigenic variation. Current opinion in microbiology 10, 357-362.

Hall, I.M., Shankaranarayana, G.D., Noma, K., Ayoub, N., Cohen, A., and Grewal, S.I. (2002). Establishment and maintenance of a heterochromatin domain. Science (New York, NY 297, 2232-2237.

Horrocks, P., Dechering, K., and Lanzer, M. (1998). Control of gene expression in Plasmodium falciparum. Molecular & Biochemical Parasitology 95, 171-181.

Ishii, K., Arib, G., Lin, C., Van Houwe, G., and Laemmli, U.K. (2002). Chromatin boundaries in budding yeast: the nuclear pore connection. Cell 109, 551-562.

160

Issar, N., Roux, E., Mattei, D., and Scherf, A. (2008). Identification of a novel post-translational modification in Plasmodium falciparum: Protein SUMOylation in different cellular compartments. Cell Microbiol.

Iyer, L.M., Anantharaman, V., Wolf, M.Y., and Aravind, L. (2008). Comparative genomics of transcription factors and chromatin proteins in parasitic protists and other eukaryotes. International journal for parasitology 38, 1-31.

Ji, D.D., and Arnot, D.E. (1997). A Plasmodium falciparum homologue of the ATPase subunit of a multi- protein complex involved in chromatin remodelling for transcription. Molecular and biochemical parasitology 88, 151-162.

Joshi, M.B., Lin, D.T., Chiang, P.H., Goldman, N.D., Fujioka, H., Aikawa, M., and Syin, C. (1999). Molecular cloning and nuclear localization of a histone deacetylase homologue in Plasmodium falciparum. Molecular and biochemical parasitology 99, 11-19.

Kelly, J.M., McRobert, L., and Baker, D.A. (2006). Evidence on the chromosomal location of centromeric DNA in Plasmodium falciparum from etoposide-mediated topoisomerase-II cleavage. Proceedings of the National Academy of Sciences of the United States of America 103, 6706-6711.

Khan, S.M., Franke-Fayard, B., Mair, G.R., Lasonder, E., Janse, C.J., Mann, M., and Waters, A.P. (2005). Proteome analysis of separated male and female gametocytes reveals novel sex-specific Plasmodium biology. Cell 121, 675-687.

Kobayashi, T., Horiuchi, T., Tongaonkar, P., Vu, L., and Nomura, M. (2004). SIR2 regulates recombination between different rDNA repeats, but not recombination within individual rRNA genes in yeast. Cell 117, 441-453.

Komaki-Yasuda, K., Okuwaki, M., Kano, S., Nagata, K., and Kawazu, S.I. (2008). 5' sequence- and chromatin modification-dependent gene expression in Plasmodium falciparum erythrocytic stage. Molecular and biochemical parasitology.

Kouzarides, T. (2007). Chromatin modifications and their function. Cell 128, 693-705.

Kraemer, S.M., Kyes, S.A., Aggarwal, G., Springer, A.L., Nelson, S.O., Christodoulou, Z., Smith, L.M., Wang, W., Levin, E., Newbold, C.I., et al. (2007). Patterns of gene recombination shape var gene repertoires in Plasmodium falciparum: comparisons of geographically diverse isolates. BMC genomics 8, 45.

Kyes, S., Christodoulou, Z., Pinches, R., Kriek, N., Horrocks, P., and Newbold, C. (2007a). Plasmodium falciparum var gene expression is developmentally controlled at the level of RNA polymerase II-mediated transcription initiation. Mol Microbiol 63, 1237-1247.

Kyes, S., Christodoulou, Z., Pinches, R., and Newbold, C. (2002). Stage-specific merozoite surface protein 2 antisense transcripts in Plasmodium falciparum. Molecular and biochemical parasitology 123, 79-83.

Kyes, S., Pinches, R., and Newbold, C. (2000). A simple RNA analysis method shows var and rif multigene family expression patterns in Plasmodium falciparum. Molecular & Biochemical Parasitology 105, 311-315.

Kyes, S.A., Christodoulou, Z., Raza, A., Horrocks, P., Pinches, R., Rowe, J.A., and Newbold, C.I. (2003). A well-conserved Plasmodium falciparum var gene shows an unusual stage-specific transcript pattern. Mol Microbiol 48, 1339-1348.

