Habilitation Thesis Functions of Plant Proteins associated with Telomeric Repeats and Telomerase Mgr. Petra Procházková Schrumpfová, Ph.D. Brno, 2023 Acknowledgement I would like to express my deep gratitude to the head of the group, Prof. Jiří Fajkus, who has not only continuously supported my research, but also motivated me and provided fruitful scientific comments and advice. Additionally, he has been the person I turned to during challenging moments of my life. I would also like to acknowledge both current and former colleagues who contributed to the friendly, tolerant, and supportive atmosphere. Thanks also go to many of my students for their enthusiasm, patience, and contribution to the successful work on my scientific project. Their efforts exceeded my expectations, especially when considering the context of my maternity leave, which made our work more demanding. The greatest thanks belong to my family, especially my husband Pavel, who has supported me in all aspects of my life and decisions, and my lovely two children, Verunka and Martínek. Contents Commentary….………………………………………………………………………………….….......................................................4 Telomeric repeats and their interactome ……………………………………………………….............................................8 1 Telomeric repeats at the physical ends of linear chromosomes - telomeres............................... 8 1.1 Proteins associated with telomeres in mammals ...............................................................12 1.2 Proteins associated with telomeres in plants .....................................................................15 1.2.1 Telomeric dsDNA associated proteins in plants..........................................................15 Smh/TRB family............................................................................................................16 TRFL family...................................................................................................................22 AID family.....................................................................................................................23 1.2.2 Telomeric ssDNA associated proteins in lants ............................................................23 Proteins with OB-fold...................................................................................................23 Non-OB-fold proteins...................................................................................................26 2 Interstitially located telomeric repeats.......................................................................................27 2.1 Proteins associated with long interstitial telomeric repeats (ITSs).....................................28 2.2 Proteins associated with short internally localized telomeric repeats (telo-boxes)...........29 3 Orchestration of telomere homeostasis.....................................................................................31 3.1 Telomerase..........................................................................................................................32 3.1.1 Telomerase reverse transcriptase (TERT) ...................................................................33 3.1.2 Telomerase RNA (TR) ..................................................................................................34 3.1.3 Telomerase-associated proteins .................................................................................35 3.1.4 Telomerase assembly..................................................................................................38 3.2 Telomere maintenance proteins.........................................................................................44 3.2.1 Mammalian telomere maintenance proteins .............................................................44 3.2.2 Plant telomere maintenance proteins ........................................................................45 3.2.3 HMG proteins..............................................................................................................46 3.3 Telomerase-independent telomere maintenance..............................................................47 Conclusion….………………………………………………………………………………….….........................................................51 Bibliography………………………………………………………………………………………………………………………………………….....52 Supplements…………………………………………………………………………………………………………………………………………….67 4 Commentary This Habilitation Thesis is a compilation of selected scientific publications in which I have contributed as a primary author or co-author. The primary goal of my research was to contribute to a fundamental understanding of the nature and characteristics of proteins and enzymes associated with long or short telomeric repeats. At the very beginning of my scientific career, I initially focused on characterizing previously unknown plant putative homologues of proteins linked to telomeric sequences in vitro, known as TRBs. Interestingly, this protein family turned out to be a fruitful discovery in my scientific journey. Over the course of nearly two decades, we have uncovered that these proteins not only interact with the physical ends of chromosomes but also serve as the first described plant interactors of telomerase, the enzyme responsible for elongating telomeres. Furthermore, I found out that TRB proteins are associated with short telomeric sequences in the promoters of various genes, which resulted in very fruitful collaborations with laboratories investigating the epigenetic regulation of gene transcription. Moreover, my exploration of telomeric sequences and associated proteins has led me to the characterization of other diverse proteins linked to telomeric sequences or telomerase, such as RUVBLs, HMGBs, POTs, the PRC2 complex, and many others. These investigations have also touched subjects like plant gametogenesis, alternative lengthening of telomeres, and even the development of novel software for detecting short regulatory motifs within gene promoters. To investigate and characterize the proteins associated with telomeric repeats and telomerase, we employed a range of general biochemical and molecular biological techniques. These included cloning, protein expression and purification, Electrophoretic mobility shift assay (EMSA), yeast two hydrid assay (Y2H), Co-Immunoprecipitation (CoIP), Bimolecular Fluorescence Complementation (BiFC), Tandem Affinity Purification (TAP-tag), Chromatin Immunoprecipitation (ChIP) followed by Next-gen sequencing or hybridization and many others. Additionally, I utilized specialized techniques focused on telomeres or telomerase, such as Telomere Restriction Fragment (TRF) analysis and the Telomerase Repeated Amplification protocol (TRAP). In order to characterize plant material, I also employed techniques necessary for the analysis of T-DNA insertion mutant plants or plant cell cultures. Furthermore, as a supervisor for several bachelor's, master's, and doctoral theses, as well as a principal investigator of grants or a member of the grant-investigating team, I took on the responsibility of establishing and conducting research on various topics. This included the characterization of protein localization using microscopic techniques, the investigation of plant gametogenesis, and even the bioinformatic analysis. 5 Our findings have already been published in total of 20 publications on WoS, including articles and reviews. Among these, there is 1 correction to a research article (Schrumpfová and Majerská et al., Protoplasma, 2017), where Schrumpfová P.P. was recognized as the first co-author. Additionally, apart from these 20 publications, there is 1 book chapter published in Methods in Molecular Biology, The Nucleus, Book Series, Springer protocols (Schořová et al., 2020) and 1 Meeting abstract (Schrumpfová et al., 2005, FEBS Journal) listed in WoS. Among these 20 publications, I have served as the primary author in 9 of them and as the corresponding author in 7. The habitation theses comprise a compilation of 18 of these publications relevant to the thesis title: *Corresponding Author [1] Teano G., Concia L., Wolff L., Carron L., Biocanin I., Adamusová K., Fojtová M., Bourge M., Kramdi A., Colot V., Grossniklaus U., Bowler Ch., Baroux C., Carbone A., Probst A.V., Schrumpfová P.P., Fajkus J., Amiard S., Grob S., Bourbousse C., and Barneche F. 2023. Histone H1 protects telomeric repeats from H3K27me3 invasion in Arabidopsis. Cell Reports, 42(8):112894. (IF 8.8, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 0% 5% 5% 10% [2 ] Kusová A., Steinbachová L., Přerovská T., Drábková Z.L., Paleček J, Khan A., Rigóová G., Gadiou Z., Jourdain C., Stricker T., Schubert D., Honys D., and Schrumpfová P.P.*. Completing the TRB family: newly characterized members show ancient evolutionary origins and distinct localization, yet similar interactions. PLANT MOLECULAR BIOLOGY 112:61–83 2023 (IF 4.076, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 5% 50% 90% 90% [3] Tomaštíková E.D., Yang F., Mlynárová K., Hafidh S., Schořová Š., Kusová A., Pernisová M., Přerovská T., Klodová B., Honys D., Fajkus J., Pecinka A., Schrumpfová P.P.*. RUVBL proteins are involved in plant gametophyte development. THE PLANT JOURNAL, Volume: 114, Issue: 2 Pages: 325–337, 2023 (IF 7.091, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 5% 90% 90% 80% [4] Pecinka A., Schrumpfova P.P., Fisher L., Tomastikova E., Mozgova I. The Czech Plant Nucleus Workshop. Biologia Plantarum, Volume: 66, Pages: 39-45, 2021 (IF 1.122, Q4) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) - - 25% [5] Schrumpfová P.P.* and Fajkus J. Composition and Function of Telomerase—A Polymerase Associated with the Origin of Eukaryotes. Biomolecules, 10(10), 2020 (IF 4.879, Q2) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) - - 80% [6] Schořová Š., Fajkus J., Drábkova L., Honys D., Schrumpfová P.P.* The plant Pontin and Reptin homologues, RuvBL1 and RuvBL2a, colocalize with TERT and TRB proteins in vivo, and participate in telomerase biogenesis. THE PLANT JOURNAL, 98(2):195-212, 2019 (IF 5.775, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) - 80% 90% 90% 6 [7] Schrumpfová P.P., Fojtova M., Fajkus J. Telomeres in Plants and Humans: Not So Different, Not So Similar. CELLS Volume: 8 Issue: 1 Article Number: 58 Published: 2019 (IF 4.366, Q2) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) - - 40% [8] Schrumpfová P.P., Majerská J., Dokládal L., Schořová Š., Stejskal K., Obořil M., Honys D., Kozáková, L., Polanská P.S., Sýkorová E. Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo. PROTOPLASMA Volume: 254 Issue: 4 Pages: 1547-1562, 2017 (IF 2.457, Q2) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 20% 70% 5% 10% [9] Schrumpfová P.P.*, Schořová Š., Fajkus J. Telomere- and telomerase-associated proteins and their functions in the plant cell. FRONTIERS IN PLANT SCIENCE Volume: 7 Article Number: 851, 2016 (IF 4.291, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) - - 80% [10] Schrumpfová P.P., Vychodilová I., Hapala J., Schořová Š., Dvořáček V., Fajkus J. Telomere binding protein TRB1 is associated with promoters of translation machinery genes in vivo. PLANT MOLECULAR BIOLOGY Volume: 90 Issue: 1-2 Pages: 189-206, 2016 (IF 3.359, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 70% 50% 90% 90% [11] Schrumpfová P.P.*, Vychodilová I., Dvořáčková M., Majerská J., Dokládal L., Schořová S., Fajkus J. Telomere repeat binding proteins are functional components of Arabidopsis telomeres and interact with telomerase. PLANT JOURNAL Volume: 77 Issue: 5 Pages: 770-781, 2014 (IF 6.815, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 80% 30% 70% 50% [12] Schrumpfová P.P.*, Fojtová M., Mokroš P., Grasser K.D., Fajkus J. Role of HMGB proteins in chromatin dynamics and telomere maintenance in Arabidopsis thaliana. CURRENT PROTEIN & PEPTIDE SCIENCE Volume: 12 Issue: 2 Pages: 105-111, 2011 (IF 3,790, Q2) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 80% - 30% 10% [13] Peška V., Schrumpfová P.P., Fajkus J. Using the telobox to search for plant telomere binding proteins. CURRENT PROTEIN & PEPTIDE SCIENCE Volume: 12 Issue: 2 Pages: 75-83, 2011 (IF 3,790, Q2) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) - - 20% - 7 [14] Hofr C., Šultesová P., Zimmermann M., Mozgová I., Schrumpfová P.P., Wimmerová M., Fajkus J. SingleMyb-histone proteins from Arabidopsis thaliana: a quantitative study of telomere-binding specificity and kinetics. BIOCHEMICAL JOURNAL Volume: 419 Pages: 221-228, 2009 (IF 3.857, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 10% - - [15] Mozgová I., Schrumpfová P.P., Hofr C., Fajkus J. Functional characterization of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. PHYTOCHEMISTRY Volume: 69 Issue: 9 Pages: 1814-1819, 2008 (IF 3.186, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 5% 50% 10% 30% [16] Schrumpfová P.P., Kuchař M., Paleček J., Fajkus J. Mapping of interaction domains of putative telomerebinding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. FEBS LETTERS Volume: 582 Issue: 10 Pages: 1400-1406, 2008 (IF 3.264, Q2) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 70% - 10% [17] Růčková E., Friml J., Schrumpfová P.P., Fajkus J. Role of alternative telomere lengthening unmasked in telomerase knock-out mutant plants. PLANT MOLECULAR BIOLOGY Volume: 66 Issue: 6 Pages: 637-646, 2008 (IF 3.543, Q1) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 20% 40% 10% [18] Schrumpfová P., Kuchař M., Miková G., Skříšovská L., Kubičárová T., Fajkus J. Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. GENOME Volume: 47 Issue: 2 Pages: 316-324, 2004 (IF 2.100, Q2) Experimental work (%) Supervision (%) Manuscript (%) Research direction (%) 80% - 20% 10% 8 Telomeric repeats and their interactome Repetitive G-rich nucleotide sequences have been detected at the ends of the chromosomes of most living organisms, and hence named telomeric DNA repeats – from the ancient Greek télos 'end' and méros 'part' (Blackburn & Gall, 1978; reviewed in Jenner et al., 2022). Subsequently, it has become clear that telomeric motifs are also present within chromosomes. These internally localized telomeric repeats can be distinguished into two groups: short telomeric DNA repeats called telo-boxes and long telomeric DNA tracts, called interstitial telomeric sequences (ITSs) (Tremousaygue et al., 1999; Uchida et al., 2002). However, the defining of these two groups is not entirely precise, it is assumed that telo-boxes are composed of one to two telomeric DNA units and ITSs contain from several units to hundreds or thousands of telomeric DNA repeats. Moreover, ITSs are not composed of absolutely pure tracts of telomeric DNA repeats, but they are generally composed of telomeric repeats interspaced with degenerate repeats. The telomeric repeats, either located at the ends of or within the chromosomes, act as binding targets for large number of proteins. Some of these proteins recognize telomeric repeats specifically, while some of the telomeric-sequence associated proteins show higher sequence variability. Despite the initial assumptions that telomere-binding proteins are exclusively localized at the terminal telomeric tracts (Palm & de Lange, 2008), nowadays it is clearer that functions of telomeric-sequence associated proteins, including enzyme elongating telomeres telomerase, is more complex and these proteins possess a broad spectrum of activities. 1 Telomeric repeats at the physical ends of linear chromosomes - telomeres Telomeres are nucleoprotein structures forming and protecting the ends of linear chromosomes. They serve at least three functions which are essential for cell viability. First, they protect chromosome physical ends from fusion, endogenous nucleases and erroneous recognition as unrepaired chromosomal breaks. Secondly, telomeres facilitate the complete replication of the physical ends of the DNA. Finally, telomeres are implicated in intranuclear chromosome localization and meiotic chromosome pairing (reviewed in Blackburn et al., 2015; Procházková Schrumpfová et al., 2019; see Suppl. M; Shay & Wright, 2019; Schmidt & Cech, 2015; Schrumpfová & Fajkus, 2020; see Suppl. O). For its potential significance to aging, cancer and cell viability serve telomeres, telomerase and telomereassociated proteins as a subject of intensive research. Barbara McClintock was the first to recognized that induced chromosome ends were distinctly different from natural ends and Hermann Müller, based in part on some of the findings of McClintock, called the ends of linear chromosomes "telomeres" (Müller, 1938; McClintock, 1942). However, the intensive research on the telomeres was started only three decades ago with a description of telomere DNA component (Blackburn & Gall, 1978), detection of telomerase 9 (ribonucleoprotein with a reverse transcriptase function) (Greider & Blackburn, 1985, 1989) and uncovering telomere binding proteins (Bianchi et al., 1997; Broccoli et al., 1997). Figure 1. The replicating DNA in eukaryotes: DNA polymerases involved in replication (adopted from Schrumpfova et al., 2020; see Suppl. O). During semiconservative DNA replication, each strand serves as a template for DNA polymerases to synthesize a new complementary strand. A specialized RNA polymerase (primase), that is a part of DNA Pol α, synthesizes the RNA primer. A single RNA primer aids DNA replication on the leading strand and multiple primers initiate Okazaki fragment synthesis on the lagging strand. Further DNA synthesis is carried out by DNA Pol ε and DNA Pol δ. The newly replicated telomere resulting from the lagging strand synthesis (Lagging telomere) retains the terminal RNA primer, which is subsequently removed. Attachment of the last RNA primer more proximally on the DNA strand, together with RNA-primer removal, creates an overhang on the G-rich strand. The initial product of the leading strand DNA synthesis (Leading telomere) is a blunt terminus whose C-strand is then resected by an exonuclease to create the mature G-rich overhang. In cells with an active RNA-dependent DNA polymerase (Telomerase), the G-rich overhangs originating from Lagging or Leading telomeres, can undergo elongation. Telomerase carries its own RNA molecule, which is used as a template, and can anneal through the first few nucleotides of its template region to the distal-most nucleotides of the G-rich overhang of the telomere DNA, add a new telomere repeat (GGTTAG) sequence, translocate and then repeat the process. The complementary C-strand is then in-filled by DNA Pol α-primase. 