Next Article in Journal
Undiagnosed Cases of Human Pneumonia Following Exposure to Chlamydia psittaci from an Infected Rosella Parrot
Previous Article in Journal
Detection and Characterization of Clinical Bordetella trematum Isolates from Chronic Wounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulation of Antigenic Variation by Trypanosoma brucei Telomere Proteins Depends on Their Unique DNA Binding Activities

1
Center for Gene Regulation in Health and Disease, Department of Biological, Geological, and Environmental Sciences, College of Sciences and Health Professions, Cleveland State University, 2121 Euclid Avenue, Cleveland, OH 44115, USA
2
Case Comprehensive Cancer Center, Case Western Reserve University, 10900 Euclid Avenue, Cleveland, OH 44106, USA
3
Department of Inflammation and Immunity, Lerner Research Institute, Cleveland Clinic, 9500 Euclid Avenue, Cleveland, OH 44195, USA
4
Center for RNA Science and Therapeutics, Case Western Reserve University, 10900 Euclid Avenue, Cleveland, OH 44106, USA
5
Shenzhen Research Institute, The Hong Kong Polytechnic University, Shenzhen, China
6
State Key Laboratory of Chemical Biology and Drug Discovery, Department of Applied Biology and Chemical Technology, The Hong Kong Polytechnic University, Hung Hom, Hong Kong, China
*
Authors to whom correspondence should be addressed.
Pathogens 2021, 10(8), 967; https://doi.org/10.3390/pathogens10080967
Submission received: 8 July 2021 / Revised: 22 July 2021 / Accepted: 27 July 2021 / Published: 30 July 2021
(This article belongs to the Section Parasitic Pathogens)

Abstract

:
Trypanosoma brucei causes human African trypanosomiasis and regularly switches its major surface antigen, Variant Surface Glycoprotein (VSG), to evade the host immune response. Such antigenic variation is a key pathogenesis mechanism that enables T. brucei to establish long-term infections. VSG is expressed exclusively from subtelomere loci in a strictly monoallelic manner, and DNA recombination is an important VSG switching pathway. The integrity of telomere and subtelomere structure, maintained by multiple telomere proteins, is essential for T. brucei viability and for regulating the monoallelic VSG expression and VSG switching. Here we will focus on T. brucei TRF and RAP1, two telomere proteins with unique nucleic acid binding activities, and summarize their functions in telomere integrity and stability, VSG switching, and monoallelic VSG expression. Targeting the unique features of TbTRF and TbRAP1′s nucleic acid binding activities to perturb the integrity of telomere structure and disrupt VSG monoallelic expression may serve as potential therapeutic strategy against T. brucei.

1. Antigenic Variation in T. brucei Involves Its Major Surface Antigen, Variant SURFACE Glycoprotein (VSG)

Trypanosoma brucei is a protozoan parasite that causes human African trypanosomiasis. It is transmitted by tsetse flies (Glossina spp.) that inhabit 36 sub-Saharan African countries, thus endangering ~60 million people in this region [1]. Two other very closely related trypanosomes, Trypanosoma cruzi and Leishmania, cause debilitating Chagas disease and Leishmaniasis in humans, respectively. Together, ~20 million people worldwide are infected by these kinetoplastid parasites [2]. In addition, T. brucei can infect livestock, which severely affects the economics in sub-Sahara Africa [3]. However, few drugs are available to treat these diseases safely and effectively with easy administering, and drug resistance cases have been observed [2].
T. brucei is heteroxenous and has several key life cycle stages. After ingested by its insect vector during a blood meal on an infected mammal, T. brucei proliferates in the tsetse midgut [4]. At this stage, procyclic form (PF) T. brucei expresses procyclins as its major surface proteins, as the C-termini of procyclins are resistant to tsetse midgut enzymes [5]. Subsequently, T. brucei migrates to tsetse’s proventriculus and then salivary glands, where it differentiates into the metacyclic form (MF) in the latter organ. MF T. brucei is not proliferative and expresses VSG as its major surface antigen. When tsetse takes another blood meal, the parasites can be injected into a new mammalian host [6]. After T. brucei enters its mammalian host, it differentiates into the bloodstream form (BF), which proliferates in extra cellular spaces of its host and expresses VSG as its major surface antigen [6]. VSG is highly immunogenic [7]. However, T. brucei undergoes antigenic variation and sequentially expresses immunologically distinct VSGs, thereby effectively evading the host immune response [8]. Antigenic variation is a key pathogenesis mechanism and is essential for T. brucei to establish long-term infections.

1.1. VSG Is Expressed in a Monoallelic Manner from a Single VSG Expression Site (ES)

T. brucei has a large VSG gene pool with >2500 VSG genes and pseudogenes [9]. Most of these are located in subtelomeric gene arrays (Figure 1A) [9,10] on the 11 pairs of megabase chromosomes, which are 0.9–5.7 Mb and contain all essential genes [11,12]. In addition, T. brucei has ~100 minichromosomes, which are 30–150 kb and mainly include central 177 bp repeats and terminal telomere repeats [12,13]. Individual VSG genes can be found on two thirds of the minichromosome telomeres (Figure 1B) [9], contributing to the VSG gene pool size. In distinct contrast to the fact that protein-coding genes are transcribed by RNA polymerase II (RNAP II), VSGs are transcribed by RNA polymerase I (RNAP I) [14] exclusively from BF VSG expression sites (ESs) (Figure 1C) while growing in its mammalian host and from MF ESs (Figure 1D) while residing in the tsetse salivary glands. BF VSG ESs are polycistronic transcription units (PTUs) [15] located at subtelomeres (Figure 1C) [16] of megabase chromosomes and, in one case, at the subtelomere of an intermediate chromosome (T. brucei has four to five intermediate chromosomes that are 200–700 kb [12]). VSG is the last gene in any ES and within 2 kb from the telomere repeats, while the ES promoter is located 40–60 kb upstream [15]. In contrast, MF ES are telomeric monocistronic transcription units with the ES promoter located ~5 kb upstream of the telomere (Figure 1D) [17].
T. brucei has multiple ESs (~15 BF ESs in the Lister 427 strain) [10,15], all with the same gene organization and ~90% sequence identity [15]. However, only one ES is fully active, expressing only one type of VSG on the cell surface at any time (Figure 1E) [18]. Monoallelic VSG expression has been shown to depend on multiple factors, although the detailed mechanisms of how it is achieved are still unclear. It has been shown that defects in a telomere protein TbRAP1 [19,20,21,22], chromatin structure [20,23,24,25,26,27,28,29,30,31,32], transcription elongation [33,34,35], the inositol phosphate pathway [36,37], nuclear lamina [38,39], recruitment of sumoylated protein(s) to the active ES promoter [40], DNA replication initiation factors [41,42,43,44], and a subtelomere and VSG-associated VEX complex [45,46,47] can all abolish VSG monoallelic expression. Although monoallelic VSG expression is not essential for ex vivo cultured cells, parasites expressing multiple VSGs are more quickly eliminated by its mammalian host [48], indicating that monoallelic VSG expression is important for the parasite virulence.

1.2. Several Pathways Mediate VSG Switching

Proliferative BF T. brucei undergoes VSG switching regularly [49]. Most VSG switching events are either transcriptional or DNA recombination-mediated (Figure 2) [50,51]. In a transcriptional in situ switch, the originally active ES is silenced while an originally silent ES becomes fully expressed, and no gene rearrangement is involved. Recombination-mediated switches have two major types. In gene conversion events, a silent VSG is duplicated into the active ES to replace the active VSG, which is subsequently lost. In crossover (or telomere exchange) events, a silent VSG (frequently together with its downstream telomere sequence) exchange places with the active VSG without any loss of genetic information. In addition, fragments of several silent VSG genes can also be recombined to form a new mosaic VSG gene in the originally active ES [52,53]. Furthermore, it has been observed that the active ES can simply be lost and a different ES derepressed during a switching event, and such “ES loss + in situ” switches can be frequently detected when telomere proteins are depleted [54,55]. In many recent studies, DNA recombination-mediated switches are much more prevalent than in situ switches [54,55,56,57,58,59,60,61,62,63].
Proteins involved in DNA recombination include RAD51 that mediates strand invasion in homologous recombination (HR) [64], RAD51-3 (a RAD51-related protein) [65], and BRCA2 [66] facilitate VSG switching. Proteins involved in DNA metabolism including Topoisomerase 3 alpha [62]; the RMI1 homolog [59]; a replication origin binding factor TbORC1 [41]; and a RecQ helicase, RECQ2 [67], suppress VSG switching. Additionally, telomere proteins have been shown to suppress VSG switching [54,55,63,68], while cells harboring a critically short active VSG-adjacent telomere have an increased VSG switching rate [61]. Furthermore, cells with defective RNase H enzymes that degrade the RNA strand in the DNA:RNA hybrid appear to have a higher VSG switching frequency than WT cells [69,70].
How VSG switching is initiated and regulated is not clear [71,72]. For recombination-mediated switching events, sequences flanking the VSG gene provide the sequence homology. All VSG 3′UTRs have the conserved 16 bp and 9 bp motifs with consensus sequences, and 70 bp repeats are located upstream of 70–80% of VSG genes [9,73]. The 70 bp repeats and VSG 3′UTRs can mediate efficient HR. In addition, telomere repeats are located immediately downstream of all ES-linked VSGs and VSGs on minichromosomes [9], which can also provide good sequence homology. Since HR is an efficient DNA damage repair mechanism with high fidelity [74], DNA double-strand breaks (DSBs) within the homologous sequences flanking the active VSG have been proposed to be a potent VSG switching inducer [71,72]. Indeed, introducing an I-SceI cut between the 70 bp repeats and the active VSG gene caused a 250-fold increase in the VSG switching rate [56,57]. In addition, DSBs can be detected at subtelomeres in WT T. brucei cells [56]. However, how DSB is naturally induced in the parasite, and whether this is the only initiator for VSG switching still awaits further investigations.
As VSG is expressed exclusively from subtelomeric loci; the telomere and subtelomere integrity has been shown to be a key factor influencing VSG switching frequency. Introducing a DSB immediately upstream of the active VSG increased the switching frequency by ~250 fold in vitro with most switching events mediated by DNA recombination [56,57]. In addition, T. brucei cells carrying a critically short telomere downstream of the active VSG have a VSG switching rate ~10-fold higher than that in cells with normal sized telomeres [61]. Furthermore, telomere proteins have been shown to play important roles in VSG monoallelic expression [19,20,21,22] and affect VSG switching frequencies [21,54,55,63,68]. Depletion of telomere proteins TbTRF, TbRAP1, or TbTIF2 induces more telomere/subtelomere DNA damage, disrupts the telomere/subtelomere stability, and results in significantly increased VSG switching rates [21,22,54,55,63,68,75].

2. Telomere Functions in Antigenic Variation

Linear chromosomes in eukaryotic cells are capped by a special nucleoprotein complex called telomere. In most eukaryotes, telomere DNA consists of a simple repetitive TG-rich sequences [76]. Although kinetoplastids branched away from vertebrates more than 500 million years ago during evolution, telomeres in both vertebrates and kinetoplastids including T. brucei contain TTAGGG repeats [76]. In most eukaryotes, telomeres end in a 3′ single-stranded overhang structure [77,78,79,80,81,82,83,84,85], which can invade the duplex telomere region and form a T-loop structure that has been observed in human, mouse, chicken, T. brucei, ciliates, common garden pea, C. elegans, and K. lactis [86,87,88,89,90,91,92]. Conventional DNA polymerases cannot fully replicate the ends of linear DNA molecules, resulting in the so-called “end replication problem” [93]. Many eukaryotes use a specialized reverse transcriptase, telomerase, to synthesize the G-rich telomere strand de novo according to a short template provided by the telomerase RNA component [94,95,96,97]. In the absence of the telomerase activity, telomere can be maintained by HR (such as breakage-induced replication) in ~15% cancer cells [98,99] and in telomerase null yeast survivors [100,101].