161

Kyes, S.A., Kraemer, S.M., and Smith, J.D. (2007b). Antigenic variation in Plasmodium falciparum: gene organization and regulation of the var multigene family. Eukaryot Cell 6, 1511-1520.

LaCount, D.J., Vignali, M., Chettier, R., Phansalkar, A., Bell, R., Hesselberth, J.R., Schoenfeld, L.W., Ota, I., Sahasrabudhe, S., Kurschner, C., et al. (2005). A protein interaction network of the malaria parasite Plasmodium falciparum. Nature 438, 103-107.

Landeira, D., and Navarro, M. (2007). Nuclear repositioning of the VSG promoter during developmental silencing in Trypanosoma brucei. The Journal of cell biology 176, 133-139.

Langsley, G., Hyde, J.E., Goman, M., and Scaife, J.G. (1983). Cloning and characterisation of the rRNA genes from the human malaria parasite Plasmodium falciparum. Nucleic acids research 11, 8703-8717.

Lanzer, M., De Bruin, D., and Ravetch, J.V. (1993). Transcriptional differences in polymorphic and conserved domains of a complete cloned P. falciparum chromosome. Nature 361, 654-657.

Latham, J.A., and Dent, S.Y. (2007). Cross-regulation of histone modifications. Nature structural & molecular biology 14, 1017-1024.

Lavstsen, T., Salanti, A., Jensen, A.T., Arnot, D.E., and Theander, T.G. (2003). Sub-grouping of Plasmodium falciparum 3D7 var genes based on sequence analysis of coding and non-coding regions. Malaria journal 2, 27.

Le Roch, K.G., Johnson, J.R., Florens, L., Zhou, Y., Santrosyan, A., Grainger, M., Yan, S.F., Williamson, K.C., Holder, A.A., Carucci, D.J., et al. (2004). Global analysis of transcript and protein levels across the Plasmodium falciparum life cycle. Genome research 14, 2308-2318.

Le Roch, K.G., Zhou, Y., Blair, P.L., Grainger, M., Moch, J.K., Haynes, J.D., De La Vega, P., Holder, A.A., Batalov, S., Carucci, D.J., et al. (2003). Discovery of gene function by expression profiling of the malaria parasite life cycle. Science (New York, NY 301, 1503-1508.

Lewin, B. (2000). Genes: clusters and repeats. Genes VII, 89-115.

Li, B., Carey, M., and Workman, J.L. (2007). The role of chromatin during transcription. Cell 128, 707- 719.

Li, F., Sonbuchner, L., Kyes, S.A., Epp, C., and Deitsch, K.W. (2008). Nuclear non-coding RNAs are transcribed from the centromeres of Plasmodium falciparum and are associated with centromeric chromatin. The Journal of biological chemistry 283, 5692-5698.

Li, J., McConkey, G.A., Rogers, M.J., Waters, A.P., and McCutchan, T.R. (1994). Plasmodium: the developmentally regulated ribosome. Experimental parasitology 78, 437-441.

Li, W.B., Bzik, D.J., Gu, H.M., Tanaka, M., Fox, B.A., and Inselburg, J. (1989). An enlarged largest subunit of Plasmodium falciparum RNA polymerase II defines conserved and variable RNA polymerase domains. Nucleic acids research 17, 9621-9636.

Li, W.B., Bzik, D.J., Tanaka, M., Gu, H.M., Fox, B.A., and Inselburg, J. (1991). Characterization of the gene encoding the largest subunit of Plasmodium falciparum RNA polymerase III. Molecular and biochemical parasitology 46, 229-239.

Lopez-Rubio, J.J., Gontijo, A.M., Nunes, M.C., Issar, N., Hernandez Rivas, R., and Scherf, A. (2007a). 5' flanking region of var genes nucleate histone modification patterns linked to phenotypic inheritance of virulence traits in malaria parasites. Mol Microbiol.

162

Lopez-Rubio, J.J., Riviere, L., and Scherf, A. (2007b). Shared epigenetic mechanisms control virulence factors in protozoan parasites. Current opinion in microbiology 10, 560-568.