10 Telomeres cannot be fully replicated by enzymes that duplicate DNA. Conventional DNA polymerases cannot fully replicate telomeres all the way to the end of a chromosome. The synthesis of Okazaki fragments on the lagging strand requires RNA primers attaching ahead, resulting in shortening of the chromosome's end with each duplication (Olovnikov, 1973). Moreover, the product of the leading strand DNA synthesis is a blunt terminus whose C-strand is then resected by an exonuclease to create the mature G-rich overhang (see Figure 1). Telomeric DNA in most organisms consists of tandem arrays of a short repetitive sequence. Two strands are recognized: one strand of the telomeric repeat tract running towards the 3′ end that is rich in guanines, called G-strand, whereas the complementary strand rich in cytosines is called C-strand (Makarov et al., 1997). The telomere in most of the species terminates in a 3′ single-stranded G-rich DNA overhang. In human telomeres a G-overhang is prevalent whose length varies from several tens to 280 base pairs (bp) (Cimino-Reale et al., 2001; Makarov et al., 1997; Wright et al., 1997). Telomeric sequence is highly conserved in Unikonta, where telomere motif is composed from (TTAGGG)n (Moyzis et al., 1988). This motif is the predominant terminal repeat sequence for fungi, animals, and Amoebozoa (Fulnecková et al., 2013) and sometimes is referred to as the vertebrate telomeric sequence. However even among Unikonts, the DNA sequence at chromosome ends is not completely uniform, e.g. there was detected presence of (TTAGGC)n in Nematoda, (TGTGGG)n in Rotifera or (TTAGG)n in insect Coleoptera (Frydrychová & Marec, 2002; Mason et al., 2016; Müller et al., 1991). While most filamentous fungi use (TTAGGG)n at their chromosome ends, yeasts telomeric sequence is not regular and can be described as T(G1–3) (reviewed in Kupiec, 2014; Peska et al., 2020; Tomáška et al., 2018). Moreover in addition to 3′ G-overhangs, Caenorhabditis elegans possess telomeric 5′ C-overhangs (Oganesian & Karlseder, 2011; Raices et al., 2008). Recently, in Hymenoptera (Insecta) various sequences (e.g. TTAGGTTGGG, TTAGG, TTTAGGTTAGG) were identified in terminal regions of assembled genomes (Fajkus et al., 2023). In land plants, the telomere is mostly composed of Arabidopsis-type (TTTAGGG)n repeats (Richards & Ausubel, 1988; reviewed in Schrumpfová et al., 2016a; see Supp. J; Schrumpfova et al., 2020; see Suppl. O) (see Figure 2). Several groups of flowering plants are known in which a replacement of the plant telomere sequence has occurred. Known exceptions are species in the order Asparagales, starting from divergence of the Iridaceae family. Iridaceae family shares the human-type telomeric repeat (TTAGGG)n, probably caused by a mutation that altered the telomerase RNA subunit of telomerase ∼80 Mya (Adams et al., 2001; Sýkorová et al., 2003). The human-type telomere is also shared by species of the Allioideae subfamily (Sýkorová et al., 2006), except for the Allium genus where unusual telomeric sequence (CTCGGTTATGGG)n was detected (Fajkus et al., 2016). An unusual telomeric motif (TTTTTTAGGG)n was also found in the closely related genera Cestrum elegans (family Solanaceae) (Peška et al., 2015). Moreover, outside of land plants in red and green algae and glaucophytes (Koonin, 2010), telomere types also vary. 11 For example, in algae, not only the Arabidopsis-type of telomeric repeat, but also human-type (TTAGGG)n, the Chlamydomonas-type (TTTTAGGG)n and a (TTTTAGG)n repeat were described (Fulnecková et al., 2013). Even within one plant carnivorous genus Genlisea, the telomeric sequence can vary from the Arabidopsistype telomere repeat present in G. nigrocaulis to two variant sequences (TTCAGG)n and (TTTCAGG)n in G. hispidula and its close relative G. subglabra (Tran et al., 2015) (reviewed in Schrumpfová et al., 2016a; see Supp. J; Schrumpfova et al., 2020; see Suppl. O). In plants, as well as in most of other species, replication of chromosomal ends results in G-overhangs after degradation of the last RNA primer at the 5’ terminus of a nascent strand. In A. thaliana or Silene latifolia, relatively short (20–30 nucleotides (nt)) G-overhangs were detected. Moreover, half of the Arabidopsis and Silene telomeres showed no overhangs or overhangs less than 12 nt in length (Kazda et al., 2012; Riha et al., 2000). A substantial portion of telomeres in Arabidopsis does not apparently undergo nucleolytic resection. Říha et al. showed that A. thaliana contain blunt-ended and short (1- to 3nucleotide) G-overhang-containing telomeres (Riha et al., 2000). Figure 2. Telomeres in the evolutionary tree (adopted from Schrumpfova et al., 2020; see Suppl. O). A simplified phylogenetic tree is shown, where telomeres and telomerase evolved upon linearization of chromosomes by the insertion of Group II self-splicing introns. In the Eukaryote branch, the groupings correspond to the current ‘supergroups’ according to the recent eukaryotic Tree of Life (eToL). Unresolved branching orders among lineages are shown as multifurcations. Broken lines reflect lesser uncertainties about the monophyly of certain groups. Examples of known telomeric repeat variants are listed next to respective supergroups. The major known telomeric repeat variants in the supergroups are marked with a larger font. Last eukaryote common ancestor (LECA); last universal common ancestor (LUCA). 12 Human telomere length at birth is highly heterogeneous, ranging from roughly 5-15 kilobase pairs (kb) (Sanders & Newman, 2013). The length of the germline human telomeres varies from 15–20 kb and laboratory mice have 25–150 kb telomeres (Sanders & Newman, 2013; Shay & Wright, 2000). It is obvious that length of the telomeres can vary not only between the species but also in between one genus (Gomes et al., 2011). Estimates of telomeric bp loss vary between 30–200 bp per division (Lansdorp et al., 1996). The length of plant telomeric DNA at a single chromosomal arm can be as small as 500 bp in Physcomitrella patens (Fojtová et al., 2015; Shakirov et al., 2010) as long as 160 kb in Nicotiana tabacum (Fajkus et al., 1995) or 200 kb in N. sylvestris (Kovařik et al., 1996). Besides the remarkable variation in telomere length among diverse plant genera or orders, telomere lengths can also vary at the level of the species or ecotypes, e.g. Arabidopsis telomeres range from 1.5 to 9 kb, depending on the ecotype (Maillet et al., 2006; Shakirov & Shippen, 2004); telomeres from inbred lines of maize range from 1.8 to 40 kb (Burr et al., 1992). Additionally, in the long-living organism Betula pendula, telomeres in different genotypes varied from a minimum length of 5.9 - 9.6 kb to a maximum length of 15.3 - 22.8 kb (Aronen & Ryynänen, 2014) reviewed in Schrumpfová et al., 2016a; see Supp. J). 1.1 Proteins associated with telomeres in mammals Telomere-associated proteins can regulate the length of the telomere tract by modulating access of telomerase or affecting conventional DNA replication machinery. In mammals, telomeric DNA is maintained primarily by six-protein complex called Shelterin: Telomere Repeat Binding Factors 1 and 2 (TRF1, TRF2), Protection of telomeres 1 (POT1), TRF1- and TRF2-Interacting Nuclear Protein 2 (hTIN2), telomere protection protein 1 (TPP1) and Repressor/activator protein 1 (RAP1). Moreover, the physical ends of the chromosomes are associated with a nucleoprotein complex named CST (Cdc13/CTC1, STN1, TEN1). The specific telomeric dsDNA binding of Shelterin complex is mediated by TRF1 and TRF2 proteins (Broccoli et al., 1997) through their Myb-like domain, with an LKDKWRT amino acid motif. The Myb-like domain is conserved in telomeric sequence binding proteins not only in mammals but also in plants or yeasts (Bilaud et al., 1996; Feldbrügge et al., 1997). Myb-like domains of the TRF1 and TRF2 proteins are located at their C-terminus. Another Shelterin subunit - POT1 protein - is linked to the Shelterin complex by TPP1 protein, which in turn binds to TIN2 and RAP1 proteins and interacts with TRF2 protein (reviewed in Schmidt & Cech, 2015) (see Figure 3A). 13 TRF2 protein has a central protective role in shelterin complex because it specifically inhibits Ataxia Telangiectasia Mutated (ATM) kinase-dependent DNA damage signalling and the classical Ku70/80- and Ligase IV-mediated non-homologous enjoining pathway (NHEJ) at telomeres. The POT1 proteins (POT1a and POT1b in the mouse) associate with single-stranded (ss) telomeric DNA and POT1 protein safeguards Figure 3. An integrative schematic view of the human and plant terminal telomeric complex (adopted from Schrumpfová et al. 2019; see Suppl. M). A) Human active telomerase is associated with chaperones as well as with TR associated conserved scaffold box H/ACA of small nucleolar RNAs proteins. Mammalian shelterin proteins (TRF1/2, RAP1, TIN2, TPP1 and POT1) modulate access to the telomerase complex and the ATR/ATM-dependent DNA damage response pathway. The CST complex (CTC1-STN1-TEN1) affects telomerase and DNA polymerase α recruitment to the chromosomal termini and, thus, coordinates G-overhang extension by telomerase with fill-in synthesis of the complementary C-strand (blue dashed line). G-quadruplexes, D-loops and T-loops during telomere replication are resolved by RTEL helicase. HOT1 directly binds double strand telomere repeats and associates with the active telomerase. Telomere nucleosomes show a shorter periodicity than that in the other parts of chromosomes. B) Arabidopsis telomerase is associated with TRB proteins as well as with POT1a that interacts with the dyskerin orthologue CBF5. Plants possess all orthologue proteins of conserved scaffold box H/ACA of small nucleolar RNAs (CBF5, GAR1, NOP10, NHP2). Moreover, TRB proteins interact with the telomeric sequence due to the same Myb-like binding domain as that in mammalian TRF1/2. TRB proteins interact with TERT and POT1b and, when localized at chromosomal ends, they are eligible to function as components of the plant shelterin complex. An evolutionarily conserved CST complex is suggested to coordinate the unique requirements for efficient replication of telomeric DNA in plants as well as in other organisms. In addition, plant RTEL contributes to telomere homeostasis. For the sake of clarity, only the situation in telomere with 3' overhang is depicted. For further information and for human and plant telomere histone modifications see Schrumpfová et al. (2019; see Suppl. M). A B Telomerase Telomerase 14 telomeres against Ataxia Telangiectasia and Rad3 related (ATR) kinase pathway (Denchi & de Lange, 2007; Smogorzewska et al., 2002). A bridge between proteins directly associated with DNA-TRF1, TRF2 and POT1 is mediated by TIN2 and the oligosaccharide/oligonucleotide binding (OB)-fold domain of TPP1 protein (reviewed in (Schmidt & Cech, 2015). Moreover protein RAP1, interacts with TRF2 (Arat & Griffith, 2012) and modulates its recruitment to telomeric DNA (Janoušková et al., 2015; Nečasová et al., 2017). The most of the Shelterin complexes can be purified without dissociation, indicating they form stable complexes at least in vitro. It was published that TRF2 interacts with TIN2 with an 2:1 stoichiometry in the context of Shelterin (RAP12:TRF22:TIN21:TPP11:POT11) (Lim et al., 2017). The maintenance of telomere repeats in most eukaryotic organisms requires enzyme telomerase. Telomerase consists of a telomerase reverse transcriptase (TERT) and telomerase RNA subunits (TR) that dictates the synthesis of the G-rich strand of telomere terminal repeats and elongates telomeric tracts at the chromosomal terminus (Blackburn & Gall, 1978; Greider & Blackburn, 1985, 1989). Most enzymes encounter their substrates by simple diffusion but both telomerase and its chromosome end substrate have very low abundance and the telomerase enzyme is recruited to telomeres rather than simply encountering them by diffusion (Schmidt & Cech, 2015; Xi & Cech, 2014). The Shelterin component TPP1 is the key telomeric component necessary for telomerase recruitment to telomeres (Xin et al., 2007). In addition, TPP1 in complex with POT1 stimulates telomerase to synthesize additional telomeric repeats in vitro and has therefore been proposed to be a processivity factor for telomerase action at telomeres (Wang et al., 2007). Protein TPP1 is composed of an N-terminal OB-fold domain required for telomerase recruitment, a central domain that directly binds to POT1 and a C-terminal domain necessary for its association with TIN2. Loss of any of the members of Shelterin protein complex can result in inappropriate DNA damage response (DDR), can lead to chromosome fusion, telomere loss or activate replicative senescence or apoptosis (Sfeir, 2012). Kappei et al. identified by the proteomics of isolated chromatin segments (PICh) approach a telomeric DNA binding protein named homeobox telomere-binding protein 1 (HOT1). HOT1 directly binds ds telomere repeats and associates with the active telomerase and is required for telomerase chromatin binding. HOT1 is the telomere-binding protein that acts as a positive regulator of telomere length (Déjardin & Kingston, 2009; Kappei et al., 2013). 15 1.2 Proteins associated with telomeres in plants 1.2.1 Telomeric dsDNA associated proteins in plants In plants, telomeric dsDNA sequence binding proteins with a Myb-like domain can be classified into three main groups: (i) with a Myb-like domain at the N-terminus (Smh/TRB family) (ii) with a Myb-like domain at the C-terminus (TRFL family) (iii) with a Myb-like domain at the C-terminus (AID family) (reviewed in Du et al., 2013; Peška et al., 2011; Schrumpfová et al., 2016a; see Supp. F and J). The first group of proteins, with a Myb-like domain at the N-terminus, contain a central histone-like domain (H1/5 domain) with homology to the H1 globular domain found in the linker histones H1/H5 and is therefore called the Smh (Single myb histone) family (Marian et al., 2003; Marian & Bass, 2005; Schrumpfová et al., 2004; see Supp. A). Members belonging to Smh family are frequently named Telomere Repeat Binding (TRB) proteins so we call this family also Smh/TRB family (see Figure 3B and Figure 5). Within the second family of the proteins with a Myb-like domain - TRFL family - there were also identified several plant orthologues. In A. thaliana there were characterised six proteins with the C-terminal Myb-like domain (AtTBP1, AtTRP1 and AtTRFL1, 2, 4, 9) belonging to the subfamily of proteins named TRFL I with characteristic features. Proteins from TRFL I family can homo- and heterodimerize and they can efficiently bind to telomeric DNA in vitro (Karamysheva et al., 2004). A key feature of this subfamily is a ∼30 amino acid extension of the Myb-like domain that is likely responsible for specific binding to plant telomeric DNA. Moreover, the TRFL family includes six proteins, that are unable to bind telomeric DNA in vitro and are also unable to form homo- and heterodimers, despite possessing the C-terminal Myb-like domain. These proteins are members of subfamily named TRFL II (AtTRFL3, 5, 6, 7, 8, 10) (Karamysheva et al., 2004). The proteins from the third family contain a single Myb-like domain at the C-terminus and contains only a few described members. All three Myb-like protein subfamilies were already detected in the moss P. patens and separation of Smh/TRB and TRFL and is apparent already in unicellular algae. These data suggest ancient origin of the three protein subfamilies and their diversification early in evolution of the plant lineage (Fulcher & Riha, 2016). Especially the first (Smh/TRB) family and the second (TRFL) family contain increased number of family members. However, this observation is not surprising as whole genome duplication events (WGDs) have occurred in many plant families (Freeling, 2009; Qiao et al., 2022). These WGDs result in a multitude of genomic changes, such as deletions of large fragments of chromosomes, silencing of duplicate genes and recombining of homologous chromosomal segments, as was shown, e.g. in crucifer species (Freeling, 2009; Mandáková & Lysak, 2008). Increased numbers of genes of the same family may lead to gene sub- 16 functionalization, neo-functionalization and partial or full redundancy, and complicates assignment of an actual and specific function for individual proteins in vivo. Overall, the conserved domain composition of the plant proteins with respect to their mammalian counterparts does not guarantee conservation of their function. It seems that some proteins are involved in a similar biochemical pathways, but their interaction partners, and consequently potential regulatory factors, might slightly differ (reviewed in Schrumpfová et al., 2019; see Supp. M). Smh/TRB family Screening of Z. mays cDNA led to identification of gene coding ZmSmh1 protein (Marian et al., 2003). The Smh1 gene is expressed in leaf tissue and the ZmSmh1 protein binds ds oligonucleotide probes with at least two internal tandem copies of the maize telomere repeat, TTTAGGG. Simultaneously as Smh protein from Zea was characterized, we searched A. thaliana databases in our laboratory for putative genes coding for proteins with the Myb-like domain. This search resulted in two candidate protein sequences at that time, AtTRB2 and AtTRB3, formerly named AtTBP3 and AtTBP2, respectively (Kuchař & Fajkus, 2004; Schrumpfová et al., 2004; see Supp. A). We characterised these two candidates and we found out that AtTRB2 and AtTRB3 proteins able to bind the G-rich strand and dsDNA of plant telomeric sequence with an affinity proportional to a number of telomeric repeats. The binding of AtTRB proteins to telomeric ds telomeric oligonucleotides is highly specific, because even a 100-fold abundance of non-telomeric sequence cannot displace their binding to tetramers of the telomeric sequence (Schrumpfová et al., 2004; see Supp. A). Binding affinity to ds- and ss-oligonucleotides of the plant telomeric sequence is roughly proportional to the number of telomeric repeat. Additionally, the later identified member of Smh/TRB family - AtTRB1 protein - is able to bind related telomeric DNA sequences (plant (TTTAGGG) or human (TTAGGG)) with a certain flexibility, as well as AtTRB2 or AtTRB3 proteins. We analysed DNA-protein interaction of the full-length and truncated variants of AtTRB1. We showed that preferential interaction of AtTRB1 with ds telomeric DNA is mediated by the Myblike domain while the H1/5 domain interacts non-specifically with any DNA without preference for either telomeric or non-telomeric sequence (Ellen & van Holde, 2004; Mozgová et al., 2008; see Supp. D). The partial non-selective binding of the Myb-like domain to either plant (TTTAGGG) or human (TTAGGG) telomeric sequence appears to be a general feature of the A. thaliana Smh/TRB family proteins (Mozgová et al., 2008; Schrumpfová et al., 2004; see Supp. D and A). Recently, we have completed characterization of the TRB family as we described two novel members of the TRB family from Arabidopsis (AtTRB4 and AtTRB5) (see Figure 4). The results clearly showed that AtTRB4 and AtTRB5 do preferentially bind long arrays of telomeric sequences. However, the AtTRB minimal 17 recognition motif was newly defined as one telo-box positioned within a non-telomeric DNA sequence (Kusova et al., 2023; see Suppl. R). Figure 4. Sequence and structural alignments of TRB family proteins (Kusová et al., 2023; see Supp. R). A) Schematic representation of the conserved domains of TRBs from A. thaliana. Myb-like, Myb-like domain; H1/5-like, histone-like domain; coiled-coil, C-terminal domain. B) Unrooted Maximum likelihood (ML) phylogenetic tree of Brassicaceae TRB proteins. The length of the branches are proportional, and the black dots indicate the position of TRB1-5 from A. thaliana. C) Multiple alignments of the Myb-like, H1/5-like and coiled-coil domains. The positions of α-helices or β-sheets of the uppermost or the lowermost sequence in each alignment are highlighted: bold, experimentally determined structures (cryo-EM or X-ray crystallography); thin, AlphaFold prediction. Human Telomeric repeat-binding factor 2 (hTRF2) and Xenopus laevis histone H1.0 (Xl H1.0-B) were used to show the most conserved amino acid (aa) residues. Amino acid shading indicates the following conserved amino acids: dark green, hydrophobic and aromatic; light green, polar; blue, basic; magenta, acidic; yellow, without side chain (glycine and proline). The aa of hTRF2 that mediate intermolecular contacts between telomeric DNA and hTRF2 are marked with an asterisk. D) A certain flexibility in binding related telomeric DNA sequences was observed from the A. thaliana telomeric DNA: A. thaliana (TTTAGGG), Chlamydomonas reinhardtii (TTTTAGGG), human (TTAGGG), Bombyx mori [TTAGG]5TTAG and Ascaris lumbricoides (TTAGGC) (Niedermaier and Moritz, 2000; Okazaki et al., 1993; Petracek and Berman, 1992), however, the ability to bind variant telomere sequences decreased with sequence divergence. In addition to being able to bind ds telomeric sequences, AtTRB proteins can also bind to the G-rich ss telomeric DNA although with lower affinity compared to ds telomeric sequences. The C-rich telomeric strand is not preferentially bound (Schrumpfová et al., 2004; see Supp. A). Our observation are consistent with the findings that also other members of Smh/TRB family proteins in land plants show telomeric dsDNA binding capability, e.g. Z. mays ZmSMHs or O. sativa OsTRBFs (Byun et al., 2018; Marian et al., 2003). 