2.1. Telomeres Are Essential for Genome Integrity and Affect Nearby Gene Expression

A key function of the telomere is to mask the natural chromosome ends from being recognized by the DNA damage response machineries as a DNA damage site. Exposed telomeres are not only vulnerable to nucleolytic degradation but can also be processed to form chromosome end-to-end fusions that lead to the “breakage-fusion-bridge” cycle [102], which can induce loss of heterozygosity, non-reciprocal translocations, and gene amplification [103]. Indeed, recent studies in mammalian cells have shown that telomere end fusions can lead to chromothripsis and kataegis [104], while chromoanagenesis (including chromothripsis, chromoplexy, and chromoanasynthesis) is an important mechanism of genome instability that can contribute to tumorigenesis [105,106].
In humans, Shelterin [107]—a complex including six telomere proteins (TRF1 [108], TRF2 [109,110], RAP1 [111], TIN2 [112], TPP1 [113,114,115], and POT1 [116,117]) and the CST complex (CTC1/STN1/TEN1 in vertebrates and CDC13/STN1/TEN1 in budding yeast) [118,119] are key components of the telomere complex that are indispensable for chromosome end protection. TRF1 and TRF2 bind the duplex TTAGGG repeats [108,120,121,122,123] while POT1 [116,124] and the CST complex bind the single-stranded telomere G-overhang [125]. RAP1 interacts with TRF2 [111], and TIN2 interacts with both TRF1 and TRF2 [112,126], while TPP1 interacts with both TIN2 and POT1 [113,114,115]. The Shelterin components can help protect telomere stability by inhibition of Non-Homologous End Joining [127,128,129], HR [130,131,132,133], and Microhomology-Mediated End Joining [134,135] at the telomere. In addition, the T-loop structure buries the telomere G-overhang, which suppresses ATM activation at mammalian telomeres [136].
The telomere complex can also suppress the nearby gene expression [137]. This telomere position effect or telomeric silencing is an epigenetic phenomenon, depending on the telomere heterochromatic structure [138]. Telomeric silencing is best understood in budding yeast, where ScRAP1 is the predominant duplex telomere DNA binding factor [139,140,141]. ScRAP1 recruits SIR3 and SIR4 proteins [142,143,144,145,146], which in turn recruit SIR2, a histone deacetylase [147], to nucleate and maintain the telomere heterochromatic structure [148]. In general, genes located closer to the telomere are more strongly repressed than genes located further away [147]. In addition, longer telomere repeats have stronger telomeric silencing effects in budding yeast [149,150], presumably because more ScRAP1 proteins are recruited to the telomere DNA.
Although telomeres frequently form a heterochromatic structure, the Telomere Repeat-containing RNA (TERRA) has been detected in many organisms including T. brucei [63,75,151,152,153], several kinetoplastids and Plasmodium falciparum [151,152,154], human [155], mouse [156], fission [157] and budding yeasts [158], and birds [159]. TERRA is prone to form an R-loop structure with the telomere DNA (a three-stranded structure containing a DNA:RNA hybrid and a displaced ssDNA) [160]. Both TERRA and telomeric R-loop (TRL) have been shown to regulate telomerase-dependent and recombination-mediated telomere maintenance and also play a role in chromosome end protection [153,160,161]. T. brucei TERRA has a few unique features compared to that in human and yeast cells, where frequently multiple telomeres are transcribed by RNAP II [155,157,158,162,163,164,165,166,167,168]. First, in T. brucei, TERRA is transcribed by RNAP I as a read-through product when RNAP I continues into the telomeric repeats downstream of the active VSG [63,75,151,152]. Second, the active ES-adjacent telomere is the only TERRA transcription site [75]. Third, the number of nuclear TERRA foci is cell cycle-regulated in T. brucei [75]. Most G1 cells (~60%) have only a single TERRA focus, and the number of TERRA foci increases as cells enters S and G2/M stages [75].

2.2. DNA Binding Activities of Telomere Proteins Are Critical for Their Essential Functions

Telomere binding proteins apparently play pivotal roles in key telomere functions, as they nucleate the telomere nucleoprotein complex. Studies in yeasts, mammals, plants, ciliates, and kinetoplastids have shown that two major DNA binding activities—the Myb motif-mediated duplex telomere DNA binding [141,169,170] and the OB fold-mediated single-stranded telomere DNA binding [171]—are conserved across many species [172,173]. TRF1 and TRF2 are the first telomere proteins that have been found to bind the duplex TTAGGG repeats with their C-terminal Myb domains [108,109,110]. Frequently, two Myb domains are necessary for a robust DNA binding [174]. This was confirmed to be true in human TRF1/2 proteins [175,176]. Mammalian TRF1 and TRF2 have a TRF Homology (TRFH) domain towards their N-termini that can homodimerize, which allows TRF1 and TRF2 dimers to bind the duplex telomere DNA [108,109,110]. Hence, an elegant set of experiments were performed before the era of RNA interference, TELENs, or CRISPR/cas using dominant negative Myb domain deletion mutants of human TRF1 and TRF2 [175,176]. When overexpressed, Myb deletion TRF mutants tether the endogenous WT TRFs off the telomere through the TRFH-mediated dimerization, as the mutant-WT dimer only possesses a single Myb domain and cannot bind the telomere DNA sufficiently [175,176]. Fission yeast SpTAZ1 that binds the duplex telomere DNA was shown to be a functional homolog of TRF and to also have a C-terminal Myb motif [177,178]. Subsequently, Myb motif has been identified in a number of duplex telomere binding proteins in plants and yeasts [140,169,170]. It is worth mentioning that the TRF homolog in budding yeast, ScTBF1, does not bind the duplex telomere DNA but binds subtelomeric TTAGGG repeats [179], using its C-terminal Myb domain [180]. On the other hand, budding yeast telomeres contain imperfect repeats [(TG1-3)n in S. cerevisiae [76]] that are recognized by the RAP1 homologs [139]. Interestingly, ScRAP1 also uses Myb-type DNA binding motifs to bind the duplex telomere DNA [140,141], although this was only revealed when the crystal structure of the ScRAP1 central region was solved [140]. It turns out that ScRAP1 has a central Myb domain and a Myb-Like domain that coordinate for duplex DNA binding [140]. Therefore, even though ScRAP1 is not a homolog of mammalian TRFs, it still binds duplex telomere DNA using Myb motifs like TRFs. Both TRF and RAP1 homologs have been identified in T. brucei and are found to play important roles in VSG monoallelic expression and suppress VSG switching, which rely on their telomere DNA binding activities (see below).
Most known single-stranded telomere DNA binding proteins use OB folds to recognize the DNA [171], including hypotrichous ciliate Oxytricha nova TEBPα/TEBPβ [181,182,183,184,185], the human POT1/TPP1 heterodimer [115,186,187,188,189,190], and the CST complex [191,192,193,194,195,196,197,198,199,200], although the CST OB folds are different from the ones found in TEBP and POT1/TPP1 complexes [194], and CST has both sequence-specific and sequence-independent DNA binding activities [201,202,203]. Interestingly, although T. brucei telomere has a short single-stranded telomere G-overhang structure [84,85], no sequence-specific telomere G-overhang binding proteins have been identified, suggesting that T. brucei uses different protein(s) or mechanism(s) to protect the telomere termini.

2.3. T. brucei TRF and RAP1 Play Crucial Roles in Antigenic Variation

A number of telomere proteins have been identified in T. brucei (Figure 3) [19,55,204,205]. TbTRF is the duplex TTAGGG repeat binding factor and a TRF homolog [204]. TbRAP1 was identified as a TbTRF-interacting factor and a RAP1 homolog [19]. TbTIF2 interacts with TbTRF and is a functional homolog of TIN2 [55]. In addition, TelAP1, PPL2, and PolIE were identified to be able to bind a DNA oligo with the telomere sequence [205]. Furthermore, T. brucei telomerase have been identified to be a major mechanism of telomere maintenance [206,207,208].

2.3.1. The TbTRF Myb Domain Has Sequence-Specific Duplex Telomere DNA and TERRA Binding Activities That Are Critical for Maintaining the Telomere Integrity and for Suppressing VSG Switching

Since T. brucei and vertebrates have the exact same telomere sequence, TTAGGG repeats [76], it is expected that telomere binding factors in T. brucei and vertebrates use similar DNA binding motifs. Indeed, TbTRF was identified in silico, because the sequence of its C-terminal Myb domain is 33–38% homologous to those of mammalian TRF Myb domains [204]. TRF homologs have been identified in T. cruzi and Leishmania, two closely related kinetoplastid parasites [204,209]. While mammalian species have both TRF1 and TRF2, only one TRF homolog was identified in kinetoplastids despite an extensive sequence search [204,209].
TbTRF associates with the telomere chromatin and co-localizes with the telomere throughout the cell cycle [204]. Knockdown of TbTRF by RNAi leads to cell growth arrest and the loss of telomeric G-overhang, indicating that TbTRF is essential for the terminal telomere structure and cell proliferation [204]. TbTRF also self-dimerizes through a putative TRFH domain at its N-terminal region, a feature shared with other TRF homologs such as TRF1, TRF2, and SpTAZ1 [178,204,210]. TbTRF-depleted cells have an increased amount of DNA damage at the telomere [75] and an elevated VSG switching rate with many switching events involving the loss of the active ES [54], further indicating that defects in telomere integrity maintenance allow more VSG switching. Interestingly, depletion of TbTRF leads to higher TERRA and TRL levels [75]. The R-loop structure has a tendency to introduce DNA damage [211,212,213]. Indeed, overexpression of RNase H1 [69] partially suppresses the elevated TRL level and the increased amount of DNA damage at the telomere in TbTRF-depleted cells [75], indicating that suppressing the TERRA and TRL levels is an underlying mechanism of how TbTRF helps maintain the telomere integrity.
TbTRF uses its Myb domain to bind the duplex TTAGGG repeats in a sequence-specific manner [54,204]. Similar to the scenarios for human TRF1 and TRF2, the self-interaction of TbTRF may enhance its potency and specificity for telomere binding. TbTRF’s role in suppressing VSG switching relies on its telomere DNA binding activity, i.e., its Myb domain [54]. The structure of TbTRF Myb domain, as determined by our team, is essentially identical to the Myb structures reported for other TRF homologs [54,214,215,216]. Specifically, the TbTRF Myb domain adopts the canonical three-helix bundle architecture with the third helix predicted to insert into the major groove of telomere DNA for sequence-specific interactions [54]. NMR titration experiments and in vitro DNA binding studies led to the identification of several residues that are critical for the DNA binding activity of TbTRF Myb domain, including R348 on the third helix that is conserved among TRF homologs and a few TbTRF specific residues such as H346 and Q320 [54]. Mutational perturbations of these critical residues that completely abolish the DNA binding activity in vitro render T. brucei cells non-viable in vivo. The TbTRF’s DNA binding activity is presumably essential for telomere integrity. Depletion of TbTRF results in an increased amount of telomere DNA damage [75], and DNA damage at the active VSG vicinity results in lethality in >90% of cells [57]. Mutations that weaken but do not abolish the DNA binding activity, such as R298K, H346R, and R348K, yield cells that are viable with the VSG switching frequency elevated by ~1.6–3.3 fold [54]. These results confirm that the affinity of the TbTRF Myb-DNA interaction is specifically important for suppressing the frequency of VSG switching [54].
Additionally, the TbTRF Myb domain is found to bind to both regular and J-containing telomere DNA with similar affinity (Kd of 12 µM vs. 20 µM) [54]. J represents β-D-glucosyl(hydroxymethyl)uracil, a sugar-modified version of thymidine found in kinetoplastid flagellates [217]. In T. brucei, J is present only at the BF stage [218,219,220,221,222]. In kinetoplastids, only a small fraction (~1%) of thymidine bases in the genome are modified into J. Interestingly, J can be found in silent ESs [221] and is highly enriched in the telomere, replacing ~14% of T in (CCCTAA)n and ~36% in (TTAGGG)n [220,222]. It has been speculated that TbTRF, as the only telomere DNA binding factor, may bind to the J base differently due to the bulky sugar added. However, the finding of comparable binding affinities suggests that the TbTRF Myb domain does not differentiate between the T and J bases [54]. Structural modeling of the TbTRF Myb domain in complex with J-containing telomere DNA also shows that neither the J base in the (CCCTAA)n strand nor that in the (TTAGGG)n strand are located close enough to the third helix of the Myb domain for possible direct interactions [54]. Thus, J base is unlikely to play a significant role in telomere integrity as it is not differentially recognized by the TbTRF Myb domain. Instead, J base has been reported to regulate the transcription mediated by RNAP II as it is located at the ends of RNAP II PTUs [223,224,225].
The TbTRF Myb domain is quite unique in that it can also bind TERRA [75]. Human TRF2 is also capable of binding TERRA and suppresses its level [226]. However, TRF2 uses its N-terminal basic GAR domain to bind TERRA [226,227] and TRF2 also promotes the TRL formation [228]. In contrast, TbTRF suppresses the TRL level, and most interestingly, TbTRF binds TERRA through its Myb domain [75]. The TbTRF Myb domain also has a weak binding activity to CCCUAA repeats, but it clearly has a higher affinity to UUAGGG repeats [75]. Myb domain is well-known for its function to bind to dsDNA in a sequence-specific manner but has not been reported to possess binding activity for ssDNA or RNA [229]. Indeed, TbTRF does not bind single-stranded telomere DNA [204]. Therefore, it is surprising that the TbTRF Myb domain has a sequence-specific RNA binding activity, which is unique among all known Myb-containing telomere proteins. In addition, the R298E mutant abolishes the ds(TTAGGG)n binding activity [54,75] but enhances the (UUAGGG)n binding [75], suggesting that the DNA- and RNA-interacting interfaces overlap with each other (at least partially). Indeed, in vitro competition binding assays indicate that TbTRF has a higher affinity to ds(TTAGGG)n than (UUAGGG)n [75]. Furthermore, the TbTRF-DNA-RNA ternary complex has not been detected in in vitro analysis [75].
Overexpressing RNase H1 suppresses the increased amount of telomeric DNA damage and elevated TRL level in TbTRF-depleted cells [75], indicating that suppression of the TRL level by TbTRF is critical for telomere integrity and cell viability, as >90% cell die when a DSB is introduced in the active VSG vicinity [57]. It has been hypothesized that TbTRF suppresses the TRL level through both its nucleic acid-binding activities. First, TbTRF suppresses the TERRA level by telomeric silencing that presumably relies on its duplex telomere DNA binding activity [75]. Second, TbTRF appears to promote trans-localization of TERRA [75], which may depend on both its TERRA and ds(TTAGGG)n binding activities. It has been shown that TbTRF-depleted cells have fewer nuclear TERRA foci, indicating that TbTRF promotes trans-localization of TERRA [75]. Since TbTRF binds both duplex telomere DNA and TERRA [75,204], and TbTRF has a self-dimerization function [204], it is possible that TbTRF recruits TERRA away from its transcription site, as a telomere-binding and a TERRA-binding TRF may interact with each other. The active ES is depleted of nucleosomes due to the high level of RNAP I transcription [27,28], and little TbTRF is expected to bind the telomere that is transcribed by RNAP I. Therefore, TERRA is likely recruited away from its transcription site by TbTRF, which helps reduce the local concentration of TERRA at the active telomere and limits the chance of TRL formation. The telomere DNA and TERRA binding activities of TRF homologs are summarized in Table 1.