Mair, G.R., Braks, J.A., Garver, L.S., Wiegant, J.C., Hall, N., Dirks, R.W., Khan, S.M., Dimopoulos, G., Janse, C.J., and Waters, A.P. (2006). Regulation of sexual development of Plasmodium by translational repression. Science (New York, NY 313, 667-669.

Marty, A.J., Thompson, J.K., Duffy, M.F., Voss, T.S., Cowman, A.F., and Crabb, B.S. (2006). Evidence that Plasmodium falciparum chromosome end clusters are cross-linked by protein and are the sites of both virulence gene silencing and activation. Mol Microbiol 62, 72-83.

Matuschewski, K., and Mueller, A.K. (2007). Vaccines against malaria - an update. The FEBS journal 274, 4680-4687.

McAndrew, M.B., Read, M., Sims, P.F., and Hyde, J.E. (1993). Characterisation of the gene encoding an unusually divergent TATA-binding protein (TBP) from the extremely A+T-rich human malaria parasite Plasmodium falciparum. Gene 124, 165-171.

McCutchan, T.F., de la Cruz, V.F., Lal, A.A., Gunderson, J.H., Elwood, H.J., and Sogin, M.L. (1988). Primary sequences of two small subunit ribosomal RNA genes from Plasmodium falciparum. Molecular & Biochemical Parasitology 28, 63-68.

McCutchan, T.F., van Spaendonk, R.M., Janse, C.J., Fang, J., and Waters, A.P. (2004). Regulation of rRNA transcription and processing during the Plasmodium life cycle. Malaria Parasites: Genomes and Molecular Biology, ed AP Waters and CJ Janse, 229-247.

Mehra, P., Biswas, A.K., Gupta, A., Gourinath, S., Chitnis, C.E., and Dhar, S.K. (2005). Expression and characterization of human malaria parasite Plasmodium falciparum origin recognition complex subunit 1. Biochemical and biophysical research communications 337, 955-966.

Mellor, J. (2006). Dynamic nucleosomes and gene transcription. Trends Genet 22, 320-329.

Meneghini, M.D., Wu, M., and Madhani, H.D. (2003). Conserved histone variant H2A.Z protects euchromatin from the ectopic spread of silent heterochromatin. Cell 112, 725-736.

Mercereau-Puijalon, O., Barale, J.C., and Bischoff, E. (2002). Three multigene families in Plasmodium parasites: facts and questions. International journal for parasitology 32, 1323-1344.

Merrick, C.J., and Duraisingh, M.T. (2007). Plasmodium Falciparum Sir2: an Unusual Sirtuin with Dual Histone Deacetylase and Adp-Ribosyltransferase Activity. Eukaryot Cell.

Miao, J., Fan, Q., Cui, L., and Li, J. (2006). The malaria parasite Plasmodium falciparum histones: Organization, expression, and acetylation. Gene 369, 53-65.

Militello, K.T., Patel, V., Chessler, A.D., Fisher, J.K., Kasper, J.M., Gunasekera, A., and Wirth, D.F. (2005). RNA polymerase II synthesizes antisense RNA in Plasmodium falciparum. RNA (New York, NY 11, 365-370.

Mok, B.W., Ribacke, U., Rasti, N., Kironde, F., Chen, Q., Nilsson, P., and Wahlgren, M. (2008). Default Pathway of var2csa switching and translational repression in Plasmodium falciparum. PLoS ONE 3, e1982.

Mondoux, M.A., and Zakian, V.A. (2006). Telomere position effect: silencing near the end. Cold Spring Harbor Laboratory Press Telomeres, 2nd Ed., 261-317.

163

Mourier, T., Carret, C., Kyes, S., Christodoulou, Z., Gardner, P.P., Jeffares, D.C., Pinches, R., Barrell, B., Berriman, M., Griffiths-Jones, S., et al. (2008). Genome-wide discovery and verification of novel structured RNAs in Plasmodium falciparum. Genome research 18, 281-292.

Murayama, A., Ohmori, K., Fujimura, A., Minami, H., Yasuzawa-Tanaka, K., Kuroda, T., Oie, S., Daitoku, H., Okuwaki, M., Nagata, K., et al. (2008). Epigenetic control of rDNA loci in response to intracellular energy status. Cell 133, 627-639.