18 We performed comprehensive phylogenetic analysis and found out that TRB proteins first evolved in Streptophyta in Klebsormidiophyceae. In Klebsormidium nites only one TRB homolog was identified. Following the evolutionary tree, an increasing number of TRB homologues were found in Bryophyta and Tracheophyta. In seed plants, which have undergone more rounds of WGDs than Bryophyta and Lycophyta (Clark & Donoghue, 2018), predominantly three TRB proteins were recognized. Within Brassicaceae, which has undergone an additional recent round of WGD (Walden et al., 2020), five TRB homologs were revealed (Kusová et al., 2023; see Supp. R). The ability of AtTRB proteins to bind typical plant and human telomeric motifs with a similar affinity could be important for an easier adaptation to a change in telomere sequence from a plant to divergent telomeric motifs, which has occurred during the evolution of plants of several species as was described above. It is also consistent with the Kováč et al. that argued that there is an upper limit for the specificity of interaction between binding partners (e.g. enzyme-substrate, ligand-receptor, protein-DNA sequence), since interactions that are too specific would lack flexibility and a perfect recognition would be too rigid and possibly non-functional (Kováč, 1987), e.g. Tay1 (telomere-associated in Yarrowia lipolytica 1) protein, the double strand (ds) sequence telomere-binding protein of the yeast Y. lipolytica, exhibits lower affinity for its own telomeres (TTAGTCAGGG) than for the mammalian-type telomeric repeats (TTAGGG) (reviewed in Tomáška et al., 2018). Another typical character of a telomere dsDNA-binding proteins seems to be capability for multimerization. Dimerization has been proved to increase the efficiency of the binding of telomere-associated proteins, TRF1 and TRF2, to telomeric DNA also in mammalian cells. Mammalian TRF1 is a homodimer in vivo and its accumulation at telomeres depends on homotypic interactions. Similarly, the TRF homology (TRFH) domain near their N-terminus from TRF2 protein mediates homotypic interactions, but TRF1 and TRF2 do not form heterodimers (Bianchi et al., 1997; Fairall et al., 2001). In our studies we demonstrated that AtTRBs show strong mutual and self-interactions using yeast two hydrid assay (Y2H) assay (Schrumpfová et al., 2004; see Supp. A; Kusova et al., 2023; see Suppl. R). Additionally, we investigated the ability of the AtTRB1 fragments to form self-dimers or multimers using Poly Acrylamide Gel Electrophoresis (PAGE) with a weak detergent, perfluoro-octanoic acid (PFO) (Mozgová et al., 2008; see Supp. D). This method can be used for detection and molecular mass determination of protein complexes since (in contrast to SDS-PAGE with sodium dodecyl sulfate), PFO-PAGE preserves high affinity protein–protein interactions (Ramjeesingh et al., 1999). The results confirmed the strong tendency of the H1/5 domain to multimerize and the same holds true for all the fragments of AtTRB1 which contain the H1/5 domain. Myb-like domain of the rice RTBP, TRFL family protein, also interacts with plant telomeric DNA in the form of a homodimer (Yu et al., 2000). In contrast to H1/5 domain, the N-terminal Myb-like 19 domain itself did not form higher molecular weight complexes. According to these results we proposed model of binding of AtTRB proteins to plant telomeric DNA where the Myb-like domain primarily ensures direct sequence-specific binding of AtTRB1 to telomeric DNA, while the H1/5 domain may enhance this binding by protein dimerization and sequence-non-specific binding to DNA (Mozgová et al., 2008; see Supp. D). The study of stoichiometry and kinetics of AtTRB1 and AtTRB3 proteins binding to the telomeric DNA revealed that the affinity of AtTRB1 to telomeric substrate with four telomeric repeats is 4-fold higher than that of AtTRB3, although AtTRB1 and AtTRB3 are relatively similar in their primary sequences (Hofr et al., 2009; see Supp. E). Similarly to our results, human hTRF1 binds telomeric DNA with a 4-fold higher affinity than that of hTRF2 when interacting with human telomeric DNA (Hanaoka et al., 2005). In Mozgova et al. (2008; see Supp. D) we assumed that the non-specific interaction of H1/5 domain with any DNA without preference for either telomeric or non-telomeric sequence (Ellen & van Holde, 2004), together with the high pI of the AtTRB1 fragments, suggests that electrostatic interactions take part in the interaction of the fragments of AtTRB1 with telomeric dsDNA (Mozgová et al., 2008; see Supp. D). Interestingly, our model showing model of Myb-like domain revealed that the solution accessible surface of AtTRB4 and AtTRB5 differ to the solution accessible surface of AtTRB2/AtTRB3 and to AtTRB1 (Kusova et al., 2023; see Suppl. R). All five AtTRB members preferentially localize to the nucleus and nucleolus during interphase. Both the central H1/H5‐like domain and the Myb-like domain from AtTRB1 can direct a GFP fusion protein to the nucleus and nucleolus (Dvořáčková et al., 2010a; Kusova et al., 2023; see Suppl. R). AtTRB1-GFP localization is cell cycle‐regulated, as the level of nuclear‐associated GFP diminishes during mitotic entry and GFP progressively re‐associates with chromatin during anaphase/telophase. Although a possible association of AtTRB1-GFP with the telomere was suggested previously in Dvořáčková et al. (2010a) the small size of Arabidopsis chromosomes, in combination with short telomere lengths, precluded the authors to visualize the AtTRB-telomere association. We took advantage of the well-established protocol of Nicotiana benthamiana leaf infiltration and the fact that N. benthamiana has longer telomeres that are easier to visualize compared to Arabidopsis. In our study Schrumpfová et al. (2014; see Supp. H) we proved that AtTRB proteins are not only binding to the telomeric DNA sequence in vitro, as was described above, but that they also co-localize with telomeres in situ. Later on, localization of AtTRB1 protein at the plant telomeres in vivo was verified by independent technique by the teams of Holger Puchta and Andreas Houben. They used imaging technique based on two orthologues of the bacterial clustered regularly interspaced short palindromic repeats (CRISPR)–CRISPR associated protein 9 (Cas9). Dreissig et al. demonstrated not only that CRISPR–dCas9 can be used to 20 visualize specific DNA sequences in combination with fluorescently tagged proteins interacting with those DNA sequences but they also demonstrated that around 87.6 % of telomeres were simultaneously bound by AtTRB1 protein and CRISPR–dCas9 signals resembling telomeres (Dreissig et al., 2017). Telomere shortening was observed in attrb1 mutants in the A. thaliana ecotype Columbia, with otherwisestable telomere lengths (Shakirov & Shippen, 2004; Schrumpfová et al., 2014; see supp. H). In contrast, telomere extension was detected in attrb2 knockout mutants of the A. thaliana ecotype Wassilewskija, which exhibits telomere length polymorphism in wild-type plants (W. K. Lee & Cho, 2016; Maillet et al., 2006; Shakirov & Shippen, 2004). Triple homozygous mutant plants, containing the alleles from A. thaliana Columbia (attrb1 and attrb3) and from Wassilewskija (attrb2), exhibit telomere shortening (Zhou et al., 2016, 2018). Our suggestion, that AtTRBs are part of telomere-associated interactome was supported by the group of Simon Amiard and Charles White that used pull-down assays to identify potential telomeric interactors in the Arabidopsis. They identified several candidate proteins, including TRB1 and TRB3 proteins. The TRB proteins were enriched in pull-down with telomeric probe even more than the GH1-HMGA1 proteins that are the main objects their study (Charbonnel et al., 2018). Involvement of AtTRB proteins in telomere interactome was furthermore boosted by our detection of direct interaction between AtTERT and AtTRB proteins (Kusova et al., 2023; see Suppl. R). AtTRB proteins interact in Y2H, Co-Immunoprecipitation (Co-IP) or Bimolecular Fluorescence Complementation (BiFC) systems with the N-terminal part of AtTERT that contains telomerase-specific motifs. Moreover, AtTRB1 was, among the others, co-purified with N-terminal constructs of AtTERT from A. thaliana suspension cultures (Schrumpfová and Majerská et al., 2017, 2018; see Supp. K and L). However, neither of the AtTRB2 and AtTRB3 proteins purified from Escherichia coli, nor their mixture, had any effect on telomerase activity in vitro, measured by Telomerase Repeat Amplification Protocol (TRAP) (Schrumpfová et al., 2004; see Supp. A). Likewise, no changes in telomerase activity or processivity were observed in extracts from attrb1 mutant plants. Correspondingly, no variations in telomerase activity were detected in transformed plants (TRB1pro:TRB1-GFP) expressing higher levels of protein (Schrumpfová et al., 2014; see Supp. H). Kuchař et al. (2004) detected that AtTRB2 and AtTRB3 proteins interact with AtPOT1b protein, one of two homologues of human telomeric ssDNA binding protein POT1 (Baumann & Cech, 2001; Kuchař & Fajkus, 2004). In our study Schrumpfová et al., (2008; see Supp. C) we found out that also other member of Smh/TRB family - AtTRB1 protein - interacts with AtPOT1b. Using combination of Y2H and Co-IP we detected that AtTRB1 protein physically interacts with N-terminus AtPOT1b via its H1/5 domain (Schrumpfová et al., 2008; see Supp. C). Recently we detected also interaction between AtTRB4 and AtTRB5 and AtPOT1a (Kusova et al., 2023; see Suppl. R). 21 Moreover, proteins from Smh/TRB family physically interact with AtRUVBLs. RUVBL proteins belong to the evolutionarily highly conserved AAA+-family (ATPase Associated with various cellular Activities) that are involved in ATP binding and hydrolysis (Matias et al., 2006). AtTRBs together with AtTERT and AtRUVBLs form trimeric complex AtTERT-AtTRB-AtRUVBL (see also bellow) (Schořová et al., 2019; see Supp. N). Our results suggested that AtTRB proteins thus play a role of interaction hubs not only in telomere chromatin structure but also in telomerase biogenesis. Figure 5. Telomeric and putative telomeric dsDNA- and ssDNA-binding proteins from A. thaliana (adopted from Schrumpfová et al., 2016a, see Supp. J). Myb-like domain (Myb); Myb-extension (-ext); Histone-like domain (H1/5); Coiled Coil Domain (CCD); Oiigonucleotide/Oligosaccharide-Binding Fold domain (OB); Whirly domain (Whirly); RNA-binding domain (RB); A. thaliana (At); Telomere Repeat Binding Protein (AtTRB); TRF-like family (TRFL family); Suppressor of cdc thirteen homolog (AtStn1); Conserved telomere maintenance component 1 (AtCTC1); (CTC1-Stn1-Ten1) complex (CST); RNA recognition motifs (RRM); Protection of telomeres 1a, b, c (AtPot1 a,b,c); Whirly 1 (Why1); Single-stranded telomere-binding protein 1 (STEP1). 22 In O. sativa there were three proteins from Smh/TRB family identified (Byun et al., 2008). Proteins OsTRBF1 and OsTRBF2 are constitutively transcribed in rice plants grown under greenhouse conditions. Gel retardation assays showed that these OsTRBF proteins bind specifically to the plant double-stranded telomeric sequence, TTTAGGG, with markedly different binding affinities. Y2H and Co-IP assays indicated that both OsTRBF1 and OsTRBF2 interact with one another to form homo- and hetero-complexes, while OsTRBF3 appeared to act as a monomer (Byun et al., 2008). In an affinity pull-down technique, 80 proteins from O. sativa were identified for their ability to bind to a telomeric repeat (He et al., 2013). Among them, two of three previously reported proteins from the Smh/TRB family - OsTRBF1 and OsTRBF2 were isolated. TRFL family The second group of proteins, with a Myb-like domain at the C-terminus, is named TRFL (TRF-like). TRFL family can be divided into two subfamilies named TRFL I (possess extension of the Myb-like domain (Mybext) that is likely responsible for specific binding to plant telomeric DNA proteins in vitro) and TRFL II (unable to bind telomeric DNA in vitro) (Karamysheva et al., 2004; Ko et al., 2008) (see Figure 5). The first identification of a TRFL family protein from O. sativa - Telomere-binding protein 1 (OsRTBP1) (Yu et al., 2000) - was soon followed by numerous other TRFL members, e.g. Nicotiana glutinosa (NgTRF1) (Yang et al., 2003), Solanum lycopersicum (LeTBP1) (Moriguchi et al., 2006), A. thaliana (AtTBP1, AtTRP1, AtTRFL2- 10) (Hwang et al., 2001; Chen et al., 2001; Karamysheva et al., 2004) or Cestrum parqui (CpTBP) (Peška et al., 2011; see Supp. F). Even though O. sativa or N. glutinosa mutants for TRFL members exhibited markedly longer telomeres (Hong et al., 2007; Yang et al., 2004), in A. thaliana, a knockout of AtTRP1, member of TRFLI subfamily with a Myb-ext, did not change telomere length significantly (Chen et al., 2005). In A. thaliana even multiple knockout plant, deficient for all six proteins from TRFLI subfamily (AtTBP1, AtTRP1, AtTRFL1, AtTRFL2, AtTRFL4 and AtTRF9) did not exhibit changes in telomere length or phenotypes associated with telomere dysfunction (Fulcher & Riha, 2016; reviewed in Schrumpfová et al., 2016a; see Supp. J). A structurally related member to TRFLI subfamily was found in Cestrum parqui, CpTBP1, a plant species lacking typical telomeres and telomerase (Peška et al., 2011; see Supp. F). The protein shows nuclear localisation and association with chromatin while transiently expressed in N. benthamiana after infiltration Agrobacterium tumefaciens into young leaves. Although, no functional evidence exists for the role of AtTRFL proteins at telomeres so far, plausible involvement in telomere maintenance in plants was suggested in Kuchař and Fajkus (Kuchař & Fajkus, 2004). Kuchař and Fajkus observed a specific interaction between AtTRP1 (member of TRFLI subfamily) and AtKu70. The AtTRP1 domain responsible for AtKu70 interaction occurs between amino acid sequence positions 80 and 269. It was hypothesized that AtKu, a DNA repair factor with a high affinity for DNA ends, 23 sequesters chromosome termini within its DNA loading channel and protects them from nuclease processing (Valuchova et al., 2017). Another member of the TRFL family - ZmIBP2 (Initiator-binding protein) protein – binds not only telomeric repeats (Moore, 2009), but was originally identified as a promoter binding ligand (Lugert & Werr, 1994). AID family The third group with a Myb-like domain at the C-terminus (AID family) contains only a few described members. The AID family is named according to anther indehiscence 1 (AID) protein from O. sativa - OsAID1 (Zhu et al., 2004). OsAID1 was initially identified as being involved in anther development, however, OsAID1 also isolated in an affinity pull-down technique within 80 proteins from O. sativa showing ability to bind to a telomeric repeat, while no member with a Myb-like domain at the C-terminus of the TRFL family could be found (He et al., 2013). Another member of this family - ZmTacs1 (Terminal acidic SANT) from Z. mays - may function in chromatin remodelling within the meristem. In silico expression analysis revealed that ZmTacs1 is expressed in meristem-enriched tissues and in contrast, the Myb-like domains of known Myb-like domain such as ZmSMH1, or human TRF1 all have basic isoelectric points (Marian & Bass, 2005; reviewed in Schrumpfová et al., 2016a; see Supp. J). Marian and Bass proposed that the acidic patches observed on the surfaces of the plant TACS-type proteins are not compatible with direct DNA binding and may reflect areas for the binding of basic moieties, such as histone tails or basic regions of other proteins (Marian & Bass, 2005). 1.2.2 Telomeric ssDNA associated proteins in plants Proteins with OB-fold The majority of telomeric ssDNA binding proteins bind through OB motifs (oligonucleotide/oligosaccharide binding, OB-fold) and are required for both chromosomal end protection and regulation of telomere length, e.g., telomere-binding protein subunit alpha/beta (TEBPαβ) from Oxytricha nova; (C. M. Price & Cech, 1987), Cell division cycle 13 (Cdc13p) from S. cerevisiae (Garvik et al., 1995) and POT1, are present in diverse organisms including human, mouse, chicken or S. pombe (Baumann & Cech, 2001; Lei et al., 2002; Wei & Price, 2004; L. Wu et al., 2006). In A. thaliana, three POT-like proteins were named AtPOT1a (previously named AtPOT1-1, AtPot1), AtPOT1b (previously named AtPOT1-2, AtPot2) and AtPOT1c (Kuchař & Fajkus, 2004; Lei et al., 2002; Rossignol et al., 2007; Shakirov et al., 2005; Tani & Murata, 2005). AtPOT1a and AtPOT1b proteins contain two OB motifs as well as mammalian POT1 proteins, but share only 49 % sequence similarity, while mouse 24 proteins share 72 % similarity. AtPOT1c protein is short version of AtPOT1a and originates by gene duplication and contain only one OB motif (Rossignol et al., 2007) (see Figure 3B and Figure 5). However, descriptions of plant POT protein functions and binding properties are not unanimously agreed. While a very weak, but specific affinity of AtPOT1a and AtPOT1b expressed in E. coli for plant telomeric ssDNA was originally described (Shakirov et al., 2005), later these authors could not demonstrate AtPOT1a and AtPOT1b binding to telomeric ssDNA in vitro (Shakirov, McKnight, et al., 2009; Shakirov, Song, et al., 2009). In our laboratory, AtPOT1a and AtPOT1b proteins were expressed in bacteria or using in vitro WG transcription/translation extract. Unfortunately, none of these systems or their modifications resulted in expression of intact AtPOT1 proteins. AtPOT1 proteins were either not expressed or due to their hydrophobicity localized mainly in the bacterial inclusion bodies or they were co-purified with chaperon GroEL (Schrumpfova, dissertation thesis). There was no proof that the AtPOT1b, purified from bacterial extract or expressed in vitro translation extract, had the ability to bind telomeric ss oligonucleotides (Schrumpfová, 2008; see Supp. C). Subsequently it was demonstrated that functional human and mouse POT1 should be isolated from baculovirus-infected insect cells (Palm et al., 2009). Nevertheless, stable telomeric ssDNA binding was observed for two full-length plant POT1 proteins: OlPOT1 from the green alga O. lucimarinus as well as for ZmPOT1b from Z. mays (Shakirov, Song, et al., 2009). Although POT1 proteins from plant species as diverse as Populus trichocarpa (poplar), Hordeum vulgare (barley), Gossypium hirsutum (cotton), Helianthus argophyllus (sunflower), S. moellendorffii (spikemoss), Pinus taeda (pine), Solanum tuberosum (potato), Asparagus officinalis (garden asparagus) and Z. mays (maize) (ZmPOT1a) failed to bind telomeric DNA when expressed in a RRL expression system in vitro and subjected to an electrophoretic mobility shift assay (EMSA) (Shakirov, Song, et al., 2009), binding