2.3.2. The Electrostatics-Based Sequence-Nonspecific dsDNA Binding Activity of TbRAP1 Is Essential for VSG Silencing and Telomere Integrity

A yeast two-hybrid screen using TbTRF as bait led to the identification of TbRAP1, a homolog of yeast and mammalian RAP1s [19]. Based on sequence comparison with other RAP1 homologs, TbRAP1 is predicted to contain an N-terminal BRCT domain, a Myb domain and a MybLike domain in the middle region, and a RAP1 C-Terminus (RCT) domain [19]. However, the exact boundaries for these domains are not well defined, due to the low sequence homology between TbRAP1 and other RAP1 homologs. We have started to understand the key functions of these structural domains [22], but their precise functions still need further investigation.
TbRAP1 is confirmed to be a TbTRF-interacting factor by coimmunoprecipitation and shown to associate with the telomere by chromatin immunoprecipitation [19]. TbRAP1 is also essential for T. brucei proliferation as depletion of TbRAP1 by RNAi or conditional knockout of TbRAP1 leads to cell growth arrest [19,20,21,22,63]. Most strikingly, knockdown of TbRAP1 leads to derepression of all ES-linked silent VSGs [19,20,21,22]. Normally, only the active ES is colocalized with RNAP I at the extranucleolar ES body [230]. However, TbRAP1 depletion leads to the formation of multiple extranucleolar RNAP I foci and simultaneous expression of multiple VSGs on cell surface [19]. Such derepression is specific to silent VSGs because the mRNA level of the active VSG was subtly decreased instead [19]. The TbRAP1-mediated silencing effect spreads over the whole ES region and represses a reporter gene inserted immediately downstream of the ES promoter and 40–60 kb upstream of the telomere [19]. In addition, the TbRAP1-mediated silencing effect is stronger for the telomere-adjacent VSG and weaker for the ES promoter-adjacent reporter gene, reflecting its position-dependent characteristic [19]. On the other hand, RNAP I-mediated transcription of rRNAs and RNAP II-transcribed genes like TbTRF and histone H4 are not affected [19]. While TbRAP1 knockdown led to profound derepression of all ES-linked silent VSGs [19,21,22], the molecular mechanism of this derepression is not completely understood. RAP1 homologs have been reported to repress the transcription of subtelomeric genes by strengthening the heterochromatic structure of the telomere [145,146,149,231,232,233]. TbRAP1 also facilitates chromatin compaction at the telomere, at least in PF cells, and this appears to be a mechanism of TbRAP1-mediated VSG silencing [20]. In T. brucei, histones [26,33,234], histone chaperones [235], a histone modifier [236], and chromatin remodelers [29,30,31] have been shown to play important roles in VSG silencing. However, whether any of these factors are recruited to the telomere by TbRAP1 or other telomere proteins is unknown. Recent RNAseq analyses further showed that TbRAP1′s role on gene expression is not only limited at the telomere [21,22]. TbRAP1 depletion leads to an upregulation of more than 7000 genes, including essentially all VSG genes and many ESAGs (Figure 1) [21,22]. Interestingly, depletion of TbRAP1 also causes a downregulation of more than 2500 genes, including many ribosomal protein genes [21,22]. Both yeast and mammalian RAP1s have been shown to regulate gene expression at non-telomeric loci [237,238,239,240,241,242,243]. It appears that TbRAP1, likes its homologs, also has both activities to silence and activate gene expression at the telomere and non-telomere loci, respectively.
Association with the telomere chromatin is essential for the telomere functions of all RAP1 homologs. Interestingly, RAP1 homologs are recruited to the telomere through different means. Mammalian RAP1 does not bind the duplex TTAGGG repeat directly and is recruited to the telomere through its interaction with TRF2 [111,117]. Fission yeast RAP1 is also recruited to the telomere through its interaction with SpTAZ1 [232], the fission yeast TRF homolog [111,141]. Although RAP1 homologs all have the central Myb domain, mammalian and fission yeast RAP1s do not possess any DNA binding activity, and their Myb domains show sequence and structural features that are incompatible with direct DNA binding [244].
Recently, our team reported that a stretch of positively charged residues (737RKRRR741) in TbRAP1 is responsible for its unique DNA binding activities [21]. This R/K-patch is located within the MybLike domain and overlaps with TbRAP1′s nuclear localization signal (aa 727 to 741) [21]. In vitro biochemical studies confirm that this R/K-patch allows TbRAP1 to bind to both single- and double-stranded DNA in an electrostatics-based, sequence-nonspecific, and substrate size-dependent manner [21]. This R/K-patch is also required for TbRAP1′s localization to the telomere while TbTRF and the Myb domain of TbRAP1 are dispensable [21]. Interestingly, the dsDNA binding activity of TbRAP1 mediated by the R/K-patch appears to be sensitive to the phosphorylation status of S742 and S744 [21], two residues adjacent to this positively charged segment with their phosphorylated state detected in both BF and PF cells [245,246]. Phosphomimicking mutations S742D/S744D disrupt the telomere localization of TbRAP1, causing telomere damage and derepressed ES-linked silent VSGs [21]. Another TbRAP1 mutant, with the positively charged R/K-patch mutated to 737AAAAA741, causes the same phenotypes [21]. Thus, the electrostatics feature of the R/K-patch is of critical importance to the unique DNA binding activity of TbRAP1, which is essential for telomere integrity and VSG silencing [21]. The telomere DNA binding activities of RAP1 homologs are summarized in Table 2.

3. Conclusions

Antigenic variation is a key pathogenesis mechanism of T. brucei, enabling the parasite to establish long-term infections and rendering vaccination ineffective [247]. Work from our and others’ laboratories have shown that the telomere structure and telomere proteins are both critical for monoallelic VSG expression and affect VSG switching frequencies [19,20,21,22,54,55,56,57,61,63,68,75].
The overall telomere architecture is conserved between T. brucei and higher eukaryotes. Homologs of several Shelterin components have also been identified in T. brucei, including TbTRF and TbRAP1 [19,55,204]. Although TbTRF and TbRAP1 have conserved telomere functions as their respective homologs, the underlying mechanisms are not the same [19,21,54,75,204]. As described above, the TbTRF Myb domain binds to TERRA in addition to ds(TTAGGG)n [54,75,204], making this Myb domain unique among known telomere protein Myb domains [216], bearing both sequence-specific dsDNA and RNA binding activities. Similarly, TbRAP1′s unconventional electrostatics-based, sequence-nonspecific DNA binding activities are also unique among all RAP1 homologs [21]. Importantly, the DNA binding activities of TbTRF and TbRAP1 are required to maintain the telomere integrity (hence cell viability) and suppress VSG switching and are essential for monoallelic VSG switching [21,54].
Given the importance of their unique DNA binding activities, TbTRF and TbRAP1 can serve as potential targets to develop effective therapeutics against T. brucei. For example, molecular agents that interfere or abolish the DNA binding activities of TbTRF and TbRAP1 may reduce T. brucei viability by compromising telomere integrity and weakening the defense mechanism of antigenic variation through loss of VSG monoallelic expression. Such a drug discovery effort may bring enormous health and economic benefits to those disadvantaged regions exposed disproportionately to the risk of this parasite. In addition, as T. cruzi and Leishmania are closely related to T. brucei, and TRF and RAP1 homologs are readily identifiable in these parasites [19,204,209], knowledge gleaned from studies on T. brucei TRF and RAP1 will also help fight T. cruzi and Leishmania infections.