Nakayama, J., Rice, J.C., Strahl, B.D., Allis, C.D., and Grewal, S.I. (2001). Role of histone H3 lysine 9 methylation in epigenetic control of heterochromatin assembly. Science (New York, NY 292, 110-113.

Navadgi, V.M., Chandra, B.R., Mishra, P.C., and Sharma, A. (2006). The two Plasmodium falciparum nucleosome assembly proteins play distinct roles in histone transport and chromatin assembly. The Journal of biological chemistry 281, 16978-16984.

Navarro, M., and Gull, K. (2001). A pol I transcriptional body associated with VSG mono-allelic expression in Trypanosoma brucei. Nature 414, 759-763.

O'Donnell, R.A., Freitas-Junior, L.H., Preiser, P.R., Williamson, D.H., Duraisingh, M., McElwain, T.F., Scherf, A., Cowman, A.F., and Crabb, B.S. (2002). A genetic screen for improved plasmid segregation reveals a role for Rep20 in the interaction of Plasmodium falciparum chromosomes. The EMBO journal 21, 1231-1239.

Oki, M., and Kamakaka, R.T. (2005). Barrier function at HMR. Molecular cell 19, 707-716.

Oki, M., Valenzuela, L., Chiba, T., Ito, T., and Kamakaka, R.T. (2004). Barrier proteins remodel and modify chromatin to restrict silenced domains. Molecular and cellular biology 24, 1956-1967.

Olson, M. (2004). Nontraditional roles of the nucleolus. The Nucleolus, ed Olson M, 330-342.

Pak, D.T., Pflumm, M., Chesnokov, I., Huang, D.W., Kellum, R., Marr, J., Romanowski, P., and Botchan, M.R. (1997). Association of the origin recognition complex with heterochromatin and HP1 in higher eukaryotes. Cell 91, 311-323.

Patankar, S., Munasinghe, A., Shoaibi, A., Cummings, L.M., and Wirth, D.F. (2001). Serial analysis of gene expression in Plasmodium falciparum reveals the global expression profile of erythrocytic stages and the presence of anti-sense transcripts in the malarial parasite. Molecular biology of the cell 12, 3114-3125.

Ralph, S.A., Bischoff, E., Mattei, D., Sismeiro, O., Dillies, M.A., Guigon, G., Coppee, J.Y., David, P.H., and Scherf, A. (2005a). Transcriptome analysis of antigenic variation in Plasmodium falciparum--var silencing is not dependent on antisense RNA. Genome biology 6, R93.

Ralph, S.A., Scheidig-Benatar, C., and Scherf, A. (2005b). Antigenic variation in Plasmodium falciparum is associated with movement of var loci between subnuclear locations. Proceedings of the National Academy of Sciences of the United States of America 102, 5414-5419.

Rea, S., Eisenhaber, F., O'Carroll, D., Strahl, B.D., Sun, Z.W., Schmid, M., Opravil, S., Mechtler, K., Ponting, C.P., Allis, C.D., et al. (2000). Regulation of chromatin structure by site-specific histone H3 methyltransferases. Nature 406, 593-599.

Ridley, R.G. (2002). Medical need, scientific opportunity and the drive for antimalarial drugs. Nature 415, 686-693.

Ridley, R.G. (2003). Malaria: to kill a parasite. Nature 424, 887-889.

164

Sacarlal, J., Aponte, J.J., Aide, P., Mandomando, I., Bassat, Q., Guinovart, C., Leach, A., Milman, J., Macete, E., Espasa, M., et al. (2008). Safety of the RTS,S/AS02A malaria vaccine in Mozambican children during a Phase IIb trial. Vaccine 26, 174-184.

Sautel, C.F., Cannella, D., Bastien, O., Kieffer, S., Aldebert, D., Garin, J., Tardieux, I., Belrhali, H., and Hakimi, M.A. (2007). SET8-mediated methylations of histone H4 lysine 20 mark silent heterochromatic domains in apicomplexan genomes. Molecular and cellular biology 27, 5711-5724.

Scherf, A., Figueiredo, L.M., and Freitas-Junior, L.H. (2001). Plasmodium telomeres: a pathogen's perspective. Current opinion in microbiology 4, 409-414.