of plant POT1 proteins to telomeric DNA under native conditions cannot be excluded. Plants expressing AtPOT1a truncated by an N-terminal OB-fold, showed progressive loss of telomeric DNA. These findings denote that AtPOT1a plays role in positive regulation of telomere length (Surovtseva et al., 2007). In contrast, expression of only N-terminal part of AtPOT1b leads to severe defects in plant growth and development, telomeres are shortened and there is a high formation of anaphase bridges or defective segregation of chromosome, which means that AtPOT1b plays role in protection of chromosomal ends (Shakirov et al., 2005). POT1 proteins from A. thaliana differ not only in their functions, but also have divergent interaction partners. AtPOT1a binds AtSTN1 and AtCTC1 proteins from CST complex (Renfrew et al., 2014). AtPOT1a, but not AtPOT1b, is associated with an N-terminal part of AtTERT in nucleoplasm in vitro (Rossignol et al., 2007). Among other interactors of AtPOT1a belongs CBL-interacting protein kinase (CIPK21). This kinase belongs to the large family in A. thaliana of which several members were shown to be involved in Ca2+ 25 signalling and moreover, CIPK21 is presumed to have a function in DDR signalling (Rossignol et al., 2007). These data suggest a potential role of AtPOT1a in DDR pathway as was described to many other telomeric proteins (Gallego & White, 2005). We found AtPOT1a protein among proteins that we co-purified with Nterminal domains of AtTERT using (TAP-MS) (Schrumpfová and Majerská et al., 2017, 2018; see Supp. K and L). Using BiFC it was confirmed that AtPOT1a interacts with AtCBF5 protein (Centromere-binding factor 5; a plant homologue of dyskerin) in the cytoplasmic or nucleolus foci (Kannan et al., 2008; Lermontova et al., 2007; Schořová et al., 2019; see Supp. N). Interestingly, AtPOT1a forms weak interaction with AtRUVBL1 protein. This fact corelates with our recent observation that AtPOT1a, AtTERT, AtTRB, AtCBF5 and AtRUVBL1 proteins are involved in assembly of the plant telomerase (Schořová et al., 2019; see Supp. N). Moreover, both AtPOT proteins directly interacts with AtTRB proteins (Kuchař & Fajkus, 2004; Kusova et al., 2023; see Suppl. R), nevertheless, AtPOT1b does not seem to substantially contribute to telomere maintenance (Cifuentes-Rojas et al., 2012). Using Y2H and BiFC have also recently detected novel interaction between AtTRB4-5 and AtPOT1a (Kuchař & Fajkus, 2004; Kusova et al., 2023; see Suppl. R). CST is an evolutionarily conserved trimeric protein complex that in budding yeast is composed of the proteins Cdc13, Stn1 and Ten1, whereas in mammals the CST complex consists of the proteins CTC1, STN1 and TEN1. CST complex plays role in DNA replication and telomere maintenance through its ability to interact with ssDNA. Nevertheless, it was found out that CST is neither a nonspecific nor a telomere ssDNA specific binder, and rather CST is a tight binder of ssDNA with a preference for G-rich sequences (Hom & Wuttke, 2017). In yeast, these OB-fold proteins are required for recruitment of telomerase and DNA polymerase α to the chromosomal termini and thus coordinate G-overhang extension by telomerase with the fill-in synthesis of the complementary C-strand (Giraud-Panis et al., 2010; Grossi et al., 2004; Qi & Zakian, 2000; Wellinger & Zakian, 2012). Mammalian CST is ortholog of an archaeal RPA complex and is involved in the rescue of stalled replication forks either at the telomere or elsewhere in the genome and Cstrand fill-in. However, CST in mammals is also proposed to limit telomerase action, perhaps by competing for binding to the telomere protein TPP1 (reviewed in Lue, 2018; Rice & Skordalakes, 2016; Schrumpfova et al., 2019; see Supp. M) (see Figure 3A). CST in plants is needed for telomere integrity (Leehy et al., 2013; Surovtseva et al., 2007), however, clear evidence that would show any direct physical interaction of any component of the CST complex with plant telomeric DNA is absent. It seems that the CST complex controls access of telomerase, end-joining recombination and the ATR-dependent (ATM and Rad3-related) DNA damage response pathway at the chromosomal ends in wild-type plants (see Figure 3B) (Amiard et al., 2011; Boltz et al., 2012; Derboven et al., 2014; Leehy et al., 2013; reviewed in Schrumpfova et al., 2019; see Supp. M). 26 Non-OB-fold proteins Aside predominantly characterised proteins with OB-fold domain associated with telomeric ssDNA sequence in plants, several proteins lacking the OB-fold domain were also identified, such as Whirly proteins or proteins with RNA recognition (RRM) motifs (see Figure 5). The transcriptional activator protein Whirly 1 (AtWhy1), from a small protein family found mainly in land plants (Desveaux et al., 2000, 2002; Krause et al., 2005), was also identified in a fraction of AtTERT binding proteins in A. thaliana (Schrumpfová and Majerská et al., 2017, 2018; see Supp. K and L). Although a TDNA insertional mutation of AtWHY1 did not result in detectable abnormal phenotypes, atwhy1 mutant plants contained longer telomeres, whereas AtWHY1 overexpressing plants showed shortened telomeres and decreased telomerase activity (Yoo et al., 2007). While proteins from A. thaliana (AtWhy1) and from Hordeum vulgare (HvWhy1) (Grabowski et al., 2008) were found to bind plant telomeric repeat sequences in vitro, diverse organelle localization of other Why family members from O. sativa, A. thaliana, S. tuberosum (Krause et al., 2005; Schwacke et al., 2007) and proposed binding to ssDNA of melted promoter regions (Desveaux et al., 2002), rather indicate a role in communication between plastid and nuclear genes encoding photosynthetic proteins (Comadira et al., 2015; Foyer et al., 2014). Overall, it seems that Why proteins bind to various DNA sequences, including: telomeres; a distal element upstream of a kinesin gene; the promoter region of the early senescence marker gene AtWRKY53 (in a development-dependent manner) in Arabidopsis. It was further proposed that WHY1 proteins bind to both ssDNA and RNA in Z. mays chloroplasts, where it plays a role in intron splicing and WHY1 is associated with intron-containing RNA in barley chloroplasts (Guan et al., 2018). Among other proteins lacking OB-fold from A. thaliana, belongs truncated derivative of chloroplast RNAbinding protein (AtCP31) with RRM motif, named AtSTEP1 (single- stranded telomere-binding protein 1) (see Figure 5). AtSTEP protein localizes exclusively to the nucleus, specifically binds single-stranded G-rich plant telomeric DNA sequences and inhibits telomerase-mediated telomere extension (Kwon & Chung, 2004). A 36-kD protein identified by EMSA that specifically binds the G-strand of telomeric ssDNA from N. tabacum (NtGTBP1) also contains a tandem pair of RRM motifs (Hirata et al., 2004). NtGTBP1 is not only associated with telomeric sequences, as well as two additional GTBP paralogs (NtGTBP2 and NtGTBP3), but also inhibits telomeric strand invasion in vitro and leaves of knockdown tobacco plants contained longer telomeres with frequent formation of extrachromosomal T-circles (see bellow) (Lee & Kim, 2010). These observations correspond to a previously detected protein from tobacco nuclei that binds G-rich telomeric strands and reduces accessibility to telomerase or terminal transferase (Fulnečková & Fajkus, 2000). Fulnečková and Fajkus detected a 40 kDa polypeptide by SDS-PAGE after cross-linking the complex formed 27 by extracts from tobacco leaf nuclei. In addition to the above described proteins, various telomeric ssDNA binding proteins have also been reported in nuclear extracts from Glycine max, A. thaliana, O. sativa or Vigna radiata (Ho Lee et al., 2000; Kim et al., 1998; Kwon & Chung, 2004; Zentgraf, 1995). However, precise characterization of these proteins, identified by EMSA is mostly missing. 2 Interstitially located telomeric repeats Telomeric repeats are not exclusively located at the physical ends of chromosomes, known as telomeres. They are also present in multiple internal sites of chromosomes in many species, where they are referred to as interstitial telomeric sequences (ITSs) (also named interstitial telomeric repeats (ITRs) or even very short internally localized telomeric repeats (named telo-boxes). ITSs are relatively abundant in subtelomeric, pericentromeric, and centromeric regions of most eukaryotic organisms, but can also be found at various positions throughout chromosomes. Short internally localized telomeric repeats - called telo-boxes - are composed of one to two telomeric DNA repeats. However, the defining of these groups is not entirely precise and may vary in various scientific resources (Aksenova & Mirkin, 2019; Tremousaygue et al., 1999; Uchida et al., 2002). Most interstitial telomeric sequences studied in the human genome are short ITSs with lengths varying from 2–25 copies. They are present in all human chromosomes in subtelomeric regions as well as far from chromosomal ends (Aksenova & Mirkin, 2019; Azzalin et al., 2001; Ruiz-Herrera et al., 2009). Recent studies suggested function of ITSs in the stability of the genome and specifically at the role played by ITSs in interacting with the nuclear envelope and shaping the genome’s 3D structure. A model of mammalian chromosomal organization involves interaction of telomeres with ITSs and nuclear Lamins (Lamin A/C) (Vicari et al., 2022; A. M. Wood et al., 2014). Long ITSs represent fragile parts of chromosomes, which are prone to rearrangements and recombination’s (reviewed in Aksenova & Mirkin, 2019). In A. thaliana, 8 regions of long ITSs were described on three chromosomes, ranging from 300 bp to 1.2 kb (Uchida et al., 2002). Large blocks of telomeric repeats were found in pericentromeric regions of some chromosomes in representatives of the Solanaceae family (He et al., 2013). Interestingly, the large blocks of imperfect telomeric repeats were found as well in the proximity of centromeres of all Ballantinia antipoda (Brassicaceae) chromosomes (Mandáková et al., 2010), however, in N. tabacum, no detectable ITS regions were observed (Majerová et al., 2014) while telomere lengths ranged from 20 to 160 kb (Fajkus et al., 1995; Kovařik et al., 1996). Aside of long telomeric repeats the Arabidopsis genome contains very short interspersed segments (teloboxes) of the telomeric sequence both mainly in interstitial positions. These short telo-boxes, exhibit a nonrandom distribution. They were described in the promoters of genes coding for translation elongation 28 factor EF1a (Liboz et al., 1990), promoters of many ribosomal protein coding genes (Tremousaygue et al., 1999) and promoters of genes involved in the biogenesis of the translation machinery (Gaspin et al., 2010). We developed our own program, named Gene RegulatOry ELEMents (GOLEM) https://golem.ncbr.muni.cz/ (Nevosad et al., in preparation), to precisely localize the distribution of telo-boxes in the vicinity of the Transcription Start Site (TSS) and Translation Start Site (ATG). Using this program, we found that most of the telo-boxes in the Arabidopsis genome are located in very close proximity to the TSS. Additionally, we discovered that genes with high transcription levels in plant leaves or during certain stages of gametophyte development tend to have telo-boxes located predominantly 100 bp downstream of the TSS (Klodová et al., in preparation). 2.1 Proteins associated with long interstitial telomeric repeats (ITSs) The long extra-telomeric repeats can be recognised by the proteins that were previously characterised as telomere-binding. In yeast several proteins were found to be associated with an artificial interstitial telomeric tract or subtelomeric ITSs, e.g. Rap1, KU or Tbf1 proteins (reviewed in Aksenova & Mirkin, 2019). Also in mammals Shelterin components occupy selective ITSs in the human genome, e.g. long artificial ITSs showed enrichment in hTRF1 and hTRF2 proteins, as well as in the hTRF2-interacting partner, Apollo exonuclease (Simonet et al., 2011; Ye et al., 2010). The region 2q14 on human chromosome, containing stretches of degenerate TTAGGG repeats, binds hTRF1, hTRF2, hRAP1 and hTIN2 proteins (Fan et al., 2002). These extra-telomere located Shelterin components thus participate in additional roles, e.g. gene activation and repression, DNA replication, heterochromatin boundary-element formation, creation of hotspots for meiotic recombination and chromatin opening (reviewed in Aksenova & Mirkin, 2019; Schrumpfová et al., 2016a; see Supp. J). Using Chromatin immunoprecipitation assay combined with Next generation sequencing (ChIP-Seq) we revealed preferential association of AtTRB1 protein with long telomeric repeats, but not centromeric or 18S rDNA sequences (Schrumpfová et al., 2016b; see Supp. I). Recently we contributed to the findings that histone H1 selectively prevents accumulation of trimethylation of lysine 27 of histone H3 (H3K27me3) at telomeres and long-ITSs by restricting DNA accessibility to AtTRB proteins. It was proposed that H1 safeguards telomeres and long-ITSs against excessive H3K27me3 deposition and preserves their topological organization. Despite low protein sequence similarity of H1/H5 domain of AtTRBs and H1 (14%), AtTRBs display a typical H1/H5 domain, that may antagonize chromatin incorporation of the H1/H5 of AtTRB and H1 proteins and might modulate PRC2 recruitment at ITSs (Teano et al, 2023; see Suppl. S). 29 2.2 Proteins associated with short internally localized telomeric repeats (telo-boxes) In our study Schrumpfová et al., we were the first group to describe the association of AtTRB1 with teloboxes in the plant genome (Schrumpfová et al., 2016b; see Supp. I). Moreover, we found out that AtTRB1 is bound to telo-boxes in promoters all over the genome. Almost 28 % of telo-box sequences located in the 5' UTR region of the genes coding proteins are covered by AtTRB1. As telo-box sequences are preferentially located in the promoters of genes involved in the biogenesis of the translation machinery we proposed role of AtTRB proteins in regulation of several genes, especially genes involved in biogenesis of the translational machinery (Schrumpfová et al., 2016b; see Supp. I) (see Figure 6). Our observation that AtTRB proteins are associated with telo-box sequences located outside the telomeres was later proven by group of Franziska Turck. Zhou et. al. (2016) showed that AtTRB1 binds to thousands of genomic sites containing telo-box or related cis-elements with a significant increase of sites and strength of binding in the mutant plants for Like Heterochromatin Protein 1 (AtLHP1) (Zhou et al., 2016, 2018). AtLHP1 is a plant Polycomb Repressive Complex 1 (PRC1) component that directly binds to H3K27me3 (Turck et al., 2007). It was further shown that telo-boxes are part of the cis-regulatory elements that may relate to recruitment of Polycomb repressive complex 2 (PRC2) which may regulate transcription of target genes through histone Figure 6. An overview of telomeric and non-telomeric locations of TRB1 protein within A. thaliana nucleus, where the telomeres are clustered in a rosette-like configuration, including nucleolus-associated telomeres. Modified from Schrumpfová et al. (2016b, see Supp. I). 30 modifications. Zhou et al., 2018 have show direct interaction between AtTRB1,2,3 and CURLY LEAF (AtCLF) and SWINGER (AtSWN) subunits of PRC2 complex. Recently we have described novel interaction between AtTRBs and EMBRYONIC FLOWER 2 (AtEMF2) and VERNALIZATION 2 (AtVRN2) subunits of PRC2 complex (Kusova et al., 2023; see Suppl. R). There was also proposed role of AtTRB proteins in PEAT complex (PWWPs-EPCRs-ARIDs-TRBs). PEAT complex may mediate histone deacetylation and heterochromatin condensation and thereby facilitate Figure 7. Overview of the main Telomere repeat binding proteins (TRBs) functions (Kusova et al., 2023; see Suppl. R). A) TRBs are associated with the physical ends of chromosomes (telomeres) via their Myb-like domain (Schrumpfová et al. 2004; see Suppl. A; Mozgová et al. 2008; see Suppl. D; Dvořáčková et al. 2010; Schrumpfová et al. 2014; see Suppl. H; Dreissig et al. 2017). TRBs interact with Arabidopsis homologs of the G-overhang binding protein Protection of telomere 1a, b (POT1a, b) (Schrumpfová et al. 2008; see Suppl. D; Kusova et al., 2023; see Suppl. R). B) TRBs mediate interactions of Recombination UV B – like (RUVBL) proteins with the catalytic subunit of telomerase (TERT) (Oguchi et al. 1999), and participate in telomerase biogenesis (Schrumpfová et al. 2014; see Suppl. H; Schořová et al. 2019; see Suppl. N). TRBs are associated in the nucleus/nucleolus with POT1a (Schořová et al. 2019; see Suppl. N), and also with a plant orthologue of dyskerin, named CBF5 (Lermontova et al. 2007) that binds the RNA subunit of telomerase (TR) (Fajkus et al. 2019; Song et al. 2021). C) TRBs are associated with short telomeric sequences (telo-boxes) in the promoters of various genes in vivo, mainly with translation machinery genes (Schrumpfová et al. 2016; see Suppl. I) . ORF, Open reading frame. D) Telo-box motifs recruit Polycomb repressive complexes (PRC2) via interactions of PRC2 subunits with TRB (Zhou et al. 2016; Zhou et al. 2018, this study) CLF, CURLY LEAF; SWN, SWINGER; EMF2, EMBRYONIC FLOWER 2; VRN2, VERNALIZATION 2. E) Histone H1 prevents the invasion of H3K27me3 and TRB1 over telomeres and long interstitial telomeric regions (Teano et. al, 2023; see Suppl. S). F) TRB proteins, as subunits of the PEAT (PWO-EPCR-ARID-TRB) complex, are involved in heterochromatin formation and gene repression, but also have a locus‐specific activating role, possibly through the promotion of histone acetylation (Tan et al. 2018; Tsuzuki and Wierzbicki 2018; Mikulski et al. 2019). 31 heterochromatin silencing. PEAT complex represses in heterochromatin regions the production of small interfering RNAs (siRNAs) and DNA methylation in A. thaliana (Tan et al., 2018; Tsuzuki & Wierzbicki, 2018). On the other hand, PEAT complex may possess a locus‐specific activating role, possibly through promoting histone acetylation through two MYST-type histone acetyltransferases, AtHAM1 and AtHAM2. The composition of PEAT indicates that it binds to specific regions of chromatin, probably telo-boxes via AtTRB protein, and adds or removes acetyl groups from histones (Tan et al., 2018; Tsuzuki & Wierzbicki, 2018). Additionally, AtTRB2 directly interacts with histone deacetylases, AtHDT4 and AtHDA6, in vitro and in vivo (Lee & Cho, 2016). Deacetylase activity of AtHDT4 (W. K. Lee & Cho, 2016) and AtHDA6 (To et al., 2011) against acetylation of lysine 27 of histone H3 (H3K27ac), could be important for subsequent methylations of H3K27me3, that is among others target also for AtLHP1. Recently was identified a PWWP Interactor of Polycombs 1 (PWO1) as a novel plant-specific factor associated with chromatin and PRC (Hohenstatt et al., 2018). PWO1 associates physically with CRWN1, that is one of the Lamin-like genes in Arabidopsis forming the plant-specific CROWDED NUCLEI (CRWN) family. The authors speculated that PWO1 links H3K27me3-marked chromatin and the nuclear periphery in plants. Interestingly, AtTRB1 protein was identified as putative interactors of PWO1 in Co-IP experiments coupled with MS using the PWO1:PWO1-GFP Arabidopsis transgenic line (Mikulski et al., 2019). Very recent it was demonstrated AtTRBs also associate and colocalize with JUMONJI14 (JMJ14) and trigger H3K4me3 demethylation at some loci (Wang et al., 2023). JMJ14 is histone H3K4 demethylase regulating flowering time in Arabidopsis (Lu et al., 2010). The attrb1/2/3 triple mutant and the atjmj14-1 mutant show an increased level of H3K4me3 over AtTRB and JMJ14 binding sites, resulting in up-regulation of their target genes (Wang et al., 2023). Overall, we can hypothesise that although the TRBs were originally characterized as being associated with long arrays of telomeric repeats (see Figure 7A, E), recent observations indicate broad engagement of TRB proteins in various cellular pathways via recruiting various complexes to telo-boxes (see Figure 7C, D, F). 