Author Contributions

Writing—original draft preparation, B.L. and Y.Z.; writing—review and editing, B.L. and Y.Z.; funding acquisition, B.L. and Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by NIH, AI066095 (PI, B.L.), Research Grants Council Hong Kong, PolyU 151062/18M (PI, Y.Z.), and Shenzhen Basic Research Program, JCYJ20170818104619974 (PI, Y.Z.).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sutherland, C.S.; Yukich, J.; Goeree, R.; Tediosi, F. A Literature Review of Economic Evaluations for a Neglected Tropical Disease: Human African Trypanosomiasis (“Sleeping Sickness”). PLoS Negl. Trop. Dis. 2015, 9, e0003397. [Google Scholar] [CrossRef] [Green Version]
  2. WHO. Investing to Overcome the Global Impact of Neglected Tropical Diseases. Third WHO Report on Neglected Tropical DiseasesWHO. 2015. Available online: https://apps.who.int/iris/handle/10665/152781 (accessed on 29 July 2021).
  3. Weny, G.; Okwee-Acai, J.; Okech, S.G.; Tumwine, G.; Ndyanabo, S.; Abigaba, S.; Goldberg, T.L. Prevalence and Risk Factors Associated with Hemoparasites in Cattle and Goats at the Edge of Kibale National Park, Western Uganda. J. Parasitol. 2017, 103, 69–74. [Google Scholar] [CrossRef] [Green Version]
  4. Matthews, K.R. The developmental cell biology of Trypanosoma brucei. J. Cell Sci. 2005, 118, 283–290. [Google Scholar] [CrossRef] [Green Version]
  5. Acosta-Serrano, A.; Vassella, E.; Liniger, M.; Renggli, C.K.; Brun, R.; Roditi, I.; Englund, P.T. The surface coat of procyclic Trypanosoma brucei: Programmed expression and proteolytic cleavage of procyclin in the tsetse fly. Proc. Natl. Acad. Sci. USA 2001, 98, 1513–1518. [Google Scholar] [CrossRef] [PubMed]
  6. Quintana, J.F.; Zoltner, M.; Field, M.C. Evolving Differentiation in African Trypanosomes. Trends Parasitol. 2021, 37, 296–303. [Google Scholar] [CrossRef]
  7. Diffley, P. Trypanosoma brucei: Immunogenicity of the variant surface coat glycoprotein of virulent and avirulent subspecies. Exp. Parasitol. 1985, 59, 98–107. [Google Scholar] [CrossRef]
  8. Deitsch, K.W.; Lukehart, S.A.; Stringer, J.R. Common strategies for antigenic variation by bacterial, fungal and protozoan pathogens. Nat. Rev. Genet. 2009, 7, 493–503. [Google Scholar] [CrossRef] [Green Version]
  9. Cross, G.A.; Kim, H.-S.; Wickstead, B. Capturing the variant surface glycoprotein repertoire (the VSGnome) of Trypanosoma brucei Lister 427. Mol. Biochem. Parasitol. 2014, 195, 59–73. [Google Scholar] [CrossRef] [PubMed]
  10. Müller, L.S.; Cosentino, R.O.; Förstner, K.U.; Guizetti, J.; Wedel, C.; Kaplan, N.; Janzen, C.J.; Arampatzi, P.; Vogel, J.; Steinbiss, S.; et al. Genome organization and DNA accessibility control antigenic variation in trypanosomes. Nat. Cell Biol. 2018, 563, 121–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Melville, S.E.; Leech, V.; Navarro, M.; Cross, G. The molecular karyotype of the megabase chromosomes of Trypanosoma brucei stock Mol. Biochem. Parasitol. 2000, 111, 261–273. [Google Scholar] [CrossRef]
  12. El-Sayed, N.M.; Hegde, P.; Quackenbush, J.; Melville, S.E.; Donelson, J.E. The African trypanosome genome. Int. J. Parasitol. 2000, 30, 329–345. [Google Scholar] [CrossRef]
  13. Alsford, S.; Wickstead, B.; Ersfeld, K.; Gull, K. Diversity and dynamics of the minichromosomal karyotype in Trypanosoma brucei. Mol. Biochem. Parasitol. 2001, 113, 79–88. [Google Scholar] [CrossRef] [Green Version]
  14. Günzl, A.; Bruderer, T.; Laufer, G.; Schimanski, B.; Tu, L.-C.; Chung, H.-M.; Lee, P.-T.; Lee, M.G.-S. RNA Polymerase I Transcribes Procyclin Genes and Variant Surface Glycoprotein Gene Expression Sites in Trypanosoma brucei. Eukaryot. Cell 2003, 2, 542–551. [Google Scholar] [CrossRef] [Green Version]
  15. Hertz-Fowler, C.; Figueiredo, L.M.; Quail, M.A.; Becker, M.; Jackson, A.; Bason, N.; Brooks, K.; Churcher, C.; Fahkro, S.; Goodhead, I.; et al. Telomeric Expression Sites Are Highly Conserved in Trypanosoma brucei. PLoS ONE 2008, 3, e3527. [Google Scholar] [CrossRef] [Green Version]
  16. De Lange, T.; Borst, P. Genomic environment of the expression-linked extra copies of genes for surface antigens of Trypanosoma brucei resembles the end of a chromosome. Nature 1982, 299, 451–453. [Google Scholar] [CrossRef] [PubMed]
  17. Barry, J.; Graham, S.; Fotheringham, M.; Graham, V.S.; Kobryn, K.; Wymer, B. VSG gene control and infectivity strategy of metacyclic stage Trypanosoma brucei. Mol. Biochem. Parasitol. 1998, 91, 93–105. [Google Scholar] [CrossRef]
  18. Cross, G. Identification, purification and properties of clone-specific glycoprotein antigens constituting the surface coat of Trypanosoma brucei. Parasitology 1975, 71, 393–417. [Google Scholar] [CrossRef] [PubMed]
  19. Yang, X.; Figueiredo, L.M.; Espinal, A.; Okubo, E.; Li, B. RAP1 Is Essential for Silencing Telomeric Variant Surface Glycoprotein Genes in Trypanosoma brucei. Cell 2009, 137, 99–109. [Google Scholar] [CrossRef] [Green Version]
  20. Pandya, U.M.; Sandhu, R.; Li, B. Silencing subtelomeric VSGs by Trypanosoma brucei RAP1 at the insect stage involves chromatin structure changes. Nucleic Acids Res. 2013, 41, 7673–7682. [Google Scholar] [CrossRef] [Green Version]
  21. Afrin, M.; Gaurav, A.K.; Yang, X.; Pan, X.; Zhao, Y.; Li, B. TbRAP1 has an unusual duplex DNA binding activity required for its telomere localization and VSG silencing. Sci. Adv. 2020, 6, eabc4065. [Google Scholar] [CrossRef]
  22. Afrin, M.; Kishmiri, H.; Sandhu, R.; Rabbani, M.A.G.; Li, B. Trypanosoma brucei RAP1 Has Essential Functional Domains That Are Required for Different Protein Interactions. Msphere 2020, 5, e00027-20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Günzl, A.; Kirkham, J.; Nguyen, T.N.; Badjatia, N.; Park, S.H. Mono-allelic VSG expression by RNA polymerase I in Trypanosoma brucei: Expression site control from both ends? Gene 2015, 556, 68–73. [Google Scholar] [CrossRef] [Green Version]
  24. Rudenko, G. Epigenetics and transcriptional control in African trypanosomes. Essays Biochem. 2010, 48, 201–219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Narayanan, M.S.; Rudenko, G. TDP1 is an HMG chromatin protein facilitating RNA polymerase I transcription in African trypanosomes. Nucleic Acids Res. 2013, 41, 2981–2992. [Google Scholar] [CrossRef] [Green Version]
  26. Povelones, M.L.; Gluenz, E.; Dembek, M.; Gull, K.; Rudenko, G. Histone H1 Plays a Role in Heterochromatin Formation and VSG Expression Site Silencing in Trypanosoma brucei. PLoS Pathog. 2012, 8, e1003010. [Google Scholar] [CrossRef] [Green Version]
  27. Stanne, T.M.; Rudenko, G. Active VSG Expression Sites in Trypanosoma brucei are Depleted of Nucleosomes. Eukaryot. Cell 2010, 9, 136–147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Figueiredo, L.M.; Cross, G.A.M. Nucleosomes are Depleted at the VSG Expression Site Transcribed by RNA Polymerase I in African Trypanosomes. Eukaryot. Cell 2010, 9, 148–154. [Google Scholar] [CrossRef] [Green Version]
  29. Hughes, K.; Wand, M.; Foulston, L.; Young, R.; Harley, K.; Terry, S.; Ersfeld, K.; Rudenko, G. A novel ISWI is involved in VSG expression site downregulation in African trypanosomes. EMBO J. 2007, 26, 2400–2410. [Google Scholar] [CrossRef] [Green Version]
  30. Denninger, V.; Rudenko, G. FACT plays a major role in histone dynamics affecting VSG expression site control in Trypanosoma brucei. Mol. Microbiol. 2014, 94, 945–962. [Google Scholar] [CrossRef] [Green Version]
  31. Denninger, V.; Fullbrook, A.; Bessat, M.; Ersfeld, K.; Rudenko, G. The FACT subunit TbSpt16 is involved in cell cycle specific control of VSG expression sites in Trypanosoma brucei. Mol. Microbiol. 2010, 78, 459–474. [Google Scholar] [CrossRef] [PubMed]
  32. Navarro, M.; Cross, G.A.; Wirtz, E. Trypanosoma brucei variant surface glycoprotein regulation involves coupled activation/inactivation and chromatin remodeling of expression sites. EMBO J. 1999, 18, 2265–2272. [Google Scholar] [CrossRef] [Green Version]
  33. Schulz, D.; Zaringhalam, M.; Papavasiliou, F.N.; Kim, H.-S. Base J and H3.V Regulate Transcriptional Termination in Trypanosoma brucei. PLoS Genet. 2016, 12, e1005762. [Google Scholar] [CrossRef]
  34. Kassem, A.; Pays, E.; Vanhamme, L. Transcription is initiated on silent variant surface glycoprotein expression sites despite monoallelic expression in Trypanosoma brucei. Proc. Natl. Acad. Sci. USA 2014, 111, 8943–8948. [Google Scholar] [CrossRef] [Green Version]
  35. Vanhamme, L.; Poelvoorde, P.; Pays, A.; Tebabi, P.; Van Xong, H.; Pays, E. Differential RNA elongation controls the variant surface glycoprotein gene expression sites of Trypanosoma brucei. Mol. Microbiol. 2000, 36, 328–340. [Google Scholar] [CrossRef] [PubMed]
  36. Cestari, I.; McLeland-Wieser, H.; Stuart, K. Nuclear Phosphatidylinositol 5-Phosphatase Is Essential for Allelic Exclusion of Variant Surface Glycoprotein Genes in Trypanosomes. Mol. Cell. Biol. 2019, 39, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Cestari, I.; Stuart, K. Inositol phosphate pathway controls transcription of telomeric expression sites in trypanosomes. Proc. Natl. Acad. Sci. USA 2015, 112, E2803–E2812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Dubois, K.N.; Alsford, S.; Holden, J.M.; Buisson, J.; Swiderski, M.; Bart, J.-M.; Ratushny, A.V.; Wan, Y.; Bastin, P.; Barry, J.D.; et al. NUP-1 Is a Large Coiled-Coil Nucleoskeletal Protein in Trypanosomes with Lamin-Like Functions. PLoS Biol. 2012, 10, e1001287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Maishman, L.; Obado, S.O.; Alsford, S.; Bart, J.-M.; Chen, W.-M.; Ratushny, A.V.; Navarro, M.; Horn, D.; Aitchison, J.D.; Chait, B.T.; et al. Co-dependence between trypanosome nuclear lamina components in nuclear stability and control of gene expression. Nucleic Acids Res. 2016, 44, 10554–10570. [Google Scholar] [CrossRef] [Green Version]
  40. López-Farfán, D.C.; Bart, J.-M.; Rojas-Barros, D.I.; Navarro, M. SUMOylation by the E3 Ligase TbSIZ1/PIAS1 Positively Regulates VSG Expression in Trypanosoma brucei. PLoS Pathog. 2014, 10, e1004545. [Google Scholar] [CrossRef] [PubMed]
  41. Benmerzouga, I.; Concepcion-Acevedo, J.; Kim, H.S.; Vandoros, A.V.; Cross, G.A.; Klingbeil, M.M.; Li, B. Trypanosoma brucei Orc1 is essential for nuclear DNA replication and affects both VSG silencing and VSG switching. Mol. Microbiol. 2013, 87, 196–210. [Google Scholar] [CrossRef] [Green Version]
  42. Tiengwe, C.; Marcello, L.; Farr, H.; Dickens, N.; Kelly, S.; Swiderski, M.; Vaughan, D.; Gull, K.; Barry, J.D.; Bell, S.D.; et al. Genome-wide Analysis Reveals Extensive Functional Interaction between DNA Replication Initiation and Transcription in the Genome of Trypanosoma brucei. Cell Rep. 2012, 2, 185–197. [Google Scholar] [CrossRef] [Green Version]
  43. Kim, H.-S.; Park, S.H.; Günzl, A.; Cross, G.A.M. MCM-BP Is Required for Repression of Life-Cycle Specific Genes Transcribed by RNA Polymerase I in the Mammalian Infectious Form of Trypanosoma brucei. PLoS ONE 2013, 8, e57001. [Google Scholar] [CrossRef] [Green Version]
  44. Kim, H.-S. Genome-wide function of MCM-BP in Trypanosoma brucei DNA replication and transcription. Nucleic Acids Res. 2019, 47, 634–647. [Google Scholar] [CrossRef] [Green Version]
  45. Faria, J.; Glover, L.; Hutchinson, S.; Boehm, C.; Field, M.C.; Horn, D. Monoallelic expression and epigenetic inheritance sustained by a Trypanosoma brucei variant surface glycoprotein exclusion complex. Nat. Commun. 2019, 10, 1–14. [Google Scholar] [CrossRef]
  46. Glover, L.; Hutchinson, S.; Alsford, S.; Horn, D. VEX1 controls the allelic exclusion required for antigenic variation in trypanosomes. Proc. Natl. Acad. Sci. USA 2016, 113, 7225–7230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Faria, J.; Luzak, V.; Müller, L.S.M.; Brink, B.G.; Hutchinson, S.; Glover, L.; Horn, D.; Siegel, T.N. Spatial integration of transcription and splicing in a dedicated compartment sustains monogenic antigen expression in African trypanosomes. Nat. Microbiol. 2021, 6, 289–300. [Google Scholar] [CrossRef]