Scherf, A., Hernandez-Rivas, R., Buffet, P., Bottius, E., Benatar, C., Pouvelle, B., Gysin, J., and Lanzer, M. (1998). Antigenic variation in malaria: in situ switching, relaxed and mutually exclusive transcription of var genes during intra-erythrocytic development in Plasmodium falciparum. The EMBO journal 17, 5418- 5426.

Scherf, A., Lopez-Rubio, J.J., and Riviere, L. (2008a). Antigenic Variation in Plasmodium falciparum. Annu Rev Microbiol 62, 445-470.

Scherf, A., Riviere, L., and Lopez-Rubio, J.J. (2008b). SnapShot: var gene expression in the malaria parasite. Cell 134, 190.

Schieck, E., Pfahler, J.M., Sanchez, C.P., and Lanzer, M. (2007). Nuclear run-on analysis of var gene expression in Plasmodium falciparum. Molecular and biochemical parasitology 153, 207-212.

Schlitzer, M. (2008). Antimalarial drugs - what is in use and what is in the pipeline. Archiv der Pharmazie 341, 149-163.

Sexton, T., Schober, H., Fraser, P., and Gasser, S.M. (2007). Gene regulation through nuclear organization. Nature structural & molecular biology 14, 1049-1055.

Shankaranarayana, G.D., Motamedi, M.R., Moazed, D., and Grewal, S.I. (2003). Sir2 regulates histone H3 lysine 9 methylation and heterochromatin assembly in fission yeast. Curr Biol 13, 1240-1246.

Shareef, M.M., Badugu, R., and Kellum, R. (2003). HP1/ORC complex and heterochromatin assembly. Genetica 117, 127-134.

Shock, J.L., Fischer, K.F., and DeRisi, J.L. (2007). Whole-genome analysis of mRNA decay in Plasmodium falciparum reveals a global lengthening of mRNA half-life during the intra-erythrocytic development cycle. Genome biology 8, R134.

Shogren-Knaak, M., Ishii, H., Sun, J.M., Pazin, M.J., Davie, J.R., and Peterson, C.L. (2006). Histone H4- K16 acetylation controls chromatin structure and protein interactions. Science (New York, NY 311, 844- 847.

Smith, J.D., Chitnis, C.E., Craig, A.G., Roberts, D.J., Hudson-Taylor, D.E., Peterson, D.S., Pinches, R., Newbold, C.I., and Miller, L.H. (1995). Switches in expression of Plasmodium falciparum var genes correlate with changes in antigenic and cytoadherent phenotypes of infected erythrocytes [see comments]. Cell 82, 101-110.

Smith, J.S., and Boeke, J.D. (1997). An unusual form of transcriptional silencing in yeast ribosomal DNA. Genes & development 11, 241-254.

165

Squazzo, S.L., O'Geen, H., Komashko, V.M., Krig, S.R., Jin, V.X., Jang, S.W., Margueron, R., Reinberg, D., Green, R., and Farnham, P.J. (2006). Suz12 binds to silenced regions of the genome in a cell-type- specific manner. Genome research 16, 890-900.

Su, X.Z., Heatwole, V.M., Wertheimer, S.P., Guinet, F., Herrfeldt, J.A., Peterson, D.S., Ravetch, J.A., and Wellems, T.E. (1995). The large diverse gene family var encodes proteins involved in cytoadherence and antigenic variation of Plasmodium falciparum-infected erythrocytes [see comments]. Cell 82, 89-100.

Sullivan, W.J., Jr., Naguleswaran, A., and Angel, S.O. (2006). Histones and histone modifications in protozoan parasites. Cell Microbiol 8, 1850-1861.

Suzuki, M.M., and Bird, A. (2008). DNA methylation landscapes: provocative insights from epigenomics. Nature reviews 9, 465-476.

Taddei, A., Van Houwe, G., Hediger, F., Kalck, V., Cubizolles, F., Schober, H., and Gasser, S.M. (2006). Nuclear pore association confers optimal expression levels for an inducible yeast gene. Nature 441, 774- 778.