3 Orchestration of telomere homeostasis Regulation of the telomere length homeostasis is very complex problem and is achieved via a balance between telomere lengthening and erosion over successive cell divisions. Additionally, the processes of telomere maintenance can be orchestrated by various telomere- and telomerase-associated proteins. Mammalian telomeres are recognized not only with above mentioned proteins (Shelterin complex, POT proteins, CST complex etc.) but telomere maintenance mechanisms appear to be affected by hundreds of proteins, However, activities of these plant telomere, and telomerase-associated proteins, are only partly understood. Some of these proteins were described in several broad studies, e.g. the hTERT associated 32 proteins (proteins were detected using TAP-MS) (Fu & Collins, 2007), telomeric factors associated with human telomeric chromatin (Déjardin & Kingston, 2009) or protein network surrounding Shelterin subunits - TRF1, TRF2, POT1 and TIN2 (Giannone et al., 2010; Grolimund et al., 2013; Nittis et al., 2010). The putative partners associating with Shelterin proteins fell into functional categories such as DNA damage repair, ubiquitination, chromosome cohesion, chromatin modification/remodelling, DNA replication, cell cycle and transcription regulation, nucleotide metabolism, RNA processing and nuclear transport. These putative protein-protein associations may participate in different biological processes at telomeres or, intriguingly, outside telomeres. 3.1 Telomerase As was already described above, telomeres cannot be fully replicated by enzymes that duplicate DNA, so the telomere shortening occurs with each round of DNA replication. Critically shortened telomeres are no longer able to protect chromosome ends from DNA repair and degradation activities and these phenomena can lead to replicative senescence and finally cell death (Lundblad & Szostak, 1989). Telomerase is a ribonucleoprotein that adds a species-dependent telomere repeat sequence to the 3' end of telomeres and elongates the telomeres. In humans, telomerase activity was detected in all early developmental stages. However, just after birth, telomerase activity in somatic cells is downregulated with the exception of highly dividing cells (e.g. proliferating cells, T-lymphocytes, hair follicle bulbs) (reviewed in Schrumpfová et al., 2019; see Supp. M). However, somatic downregulation of telomerase is not conserved mechanism across species, and the presence of telomerase activity has to be individually tested in each individual species, tissue or even in different age-classes (Gomes et al., 2011; Haussmann et al., 2007; Seluanov et al., 2007). For example: most rodent species show high telomerase activity in multiple somatic tissues (Mus musculus, Rattus norvegicus, Hetrocephalus glaber etc.) and only beaver and capybara show nearly complete somatic repression of telomerase activity, similar to humans (Seluanov et al., 2007). There was shown a clear tendency for species smaller than 1 kg to have long telomeres and active telomerase, but species larger than 1 kg have tendency to have short telomeres and repress telomerase (Gomes et al., 2011). Also plant cells possess telomerase which is used for maintenance of their telomeres (Fajkus et al., 1996; Heller et al., 1996). Active telomerase was detected in organs and tissues containing highly dividing meristem cells such as seedlings, young and middle-age leaves, root tips, floral buds and flowers (Fajkus et al., 1998; Fitzgerald et al., 1996). In terminally differentiated tissues (stems, mature leaves), telomerase activity is suppressed (Jurečková et al., 2017; Ogrocká et al., 2012; Riha et al., 1998; reviewed in Schrumpfová et al., 2019; see Supp. M ). 33 As was already mentioned above, the two core subunits of telomerase are telomerase reverse transcriptase (TERT), which possesses catalytic activity, and telomerase RNA subunits (TR), which contain a template region directing the synthesis of DNA repeats at the ends of chromosomes. 3.1.1 Telomerase reverse transcriptase (TERT) Telomerase protein catalytic subunit (TERT) contains several conserved motifs and domains. The TERT protein contains N-terminally located telomerase-specific motifs important for binding the telomerase RNA subunit (TRBD), centrally located catalytic domains with the RT motifs essential for enzyme activity (RT) and the C-terminal extension (CTE), which is highly conserved among vertebrates as well as among plants (see Figure 8B). The motifs localized at the N-terminus are telomerase-specific (T2, CP, QFP and T) and are A B Figure 8. Conservation of functional domains of two core telomerase subunits – TERT and TR (adopted from Schrumpfová et al., 2020; see Supp. O) . A) Models of secondary structures of human, Tetrahymena and Arabidopsis TRs suggest conservation of several structural motives including pseudoknot in the vicinity of the template (t/PK domain) and stem-loop region. In humans the stem-loop region contains the conserved 4/5 (CR4/5) region, the H (AnAnnA) and ACA-boxes (H/ACA) domains and the Cajal body box (CAB-box) motif that serve as binding sites for other protein components of the telomerase holoenzyme complex (dyskerin, NOP10, NHP2, and GAR1). In Tetrahymena the stem-loop 4 (SL4) is directly bound by p65 protein. To date, particular interactors and their binding sites have not been demonstrated directly in Arabidopsis. B) Domain arrangement of human (Animals), Tetrahymena (Ciliates) and Arabidopsis (Plants) TERTs. The supergroup for each species is given. N-terminus: telomerase essential N-terminal (TEN) domain and RNA‐binding domain (TRBD domain) are separated by Linker that contains a nucleus localization‐like signal (NLS). The central RT domain: catalytical part of the enzyme that contains seven evolutionary-conserved RT motifs (1, 2, A, B′, C, D and E motifs) and also telomerase specific 3 motif. C-terminus: C‐terminal extension (CTE) domain. 34 important for binding the telomerase RNA subunit. The centrally located RT motifs (1, 2 and A–E) are essential for enzyme activity (reviewed in Sýkorová & Fajkus, 2009). The human telomerase complex purified from human cell line overexpressing hTERT and hTR forms a dimeric structure (Sauerwald et al., 2013). However, the presence of two catalytically active hTERT subunits has been a topic of controversy, as indicated by other studies. Although, the biological significance of a dimeric telomerase RNP is unclear, it could perhaps facilitate telomerase recruitment to telomeres by providing multiple binding sites, thus increasing the affinity for its telomeric receptor (reviewed in Schmidt & Cech, 2015). We performed Y2H screening of several AtTERT fragments. These fragments of AtTERT were previously designed to variously cover N-terminal, TRBD, RT or CTE domains (Zachová et al., 2013). According our results, dimerization of AtTERT in A. thaliana can be mediated by the RNA binding domain (TRBD) that is able to interact separately with the N-terminal fragments and itself (Schrumpfová and Majerská et al., 2017, 2018; see Supp. K and L). In AtTERT, multiple nuclear localization signals (NLS), nuclear export signal (NES) or a mitochondrial targeting signal were reported (Zachová et al., 2013). Due to the presence of these signals, AtTERT protein and its domains localize mainly within the nucleus and the nucleolus of A. thaliana (Rossignol et al., 2007; Zachová et al., 2013). Similarly in our study we localised AtTERT domains in the nucleolus. According to our observation that AtTERT domains can be colocalized together with the AtRUVBL, AtTRB and AtCBF5 proteins in the nucleolus, we hypothesised that AtTERT nucleolus localisation may be part of the telomerase assembly pathway (Schořová et al., 2019; see Supp. N). Apart from telomeric functions of telomerase in the nucleus, there was reported the presence of telomerase in other subcellular compartments or telomerase putative involvement in signalling pathways, transcriptional regulation and stress protection (reviewed in Majerská et al., 2011). It has been proposed that human telomerase is subjected to posttranslational regulation such as phosphorylation (Kang et al., 1999). Putative phosphorylation sites were also detected in the TERT sequences from O. sativa or N. tabacum BY-2 cells but not in AtTERT from A. thaliana (Oguchi et al., 2004; Yang et al., 2002). Moreover, in tobacco cell culture, phytohormones such as auxin or abscisic acid regulate phosphorylation of telomerase protein, which is required for the generation of a functional telomerase complex (Tamura et al., 1999; Yang et al., 2002). 3.1.2 Telomerase RNA (TR) Compared to the conserved structure of the TERT subunit, TRs show high sequence diversity among more distant organisms, as exemplified by the length differences of TRs in protozoa (159 nt in ciliate Tetrahymena, 2200 nt in Plasmodium), zebrafish (317 nt), mouse (397 nt), human (451 nt) and budding 35 yeasts (1160 nt). Even within yeasts, the homology among TRs is rather low and their lengths range from 930 nt to more than 2000 nt (see Figure 8B) (reviewed in Schrumpfová et al., 2019; see Supp. M; Webb & Zakian, 2016; Zhang et al., 2011). In A. thaliana, there were earlier reported two AtTR candidates, named AtTER1 and AtTER2 (CifuentesRojas et al., 2011, 2012). It was shown that AtTER1 is able to provide a templating function in telomerase reconstitution experiments in vitro but direct evidence of its in vivo function were missing (Fajkus et al., 2019). However, later it was found out that neither AtTER1 nor AtTER2 serve as RNA subunits of active telomerase and the article Cifuentes-Rojas et 2011 was retracted. It seems that the natural templating subunit of telomerase in Arabidopsis, as well in other land plants, are TRs identified by our group (Fajkus et al., 2019). My colleagues used unusually large length of the Allium telomere repeat unit (12 nt) and identified the candidate TRs in transcriptomes. Based on the Allium TRs, they consequently identified TRs orthologs in the other land plants. AtTR has been characterized earlier as a hypoxic stress-responsive long non-coding RNA (lncRNA) transcribed by RNA polymerase III (Pol III) in A. thaliana (AtR8) and related Brassicaceae species. All AtTR identified homologs in other plant species possess the conserved Pol III type 3 promotor with specific localization of USE and TATA boxes and poly-U terminator elements (Wu et al., 2012, 2019). It seems that land plant TR gene is highly conserved in contrary to the very divergent TR genes found in animal, yeast or protozoan models (Fajkus et al., 2019). Models of secondary structures of human, Tetrahymena and Arabidopsis TRs suggest conservation of several structural motives. The most prominent are pseudoknot in the vicinity of the template (t/PK domain), stem-loop regions and template boundary element (TBE). In humans the stem-loop region contains the conserved 4/5 (CR4/5) region, the H (AnAnnA) and ACA-boxes (H/ACA) domains and the Cajal body box (CAB-box) motif that serve as binding sites for other protein components of the telomerase holoenzyme complex. The TBE defines the end of the sequence recognized by TERT as a template (reviewed in Schrumpfová et al., 2020; see Supp. O). Evolution of both subunits of telomerase, TERT and TR, were discussed at The Czech Plant Nucleus Workshop 2021. The results, together with other results focused on maintenance of the chromosome ends, were summarized in the Conference report named The Czech Plant Nucleus Workshop 2021 (Pecinka et al., 2022; see Supp. P). 3.1.3 Telomerase-associated proteins Besides these two core subunits, TERT and TR, the telomerase complex comprises several other accessory proteins with diverse roles in telomerase assembly, trafficking, localization, recruitment to telomeres or the processivity of telomere synthesis (Chan et al., 2017; Nguyen et al., 2018) (see Figure 9 and Figure 10). 36 In humans the active telomerase is associated with Hsp90 and p23 chaperones as well as with TR associated with conserved scaffold proteins of box H/ACA small nucleolar RNAs (dyskerin, Non-histone protein 2 (NHP2), Nucleolar protein 10 (NOP10), Glycine arginine rich 1 (GAR1)). The telomerase RNP is probably retained into the nucleoli through the interaction between TERT and nucleolin. Assembly of TR and TERT into catalytically active telomerase is aided by Pontin (RUVBL1) and Reptin (RUVBL2) (reviewed in Schořová et al., 2019; see Supp. N) (see Figure 9A). In plants, a limited number of proteins that directly interact with TERT were described. Using Tandem Affinity Purification coupled to Mass Spectrometry (TAP-MS) we co-purified and identified several putative AtTERT interaction partners (Schrumpfová and Majerská et al., 2017, 2018; see Supp. K and L). To confirm putative protein-protein interactions between AtTERT and proteins of interest, we used Y2H, Co-IP and BiFC systems. As some of the proteins of interest showed indirect interaction with AtTERT, to achieve reproducible results we used many modifications, improvements and mutual combination of Y2H, BiFC and Co-IP. Our optimized BiFC protocol in A. thaliana protoplasts provided us a robust tool to observe direct or even indirect interactions of (not only) telomere- and telomerase-associated proteins and to distinguish nucleus, nucleolus or cytoplasmic localization of these interactions. Our modification of Co-IP technique (Co-Immunoprecipitation with Three Proteins of Interest) allowed detection not only of two proteins of interest, as is common, but also detection of trimeric complexes, where two proteins of interest interact indirectly via a protein sandwiched in between them and mediating the interaction. Our improvements and modifications of these protein-protein interaction techniques were described in the book chapter named Figure 9. Comparative model of telomerase in human and Arabidopsis localised in the nucleolus. A) Human active telomerase is associated with Hsp90 and p23 chaperones as well as with TR associated with conserved scaffold proteins of box H/ACA small nucleolar RNAs (dyskerin, NHP2, NOP10, GAR1). The telomerase RNP is retained into the nucleoli through the interaction between TERT and nucleolin. Assembly of TR and TERT into catalytically active telomerase is aided by Pontin (RUVBL1) and Reptin (RUVBL2). B) TERT colocalize with RUVBL proteins, bridged by telomeric TRB proteins, in the nucleolus as well as the interaction of telomeric protein POT1a with Arabidopsis CBF5 (dyskerin). CBF5 together with GAR1, NOP10, NHP2, but in contrast with human cells also NAF1, were localized in the plant nucleolus, however entire association with active telomerase holoenzyme has to be elucidated. Modified from Schořová et al., 2019; see Supp. N. A B 37 'Analysis of direct and indirect protein-protein interactions of telomere-associated proteins' (Methods in Molecular Biology, The Nucleus, Book Series, Springer protocols (Schořová et al., 2020). In our laboratory, we have demonstrated that AtTRB proteins, physically interact with N-terminal domains of AtTERT (see Figure 3B and 9B). We also suggested a mediated interaction between Telomeric Repeat Binding Protein 1 (AtTRP1) protein and AtTERT (Schrumpfová et al., 2014; see Supp. H). Rossignol et al. observed that the N-terminal part of AtTERT exclusively interacts with AtPOT1a but not AtPOT1b (Rossignol et al., 2007). As well various other proteins from A. thaliana were shown to be associated with AtTERT: AtRRM (RNA recognition motif (RRM)), AtARM (armadillo/β-catenin-like repeat-containing protein), AtCHR19 (chromatin remodeling protein), AtMT2A (Metallothionein-like), AtG2p (RNA-binding), AtPURα1 (Pur-alpha 1), AtNUC-L1 (Nucleolin like 1) or Importin4 (ImpA4) (Dokládal et al., 2015, 2018; Fulnečková et al., 2022; Pontvianne et al., 2010, 2016; reviewed in Schrumpfová et al., 2019; see Supp. M). Some of these AtTERT partners that we co-purified with N-terminal fragments of AtTERT are possibly involved also in nontelomeric functions of telomerase, e.g. the human homologue of the AtPURα1 protein, named PURα, has been implicated in the control of gene transcription (Safak et al., 1999) and DNA replication (Bergemann & Johnson, 1992). Among proteins co-purified with AtTERT fragments using TAP-MS we identified also AtRUVBL1 and AtRUVBL2a proteins (plant homologues of human Pontin and Reptin) (Schrumpfová and Majerská et al., 2017, 2018; see Supp. K and L). We closely characterised AtRUVBL1-AtTERT and AtRUVBL2a-AtTERT interactions in the plant cell and found out that, against mammalian counterparts, interaction between AtRUVBLs and AtTERT proteins in A. thaliana is not direct and is more likely mediated by one of the AtTRB proteins. Our data show that AtRUVBLs, together with AtTRBs protein, colocalize with N-terminal part of AtTERT subunit of plant telomerase in the plant nucleolus. It seems that AtRUVBLs are recruited into the AtTERT complex through an interaction with AtTRBs protein, which mediate interaction with both proteins: AtTERT and also with AtRUVBLs. Our data indicate the presence of AtTERT-AtTRB-AtRUVBL complex in the plant nucleolus (Schořová et al., 2019; see Supp. N). In humans, proper catalysis, accumulation, 3' end processing, and localization of hTR are necessary for the creation of functional mature hTR, which provides the template for the synthesis of telomere DNA repeats. Human TR is associated with dyskerin, NOP10, NHP2 and GAR1, that displaces previously bound Nuclear assembly factor 1 (hNAF1) in the hTR RNP. In A. thaliana, expression of putative AtGAR1, AtNOP10, AtNHP2 genes encoding protein components of the H/ACA box snoRNP complex correlate with that of AtCBF5 - plant homologue of dyskerin (Lermontova et al., 2007). AtCBF5 has been identified as a component of the enzymatically active A. thaliana telomerase RNP (Kannan et al., 2008; Lermontova et al., 2007). Scaffold proteins AtCBF5, AtGAR1, AtNOP10, AtNHP2, 38 but in contrast to human cells also AtNAF1, were localized into the plant nucleolus (Lermontova et al., 2007; Pendle et al., 2005). The association of TRs with dyskerin appears to be conserved between plant and animal kingdoms as telomerase activity was immunoprecipitated with the anti-plant dyskerin antibody from protein extract from Allium cepa seedlings (Fajkus et al., 2019). Moreover, despite the absence of a canonical H/ACA binding motif within AtTR, dyskerin binds AtTR with high affinity and specificity in vitro via a plant specific three-way junction (Song et al., 2021). However, it has not yet been elucidated whether plant homologues of human GAR1, NOP10, NHP2, or NAF1 are also part of the active holoenzyme of telomerase in plants. Comparative overview of human and plant homologues of proteins associated either with the telomerase catalytic subunit TERT or with the RNA component of telomerase is reviewed in Schrumpfová et al., 2019 (see Supp. M). 