  48. Aresta-Branco, F.; Sanches-Vaz, M.; Bento, F.; Rodrigues, J.A.; Figueiredo, L.M. African trypanosomes expressing multiple VSGs are rapidly eliminated by the host immune system. Proc. Natl. Acad. Sci. USA 2019, 116, 20725–20735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Batram, C.; Jones, N.G.; Janzen, C.J.; Markert, S.M.; Engstler, M. Expression site attenuation mechanistically links antigenic variation and development in Trypanosoma brucei. Elife 2014, 3, e02324. [Google Scholar] [CrossRef]
  50. Myler, P.; Allison, J.; Agabian, N.; Stuart, K. Antigenic variation in African trypanosomes by gene replacement or activation of alternate telomeres. Cell 1984, 39, 203–211. [Google Scholar] [CrossRef]
  51. Myler, P.; Nelson, R.G.; Agabian, N.; Stuart, K. Two mechanisms of expression of a predominant variant antigen gene of Trypanosoma brucei. Nat. Cell Biol. 1984, 309, 282–284. [Google Scholar] [CrossRef]
  52. Kamper, S.M.; Barbet, A.F. Surface epitope variation via mosaic gene formation is potential key to long-term survival of Trypanosoma brucei. Mol. Biochem. Parasitol. 1992, 53, 33–44. [Google Scholar] [CrossRef]
  53. Marcello, L.; Barry, J.D. Analysis of the VSG gene silent archive in Trypanosoma brucei reveals that mosaic gene expression is prominent in antigenic variation and is favored by archive substructure. Genome Res. 2007, 17, 1344–1352. [Google Scholar] [CrossRef] [Green Version]
  54. Jehi, S.E.; Li, X.; Sandhu, R.; Ye, F.; Benmerzouga, I.; Zhang, M.; Zhao, Y.; Li, B. Suppression of subtelomeric VSG switching by Trypanosoma brucei TRF requires its TTAGGG repeat-binding activity. Nucleic Acids Res. 2014, 42, 12899–12911. [Google Scholar] [CrossRef]
  55. Jehi, S.E.; Wu, F.; Li, B. Trypanosoma brucei TIF2 suppresses VSG switching by maintaining subtelomere integrity. Cell Res. 2014, 24, 870–885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Boothroyd, C.E.; Dreesen, O.; Leonova, T.; Ly, K.I.; Figueiredo, L.M.; Cross, G.; Papavasiliou, F.N. A yeast-endonuclease-generated DNA break induces antigenic switching in Trypanosoma brucei. Nat. Cell Biol. 2009, 459, 278–281. [Google Scholar] [CrossRef] [PubMed]
  57. Glover, L.; Alsford, S.; Horn, D. DNA Break Site at Fragile Subtelomeres Determines Probability and Mechanism of Antigenic Variation in African Trypanosomes. PLoS Pathog. 2013, 9, e1003260. [Google Scholar] [CrossRef]
  58. Cross, M.; Taylor, M.C.; Borst, P. Frequent Loss of the Active Site during Variant Surface Glycoprotein Expression Site Switching In Vitro in Trypanosoma brucei. Mol. Cell. Biol. 1998, 18, 198–205. [Google Scholar] [CrossRef] [Green Version]
  59. Kim, H.-S.; Cross, G.A.M. Identification of Trypanosoma brucei RMI1/BLAP75 Homologue and Its Roles in Antigenic Variation. PLoS ONE 2011, 6, e25313. [Google Scholar] [CrossRef] [Green Version]
  60. Robinson, N.; Burman, N.; Melville, S.E.; Barry, J.D. Predominance of Duplicative VSG Gene Conversion in Antigenic Variation in African Trypanosomes. Mol. Cell. Biol. 1999, 19, 5839–5846. [Google Scholar] [CrossRef] [Green Version]
  61. Hovel-Miner, G.; Boothroyd, C.E.; Mugnier, M.; Dreesen, O.; Cross, G.; Papavasiliou, F.N. Telomere Length Affects the Frequency and Mechanism of Antigenic Variation in Trypanosoma brucei. PLoS Pathog. 2012, 8, e1002900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Kim, H.-S.; Cross, G.A.M. TOPO3α Influences Antigenic Variation by Monitoring Expression-Site-Associated VSG Switching in Trypanosoma brucei. PLoS Pathog. 2010, 6, e1000992. [Google Scholar] [CrossRef] [Green Version]
  63. Nanavaty, V.; Sandhu, R.; Jehi, S.E.; Pandya, U.; Li, B. Trypanosoma brucei RAP1 maintains telomere and subtelomere integrity by suppressing TERRA and telomeric RNA:DNA hybrids. Nucleic Acids Res. 2017, 45, 5785–5796. [Google Scholar] [CrossRef]
  64. McCulloch, R.; Barry, J.D. A role for RAD51 and homologous recombination in Trypanosoma brucei antigenic variation. Genes Dev. 1999, 13, 2875–2888. [Google Scholar] [CrossRef] [Green Version]
  65. Proudfoot, C.; McCulloch, R. Distinct roles for two RAD51-related genes in Trypanosoma brucei antigenic variation. Nucleic Acids Res. 2005, 33, 6906–6919. [Google Scholar] [CrossRef]
  66. Hartley, C.L.; McCulloch, R. Trypanosoma brucei BRCA2 acts in antigenic variation and has undergone a recent expansion in BRC repeat number that is important during homologous recombination. Mol. Microbiol. 2008, 68, 1237–1251. [Google Scholar] [CrossRef] [Green Version]
  67. Devlin, R.; Marques, C.A.; Paape, D.; Prorocic, M.; Zurita-Leal, A.; Campbell, S.J.; Lapsley, C.; Dickens, N.; McCulloch, R. Mapping replication dynamics in Trypanosoma brucei reveals a link with telomere transcription and antigenic variation. Elife 2016, 5, e12765. [Google Scholar] [CrossRef]
  68. Jehi, S.E.; Nanavaty, V.; Li, B. Trypanosoma brucei TIF2 and TRF Suppress VSG Switching Using Overlapping and Independent Mechanisms. PLoS ONE 2016, 11, e0156746. [Google Scholar] [CrossRef]
  69. Briggs, E.; Crouch, K.; Lemgruber, L.; Lapsley, C.; McCulloch, R. Ribonuclease H1-targeted R-loops in surface antigen gene expression sites can direct trypanosome immune evasion. PLoS Genet. 2018, 14, e1007729. [Google Scholar] [CrossRef] [Green Version]
  70. Briggs, E.; Crouch, K.; Lemgruber, L.; Hamilton, G.; Lapsley, C.; McCulloch, R. Trypanosoma brucei ribonuclease H2A is an essential R-loop processing enzyme whose loss causes DNA damage during transcription initiation and antigenic variation. Nucleic Acids Res. 2019, 47, 9180–9197. [Google Scholar] [CrossRef] [Green Version]
  71. Da Silva, M.S.; Hovel-Miner, G.A.; Briggs, E.M.; Elias, M.C.; McCulloch, R. Evaluation of mechanisms that may generate DNA lesions triggering antigenic variation in African trypanosomes. PLoS Pathog. 2018, 14, e1007321. [Google Scholar] [CrossRef] [Green Version]
  72. McCulloch, R.; Cobbold, C.A.; Figueiredo, L.; Jackson, A.; Morrison, L.J.; Mugnier, M.R.; Papavasiliou, N.; Schnaufer, A.; Matthews, K. Emerging challenges in understanding trypanosome antigenic variation. Emerg. Top. Life Sci. 2017, 1, 585–592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ridewood, S.; Ooi, C.; Hall, B.; Trenaman, A.; Wand, N.V.; Sioutas, G.; Scherwitzl, I.; Rudenko, G. The role of genomic location and flanking 3′UTR in the generation of functional levels of variant surface glycoprotein in Trypanosoma brucei. Mol. Microbiol. 2017, 106, 614–634. [Google Scholar] [CrossRef] [Green Version]
  74. Haber, J.E. Partners and pathwaysrepairing a double-strand break. Trends Genet. 2000, 16, 259–264. [Google Scholar] [CrossRef]
  75. Saha, A.; Gaurav, A.K.; Pandya, U.M.; Afrin, M.; Sandhu, R.; Nanavaty, V.; Schnur, B.; Li, B. TbTRF suppresses the TERRA level and regulates the cell cycle-dependent TERRA foci number with a TERRA binding activity in its C-terminal Myb domain. Nucleic Acids Res. 2021, 49, 5637–5653. [Google Scholar] [CrossRef] [PubMed]
  76. Podlevsky, J.D.; Bley, C.J.; Omana, R.V.; Qi, X.; Chen, J.J.-L. The Telomerase Database. Nucleic Acids Res. 2007, 36, D339–D343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Henderson, E.R.; Blackburn, E.H. An overhanging 3’ terminus is a conserved feature of telomeres. Mol. Cell Biol. 1989, 9, 345–348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Wellinger, R.J.; Wolf, A.J.; Zakian, V.A. Saccharomyces telomeres acquire single-strand TG1–3 tails late in S phase. Cell 1993, 72, 51–60. [Google Scholar] [CrossRef]
  79. Makarov, V.L.; Hirose, Y.; Langmore, J.P. Long G Tails at Both Ends of Human Chromosomes Suggest a C Strand Degradation Mechanism for Telomere Shortening. Cell 1997, 88, 657–666. [Google Scholar] [CrossRef] [Green Version]
  80. Dionne, I.; Wellinger, R.J. Cell cycle-regulated generation of single-stranded G-rich DNA in the absence of telomerase. Proc. Natl. Acad. Sci. USA 1996, 93, 13902–13907. [Google Scholar] [CrossRef] [Green Version]
  81. Jacob, N.K.; Skopp, R.; Price, C.M. G-overhang dynamics at Tetrahymena telomeres. EMBO J. 2001, 20, 4299–4308. [Google Scholar] [CrossRef] [Green Version]
  82. Chai, W.; Du, Q.; Shay, J.W.; Wright, W.E. Human Telomeres Have Different Overhang Sizes at Leading versus Lagging Strands. Mol. Cell 2006, 21, 427–435. [Google Scholar] [CrossRef]
  83. Tomita, K.; Kibe, T.; Kang, H.-Y.; Seo, Y.-S.; Uritani, M.; Ushimaru, T.; Ueno, M. Fission Yeast Dna2 Is Required for Generation of the Telomeric Single-Strand Overhang. Mol. Cell. Biol. 2004, 24, 9557–9567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Sandhu, R.; Li, B. Examination of the Telomere G-overhang Structure in Trypanosoma brucei. J. Vis. Exp. 2011, 47, e1959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Sandhu, R.; Li, B. Telomerase activity is required for the telomere G-overhang structure in Trypanosoma brucei. Sci. Rep. 2017, 7, 15983. [Google Scholar] [CrossRef] [PubMed]
  86. Griffith, J.D.; Comeau, L.; Rosenfield, S.; Stansel, R.M.; Bianchi, A.; Moss, H.; de Lange, T. Mammalian Telomeres End in a Large Duplex Loop. Cell 1999, 97, 503–514. [Google Scholar] [CrossRef] [Green Version]
  87. Murti, K.G.; Prescott, D.M. Telomeres of polytene chromosomes in a ciliated protozoan terminate in duplex DNA loops. Proc. Natl. Acad. Sci. USA 1999, 96, 14436–14439. [Google Scholar] [CrossRef] [Green Version]
  88. Nikitina, T.; Woodcock, C.L. Closed chromatin loops at the ends of chromosomes. J. Cell Biol. 2004, 166, 161–165. [Google Scholar] [CrossRef] [Green Version]
  89. Muñoz-Jordán, J.L.; Cross, G.; De Lange, T.; Griffith, J.D. t-loops at trypanosome telomeres. EMBO J. 2001, 20, 579–588. [Google Scholar] [CrossRef]
  90. Cesare, A.J.; Quinney, N.; Willcox, S.; Subramanian, D.; Griffith, J.D. Telomere looping in P. sativum (common garden pea). Plant J. 2003, 36, 271–279. [Google Scholar] [CrossRef] [PubMed]
  91. Raices, M.; Verdun, R.E.; Compton, S.A.; Haggblom, C.I.; Griffith, J.D.; Dillin, A.; Karlseder, J. C. elegans Telomeres Contain G-Strand and C-Strand Overhangs that Are Bound by Distinct Proteins. Cell 2008, 132, 745–757. [Google Scholar] [CrossRef] [Green Version]
  92. Cesare, A.J.; Groff-Vindman, C.; Compton, S.A.; McEachern, M.J.; Griffith, J.D. Telomere Loops and Homologous Recombination-Dependent Telomeric Circles in a Kluyveromyces lactis Telomere Mutant Strain. Mol. Cell. Biol. 2008, 28, 20–29. [Google Scholar] [CrossRef] [Green Version]
  93. Wynford-Thomas, D.; Kipling, D. Telomerase. Cancer and the knockout mouse. Nature 1997, 389, 551–552. [Google Scholar] [PubMed]
  94. Autexier, C.; Lue, N.F. The Structure and Function of Telomerase Reverse Transcriptase. Annu. Rev. Biochem. 2006, 75, 493–517. [Google Scholar] [CrossRef] [PubMed]
  95. Greider, C.W.; Blackburn, E.H. Identification of a specific telomere terminal transferase activity in tetrahymena extracts. Cell 1985, 43, 405–413. [Google Scholar] [CrossRef]
  96. Greider, C.; Blackburn, E.H. The telomere terminal transferase of tetrahymena is a ribonucleoprotein enzyme with two kinds of primer specificity. Cell 1987, 51, 887–898. [Google Scholar] [CrossRef]
  97. Greider, C.; Blackburn, E.H. A telomeric sequence in the RNA of Tetrahymena telomerase required for telomere repeat synthesis. Nat. Cell Biol. 1989, 337, 331–337. [Google Scholar] [CrossRef]
  98. Sobinoff, A.; Pickett, H.A. Alternative Lengthening of Telomeres: DNA Repair Pathways Converge. Trends Genet. 2017, 33, 921–932. [Google Scholar] [CrossRef]
  99. Sobinoff, A.; Pickett, H.A. Mechanisms that drive telomere maintenance and recombination in human cancers. Curr. Opin. Genet. Dev. 2020, 60, 25–30. [Google Scholar] [CrossRef] [PubMed]
  100. Hu, Y.; Tang, H.-B.; Liu, N.-N.; Tong, X.-J.; Dang, W.; Duan, Y.-M.; Fu, X.-H.; Zhang, Y.; Peng, J.; Meng, F.; et al. Telomerase-Null Survivor Screening Identifies Novel Telomere Recombination Regulators. PLoS Genet. 2013, 9, e1003208. [Google Scholar] [CrossRef] [Green Version]
  101. Lundblad, V. Telomere maintenance without telomerase. Oncogene 2002, 21, 522–531. [Google Scholar] [CrossRef] [Green Version]
  102. De Lange, T. Shelterin-Mediated Telomere Protection. Annu. Rev. Genet. 2018, 52, 223–247. [Google Scholar] [CrossRef]
  103. Maciejowski, J.; De Lange, T. Telomeres in cancer: Tumour suppression and genome instability. Nat. Rev. Mol. Cell Biol. 2017, 18, 175–186. [Google Scholar] [CrossRef] [Green Version]