Templeton, T.J., Iyer, L.M., Anantharaman, V., Enomoto, S., Abrahante, J.E., Subramanian, G.M., Hoffman, S.L., Abrahamsen, M.S., and Aravind, L. (2004). Comparative analysis of apicomplexa and genomic diversity in eukaryotes. Genome research 14, 1686-1695.

Todryk, S.M., and Hill, A.V. (2007). Malaria vaccines: the stage we are at. Nat Rev Microbiol 5, 487-489.

Turner, B.M. (2000). Histone acetylation and an epigenetic code. Bioessays 22, 836-845.

Vakoc, C.R., Mandat, S.A., Olenchock, B.A., and Blobel, G.A. (2005). Histone H3 lysine 9 methylation and HP1gamma are associated with transcription elongation through mammalian chromatin. Molecular cell 19, 381-391.

Vakoc, C.R., Sachdeva, M.M., Wang, H., and Blobel, G.A. (2006). Profile of histone lysine methylation across transcribed mammalian chromatin. Molecular and cellular biology 26, 9185-9195. van Noort, V., and Huynen, M.A. (2006). Combinatorial gene regulation in Plasmodium falciparum. Trends Genet 22, 73-78. van Spaendonk, R.M., Ramesar, J., van Wigcheren, A., Eling, W., Beetsma, A.L., van Gemert, G.J., Hooghof, J., Janse, C.J., and Waters, A.P. (2001). Functional equivalence of structurally distinct ribosomes in the malaria parasite, Plasmodium berghei. The Journal of biological chemistry 276, 22638-22647.

Voss, T.S., Healer, J., Marty, A.J., Duffy, M.F., Thompson, J.K., Beeson, J.G., Reeder, J.C., Crabb, B.S., and Cowman, A.F. (2006). A var gene promoter controls allelic exclusion of virulence genes in Plasmodium falciparum malaria. Nature 439, 1004-1008.

Voss, T.S., Tonkin, C.J., Marty, A.J., Thompson, J.K., Healer, J., Crabb, B.S., and Cowman, A.F. (2007). Alterations in local chromatin environment are involved in silencing and activation of subtelomeric var genes in Plasmodium falciparum. Mol Microbiol 66, 139-150.

Wang, Z., Zang, C., Rosenfeld, J.A., Schones, D.E., Barski, A., Cuddapah, S., Cui, K., Roh, T.Y., Peng, W., Zhang, M.Q., et al. (2008). Combinatorial patterns of histone acetylations and methylations in the human genome. Nature genetics 40, 897-903.

Waters, A.P., Mota, M.M., van Dijk, M.R., and Janse, C.J. (2005). Parasitology. Malaria vaccines: back to the future? Science (New York, NY 307, 528-530.

166

Waters, A.P., Syin, C., and McCutchan, T.F. (1989). Developmental regulation of stage-specific ribosome populations in Plasmodium. Nature 342, 438-440.

Wellems, T.E., Walliker, D., Smith, C.L., do Rosario, V.E., Maloy, W.L., Howard, R.J., Carter, R., and McCutchan, T.F. (1987). A histidine-rich protein gene marks a linkage group favored strongly in a genetic cross of Plasmodium falciparum. Cell 49, 633-642.

WHO (2006). Guidelines for the treatment of malaria. WHO.

WHO (2008). Global Malaria Control and Elimination. WHO.

Williams, S.K., and Tyler, J.K. (2007). Transcriptional regulation by chromatin disassembly and reassembly. Current opinion in genetics & development 17, 88-93.

Wirth, D.F. (2002). Biological revelations. Nature 419, 495-496.

Young, J.A., Fivelman, Q.L., Blair, P.L., de la Vega, P., Le Roch, K.G., Zhou, Y., Carucci, D.J., Baker, D.A., and Winzeler, E.A. (2005). The Plasmodium falciparum sexual development transcriptome: a microarray analysis using ontology-based pattern identification. Molecular and biochemical parasitology 143, 67-79.

Young, J.A., Johnson, J.R., Benner, C., Yan, S.F., Chen, K., Le Roch, K.G., Zhou, Y., and Winzeler, E.A. (2008). In silico discovery of transcription regulatory elements in Plasmodium falciparum. BMC genomics 9, 70.

167