3.1.4 Telomerase assembly Proper assembly of TERT with TR into active and functional complex is stepwise regulated procedure governed also by multiple associated proteins (reviewed in Shepelev et al., 2023; reviewed in Schrumpfova et al., 2020; see Suppl. O). Telomerase and its chromosome end substrate have very low abundance (∼250 telomerases/184 telomeres in a human cell in late S phase) thus, it is perhaps not surprising that the telomerase enzyme is recruited to telomeres rather than simply encountering them by diffusion (Xi & Cech, 2014). Transcription of the human TERT gene by RNA polymerase II (RNA Pol II) is regulated by several activators and repressors acting at the promoter level (e.g., c-MYC, Nuclear Factor κB (NF- κB), Signal Transducer and Activator of Transcription 3 (hSTAT3), Specificity Protein 1/3 (SP1/3). Histone modification H3K27me3 often silences hTERT, however the mutated hTERT allele is marked by the active histone marks H3K4me2, H3K4me3 and H3K9ac. All Pol II transcripts undergo processing events that are essential for their function. The hTERT pre-mRNA with a 5’ mono-methylguanosine (MMG) cap and poly(A) 3’ tail can be spliced into full-length (FL) or multiple alternative isoforms (Alternative splicing) that are catalytically inactive or even inhibit telomerase activity. The binding of heat shock protein 90 (hHsp90) with its co-chaperone (p23) in the cytoplasm enables hTERT phosphorylation (P). hTERT is further imported back to the nucleus by Importin α or β1 (hImp) via nuclear pores (n.p.), while the export of hTERT may be mediated by the chromosome region maintenance 1 protein homolog (hCRM1, also known as exportin-1). The ubiquitin (Ubq)-proteasomal degradation of hTERT is driven by E3 ubiquitin-protein ligase makorin-1 (MKRN1), heat shock protein 70 (Hsp70) and carboxyl-terminus of Hsp70 Interacting Protein (CHIP) (see Figure 10A, for references see Schrumpfova et al., 2020; see Suppl. O). 39 Histone modifications H3K4me2/3 or H3K9Ac help to regulate read-through of the human hTR gene by RNA Pol II in telomerase-positive cell lines. SHQ1 chaperone and RUVBLs facilitate the assembly of nascent RNA with RNA scaffold proteins (dyskerin, hNOP10, hNHP2, and hNAF1). Mature hTR is capped with a trimethylguanosine (TMG) cap at the 5’ end, polyadenylated at the 3’ end and co-transcriptionally associated with scaffold proteins. The hTR variants with shorter or longer 3’ ends, or those associated with variant proteins, may lead to the degradation of hTR. hNAF1 is replaced by hGAR1 before the hTR ribonucleoprotein complex reaches the nucleolus (see Figure 10B). RUVBLs (Pontin and Reptin) enable telomerase assembly and allow hTERT recruitment to the nucleolus to form a mature telomerase complex while bound by nucleolin (hNCL). PIN2/TERF1-interacting telomerase inhibitor 1 (hPINX1), together with nucleophosmin (hNPM) and microspherule protein 2 (hMCRS2), regulate hTERT availability in a cell cycle-dependent manner. Telomere Cajal body protein 1 (hTCAB1, also known as hWRAP53) recognizes the Cajal body box (CAB-box) of the hTR in the mature telomerase complex and recruits it to the Cajal bodies (CBs). In CBs, hTR interacts with local proteins such as coilin while survival motor neuron protein (hSMN) binds hTERT (see Figure 10C and Figure 11A). In the S-phase, the CBs colocalize with telomeres and facilitate the recruitment of the mature telomerase complex to the telomeres via interaction with hTPP1 protein, which is one of the subunits of a protein complex localized at telomeres, termed as Shelterin. The presence of Shelterin proteins (hTRF1/2, hPOT1, hTIN2, hRAP1 and TPP1) helps distinguish chromosomal ends (telomeres) from DNA breaks see Figure 10D). Despite the fact that the entire TERT subunit is highly conserved across the phylogenetic tree and shows significant sequence homology between humans and plants (as reviewed in Schrumpfova et al., 2020; see Suppl. O), the assembly pathway of plant telomerase holoenzyme is not fully understood. However, our research has helped to partially elucidate the proteins associated with plant telomerase and their possible involvement in telomerase assembly. In humans, the production of hTERT is highly regulated at the transcriptional levels and also posttranscriptional levels, whereas the hTR transcript is constitutively produced (Gladych et al., 2011). However, in plants, the transcription of both telomerase subunits (AtTERT and AtTR) is regulated during the plant development, as both subunits show high transcription in seedlings and young leaves, but diminished transcription in fully maturated leaves (Jurečková et al., 2017; Ogrocká et al., 2012; Riha et al., 1998; Fajkus et al., 2019; reviewed in Schrumpfova et al., 2020; see Suppl. O). Plant TERT gene has a weak promoter. Fojtova et al. identified region 271 bp upstream of ATG as an putative ’minimal promoter’ able to drive sufficient transcription of the telomerase protein subunit gene, resulting in normal telomerase function (Fojtová et al., 2011). In Crhak et al. it was proposed that unknown factors necessary for tissue-specific expression of telomerase activity and restoration of telomerase function in the 40 maintenance of telomere are needed (Crhák et al., 2019). Additionally, AtTERT gene might be regulated by regulatory element at the 5′ end, e.g. within the intron 1, that has function at the level of transcription, while it is not involved in tissue-specific regulation (Fojtová et al., 2011). Figure 10. Regulation of human telomerase biogenesis (for description see text) (Schrumpfova et al., 2020; see Suppl. O). A) Transcription of the hTERT is regulated by several activators and repressors acting at the promoter level. B) Histone modifications help to regulate read-through of the human telomerase RNA (hTR) gene. Mature hTR is capped and recognised by several associated proteins. C) RuvBLs (pontin and reptin) enable telomerase assembly and allow hTERT recruitment to the nucleolus to form a mature telomerase complex while bound by several other proteins. D) In the S-phase, the CBs colocalize with telomeres and facilitate the recruitment of the mature telomerase complex to the telomeres. A B C D 41 In A. thaliana, mutation in the attac1 (Telomerase activator 1) gene led to the induction of telomerase in fully differentiated leaves without stimulating progression through the cell cycle (Ren et al., 2004). However, AtTAC1 protein does not directly bind the AtTERT promoter and rather regulates telomerase activity through regulation of the AtBT2 (protein with BTB,TAZ and calmodulin binding domains) gene expression (Ren et al., 2007). Alternatively spliced variants of TERT transcripts were also described in many plant species, e.g. A. thaliana (AtTERT), Zea mays (ZmTERT), Oryza sativa (OsTERT), Iris tectorum and tobacco (Rossignol et al., 2007; Sýkorová & Fajkus, 2009). We have characterised RUVBL homologues in A. thaliana and outlined plausible conservation of the telomerase trafficking pathway in the land plants. We showed, that RUVBL1 and RUVBL2 proteins from A. thaliana are able to form either homo- or heteromers as well as their homologues in diverse organisms, although they preferably form mutual heteromers (Schrumpfová and Majerská et al., 2017; Schořová et al., 2019; see Supp. K and N). Our experiments with plant RUVBL proteins showed that depletion of AtRUVBL1 and especially of AtRUVBL2a protein, reduced telomerase activity in plants with T-DNA insertion in AtRUVBL1 or AtRUVBL2a genes, respectively. We did not observe significant changes in transcripts of AtTERT gene in AtRUVBL1 heterozygous mutant plants and very slight, though significant, increase, in transcripts of AtTERT gene in AtRUVBL2 heterozygous mutant plant lines (Schořová et al., 2019; see Supp. N). Similarly to our results, both human RUVBL proteins, hRUVBL1 and hRUVBL2, regulate hTERT both on the gene and protein levels, only hRUVBL2 depletion inhibits hTERT promoter activity through the regulation of c-Myc (Mao & Houry, 2017; Venteicher et al., 2008). It was already mentioned that our data indicate AtRUVBL1 recruitment into the AtTERT complex through an interaction with AtTRB3 protein (see above). Formation of AtRUVBL1-AtRUVBL2 heteromer is distributed in whole nucleus but the localization of protein complex AtRUVBL1-AtTRB3-AtTERT occurs in nucleolus. We showed, that depletion of AtRUVBL1 and especially of AtRUVBL2 proteins causes reduced telomerase activity and suggests conserved role of AtRUVBL proteins in maturation of functional telomerase complex across the mammals and also plant species (see Figure 11B) (Schořová et al., 2019; see Supp. N). Very recently, we have shown that AtRUVBL1 and AtRUVBL2A play roles in reproductive development. We showed that mutant plants produce embryo sacs with abnormal structure or with various numbers of nuclei and pollen grains of heterozygous mutant plants exhibit reduced viability and reduced pollen tube growth in vitro. The activity of the AtRUVBL1 and AtRUVBL2A promoters was observed in the embryo sac, pollen grains, and tapetum cells, and for AtRUVBL2A also in developing ovules. It seems that RUVBL proteins are essential for the proper development of both male and particularly female gametophytes in Arabidopsis (Tomaštíková et al., 2023, see Supp. Q). 42 Plant homologue of dyskerin, named AtCBF5 (or AtNAP57), is localized within nucleoli and Cajal bodies (Lermontova et al., 2007) and associates with enzymatically active telomerase RNP particles in an RNAdependent fashion (Kannan et al., 2008). We observed indirect interaction of AtTRBs with AtCBF5 in plant nucleus. Moreover, we detected that the AtCBF5 is interacting with AtPOT1a not only in Y2H and Co-IP as was shown in Kannan et al. (2008) but we also showed nucleolar and partly cytoplasmic localization using BiFC assay. In addition, we observed weak interaction between AtPOT1a-AtRUVBL1 proteins in Y2H and CoFigure 11. Comparative model of telomerase assembly in human and Arabidopsis (adopted from Schořová et al., 2019; see Supp. N) A) Human TR, located in the nucleolus, is bound by dyskerin, NHP2, NOP10 and GAR1 and human TERT associates with the chaperones Hsp90 and p23. Assembly of TR and TERT into catalytically active telomerase is aided by RUVBL1 (Pontin) and RUVBL2 (Reptin)) AAA+ ATPases. Telomerase is recruited to Cajal bodies by its interaction with TCAB1. The CBs will colocalize with telomeres and telomerase is recruited to telomeres by the interaction with the shelterin component TPP1. B) Arabidopsis CBF5, GAR1, NOP10, NHP2 and also NAF1, were localized into the plant nucleolus. TERT interaction with RUVBL proteins is bridged by telomeric TRBs. Arabidopsis telomeres cluster at the periphery of the nucleolus which is mediated by the presence of nucleolin. Recruitment of the mature telomerase complex to telomeres with or without commitment of Cajal bodies in Arabidopsis needs further investigation. Proteins already proven as associated with Cajal bodies are highlighted in Cajal bodies in color. Proteins that have not yet been experimentally proven as Cajal bodies associated are marked with black and white. A B 43 IP assays (Schořová et al., 2019; see Supp. N). Additionally, Arabidopsis GAR1, NOP10, NHP2 and also NAF1 homologues, were localized into the plant nucleolus (Lermontova et al., 2007; Pendle et al., 2005). The telomerase trafficking pathway during the telomerase maturation, which comprises movement of maturating telomerase complex through nucleolus to CBs and finally to the telomeres, may be conserved also in land plants. Dvořáčkova et al. observed that AtTRBs are located not only in the nucleolus but also in nuclear bodies of different size, some of which might be CBs adjacent to the nucleolus (visualized by a marker protein Coilin) (Dvořáčková et al., 2010b). Furthermore, plant dyskerin, AtCBF5, indirectly interacts with AtTRB proteins not only in the plant nucleolus but also in other nuclear bodies that might be CBs (Schořová et al., 2019; see Supp. N). Notably, not all the organisms (e.g., budding yeast and ciliates) rely on the CBs trafficking since telomerase RNAs from these species do not have H/ACA or CAB box motifs, e.g., in Saccharomyces cerevisiae, telomerase assembly requires export of the TR out of the nucleus and is regulated in a cell cycle-dependent manner. RNA component of S. cerevisiae telomerase, named TLC1, is assembled with Sm proteins in nucleoplasm, 5′ TMG cap is added to the TLC1 in nucleoli, TLC1 is assembled in cytoplasm with holoenzyme proteins and consequently telomerase holoenzyme is transported again in the nucleoplasm, where telomerase can be recruited to telomeres (reviewed in Shepelev et al., 2023; R. A. Wu et al., 2017). As was already mentioned above, we used Y2H, BiFC and Co-IP techniques to detect and characterise protein-protein interactions of the telomere and telomerase associated proteins. The Co-IP technique is based on precipitation a of intact protein complexes formed by proteins usually produced in in vitro transcription/translation systems and using an antibody that specifically binds to the particular protein antigen. Interestingly, we chose mammalian Reticulocyte lysate (RRL) instead of plant Wheat germ (WG) system to express plant AtTRB proteins in vitro (Schořová et al., 2019; Schrumpfová et al., 2014; see Supp. N and H). Wheat germ extract is isolated from embryos of dry wheat seeds while Rabbit reticulocyte lysate is prepared from anaemic rabbits that are stimulated for production of immature red blood cells responsible for the synthesis of haemoglobin that have already lost their nuclei (reticulocytes). When we used for protein expression WG transcription/translation system (TnT Coupled Wheat Germ Extract System, Promega) instead of RRL (TNT T7 Quick Coupled Transcription/Translation System (Promega)), AtTRB proteins were successfully expressed but revealed no interactions, including well established positive controls (Schořová et al., 2020: Methods in Molecular Biology - The Nucleus, Book Series, Springer). This observation might relate to the HSP90 chaperone. HSP90 chaperone is present in WGE extract but is a deficient in its function (Antonsson et al., 1995). The addition of purified human or yeast co-chaperone p23 to WGE fully reconstitutes HSP90 chaperone activity (Dittmar et al., 1997; Hutchison et al., 1995). Zhang et al. showed that p23-like proteins are present in plants, they are capable of binding HSP90, but unlike human 44 p23, the plant p23-like proteins do not reconstitute HSP90 chaperone activity (Zhang et al., 2010). As human chaperone HSP90 and its co-chaperone p23 participate in the folding of a number of cell regulatory proteins, stably associate with hTERT and remain associated also with active telomerase (Forsythe et al., 2001; Holt et al., 1999) it will be interesting to learn whether and how are these chaperones in the plant cells involved in telomere- and telomerase-associated proteins folding and telomerase assembly. Generally, assembly of functional AtTR RNP, as well as the assembly of mammalian hTR RNP, is certainly a multistep process that may include AtTR, AtCBF5, AtTRBs, AtRUVBLs, AtPOT1a and many other factors, whose presence/participation/mutual interactions will be the subjects of our future research. Dynamics and complexity of mutual interactions can be demonstrated by the fact that we detect the interacting complex of AtCBF5-AtPOT1a in the nucleolus or in the cytoplasmic and nuclear foci, while AtCBF5-AtTRBs interactions are localized entirely to the nucleoli and additional nuclear bodies (Schořová et al., 2019; see Supp. N). Our first model of plant telomerase holoenzyme assembly was achieved by editors of The Plant Journal, who wrote a special article named “RESEARCH HIGHLIGHT: The journey to the end of the chromosome: delivering active telomerase to telomeres in plants.”(Sweetlove & Gutierrez, 2019; see Supp. N). As we already mentioned - regulation of telomerase assembly, maturation and trafficking is a very complex process, involving a wide range of co-factors. Moreover, these co-factors are not involved exclusively into the telomerase assembly but they also participate in various other biochemical pathways. Although in mammals the telomerase assembly pathway has been partially described, our understanding of telomerase assembly in plants is still far to be perfect and is still ongoing process. 3.2 Telomere maintenance proteins 3.2.1 Mammalian telomere maintenance proteins The mammalian Shelterin complex is involved in the repression of the primary signal transducers of DNA breakage, two phosphatidylinositol-3-kinase-like (PI3K) protein kinases: ataxia telangiectasia mutated (ATM) and ATM- and RAD3-related (ATR) kinases. Mice TRF2 acts mainly to protect telomeres against ATM activation (Celli & de Lange, 2005) and POT1 is principally involved in repression of the ATR pathway (Denchi & de Lange, 2007; Guo et al., 2007) (see Figure 3A). In mammals as well as in other organisms, DSBs activate ATM kinase in a manner dependent on the meiotic recombination 11 (MRE11), DNA repair protein 50 (RAD50) and Nijmegen breakage syndrome 1 (NBS1) named MRN complex. The MRN complex has been found to associate with telomeres and contributes to their maintenance (reviewed in Lamarche et al., 2010). Other proteins involved in DDR machinery are Ku proteins. Human Ku70 protein directly interacts 45 not only with the Shelterin proteins hTRF1, hTRF2 and hRAP1, but also with telomerase subunits hTERT and hTR (reviewed in Fell & Schild-Poulter, 2015; Schrumpfová et al., 2019; see Supp. M). Aside DNA damage factors, the mammalian telomere proteome comprises additional telomere-associated proteins, e.g. regulator of telomere elongation helicase 1 (RTEL1) (see Figure 3A) and many other proteins interactors (reviewed in Ghisays et al., 2021; Lazzerini-Denchi & Sfeir, 2016). RTEL1 helicase connects telomeric loops and circles with DNA recombination and telomere replication. RTEL1 play role in dissolving higher-order structures referred as the telomeric loops (T-loops). These lariat structures are composed of each chromosome terminus being folded back upon itself, which enables the G-rich DNA overhang to invade and base-pair with the complementary strand (Griffith et al., 1999). The 3′ G-strand extension that invades the duplex telomeric repeats forms a D-loop (displacement loop, ~150 bp) (Greider, 1999). In addition to its role in T-loop stability, mouse RTEL1 can dissolve G4-DNA structures (quadruplexes), which are predicted to form in the G-rich telomeric regions and might block replication fork progression and the extension of telomeres by telomerase. 3.2.2 Plant telomere maintenance proteins In A. thaliana short telomeres in telomerase-deficient plants activate both the ATM and ATR, whereas absence of members of the plant CST complex initiates only AtATR-dependent, but not AtATM-dependent DNA damage response (see Figure 3B) (Amiard et al., 2011; Boltz et al., 2012). In contrast to a massive loss of telomeric DNA that was observed in human cells (Wang et al., 2009), mutations in Ku70 and Ku80 in the dicotyledonous A. thaliana, as well as in the monocotyledonous O. sativa, resulted in longer telomeres, suggesting their conserved role in the negative regulation of plant telomerase (Bundock et al., 2002; Gallego et al., 2003; Hong et al., 2010; Riha et al., 2002). A. thaliana RTEL1 homolog suppresses HR and is involved in processing DNA replication intermediates and interstrand and intrastrand DNA cross-links. Deficiency of the AtRTEL1 triggers a SOG1-dependent replication checkpoint in response to DNA crosslinks. AtSOG1 targets numerous genes required for repair by HR, including AtRAD51 (Ogita et al., 2018). Similarly to the situation in mammals, the Arabidopsis RTEL1 contributes to telomere homeostasis (Recker et al., 2014). In contrast with the effects of the loss of function of HR factors, the loss of key factors of NHEJ (MRE11, RAD50, NBS1, KU70 and LIG4) has little or no impact on growth phenotype, overall DSB repair and telomere maintenance in P. patens, while a clear telomere phenotype can be seen in the corresponding A. thaliana mutants. Therefore, it is not possible to simply generalize the results obtained in only one of these model plants as applying to DNA repair and telomere biology in all plants (Fojtová et al., 2015; Goffová et al., 2019; Holá et al., 2013). 