  104. Maciejowski, J.; Li, Y.; Bosco, N.; Campbell, P.J.; de Lange, T. Chromothripsis and Kataegis Induced by Telomere Crisis. Cell 2015, 163, 1641–1654. [Google Scholar] [CrossRef] [Green Version]
  105. Pellestor, F.; Gaillard, J.; Schneider, A.; Puechberty, J.; Gatinois, V. Chromoanagenesis, the mechanisms of a genomic chaos. Semin. Cell Dev. Biol. 2021, 1084. [Google Scholar] [CrossRef]
  106. Cleal, K.; Baird, D.M. Catastrophic Endgames: Emerging Mechanisms of Telomere-Driven Genomic Instability. Trends Genet. 2020, 36, 347–359. [Google Scholar] [CrossRef]
  107. De Lange, T. Shelterin: The protein complex that shapes and safeguards human telomeres. Genes Dev. 2005, 19, 2100–2110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Chong, L.; Van Steensel, B.; Broccoli, D.; Erdjument-Bromage, H.; Hanish, J.; Tempst, P.; De Lange, T. A Human Telomeric Protein. Science 1995, 270, 1663–1667. [Google Scholar] [CrossRef] [PubMed]
  109. Broccoli, D.; Smogorzewska, A.; Chong, L.; de Lange, T. Human telomeres contain two distinct Myb–related proteins, TRF1 and TRF1 and TRF2. Nat. Genet. 1997, 17, 231–235. [Google Scholar] [CrossRef] [PubMed]
  110. Bilaud, T.; Brun, C.; Ancelin, K.; Koering, C.E.; Laroche, T.; Gilson, E. Telomeric localization of TRF2, a novel human telobox protein. Nat. Genet. 1997, 17, 236–239. [Google Scholar] [CrossRef] [PubMed]
  111. Li, B.; Oestreich, S.; de Lange, T. Identification of Human Rap1: Implications for Telomere Evolution. Cell 2000, 101, 471–483. [Google Scholar] [CrossRef] [Green Version]
  112. Kim, S.-H.; Kaminker, P.; Campisi, J. TIN2, a new regulator of telomere length in human cells. Nat. Genet. 1999, 23, 405–412. [Google Scholar] [CrossRef] [Green Version]
  113. Houghtaling, B.R.; Cuttonaro, L.; Chang, W.; Smith, S. A Dynamic Molecular Link between the Telomere Length Regulator TRF1 and the Chromosome End Protector TRF2. Curr. Biol. 2004, 14, 1621–1631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Liu, D.; Safari, A.; O’Connor, M.S.; Chan, D.W.; Laegeler, A.; Qin, J.; Songyang, Z. PTOP interacts with POT1 and regulates its localization to telomeres. Nat. Cell Biol. 2004, 6, 673–680. [Google Scholar] [CrossRef] [PubMed]
  115. Ye, J.Z.-S.; Hockemeyer, D.; Krutchinsky, A.N.; Loayza, D.; Hooper, S.M.; Chait, B.T.; De Lange, T. POT1-interacting protein PIP1: A telomere length regulator that recruits POT1 to the TIN2/TRF1 complex. Genes Dev. 2004, 18, 1649–1654. [Google Scholar] [CrossRef] [Green Version]
  116. Baumann, P.; Cech, T.R. Pot1, the putative telomere end-binding protein in fission yeast and humans. Science 2001, 292, 1171–1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Loayza, D.; De Lange, T. POT1 as a terminal transducer of TRF1 telomere length control. Nature 2003, 424, 1013–1018. [Google Scholar] [CrossRef]
  118. Surovtseva, Y.V.; Churikov, D.; Boltz, K.A.; Song, X.; Lamb, J.C.; Warrington, R.; Leehy, K.; Heacock, M.; Price, C.M.; Shippen, D.E. Conserved Telomere Maintenance Component 1 Interacts with STN1 and Maintains Chromosome Ends in Higher Eukaryotes. Mol. Cell 2009, 36, 207–218. [Google Scholar] [CrossRef] [Green Version]
  119. Feng, X.; Hsu, S.-J.; Bhattacharjee, A.; Wang, Y.; Diao, J.; Price, C.M. CTC1-STN1 terminates telomerase while STN1-TEN1 enables C-strand synthesis during telomere replication in colon cancer cells. Nat. Commun. 2018, 9, 2827. [Google Scholar] [CrossRef] [Green Version]
  120. Zhong, Z.; Shiue, L.; Kaplan, S.; De Lange, T. A mammalian factor that binds telomeric TTAGGG repeats in vitro. Mol. Cell Biol. 1992, 12, 4834–4843. [Google Scholar] [CrossRef] [Green Version]
  121. Bianchi, A.; Smith, S.; Chong, L.; Elias, P.; De Lange, T. TRF1 is a dimer and bends telomeric DNA. EMBO J. 1997, 16, 1785–1794. [Google Scholar] [CrossRef] [Green Version]
  122. Bianchi, A.; Stansel, R.M.; Fairall, L.; Griffith, J.D.; Rhodes, D.; de Lange, T. TRF1 binds a bipartite telomeric site with extreme spatial flexibility. EMBO J. 1999, 18, 5735–5744. [Google Scholar] [CrossRef]
  123. Hanaoka, S.; Nagadoi, A.; Nishimura, Y. Comparison between TRF2 and TRF1 of their telomeric DNA-bound structures and DNA-binding activities. Protein Sci. 2009, 14, 119–130. [Google Scholar] [CrossRef] [Green Version]
  124. Loayza, D.; Parsons, H.; Donigian, J.; Hoke, K.; De Lange, T. DNA binding features of human POT1: A nonamer 5’-TAGGGTTAG-3’ minimal binding site, sequence specificity, and internal binding to multimeric sites. J. Biol. Chem. 2004, 279, 13241–13248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Rice, C.; Skordalakes, E. Structure and function of the telomeric CST complex. Comput. Struct. Biotechnol. J. 2016, 14, 161–167. [Google Scholar] [CrossRef] [Green Version]
  126. Ye, J.Z.-S.; Donigian, J.R.; van Overbeek, M.; Loayza, D.; Luo, Y.; Krutchinsky, A.N.; Chait, B.T.; de Lange, T. TIN2 Binds TRF1 and TRF2 Simultaneously and Stabilizes the TRF2 Complex on Telomeres. J. Biol. Chem. 2004, 279, 47264–47271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Smogorzewska, A.; Karlseder, J.; Holtgreve-Grez, H.; Jauch, A.; de Lange, T. DNA Ligase IV-Dependent NHEJ of Deprotected Mammalian Telomeres in G1 and G2. Curr. Biol. 2002, 12, 1635–1644. [Google Scholar] [CrossRef] [Green Version]
  128. Celli, G.B.; De Lange, T. DNA processing is not required for ATM-mediated telomere damage response after TRF2 deletion. Nat. Cell Biol. 2005, 7, 712–718. [Google Scholar] [CrossRef]
  129. Deng, Y.; Guo, X.; Ferguson, D.O.; Chang, S. Multiple roles for MRE11 at uncapped telomeres. Nat. Cell Biol. 2009, 460, 914–918. [Google Scholar] [CrossRef] [PubMed]
  130. Wang, R.C.; Smogorzewska, A.; de Lange, T. Homologous Recombination Generates T-Loop-Sized Deletions at Human Telomeres. Cell 2004, 119, 355–368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Sfeir, A.; Kabir, S.; Van Overbeek, M.; Celli, G.B.; De Lange, T. Loss of Rap1 Induces Telomere Recombination in the Absence of NHEJ or a DNA Damage Signal. Science 2010, 327, 1657–1661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Chen, Y.; Rai, R.; Zhou, Z.-R.; Kanoh, J.; Ribeyre, C.; Yang, Y.; Zheng, H.; Damay, P.; Wang, F.; Tsujii, H.; et al. A conserved motif within RAP1 has diversified roles in telomere protection and regulation in different organisms. Nat. Struct. Mol. Biol. 2011, 18, 213–221. [Google Scholar] [CrossRef] [Green Version]
  133. Guo, X.; Deng, Y.; Lin, Y.; Cosme-Blanco, W.; Chan, S.; He, H.; Yuan, G.; Brown, E.J.; Chang, S. Dysfunctional telomeres activate an ATM-ATR-dependent DNA damage response to suppress tumorigenesis. EMBO J. 2007, 26, 4709–4719. [Google Scholar] [CrossRef] [PubMed]
  134. Rai, R.; Zheng, H.; He, H.; Luo, Y.; Multani, A.; Carpenter, P.B.; Chang, S. The function of classical and alternative non-homologous end-joining pathways in the fusion of dysfunctional telomeres. EMBO J. 2010, 29, 2598–2610. [Google Scholar] [CrossRef] [Green Version]
  135. Sfeir, A.; De Lange, T. Removal of shelterin reveals the telomere end-protection problem. Science 2012, 336, 593–597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Van Ly, D.; Low, R.R.J.; Frölich, S.; Bartolec, T.; Kafer, G.; Pickett, H.A.; Gaus, K.; Cesare, A.J. Telomere Loop Dynamics in Chromosome End Protection. Mol. Cell 2018, 71, 510–525.e6. [Google Scholar] [CrossRef] [Green Version]
  137. Ottaviani, A.; Gilson, E.; Magdinier, F. Telomeric position effect: From the yeast paradigm to human pathologies? Biochimie 2008, 90, 93–107. [Google Scholar] [CrossRef] [Green Version]
  138. O’Kane, C.J.; Hyland, E.M. Yeast epigenetics: The inheritance of histone modification states. Biosci. Rep. 2019, 39. [Google Scholar] [CrossRef] [Green Version]
  139. Longtine, M.S.; Wilson, N.M.; Petracek, M.E.; Berman, J. A yeast telomere binding activity binds to two related telomere sequence motifs and is indistinguishable from RAPT. Curr. Genet. 1989, 16, 225–239. [Google Scholar] [CrossRef]
  140. König, P.; Giraldo, R.; Chapman, L.; Rhodes, D. The Crystal Structure of the DNA-Binding Domain of Yeast RAP1 in Complex with Telomeric DNA. Cell 1996, 85, 125–136. [Google Scholar] [CrossRef] [Green Version]
  141. König, P.; Rhodes, D. Recognition of telomeric DNA. Trends Biochem. Sci. 1997, 22, 43–47. [Google Scholar] [CrossRef]
  142. Luo, K.; Vega-Palas, M.A.; Grunstein, M. Rap1-Sir4 binding independent of other Sir, yKu, or histone interactions initiates the assembly of telomeric heterochromatin in yeast. Genes Dev. 2002, 16, 1528–1539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Moretti, P.; Shore, D. Multiple Interactions in Sir Protein Recruitment by Rap1p at Silencers and Telomeres in Yeast. Mol. Cell. Biol. 2001, 21, 8082–8094. [Google Scholar] [CrossRef] [Green Version]
  144. Cockell, M.; Palladino, F.; Laroche, T.; Kyrion, G.; Liu, C.; Lustig, A.J.; Gasser, S. The carboxy termini of Sir4 and Rap1 affect Sir3 localization: Evidence for a multicomponent complex required for yeast telomeric silencing. J. Cell Biol. 1995, 129, 909–924. [Google Scholar] [CrossRef] [PubMed]
  145. Liu, C.; Lustig, A.J. Genetic Analysis of Rap1p/Sir3p Interactions in Telomeric and HML Silencing in Saccharomyces cerevisiae. Genetics 1996, 143, 81–93. [Google Scholar] [CrossRef] [PubMed]
  146. Lustig, A.J.; Liu, C.; Zhang, C.; Hanish, J.P. Tethered Sir3p nucleates silencing at telomeres and internal loci in Saccharomyces cerevisiae. Mol. Cell. Biol. 1996, 16, 2483–2495. [Google Scholar] [CrossRef] [Green Version]
  147. Smith, J.S.; Brachmann, C.B.; Celic, I.; Kenna, M.A.; Muhammad, S.; Starai, V.J.; Avalos, J.L.; Escalante-Semerena, J.C.; Grubmeyer, C.; Wolberger, C.; et al. A phylogenetically conserved NAD+-dependent protein deacetylase activity in the Sir2 protein family. Proc. Natl. Acad. Sci. USA 2000, 97, 6658–6663. [Google Scholar] [CrossRef] [Green Version]
  148. Gartenberg, M.; Smith, J.S. The Nuts and Bolts of Transcriptionally Silent Chromatin in Saccharomyces cerevisiae. Genetics 2016, 203, 1563–1599. [Google Scholar] [CrossRef] [Green Version]
  149. Kyrion, G.; Liu, K.; Liu, C.; Lustig, A.J. RAP1 and telomere structure regulate telomere position effects in Saccharomyces cerevisiae. Genes Dev. 1993, 7, 1146–1159. [Google Scholar] [CrossRef] [PubMed]
  150. Buck, S.W.; Shore, D. Action of a RAP1 carboxy-terminal silencing domain reveals an underlying competition between HMR and telomeres in yeast. Genes Dev. 1995, 9, 370–384. [Google Scholar] [CrossRef] [Green Version]
  151. Rudenko, G.; Van Der Ploeg, L. Transcription of telomere repeats in protozoa. EMBO J. 1989, 8, 2633–2638. [Google Scholar] [CrossRef]
  152. Damasceno, J.D.; La Silva, G.; Tschudi, C.; Tosi, L.R. Evidence for regulated expression of Telomeric Repeat-containing RNAs (TERRA) in parasitic trypanosomatids. Memórias Do Inst. Oswaldo Cruz 2017, 112, 572–576. [Google Scholar] [CrossRef] [Green Version]
  153. Saha, A.; Nanavaty, V.P.; Li, B. Telomere and Subtelomere R-loops and Antigenic Variation in Trypanosomes. J. Mol. Biol. 2020, 432, 4167–4185. [Google Scholar] [CrossRef]
  154. Morea, E.G.O.; Vasconcelos, E.J.R.; de Santis Alves, C.; Giorgio, S.; Myler, P.J.; Langoni, H.; Azzalin, C.M.; Cano, M.I.N. Exploring TERRA during Leishmania major developmental cycle and continuous in vitro passages. Int. J. Biol. Macromol. 2021, 174, 573–586. [Google Scholar] [CrossRef]
  155. Azzalin, C.M.; Reichenbach, P.; Khoriauli, L.; Giulotto, E.; Lingner, J. Telomeric Repeat Containing RNA and RNA Surveillance Factors at Mammalian Chromosome Ends. Science 2007, 318, 798–801. [Google Scholar] [CrossRef]
  156. Schoeftner, S.; Blasco, M.A. Developmentally regulated transcription of mammalian telomeres by DNA-dependent RNA polymerase II. Nat. Cell Biol. 2007, 10, 228–236. [Google Scholar] [CrossRef]
  157. Bah, A.; Wischnewski, H.; Shchepachev, V.; Azzalin, C.M. The telomeric transcriptome of Schizosaccharomyces pombe. Nucleic Acids Res. 2011, 40, 2995–3005. [Google Scholar] [CrossRef] [Green Version]
  158. Luke, B.; Panza, A.; Redon, S.; Iglesias, N.; Li, Z.; Lingner, J. The Rat1p 5′ to 3′ Exonuclease Degrades Telomeric Repeat-Containing RNA and Promotes Telomere Elongation in Saccharomyces cerevisiae. Mol. Cell 2008, 32, 465–477. [Google Scholar] [CrossRef]
  159. Solovei, I.; Gaginskaya, E.; MacGregor, H.C. The arrangement and transcription of telomere DNA sequences at the ends of lampbrush chromosomes of birds. Chromosom. Res. 1994, 2, 460–470. [Google Scholar] [CrossRef] [PubMed]
  160. Toubiana, S.; Selig, S. DNA: RNA hybrids at telomeres–When it is better to be out of the (R) loop. FEBS J. 2018, 285, 2552–2566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Li, B. Keeping Balance Between Genetic Stability and Plasticity at the Telomere and Subtelomere of Trypanosoma brucei. Front. Cell Dev. Biol. 2021, 9, 1722. [Google Scholar] [CrossRef] [PubMed]
  162. Nergadze, S.G.; Farnung, B.O.; Wischnewski, H.; Khoriauli, L.; Vitelli, V.; Chawla, R.; Giulotto, E.; Azzalin, C.M. CpG-island promoters drive transcription of human telomeres. RNA 2009, 15, 2186–2194. [Google Scholar] [CrossRef] [Green Version]
  163. Iglesias, N.; Redon, S.; Pfeiffer, V.; Dees, M.; Lingner, J.; Luke, B. Subtelomeric repetitive elements determine TERRA regulation by Rap1/Rif and Rap1/Sir complexes in yeast. EMBO Rep. 2011, 12, 587–593. [Google Scholar] [CrossRef] [PubMed]
  164. Porro, A.; Feuerhahn, S.; Delafontaine, J.; Riethman, H.; Rougemont, J.; Lingner, J. Functional characterization of the TERRA transcriptome at damaged telomeres. Nat. Commun. 2014, 5, 1–13. [Google Scholar] [CrossRef]
  165. Feretzaki, M.; Nunes, P.R.; Lingner, J. Expression and differential regulation of human TERRA at several chromosome ends. RNA 2019, 25, 1470–1480. [Google Scholar] [CrossRef]
  166. Deng, Z.; Wang, Z.; Stong, N.; Plasschaert, R.; Moczan, A.; Chen, H.-S.; Hu, S.; Wikramasinghe, P.; Davuluri, R.V.; Bartolomei, M.S.; et al. A role for CTCF and cohesin in subtelomere chromatin organization, TERRA transcription, and telomere end protection. EMBO J. 2012, 31, 4165–4178. [Google Scholar] [CrossRef] [Green Version]
  167. Feretzaki, M.; Lingner, J. A practical qPCR approach to detect TERRA, the elusive telomeric repeat-containing RNA. Methods 2017, 114, 39–45. [Google Scholar] [CrossRef] [PubMed]
  168. Arora, R.; Brun, C.M.; Azzalin, C.M. Transcription regulates telomere dynamics in human cancer cells. RNA 2012, 18, 684–693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Peška, V.; Schrumpfová, P.P.; Fajkus, J. Using the telobox to search for plant telomere binding proteins. Curr. Protein Pept. Sci. 2011, 12, 75–83. [Google Scholar] [CrossRef] [PubMed]
  170. Lue, N.F. Duplex Telomere-Binding Proteins in Fungi with Canonical Telomere Repeats: New Lessons in the Rapid Evolution of Telomere Proteins. Front. Genet. 2021, 12, 238. [Google Scholar] [CrossRef]
  171. Horvath, M.P. Structural anatomy of telomere OB proteins. Crit. Rev. Biochem. Mol. Biol. 2011, 46, 409–435. [Google Scholar] [CrossRef] [Green Version]
  172. Amir, M.; Khan, P.; Queen, A.; Dohare, R.; Alajmi, M.F.; Hussain, A.; Islam, A.; Ahmad, F.; Hassan, I. Structural Features of Nucleoprotein CST/Shelterin Complex Involved in the Telomere Maintenance and Its Association with Disease Mutations. Cells 2020, 9, 359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Lewis, K.A.; Wuttke, D.S. Telomerase and Telomere-Associated Proteins: Structural Insights into Mechanism and Evolution. Structure 2012, 20, 28–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Ogata, K.; Morikawa, S.; Nakamura, H.; Sekikawa, A.; Inoue, T.; Kanai, H.; Sarai, A.; Ishii, S.; Nishimura, Y. Solution structure of a specific DNA complex of the Myb DNA-binding domain with cooperative recognition helices. Cell 1994, 79, 639–648. [Google Scholar] [CrossRef]
  175. Van Steensel, B.; de Lange, T. Control of telomere length by the human telomeric protein TRF1. Nat. Cell Biol. 1997, 385, 740–743. [Google Scholar] [CrossRef]
  176. Van Steensel, B.; Smogorzewska, A.; de Lange, T. TRF2 Protects Human Telomeres from End-to-End Fusions. Cell 1998, 92, 401–413. [Google Scholar] [CrossRef] [Green Version]
  177. Cooper, J.P.; Nimmo, E.R.; Allshire, R.; Cech, T.R. Regulation of telomere length and function by a Myb-domain protein in fission yeast. Nat. Cell Biol. 1997, 385, 744–747. [Google Scholar] [CrossRef]
  178. Spink, K.G.; Evans, R.J.; Chambers, A. Sequence-specific binding of Taz1p dimers to fission yeast telomeric DNA. Nucleic Acids Res 2000, 28, 527–533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Koering, C.E.; Fourel, G.; Binet-Brasselet, E.; Laroche, T.; Klein, F.; Gilson, E. Identification of high affinity Tbf1p-binding sites within the budding yeast genome. Nucleic Acids Res. 2000, 28, 2519–2526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Berthiau, A.-S.; Yankulov, K.; Bah, A.; Revardel, E.; Luciano, P.; Wellinger, R.J.; Géli, V.; Gilson, E. Subtelomeric proteins negatively regulate telomere elongation in budding yeast. EMBO J. 2006, 25, 846–856. [Google Scholar] [CrossRef] [Green Version]
  181. Gottschling, D.E.; Zakian, V.A. Telomere proteins: Specific recognition and protection of the natural termini of Oxytricha macronuclear DNA. Cell 1986, 47, 195–205. [Google Scholar] [CrossRef]
  182. Gray, J.T.; Celander, D.W.; Price, C.; Cech, T.R. Cloning and expression of genes for the Oxytricha telomere-binding protein: Specific subunit interactions in the telomeric complex. Cell 1991, 67, 807–814. [Google Scholar] [CrossRef]
  183. Horvath, M.P.; Schweiker, V.L.; Bevilacqua, J.M.; Ruggles, J.A.; Schultz, S.C. Crystal Structure of the Oxytricha nova Telomere End Binding Protein Complexed with Single Strand DNA. Cell 1998, 95, 963–974. [Google Scholar] [CrossRef] [Green Version]
  184. Classen, S.; Ruggles, J.A.; Schultz, S.C. Crystal structure of the N-terminal domain of Oxytricha nova telomere end-binding protein α subunit both uncomplexed and complexed with telomeric ssDNA. J. Mol. Biol. 2001, 314, 1113–1125. [Google Scholar] [CrossRef]
  185. Buczek, P.; Horvath, M.P. Structural Reorganization and the Cooperative Binding of Single-stranded Telomere DNA in Sterkiella nova. J. Biol. Chem. 2006, 281, 40124–40134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Lei, M.; Podell, E.R.; Cech, T.R. Structure of human POT1 bound to telomeric single-stranded DNA provides a model for chromosome end-protection. Nat. Struct. Mol. Biol. 2004, 11, 1223–1229. [Google Scholar] [CrossRef]
  187. Nandakumar, J.; Podell, E.R.; Cech, T.R. How telomeric protein POT1 avoids RNA to achieve specificity for single-stranded DNA. Proc. Natl. Acad. Sci. USA 2010, 107, 651–656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Hockemeyer, D.; Palm, W.; Else, T.; Daniels, J.-P.; Takai, K.K.; Ye, J.Z.-S.; Keegan, C.E.; de Lange, T.; Hammer, G.D. Telomere protection by mammalian Pot1 requires interaction with Tpp1. Nat. Struct. Mol. Biol. 2007, 14, 754–761. [Google Scholar] [CrossRef]
  189. Xin, H.; Liu, D.; Wan, M.; Safari, A.; Kim, H.; Sun, W.; O’Connor, M.S.; Songyang, Z. TPP1 is a homologue of ciliate TEBP-β and interacts with POT1 to recruit telomerase. Nat. Cell Biol. 2007, 445, 559–562. [Google Scholar] [CrossRef]
  190. Wang, F.; Podell, E.R.; Zaug, A.J.; Yang, Y.; Baciu, P.; Cech, T.R.; Lei, M. The POT1–TPP1 telomere complex is a telomerase processivity factor. Nat. Cell Biol. 2007, 445, 506–510. [Google Scholar] [CrossRef]
  191. Gao, H.; Cervantes, R.B.; Mandell, E.K.; Otero, J.H.; Lundblad, V. RPA-like proteins mediate yeast telomere function. Nat. Struct. Mol. Biol. 2007, 14, 208–214. [Google Scholar] [CrossRef]
  192. Sun, J.; Yu, E.Y.; Yang, Y.; Confer, L.A.; Sun, S.H.; Wan, K.; Lue, N.F.; Lei, M. Stn1-Ten1 is an Rpa2-Rpa3-like complex at telomeres. Genes Dev. 2009, 23, 2900–2914. [Google Scholar] [CrossRef] [Green Version]
  193. Bryan, C.; Rice, C.; Harkisheimer, M.; Schultz, D.C.; Skordalakes, E. Structure of the Human Telomeric Stn1-Ten1 Capping Complex. PLoS ONE 2013, 8, e66756. [Google Scholar] [CrossRef] [PubMed]
  194. Anderson, E.M.; Halsey, W.A.; Wuttke, D.S. Delineation of the high-affinity single-stranded telomeric DNA-binding domain of Saccharomyces cerevisiae Cdc13. Nucleic Acids Res. 2002, 30, 4305–4313. [Google Scholar] [CrossRef] [Green Version]
  195. Eldridge, A.M.; Wuttke, D.S. Probing the mechanism of recognition of ssDNA by the Cdc13-DBD. Nucleic Acids Res. 2008, 36, 1624–1633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Mitton-Fry, R.M.; Anderson, E.M.; Hughes, T.R.; Lundblad, V.; Wuttke, D.S. Conserved Structure for Single-Stranded Telomeric DNA Recognition. Science 2002, 296, 145–147. [Google Scholar] [CrossRef] [PubMed]
  197. Mitton-Fry, R.M.; Anderson, E.M.; Theobald, D.L.; Glustrom, L.W.; Wuttke, D.S. Structural Basis for Telomeric Single-stranded DNA Recognition by Yeast Cdc13. J. Mol. Biol. 2004, 338, 241–255. [Google Scholar] [CrossRef]
  198. Theobald, D.; Wuttke, D.S. Prediction of Multiple Tandem OB-Fold Domains in Telomere End-Binding Proteins Pot1 and Cdc13. Structure 2004, 12, 1877–1879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Sun, J.; Yang, Y.; Wan, K.; Mao, N.; Yu, T.-Y.; Lin, Y.-C.; DeZwaan, D.C.; Freeman, B.C.; Lin, J.-J.; Lue, N.F.; et al. Structural bases of dimerization of yeast telomere protein Cdc13 and its interaction with the catalytic subunit of DNA polymerase α. Cell Res. 2010, 21, 258–274. [Google Scholar] [CrossRef] [Green Version]
  200. Yu, E.Y.; Sun, J.; Lei, M.; Lue, N.F. Analyses of Candida Cdc13 Orthologues Revealed a Novel OB Fold Dimer Arrangement, Dimerization-Assisted DNA Binding, and Substantial Structural Differences between Cdc13 and RPA. Mol. Cell. Biol. 2012, 32, 186–198. [Google Scholar] [CrossRef] [Green Version]
  201. Bhattacharjee, A.; Stewart, J.; Chaiken, M.; Price, C.M. STN1 OB Fold Mutation Alters DNA Binding and Affects Selective Aspects of CST Function. PLoS Genet. 2016, 12, e1006342. [Google Scholar] [CrossRef] [Green Version]
  202. Bhattacharjee, A.; Wang, Y.; Diao, J.; Price, C.M. Dynamic DNA binding, junction recognition and G4 melting activity underlie the telomeric and genome-wide roles of human CST. Nucleic Acids Res. 2017, 45, 12311–12324. [Google Scholar] [CrossRef] [Green Version]
  203. Hom, R.A.; Wuttke, D.S. Human CST Prefers G-Rich but Not Necessarily Telomeric Sequences. Biochemistry 2017, 56, 4210–4218. [Google Scholar] [CrossRef] [PubMed]
  204. Li, B.; Espinal, A.; Cross, G.A.M. Trypanosome Telomeres Are Protected by a Homologue of Mammalian TRF. Mol. Cell. Biol. 2005, 25, 5011–5021. [Google Scholar] [CrossRef] [Green Version]
  205. Reis, H.; Schwebs, M.; Dietz, S.; Janzen, C.J.; Butter, F. TelAP1 links telomere complexes with developmental expression site silencing in African trypanosomes. Nucleic Acids Res. 2018, 46, 2820–2833. [Google Scholar] [CrossRef]
  206. Dreesen, O.; Li, B.; Cross, G.A.M. Telomere structure and shortening in telomerase-deficient Trypanosoma brucei. Nucleic Acids Res. 2005, 33, 4536–4543. [Google Scholar] [CrossRef]
  207. Sandhu, R.; Sanford, S.; Basu, S.; Park, M.; Pandya, U.M.; Li, B.; Chakrabarti, K. A trans-spliced telomerase RNA dictates telomere synthesis in Trypanosoma brucei. Cell Res. 2013, 23, 537–551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Gupta, S.K.; Kolet, L.; Doniger, T.; Biswas, V.K.; Unger, R.; Tzfati, Y.; Michaeli, S. The Trypanosoma brucei telomerase RNA (TER) homologue binds core proteins of the C/D snoRNA family. FEBS Lett. 2013, 587, 1399–1404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Da Silva, M.S.; Perez, A.M.; Rita de Cássia, V.; de Moraes, C.E.; Siqueira-Neto, J.L.; Freitas-Junior, L.H.; Cano, M.I.N. The Leishmania amazonensis TRF (TTAGGG repeat-binding factor) homologue binds and co-localizes with telomeres. BMC Microbiol. 2010, 10, 136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Fairall, L.; Chapman, L.; Moss, H.; de Lange, T.; Rhodes, D. Structure of the TRFH Dimerization Domain of the Human Telomeric Proteins TRF1 and TRF2. Mol. Cell 2001, 8, 351–361. [Google Scholar] [CrossRef]