46 3.2.3 HMG proteins Proteins classified within the High Mobility Group (HMG) family have been observed to exhibit the capacity to impact the maintenance of telomeres. These HMG proteins constitute a diverse cohort of non-histone proteins that are comparatively small in size, and are relatively abundant within the chromatin of eukaryotic organisms. There are three structurally distinct classes of HMG proteins: the HMG-nucleosome binding subfamily (HMGN), the HMG-AT-hook subfamily (HMGA) and the HMG-box subfamily (HMGB) (reviewed in Reeves, 2015). In mammals, the HMGA subfamily is composed of two proteins: HMGA1 and HMGA2. Both proteins are expressed in embryonic tissues and embryonic stem cells, are absent in most somatic adult cells and, interestingly, are highly abundant in tumorigenic cells. HMGA proteins are believed to play a role in transcription by promoting the joining of regulatory elements and were shown to have a clear role in development (Ozturk et al., 2014). There was indicated a role for HMGA1 in TERRA (TElomeric Repeatcontaining RNA) localization to the telomeres (Scheibe et al., 2013). HMGA2 positively regulates the transcription of the catalytic subunit of the telomerase in human HeLa cells (Li et al., 2011) and increases telomere stability in cancer cells (Natarajan et al., 2016). In Arabidopsis, several uncharacterized HMGA proteins are present, including AtGH1-HMGA1 (reviewed in Kotliński et al., 2017). Chabonnel et al. performed a label-free quantitative proteomics analysis of a telomere pull-down with either the Arabidopsis TTTAGGG repeat sequence or a shuffled DNA control. They identified several candidate proteins, including AtTRB1, AtTRB3 and AtGH1-HMGA1 enriched with the telomeric bait. AtGH1-HMGA1 can be present at some DNA extremities but is not associated exclusively with the telomeres. AtGH1-HMGA1 is required for efficient DNA damage repair and telomere integrity in Arabidopsis. AtGH1-HMGA1 mutants exhibit developmental and growth defects, accompanied by ploidy defects, increased telomere dysfunction-induced foci, mitotic anaphase bridges and degraded telomeres. It seems that GH1-HMGA1 in A. thaliana is involved directly in the repair process by allowing the completion of homologous recombination (Charbonnel et al., 2018). Interestingly, AtTRB proteins, associated with telomeres, possess centrally located H1/H5 domain (Schrumpfová et al., 2004; see Supp. A) (see Figure 5) that are evolutionary related to the H1/H5 domain located at the N-terminus of the AtGH1-HMGBA proteins in A. thaliana (Kotliński et al., 2017). In mammals, it has been observed that additional members of the HMG family originating from the HMGB subfamily, particularly HMGB1, are capable of regulating the activity of telomerase. However, this effect was not due to changes in expression of either of the telomerase subunits, but rather through the involvement of the HMGB1 in assembly of telomerase nucleoprotein complex. Accordingly, HMGB1 47 physically interacts with both mouse TERT and TR, as well as with active telomerase complex in vitro (Polanská et al., 2012). In contrast to mammalian HMGB proteins, which contain two HMG-box domains, the typical plant HMGBtype proteins have a single HMG-box domain, which is flanked by a basic N-terminal domain and an acidic C-terminal domain. The HMG-box domains of the various plant HMGB proteins are relatively conserved, but compared to the mammalian homologues the basic and acidic flanking regions vary considerably in length and sequence (Pedersen & Grasser, 2010). According to results of in vitro studies, plant HMGB proteins bind linear DNA in non-sequence-specifical manner with moderate affinity. They also recognise specifically certain DNA structures such as minicircles and four-way junctions and they severely bend linear DNA upon binding. In Arabidopsis is complicated by the existence of seven proteins that contain HMG-box domain flanked by a basic and acidic domain and thus can be classified as HMGB-type proteins (Lildballe et al., 2008). Most of the AtHMGB proteins were shown to be involved in various stress-response pathways (Roy et al., 2016). In our study Schrumpfová et al., 2011, T-DNA insertion lines with athmgb1 gene knockout were characterised. AtHMGB1 protein appears as a typical member of the plant HMGB-type proteins in A. thaliana and could be regarded as the ortholog of mammalian HMGB1, but not necessarily performing the equivalent functions. Similarly to mammals, general telomere lengths were significantly shortened in mutant athmgb1 plants compared to wild-type plants. In accordance with these results, in the plant lines overexpressing AtHMGB1, elongated telomeres are not dispersed continuously but they rather migrate on agarose gel as discrete bands, which is typical for telomeres generated by alternative lengthening of telomeres (ALT) (see below). These observations were proven by fluorescence in situ hybridisation on metaphase chromosomes where moderate but significant increase of telomeric signal in the AtHMGB1 overexpressing line samples as compared to the wild type (Schrumpfová et al., 2011; see Supp. G). However, the pathway mediating this effect seems to be in different between plant and mammals. While the telomere shortening in mouse cells lacking mHMGB1 can be attributed to the insufficient telomerase activity, no changes in telomerase activity and telomerase processivity could be observed in either athmgb1 or AtHMGB1 overexpressing plants. From our results we can conclude that AtHMGB1 protein does not exert its effect on telomere length via direct regulation of telomerase, however, AtHMGB1 is involved in the stress- or stimulus-responsive pathways affecting telomere length (Schrumpfová et al., 2011; see Supp. G). 3.3 Telomerase-independent telomere maintenance Besides the telomerase-based mechanism of telomere elongation, various organisms as well as plants, utilize a telomerase-independent telomere maintenance mechanism: alternative lengthening of telomeres (ALT). The exact mechanism behind telomere maintenance in the ALT pathway is unclear, but likely is based 48 on homologous recombination (HR) and may become active upon the loss of telomerase (Dunham et al., 2000; Min et al., 2017; Zhang et al., 2019). ALT relies on the formation of terminal T-loops, which parallels the first steps of HR (see Figure 12). The eventual resolution of these T-loops and aberrant HR at telomeres generates not only telomeres of highly heterogeneous lengths but also extrachromosomal T-circles, which are the known hallmarks of ALT. These ALT hallmarks include not only already mentioned heterogenous distribution of the telomere lengths and several classes of extrachromosomal telomeric repeats in the nucleus. ALT-positive cells show also increased telomere sister chromatid exchanges (T-SCE) and presence of promyelocytic leukemia nuclear bodies (APBs) associated with some telomeres. APBs are special variety of PML (promyelocytic leukemia) nuclear bodies found in the normal interphase nucleus. PML bodies are donut-shaped nuclear domains composed of PML and SP100 proteins, which are stabilized by non-covalent interactions of the posttranslational modification SUMO but they do not contain nucleic acids in normal cells (reviewed in Corpet et al., 2020). However, in ALT-positive cells, a subset of PML nuclear bodies, APBs, co-localizes with telomeric DNA. APBs contain PML nuclear bodies components such as PML, SP100 and SUMO and, moreover, telomeric DNA and telomere associated proteins including the Shelterin components TRF1, TRF2, POT1, TIN2, TPP1 and RAP1. Additionally, APBs contain factors that are involved in DNA damage response (DDR) and repair reviewed in Corpet et al., 2020). ALT mechanism is predominantly activated in a Figure 12. T-loops and Alternative lengthening of telomeres (ALT). A) Specific telomeric structure T-loop, where the 3′ G-strand extension invades the duplex telomeric repeats and forms a D-loop (displacement loop), prevent telomerase access to the telomeres. Figure modified from de Lange, 2004. B) Telomeres progressively shorten in normal cells with each division in the absence of a telomere maintenance mechanism. In in ~10–15% of tumours, a DNA homologous recombination mechanism, instead of telomerase activation, can be engaged. ALT cells use a telomeric DNA template that is copied to a telomere of a non-homologous chromosome, This telomeric DNA could add telomeric repetitive sequences to another region of the same telomere via loop formation or to the telomere of a sister chromatid (Shay and Wright, 2019). A B 49 number of human tumours and in human cells immortalized in culture but also was observed in normal somatic tissues (Neumann et al., 2013). ALT is active in about 10-15 % of cancers (Heaphy et al., 2011). Telomerase-mediated synthesis of telomeres is also essential for sustained growth and propagation in plants. Inactivation of a gene coding for catalytical subunit of telomerase, AtTERT, leads to a gradual shortening of telomeres by 200-500 nt per generation (Fitzgerald et al., 1999; Riha et al., 2001). After 6-8 generations, some telomeres in attert mutants shorten to ∼300-400 nt and start fusing with other chromosome ends (Heacock et al., 2007; Riha et al., 2001). Plants with such dysfunctional telomeres exhibit developmental defects and reduced fertility. The severity of these phenotypes worsens with progressive telomere shortening and mutant populations cannot be propagated beyond 8-10 generations (Riha et al., 2001). However, while yeast and human telomerase-deficient cell lines appear to readily adopt ALT for telomere maintenance, extensive selection of cells derived from Arabidopsis attert mutants failed to recover cultures featuring hallmarks of ALT (Watson et al., 2005; Zellinger et al., 2007; reviewed in Schrumpfová et al., 2019; see Supp. M). In our study Růčková et al. (2008) we described that the ALT mechanism is activated not only in mutant plants with telomerase dysfunction but possibly also during the earliest stages of normal plant development (Růčková et al., 2008; see Supp. B). In this study we hypothesised that extremely low rates of telomere shortening per plant generation (250-500 nt) in telomerase-deficient A. thaliana mutants (attert) does not correspond to the expected outcome of replicative telomere shortening. The meristem cells in A. thaliana, which give rise to all tissues including germ-line cells, undergo many divisions, calculated by Andrew Leitch (Queen Mary, University of London) and Jiří Friml (Institute of Science and Technology Austria, Austria) at approximately 1000 divisions from seed to seed. When considering only 5-10 nt lost per cell division (the average length of RNA primer for synthesis of Okazaki fragments) as the minimum plausible loss of telomeric DNA at each round of replication (under the very improbable scenario that the primer sits exactly at the 3’ end of the parental DNA strand), then the number of cell divisions accounting for the observed telomere erosion per generation in attert mutants would be only 25-50 cell divisions. Moreover, in mammalian cells the primer does not sit exactly at the 3’ end of the parental DNA strand and the loss at telomere is between 50-100 nt per cell division. Then it would be only 5-10 cell divisions per plant generation but not already 1000 division as was stated for A. thaliana (Fajkus et al., 2005). We propagated attert mutant plants from seeds coming either from the Lower-most or the Upper-most siliques and we followed the length of their telomeres over several generations. We proved that in the absence of telomerase, the number of cell divisions within one generation influences the control of telomere lengths. Our data showed a fast and efficient activation of a telomerase-independent mechanism 50 in response to the loss of telomerase activity and imply that ALT is probably involved also in normal plant development (Fajkus et al., 2005; Růčková et al., 2008; see Supp. B). The group of Karel Říha speculated that the meristem cells, however, do not undergo so many cell divisions as was proposed above, as they observed that the number of DNA replications is only slightly increased in plant growing under long-lived conditions in comparison to the plant growing in short-day conditions. They showed that the cell depth of gametes is not linearly proportional to the vegetative growth period and suggested that older plants may not be passing on more mutations to their offspring relative to younger plants (Watson et al., 2016). The involvement of ALT in the earliest stages of normal plant development is still questionable and needs further investigation. 51 Conclusion The ends of the linear chromosomes, called telomeres, are shields that protect the exposed chromosome ends from DNA damage machinery. Due to their significance in cell viability, cancer, and ageing, there has been intensive research on telomeric DNA, telomere-associated proteins, and telomere-related proteins for over four decades. However, the protein interactome associated with plant telomeres and telomerase is not as well-studied as the mammalian telomeric proteome. It is interesting to note that telomeric repeats can also be found dispersed throughout the genome as interstitial telomeric tracts or short telo-boxes. In plants, telomeres are primarily composed of short tandem repeats that are associated with various proteins involved in regulating telomere maintenance and the telomerase holoenzyme complex's access. Telomere Repeat Binding proteins (TRBs) play a crucial role in telomere maintenance, and they are associated not only with terminally located telomeric repeats but also with telo-boxes, which are mainly found in gene promoters. These TRBs can recruit and regulate various protein complexes and significantly influence the chromatin's epigenetic status. In my habilitation thesis, I presented data that expand our understanding of plant telomere biology, telomeric sequence- or telomerase-associated proteins, and their roles in telomere homeostasis maintenance. I discussed the involvement of these proteins in telomerase assembly, recruitment, and activity, as well as their role in regulating and protecting the chromosomes' physical ends' genomic integrity. Additionally, I comment on the involvement of these proteins in non-telomeric functions in epigenetic regulations. 52 Bibliography Adams, S. P., Hartman, T. P. V., Lim, K. Y., Chase, M. W., Bennett, M. D., Leitch, I. J., & Leitch, A. R. (2001). Loss and recovery of Arabidopsis–type telomere repeat sequences 5′–(TTTAGGG)n–3′ in the evolution of a major radiation of flowering plants. Proceedings of the Royal Society of London. Series B: Biological Sciences, 268(1476), 1541– 1546. https://doi.org/10.1098/rspb.2001.1726 Akimcheva, S., Zellinger, B., & Riha, K. (2009). Genome stability in Arabidopsis cells exhibiting alternative lengthening of telomeres. Cytogenetic and Genome Research, 122(3–4), 388–395. https://doi.org/10.1159/000167827 Aksenova, A. Y., & Mirkin, S. M. (2019). At the Beginning of the End and in the Middle of the Beginning: Structure and Maintenance of Telomeric DNA Repeats and Interstitial Telomeric Sequences. Genes, 10(2), Article 2. https://doi.org/10.3390/genes10020118 Amiard, S., Depeiges, A., Allain, E., White, C. I., & Gallego, M. E. (2011). Arabidopsis ATM and ATR Kinases Prevent Propagation of Genome Damage Caused by Telomere Dysfunction. The Plant Cell, 23(12), 4254–4265. https://doi.org/10.1105/tpc.111.092387 Antonsson, C., Whitelaw, M. L., McGuire, J., Gustafsson, J.-Å., & Poellinger, L. (1995). Distinct Roles of the Molecular Chaperone hsp90 in Modulating Dioxin Receptor Function via the Basic Helix-Loop-Helix and PAS Domains. Molecular and Cellular Biology, 15(2), 756–765. https://doi.org/10.1128/MCB.15.2.756 Arat, N. Ö., & Griffith, J. D. (2012). Human Rap1 Interacts Directly with Telomeric DNA and Regulates TRF2 Localization at the Telomere*. Journal of Biological Chemistry, 287(50), 41583–41594. https://doi.org/10.1074/jbc.M112.415984 Aronen, T., & Ryynänen, L. (2014). Silver birch telomeres shorten in tissue culture. Tree Genetics & Genomes, 10(1), 67–74. https://doi.org/10.1007/s11295-013-0662-4 Arora, A., Beilstein, M. A., & Shippen, D. E. (2016). Evolution of Arabidopsis protection of telomeres 1 alters nucleic acid recognition and telomerase regulation. Nucleic Acids Research, 44(20), 9821–9830. https://doi.org/10.1093/nar/gkw807 Azzalin, C. M., Nergadze, S. G., & Giulotto, E. (2001). Human intrachromosomal telomeric-like repeats: Sequence organization and mechanisms of origin. Chromosoma, 110(2), 75–82. https://doi.org/10.1007/s004120100135 Baumann, P., & Cech, T. R. (2001). Pot1, the Putative Telomere End-Binding Protein in Fission Yeast and Humans. Science, 292(5519), 1171–1175. https://doi.org/10.1126/science.1060036 Bergemann, A. D., & Johnson, E. M. (1992). The HeLa Pur factor binds single-stranded DNA at a specific element conserved in gene flanking regions and origins of DNA replication. Molecular and Cellular Biology, 12(3), 1257– 1265. https://doi.org/10.1128/mcb.12.3.1257-1265.1992 Bianchi, A., Smith, S., Chong, L., Elias, P., & de Lange, T. (1997). TRF1 is a dimer and bends telomeric DNA. The EMBO Journal, 16(7), 1785–1794. https://doi.org/10.1093/emboj/16.7.1785 Bilaud, T., Koering, C. E., Binet-Brasselet, E., Ancelin, K., Pollice, A., Gasser, S. M., & Gilson, E. (1996). The Telobox, a Myb-Related Telomeric DNA Binding Motif Found in Proteins from Yeast, Plants and Human. Nucleic Acids Research, 24(7), 1294–1303. https://doi.org/10.1093/nar/24.7.1294 Blackburn, E. H., Epel, E. S., & Lin, J. (2015). Human telomere biology: A contributory and interactive factor in aging, disease risks, and protection. Science, 350(6265), 1193–1198. https://doi.org/10.1126/science.aab3389 Blackburn, E. H., & Gall, J. G. (1978). A tandemly repeated sequence at the termini of the extrachromosomal ribosomal RNA genes in Tetrahymena. Journal of Molecular Biology, 120(1), 33–53. https://doi.org/10.1016/0022- 2836(78)90294-2 Boltz, K. A., Leehy, K., Song, X., Nelson, A. D., & Shippen, D. E. (2012). ATR cooperates with CTC1 and STN1 to maintain telomeres and genome integrity in Arabidopsis. Molecular Biology of the Cell, 23(8), 1558–1568. https://doi.org/10.1091/mbc.e11-12-1002 Broccoli, D., Smogorzewska, A., Chong, L., & de Lange, T. (1997). Human telomeres contain two distinct Myb–related proteins, TRF1 and TRF2. Nature Genetics, 17(2), Article 2. https://doi.org/10.1038/ng1097-231 Bundock, P., van Attikum, H., & Hooykaas, P. (2002). Increased telomere length and hypersensitivity to DNA damaging agents in an Arabidopsis KU70 mutant. Nucleic Acids Research, 30(15), 3395–3400. https://doi.org/10.1093/nar/gkf445 Burr, B., Burr, F. A., Matz, E. C., & Romero-Severson, J. (1992). Pinning down loose ends: Mapping telomeres and factors affecting their length. The Plant Cell, 4(8), 953–960. https://doi.org/10.1105/tpc.4.8.953 Byun, M. Y., Hong, J.