  211. Crossley, M.P.; Bocek, M.; Cimprich, K.A. R-Loops as Cellular Regulators and Genomic Threats. Mol. Cell 2019, 73, 398–411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Hegazy, Y.A.; Fernando, C.M.; Tran, E.J. The balancing act of R-loop biology: The good, the bad, and the ugly. J. Biol. Chem. 2020, 295, 905–913. [Google Scholar] [CrossRef]
  213. Brambati, A.; Zardoni, L.; Nardini, E.; Pellicioli, A.; Liberi, G. The dark side of RNA: DNA hybrids. Mutat. Res. Rev. Mutat. Res. 2020, 784, 108300. [Google Scholar] [CrossRef]
  214. König, P.; Fairall, L.; Rhodes, D. Sequence-specific DNA recognition by the Myb-like domain of the human telomere binding protein TRF1: A model for the protein-DNA complex. Nucleic Acids Res. 1998, 26, 1731–1740. [Google Scholar] [CrossRef] [Green Version]
  215. Nishikawa, T.; Nagadoi, A.; Yoshimura, S.; Aimoto, S.; Nishimura, Y. Solution structure of the DNA-binding domain of human telomeric protein, hTRF1. Structure 1998, 6, 1057–1065. [Google Scholar] [CrossRef] [Green Version]
  216. Court, R.; Chapman, L.; Fairall, L.; Rhodes, D. How the human telomeric proteins TRF1 and TRF2 recognize telomeric DNA: A view from high-resolution crystal structures. EMBO Rep. 2005, 6, 39–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Borst, P.; Sabatini, R. Base J: Discovery, Biosynthesis, and Possible Functions. Annu. Rev. Microbiol. 2008, 62, 235–251. [Google Scholar] [CrossRef] [PubMed]
  218. Gommers-Ampt, J.H.; Van Leeuwen, F.; De Beer, A.L.; Vliegenthart, J.F.; Dizdaroglu, M.; Kowalak, J.A.; Crain, P.F.; Borst, P. β-d-glucosyl-hydroxymethyluracil: A novel modified base present in the DNA of the parasitic protozoan T. brucei. Cell 1993, 75, 1129–1136. [Google Scholar] [CrossRef]
  219. Van Leeuwen, F.; Wijsman, E.R.; Kuyl-Yeheskiely, E.; Van Der Marel, G.A.; Van Boom, J.H.; Borst, P. The telomeric GGGTTA repeats of Trypanosoma brucei contain the hypermodified base J in both strands. Nucleic Acids Res. 1996, 24, 2476–2482. [Google Scholar] [CrossRef] [Green Version]
  220. Van Leeuwen, F.; Wijsman, E.R.; Kieft, R.; Van Der Marel, G.A.; Van Boom, J.H.; Borst, P. Localization of the modified base J in telomeric VSG gene expression sites of Trypanosoma brucei. Genes Dev. 1997, 11, 3232–3241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Borst, P.; Van Leeuwen, F. ß-D-glucosyl-hydroxymethyluracil, a novel base in African trypanosomes and other Kinetoplastida. Mol. Biochem. Parasitol. 1997, 90, 1–8. [Google Scholar] [CrossRef]
  222. Van Leeuwen, F.; Taylor, M.C.; Mondragon, A.; Moreau, H.; Gibson, W.; Kieft, R.; Borst, P. ß-d-glucosyl-hydroxymethyluracil is a conserved DNA modification in kinetoplastid protozoans and is abundant in their telomeres. Proc. Natl. Acad. Sci. USA 1998, 95, 2366–2371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Ekanayake, D.; Sabatini, R. Epigenetic Regulation of Polymerase II Transcription Initiation in Trypanosoma cruzi: Modulation of Nucleosome Abundance, Histone Modification, and Polymerase Occupancy by O-Linked Thymine DNA Glucosylation. Eukaryot. Cell 2011, 10, 1465–1472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Reynolds, D.; Cliffe, L.; Förstner, K.U.; Hon, C.-C.; Siegel, T.N.; Sabatini, R. Regulation of transcription termination by glucosylated hydroxymethyluracil, base J, in Leishmania major and Trypanosoma brucei. Nucleic Acids Res. 2014, 42, 9717–9729. [Google Scholar] [CrossRef]
  225. Reynolds, D.; Hofmeister, B.T.; Cliffe, L.; Alabady, M.; Siegel, T.N.; Schmitz, R.J.; Sabatini, R. Histone H3 Variant Regulates RNA Polymerase II Transcription Termination and Dual Strand Transcription of siRNA Loci in Trypanosoma brucei. PLoS Genet. 2016, 12, e1005758. [Google Scholar] [CrossRef] [Green Version]
  226. Deng, Z.; Norseen, J.; Wiedmer, A.; Riethman, H.; Lieberman, P.M. TERRA RNA Binding to TRF2 Facilitates Heterochromatin Formation and ORC Recruitment at Telomeres. Mol. Cell 2009, 35, 403–413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Mei, Y.; Deng, Z.; Vladimirova, O.; Gulve, N.; Johnson, F.B.; Drosopoulos, W.C.; Schildkraut, C.L.; Lieberman, P.M. TERRA G-quadruplex RNA interaction with TRF2 GAR domain is required for telomere integrity. Sci. Rep. 2021, 11, 3509. [Google Scholar] [CrossRef] [PubMed]
  228. Lee, Y.W.; Arora, R.; Wischnewski, H.; Azzalin, C.M. TRF1 participates in chromosome end protection by averting TRF2-dependent telomeric R loops. Nat. Struct. Mol. Biol. 2018, 25, 147–153. [Google Scholar] [CrossRef]
  229. Prouse, M.B.; Campbell, M. The interaction between MYB proteins and their target DNA binding sites. Biochim. Biophys. Acta (BBA) Bioenerg. 2012, 1819, 67–77. [Google Scholar] [CrossRef]
  230. Navarro, M.; Gull, K. A pol I transcriptional body associated with VSG mono-allelic expression in Trypanosoma brucei. Nat. Cell Biol. 2001, 414, 759–763. [Google Scholar] [CrossRef]
  231. Liu, C.; Mao, X.; Lustig, A.J. Mutational analysis defines a C-terminal tail domain of RAP1 essential for Telomeric silencing in Saccharomyces cerevisiae. Genetics 1994, 138, 1025–1040. [Google Scholar] [CrossRef]
  232. Kanoh, J.; Ishikawa, F. spRap1 and spRif1, recruited to telomeres by Taz1, are essential for telomere function in fission yeast. Curr. Biol. 2001, 11, 1624–1630. [Google Scholar] [CrossRef] [Green Version]
  233. Park, M.J.; Jang, Y.K.; Choi, E.S.; Kim, H.S.; Park, S.D. Fission yeast Rap1 homolog is a telomere-specific silencing factor and interacts with Taz1p. Mol. Cells 2002, 13, 327–333. [Google Scholar]
  234. Pena, A.C.; Pimentel, M.R.; Manso, H.; Vaz-Drago, R.; Pinto-Neves, D.; Aresta-Branco, F.; Rijo-Ferreira, F.; Guegan, F.; Coelho, L.P.; Carmo-Fonseca, M. Trypanosoma brucei histone H1 inhibits RNA polymerase I transcription and is important for parasite fitness in vivo. Mol. Microbiol. 2014, 93, 645–663. [Google Scholar] [CrossRef] [Green Version]
  235. Alsford, S.; Horn, D. Cell-cycle-regulated control of VSG expression site silencing by histones and histone chaperones ASF1A and CAF-1b in Trypanosoma brucei. Nucleic Acids Res. 2012, 40, 10150–10160. [Google Scholar] [CrossRef] [PubMed]
  236. Figueiredo, L.M.; Janzen, C.J.; Cross, G.A.M. A Histone Methyltransferase Modulates Antigenic Variation in African Trypanosomes. PLoS Biol. 2008, 6, e161. [Google Scholar] [CrossRef] [Green Version]
  237. Shore, D.; Nasmyth, K. Purification and cloning of a DNA binding protein from yeast that binds to both silencer and activator elements. Cell 1987, 51, 721–732. [Google Scholar] [CrossRef]
  238. Azad, G.K.; Tomar, R.S. The multifunctional transcription factor Rap1: A regulator of yeast physiology. Front. Biosci. 2016, 21, 918–930. [Google Scholar]
  239. Martinez, P.; Thanasoula, M.; Carlos, A.R.; López, G.G.; Tejera, A.M.; Schoeftner, S.; Dominguez, O.; Pisano, D.; Tarsounas, M.; Blasco, M.A. Mammalian Rap1 controls telomere function and gene expression through binding to telomeric and extratelomeric sites. Nat. Cell Biol. 2010, 12, 768–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Yeung, F.; Ramírez, C.M.; Mateos-Gomez, P.A.; Pinzaru, A.; Ceccarini, G.; Kabir, S.; Fernández-Hernando, C.; Sfeir, A. Nontelomeric Role for Rap1 in Regulating Metabolism and Protecting against Obesity. Cell Rep. 2013, 3, 1847–1856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Yang, D.; Xiong, Y.; Kim, H.; He, Q.; Li, Y.; Chen, R.; Songyang, Z. Human telomeric proteins occupy selective interstitial sites. Cell Res. 2011, 21, 1013–1027. [Google Scholar] [CrossRef] [Green Version]
  242. Teo, H.; Ghosh, S.; Luesch, H.; Ghosh, A.; Wong, E.T.; Malik, N.; Orth, A.; De Jesus, P.; Pery, A.S.; Oliver, J.D. Telomere-independent Rap1 is an IKK adaptor and regulates NF-kappaB-dependent gene expression. Nat. Cell Biol. 2010, 12, 758–767. [Google Scholar] [CrossRef] [PubMed]
  243. Kabir, S.; Sfeir, A.; De Lange, T. Taking apart Rap1: An adaptor protein with telomeric and non-telomeric functions. Cell Cycle 2010, 9, 4061–4067. [Google Scholar] [CrossRef] [Green Version]
  244. Hanaoka, S.; Nagadoi, A.; Yoshimura, S.; Aimoto, S.; Li, B.; de Lange, T.; Nishimura, Y. NMR structure of the hRap1 myb motif reveals a canonical three-helix bundle lacking the positive surface charge typical of myb DNA-binding domains. J. Mol. Biol. 2001, 312, 167–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Nett, I.R.E.; Martin, D.; Miranda-Saavedra, D.; Lamont, D.; Barber, J.D.; Mehlert, A.; Ferguson, M.A.J. The Phosphoproteome of Bloodstream Form Trypanosoma brucei, Causative Agent of African Sleeping Sickness. Mol. Cell. Proteom. 2009, 8, 1527–1538. [Google Scholar] [CrossRef] [Green Version]
  246. Urbaniak, M.; Martin, D.M.A.; Ferguson, M.A.J. Global Quantitative SILAC Phosphoproteomics Reveals Differential Phosphorylation Is Widespread between the Procyclic and Bloodstream Form Lifecycle Stages of Trypanosoma brucei. J. Proteome Res. 2013, 12, 2233–2244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Barry, J.D.; McCulloch, R. Antigenic variation in trypanosomes: Enhanced phenotypic variation in a eukaryotic parasite. Adv. Parasitol. 2001, 49, 1–70. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Representative VSG loci in T. brucei. (A) A VSG gene array. (B) A minichromosome with a subtelomeric VSG gene. (C) A typical BF VSG ES. ESAG: ES-Associated Gene. (D) A typical metacyclic VSG ES. (E) Only one VSG ES is fully active at any time. Dots at the bottom represent multiple similar BF ESs of which details are not shown.
Figure 1. Representative VSG loci in T. brucei. (A) A VSG gene array. (B) A minichromosome with a subtelomeric VSG gene. (C) A typical BF VSG ES. ESAG: ES-Associated Gene. (D) A typical metacyclic VSG ES. (E) Only one VSG ES is fully active at any time. Dots at the bottom represent multiple similar BF ESs of which details are not shown.
Pathogens 10 00967 g001
Figure 2. Major VSG switching pathways. VSGac, the active VSG; VSGsil, a silent VSG. Long green arrow represents the active ES promoter, and short red arrow represents a silent ES promoter.
Figure 2. Major VSG switching pathways. VSGac, the active VSG; VSGsil, a silent VSG. Long green arrow represents the active ES promoter, and short red arrow represents a silent ES promoter.
Pathogens 10 00967 g002
Figure 3. The T. brucei telomere protein complex. Core telomere protein components are shown.
Figure 3. The T. brucei telomere protein complex. Core telomere protein components are shown.
Pathogens 10 00967 g003
Table 1. Summary of nucleic acid binding activities of TRF homologs.
Table 1. Summary of nucleic acid binding activities of TRF homologs.
TRF Homologs
SpeciesProteinDNA BindingTERRA BindingHomodimerization
Duplex Telomere DNA BindingBinding DomainTERRA BindingBinding DomainActivityDomain
HumanhTRF1binds ds(TTAGGG)nC-terminal MybNot reportedN/AYesTRFH
hTRF2binds ds(TTAGGG)nC-terminal Mybbinds (UUAGGG)nN-terminal GAR domainYesTRFH
Budding yeastScTBF1binds ds(TTAGGG)nC-terminal MybNot reportedN/A
Fission yeastSpTAZ1binds ds[G2–8TTAC(A)]nC-terminal MybNot reportedN/AYes
SpTBF1binds ds[G2–8TTAC(A)]nC-terminal MybNot reportedN/A
T. bruceiTbTRFbinds ds(TTAGGG)nC-terminal Mybbinds (UUAGGG)nC-terminal MybYesTRFH
Table 2. Summary of DNA binding activities of RAP1 homologs.
Table 2. Summary of DNA binding activities of RAP1 homologs.
RAP1 Homologs
SpeciesProteinDNA BindingDNA Binding Domain
HumanhRAP1NoN/A
Budding yeastScRAP1binds ds(TG1–3)nMyb and Myb-Like
Fission yeastSpRAP1NoN/A
T. bruceiTbRAP1binds ds- & ssDNA737RKRRR741
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, B.; Zhao, Y. Regulation of Antigenic Variation by Trypanosoma brucei Telomere Proteins Depends on Their Unique DNA Binding Activities. Pathogens 2021, 10, 967. https://doi.org/10.3390/pathogens10080967

AMA Style

Li B, Zhao Y. Regulation of Antigenic Variation by Trypanosoma brucei Telomere Proteins Depends on Their Unique DNA Binding Activities. Pathogens. 2021; 10(8):967. https://doi.org/10.3390/pathogens10080967

Chicago/Turabian Style

Li, Bibo, and Yanxiang Zhao. 2021. "Regulation of Antigenic Variation by Trypanosoma brucei Telomere Proteins Depends on Their Unique DNA Binding Activities" Pathogens 10, no. 8: 967. https://doi.org/10.3390/pathogens10080967

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

Article Metrics

Back to TopTop