-P., & Kim, W. T. (2008). Identification and characterization of three telomere repeat-binding factors in rice. Biochemical and Biophysical Research Communications, 372(1), 85–90. https://doi.org/10.1016/j.bbrc.2008.04.181 53 Celli, G. B., & de Lange, T. (2005). DNA processing is not required for ATM-mediated telomere damage response after TRF2 deletion. Nature Cell Biology, 7(7), Article 7. https://doi.org/10.1038/ncb1275 Cifuentes-Rojas, C., Kannan, K., Tseng, L., & Shippen, D. E. (2011). Two RNA subunits and POT1a are components of Arabidopsis telomerase. Proceedings of the National Academy of Sciences, 108(1), 73–78. https://doi.org/10.1073/pnas.1013021107 Cifuentes-Rojas, C., Nelson, A. D. L., Boltz, K. A., Kannan, K., She, X., & Shippen, D. E. (2012). An alternative telomerase RNA in Arabidopsis modulates enzyme activity in response to DNA damage. Genes & Development, 26(22), 2512– 2523. https://doi.org/10.1101/gad.202960.112 Cimino-Reale, G., Pascale, E., Battiloro, E., Starace, G., Verna, R., & D’Ambrosio, E. (2001). The length of telomeric Grich strand 3′-overhang measured by oligonucleotide ligation assay. Nucleic Acids Research, 29(7), e35. Clark, J. W., & Donoghue, P. C. J. (2018). Whole-Genome Duplication and Plant Macroevolution. Trends in Plant Science, 23(10), 933–945. https://doi.org/10.1016/j.tplants.2018.07.006 Colgin, L. M., Wilkinso, C., Englezou, A., Kilian, A., Robinson, M. O., & Reddel, R. R. (2000). The hTERTα Splice Variant is a Dominant Negative Inhibitor of Telomerase Activity. Neoplasia, 2(5), 426–432. https://doi.org/10.1038/sj.neo.7900112 Comadira, G., Rasool, B., Kaprinska, B., García, B. M., Morris, J., Verrall, S. R., Bayer, M., Hedley, P. E., Hancock, R. D., & Foyer, C. H. (2015). WHIRLY1 Functions in the Control of Responses to Nitrogen Deficiency But Not Aphid Infestation in Barley. Plant Physiology, 168(3), 1140–1151. https://doi.org/10.1104/pp.15.00580 Corpet, A., Kleijwegt, C., Roubille, S., Juillard, F., Jacquet, K., Texier, P., & Lomonte, P. (2020). PML nuclear bodies and chromatin dynamics: Catch me if you can! Nucleic Acids Research, 48(21), 11890–11912. https://doi.org/10.1093/nar/gkaa828 Crhák, T., Zachová, D., Fojtová, M., & Sýkorová, E. (2019). The region upstream of the telomerase reverse transcriptase gene is essential for in planta telomerase complementation. Plant Science, 281, 41–51. https://doi.org/10.1016/j.plantsci.2019.01.001 Déjardin, J., & Kingston, R. E. (2009). Purification of Proteins Associated with Specific Genomic Loci. Cell, 136(1), 175– 186. https://doi.org/10.1016/j.cell.2008.11.045 Denchi, E. L., & de Lange, T. (2007). Protection of telomeres through independent control of ATM and ATR by TRF2 and POT1. Nature, 448(7157), Article 7157. https://doi.org/10.1038/nature06065 Derboven, E., Ekker, H., Kusenda, B., Bulankova, P., & Riha, K. (2014). Role of STN1 and DNA Polymerase α in Telomere Stability and Genome-Wide Replication in Arabidopsis. PLOS Genetics, 10(10), e1004682. https://doi.org/10.1371/journal.pgen.1004682 Desveaux, D., Allard, J., Brisson, N., & Sygusch, J. (2002). A new family of plant transcription factors displays a novel ssDNA-binding surface. Nature Structural Biology, 9(7), Article 7. https://doi.org/10.1038/nsb814 Desveaux, D., Després, C., Joyeux, A., Subramaniam, R., & Brisson, N. (2000). PBF-2 Is a Novel Single-Stranded DNA Binding Factor Implicated in PR-10a Gene Activation in Potato. The Plant Cell, 12(8), 1477–1489. https://doi.org/10.1105/tpc.12.8.1477 Dittmar, K. D., Demady, D. R., Stancato, L. F., Krishna, P., & Pratt, W. B. (1997). Folding of the Glucocorticoid Receptor by the Heat Shock Protein (hsp) 90-based Chaperone Machinery: THE ROLE OF p23 IS TO STABILIZE RECEPTOR·hsp90 HETEROCOMPLEXES FORMED BY hsp90·p60·hsp70*. Journal of Biological Chemistry, 272(34), 21213–21220. https://doi.org/10.1074/jbc.272.34.21213 Dokládal, L., Benková, E., Honys, D., Dupľáková, N., Lee, L.-Y., Gelvin, S. B., & Sýkorová, E. (2018). An armadillo-domain protein participates in a telomerase interaction network. Plant Molecular Biology, 97(4), 407–420. https://doi.org/10.1007/s11103-018-0747-4 Dokládal, L., Honys, D., Rana, R., Lee, L.-Y., Gelvin, S. B., & Sýkorová, E. (2015). CDNA Library Screening Identifies Protein Interactors Potentially Involved in Non-Telomeric Roles of Arabidopsis Telomerase. Frontiers in Plant Science, 6. https://www.frontiersin.org/articles/10.3389/fpls.2015.00985 Dreissig, S., Schiml, S., Schindele, P., Weiss, O., Rutten, T., Schubert, V., Gladilin, E., Mette, M. F., Puchta, H., & Houben, A. (2017). Live‐cell CRISPR imaging in plants reveals dynamic telomere movements. The Plant Journal, 91(4), 565– 573. https://doi.org/10.1111/tpj.13601 Du, H., Wang, Y.-B., Xie, Y., Liang, Z., Jiang, S.-J., Zhang, S.-S., Huang, Y.-B., & Tang, Y.-X. (2013). Genome-Wide Identification and Evolutionary and Expression Analyses of MYB-Related Genes in Land Plants. DNA Research, 20(5), 437–448. https://doi.org/10.1093/dnares/dst021 Dunham, M. A., Neumann, A. A., Fasching, C. L., & Reddel, R. R. (2000). Telomere maintenance by recombination in human cells. Nature Genetics, 26(4), 447–450. https://doi.org/10.1038/82586 Dvořáčková, M., Rossignol, P., Shaw, P. J., Koroleva, O. A., Doonan, J. H., & Fajkus, J. (2010a). AtTRB1, a telomeric DNAbinding protein from Arabidopsis, is concentrated in the nucleolus and shows highly dynamic association with chromatin. The Plant Journal, 61(4), 637–649. https://doi.org/10.1111/j.1365-313X.2009.04094.x 54 Dvořáčková, M., Rossignol, P., Shaw, P. J., Koroleva, O. A., Doonan, J. H., & Fajkus, J. (2010b). AtTRB1, a telomeric DNAbinding protein from Arabidopsis, is concentrated in the nucleolus and shows highly dynamic association with chromatin. The Plant Journal, 61(4), 637–649. https://doi.org/10.1111/j.1365-313X.2009.04094.x Ellen, T. P., & van Holde, K. E. (2004). Linker Histone Interaction Shows Divalent Character with both Supercoiled and Linear DNA. Biochemistry, 43(24), 7867–7872. https://doi.org/10.1021/bi0497704 Fairall, L., Chapman, L., Moss, H., de Lange, T., & Rhodes, D. (2001). Structure of the TRFH Dimerization Domain of the Human Telomeric Proteins TRF1 and TRF2. Molecular Cell, 8(2), 351–361. https://doi.org/10.1016/S1097- 2765(01)00321-5 Fajkus, J., Fulnečková, J., Hulánová, M., Berková, K., Říha, K., & Matyášek, R. (1998). Plant cells express telomerase activity upon transfer to callus culture, without extensively changing telomere lengths. Molecular and General Genetics MGG, 260(5), 470–474. https://doi.org/10.1007/s004380050918 Fajkus, J., Kovařík, A., & Královics, R. (1996). Telomerase activity in plant cells. FEBS Letters, 391(3), 307–309. https://doi.org/10.1016/0014-5793(96)00757-0 Fajkus, J., Kovařík, A., mKrálovics, R., & Bezděk, M. (1995). Organization of telomeric and subtelomeric chromatin in the higher plant Nicotiana tabacum. Molecular and General Genetics MGG, 247(5), 633–638. https://doi.org/10.1007/BF00290355 Fajkus, J., Sýkorová, E., & Leitch, A. R. (2005). Telomeres in evolution and evolution of telomeres. Chromosome Research, 13(5), 469–479. https://doi.org/10.1007/s10577-005-0997-2 Fajkus, P., Adámik, M., Nelson, A. D. L., Kilar, A. M., Franek, M., Bubeník, M., Frydrychová, R. Č., Votavová, A., Sýkorová, E., Fajkus, J., & Peška, V. (2023). Telomerase RNA in Hymenoptera (Insecta) switched to plant/ciliate-like biogenesis. Nucleic Acids Research, 51(1), 420–433. https://doi.org/10.1093/nar/gkac1202 Fajkus, P., Peška, V., Sitová, Z., Fulnečková, J., Dvořáčková, M., Gogela, R., Sýkorová, E., Hapala, J., & Fajkus, J. (2016). Allium telomeres unmasked: The unusual telomeric sequence (CTCGGTTATGGG)n is synthesized by telomerase. The Plant Journal, 85(3), 337–347. https://doi.org/10.1111/tpj.13115 Fajkus, P., Peška, V., Závodník, M., Fojtová, M., Fulnečková, J., Dobias, Š., Kilar, A., Dvořáčková, M., Zachová, D., Nečasová, I., Sims, J., Sýkorová, E., & Fajkus, J. (2019). Telomerase RNAs in land plants. Nucleic Acids Research, 47(18), 9842–9856. https://doi.org/10.1093/nar/gkz695 Fan, Y., Linardopoulou, E., Friedman, C., Williams, E., & Trask, B. J. (2002). Genomic Structure and Evolution of the Ancestral Chromosome Fusion Site in 2q13–2q14.1 and Paralogous Regions on Other Human Chromosomes. Genome Research, 12(11), 1651–1662. https://doi.org/10.1101/gr.337602 Feldbrügge, M., Sprenger, M., Hahlbrock, K., & Weisshaar, B. (1997). PcMYB1, a novel plant protein containing a DNAbinding domain with one MYB repeat, interacts in vivo with a light-regulatory promoter unit. The Plant Journal, 11(5), 1079–1093. https://doi.org/10.1046/j.1365-313X.1997.11051079.x Fell, V. L., & Schild-Poulter, C. (2015). The Ku heterodimer: Function in DNA repair and beyond. Mutation Research/Reviews in Mutation Research, 763, 15–29. https://doi.org/10.1016/j.mrrev.2014.06.002 Fitzgerald, M. S., McKnight, T. D., & Shippen, D. E. (1996). Characterization and developmental patterns of telomerase expression in plants. Proceedings of the National Academy of Sciences, 93(25), 14422–14427. https://doi.org/10.1073/pnas.93.25.14422 Fitzgerald, M. S., Riha, K., Gao, F., Ren, S., McKnight, T. D., & Shippen, D. E. (1999). Disruption of the telomerase catalytic subunit gene from Arabidopsis inactivates telomerase and leads to a slow loss of telomeric DNA. Proceedings of the National Academy of Sciences, 96(26), 14813–14818. https://doi.org/10.1073/pnas.96.26.14813 Fojtová, M., Peška, V., Dobšáková, Z., Mozgová, I., Fajkus, J., & Sýkorová, E. (2011). Molecular analysis of T-DNA insertion mutants identified putative regulatory elements in the AtTERT gene. Journal of Experimental Botany, 62(15), 5531–5545. https://doi.org/10.1093/jxb/err235 Fojtová, M., Sýkorová, E., Najdekrová, L., Polanská, P., Zachová, D., Vagnerová, R., Angelis, K. J., & Fajkus, J. (2015). Telomere dynamics in the lower plant Physcomitrella patens. Plant Molecular Biology, 87(6), 591–601. https://doi.org/10.1007/s11103-015-0299-9 Forsythe, H. L., Jarvis, J. L., Turner, J. W., Elmore, L. W., & Holt, S. E. (2001). Stable Association of hsp90 and p23, but Not hsp70, with Active Human Telomerase*. Journal of Biological Chemistry, 276(19), 15571–15574. https://doi.org/10.1074/jbc.C100055200 Foyer, C. H., Karpinska, B., & Krupinska, K. (2014). The functions of WHIRLY1 and REDOX-RESPONSIVE TRANSCRIPTION FACTOR 1 in cross tolerance responses in plants: A hypothesis. Philosophical Transactions of the Royal Society B: Biological Sciences, 369(1640), 20130226. https://doi.org/10.1098/rstb.2013.0226 Freeling, M. (2009). Bias in Plant Gene Content Following Different Sorts of Duplication: Tandem, Whole-Genome, Segmental, or by Transposition. Annual Review of Plant Biology, 60(1), 433–453. https://doi.org/10.1146/annurev.arplant.043008.092122 55 Frydrychová, R., & Marec, F. (2002). Repeated Losses of TTAGG Telomere Repeats in Evolution of Beetles (Coleoptera). Genetica, 115(2), 179–187. https://doi.org/10.1023/A:1020175912128 Fu, D., & Collins, K. (2007). Purification of Human Telomerase Complexes Identifies Factors Involved in Telomerase Biogenesis and Telomere Length Regulation. Molecular Cell, 28(5), 773–785. https://doi.org/10.1016/j.molcel.2007.09.023 Fulcher, N., & Riha, K. (2016). Using Centromere Mediated Genome Elimination to Elucidate the Functional Redundancy of Candidate Telomere Binding Proteins in Arabidopsis thaliana. Frontiers in Genetics, 6, 349. https://doi.org/10.3389/fgene.2015.00349 Fulnecková, J., & Fajkus, J. (2000). Inhibition of plant telomerase by telomere-binding proteins from nuclei of telomerase-negative tissues. FEBS Letters, 467(2–3), 305–310. https://doi.org/10.1016/s0014-5793(00)01178-9 Fulnecková, J., Sevcíková, T., Fajkus, J., Lukesová, A., Lukes, M., Vlcek, C., Lang, B. F., Kim, E., Eliás, M., & Sykorová, E. (2013). A broad phylogenetic survey unveils the diversity and evolution of telomeres in eukaryotes. Genome Biology and Evolution, 5(3), 468–483. https://doi.org/10.1093/gbe/evt019 Fulnečková, J., Dokládal, L., Kolářová, K., Nešpor Dadejová, M., Procházková, K., Gomelská, S., Sivčák, M., Adamusová, K., Lyčka, M., Peska, V., Dvořáčková, M., & Sýkorová, E. (2022). Telomerase Interaction Partners–Insight from Plants. International Journal of Molecular Sciences, 23(1), Article 1. https://doi.org/10.3390/ijms23010368 Gallego, M. E., Jalut, N., & White, C. I. (2003). Telomerase Dependence of Telomere Lengthening in ku80 Mutant Arabidopsis. The Plant Cell, 15(3), 782–789. https://doi.org/10.1105/tpc.008623 Gallego, M. E., & White, C. I. (2005). DNA repair and recombination functions in Arabidopsis telomere maintenance. Chromosome Research: An International Journal on the Molecular, Supramolecular and Evolutionary Aspects of Chromosome Biology, 13(5), 481–491. https://doi.org/10.1007/s10577-005-0995-4 Garvik, B., Carson, M., & Hartwell, L. (1995). Single-Stranded DNA Arising at Telomeres in cdc13 Mutants May Constitute a Specific Signal for the RAD9 Checkpoint. Molecular and Cellular Biology, 15(11), 6128–6138. https://doi.org/10.1128/MCB.15.11.6128 Gaspin, C., Rami, J.-F., & Lescure, B. (2010). Distribution of short interstitial telomere motifs in two plant genomes: Putative origin and function. BMC Plant Biology, 10(1), 283. https://doi.org/10.1186/1471-2229-10-283 Ghisays, F., Garzia, A., Wang, H., Canasto-Chibuque, C., Hohl, M., Savage, S. A., Tuschl, T., & Petrini, J. H. J. (2021). RTEL1 influences the abundance and localization of TERRA RNA. Nature Communications, 12(1), Article 1. https://doi.org/10.1038/s41467-021-23299-2 Giannone, R. J., McDonald, H. W., Hurst, G. B., Shen, R.-F., Wang, Y., & Liu, Y. (2010). The Protein Network Surrounding the Human Telomere Repeat Binding Factors TRF1, TRF2, and POT1. PLOS ONE, 5(8), e12407. https://doi.org/10.1371/journal.pone.0012407 Giraud-Panis, M.-J., Teixeira, M. T., Géli, V., & Gilson, E. (2010). CST Meets Shelterin to Keep Telomeres in Check. Molecular Cell, 39(5), 665–676. https://doi.org/10.1016/j.molcel.2010.08.024 Gladych, M., Wojtyla, A., & Rubis, B. (2011). Human telomerase expression regulation. Biochemistry and Cell Biology, 89(4), 359–376. https://doi.org/10.1139/o11-037 Goffová, I., Vágnerová, R., Peška, V., Franek, M., Havlová, K., Holá, M., Zachová, D., Fojtová, M., Cuming, A., Kamisugi, Y., Angelis, K. J., & Fajkus, J. (2019). Roles of RAD51 and R℡1 in telomere and rDNA stability in Physcomitrella patens. The Plant Journal, 98(6), 1090–1105. https://doi.org/10.1111/tpj.14304 Gomes, N. M. V., Ryder, O. A., Houck, M. L., Charter, S. J., Walker, W., Forsyth, N. R., Austad, S. N., Venditti, C., Pagel, M., Shay, J. W., & Wright, W. E. (2011). Comparative biology of mammalian telomeres: Hypotheses on ancestral states and the roles of telomeres in longevity determination. Aging Cell, 10(5), 761–768. https://doi.org/10.1111/j.1474-9726.2011.00718.x Grabowski, E., Miao, Y., Mulisch, M., & Krupinska, K. (2008). Single-Stranded DNA-Binding Protein Whirly1 in Barley Leaves Is Located in Plastids and the Nucleus of the Same Cell. Plant Physiology, 147(4), 1800–1804. https://doi.org/10.1104/pp.108.122796 Grandin, N., & Charbonneau, M. (2001). Hsp90 levels affect telomere length in yeast. Molecular Genetics and Genomics: MGG, 265(1), 126–134. https://doi.org/10.1007/s004380000398 Greider, C. W. (1999). Telomeres Do D-Loop–T-Loop. Cell, 97(4), 419–422. https://doi.org/10.1016/S0092- 8674(00)80750-3 Greider, C. W., & Blackburn, E. H. (1985). Identification of a specific telomere terminal transferase activity in tetrahymena extracts. Cell, 43(2, Part 1), 405–413. https://doi.org/10.1016/0092-8674(85)90170-9 Greider, C. W., & Blackburn, E. H. (1989). A telomeric sequence in the RNA of Tetrahymena telomerase required for telomere repeat synthesis. Nature, 337(6205), 331–337. https://doi.org/10.1038/337331a0 Griffith, J. D., Comeau, L., Rosenfield, S., Stansel, R. M., Bianchi, A., Moss, H., & de Lange, T. (1999). Mammalian Telomeres End in a Large Duplex Loop. Cell, 97(4), 503–514. https://doi.org/10.1016/S0092-8674(00)80760-6 56 Grolimund, L., Aeby, E., Hamelin, R., Armand, F., Chiappe, D., Moniatte, M., & Lingner, J. (2013). A quantitative telomeric chromatin isolation protocol identifies different telomeric states. Nature Communications, 4(1), Article 1. https://doi.org/10.1038/ncomms3848 Grossi, S., Puglisi, A., Dmitriev, P. V., Lopes, M., & Shore, D. (2004). Pol12, the B subunit of DNA polymerase α, functions in both telomere capping and length regulation. Genes & Development, 18(9), 992–1006. https://doi.org/10.1101/gad.300004 Guan, Z., Wang, W., Yu, X., Lin, W., & Miao, Y. (2018). Comparative Proteomic Analysis of Coregulation of CIPK14 and WHIRLY1/3 Mediated Pale Yellowing of Leaves in Arabidopsis. International Journal of Molecular Sciences, 19(8), Article 8. https://doi.org/10.3390/ijms19082231 Guo, X., Deng, Y., Lin, Y., Cosme-Blanco, W., Chan, S., He, H., Yuan, G., Brown, E. J., & Chang, S. (2007). Dysfunctional telomeres activate an ATM-ATR-dependent DNA damage response to suppress tumorigenesis. The EMBO Journal, 26(22), 4709–4719. https://doi.org/10.1038/sj.emboj.7601893 Harrington, L., McPhail, T., Mar, V., Zhou, W., Oulton, R., Program, A. E., Bass, M. B., Arruda, I., & Robinson, M. O. (1997). A Mammalian Telomerase-Associated Protein. Science, 275(5302), 973–977. https://doi.org/10.1126/science.275.5302.973 Haussmann, M. F., Winkler, D. W., Huntington, C. E., Nisbet, I. C. T., & Vleck, C. M. (2007). Telomerase activity is maintained throughout the lifespan of long-lived birds. Experimental Gerontology, 42(7), 610–618. https://doi.org/10.1016/j.exger.2007.03.004 He, L., Liu, J., Torres, G. A., Zhang, H., Jiang, J., & Xie, C. (2013). Interstitial telomeric repeats are enriched in the centromeres of chromosomes in Solanum species. Chromosome Research, 21(1), 5–13. https://doi.org/10.1007/s10577-012-9332-x He, Q., Chen, L., Xu, Y., & Yu, W. (2013). Identification of centromeric and telomeric DNA-binding proteins in rice. Proteomics, 13(5), 826–832. https://doi.org/10.1002/pmic.201100416 Heacock, M. L., Idol, R. A., Friesner, J. D., Britt, A. B., & Shippen, D. E. (2007). Telomere dynamics and fusion of critically shortened telomeres in plants lacking DNA ligase IV. Nucleic Acids Research, 35(19), 6490–6500. https://doi.org/10.1093/nar/gkm472 Heaphy, C. M., Subhawong, A. P., Hong, S.-M., Goggins, M. G., Montgomery, E. A., Gabrielson, E., Netto, G. J., Epstein, J. I., Lotan, T. L., Westra, W. H., Shih, I.-M., Iacobuzio-Donahue, C. A., Maitra, A., Li, Q. K., Eberhart, C. G., Taube, J. M., Rakheja, D., Kurman, R. J., Wu, T. C., … Meeker, A. K. (2011). Prevalence of the Alternative Lengthening of Telomeres Telomere Maintenance Mechanism in Human Cancer Subtypes. The American Journal of Pathology, 179(4), 1608–1615. https://doi.org/10.1016/j.ajpath.2011.06.018 Heller, K., Kilian, A., Piatyszek, M. A., & Kleinhofs, A. (1996). Telomerase activity in plant extracts. Molecular & General Genetics: MGG, 252(3), 342–345. https://doi.org/10.1007/BF02173780 Hirata, Y., Suzuki, C., & Sakai, S. (2004). Characterization and gene cloning of telomere-binding protein from tobacco BY-2 cells. Plant Physiology and Biochemistry, 42(1), 7–14. https://doi.org/10.1016/j.plaphy.2003.10.002 Ho Lee, J., Hyun Kim, J., Taek Kim, W., Kang, B. G., & Kwon Chung, I. (2000). Characterization and developmental expression of single-stranded telomeric DNA-binding proteins from mung bean (Vigna radiata). Plant Molecular Biology, 42(4), 547–557. https://doi.org/10.1023/A:1006373917321 Hofr, C., Šultesová, P., Zimmermann, M., Mozgová, I., Procházková Schrumpfová, P., Wimmerová, M., & Fajkus, J. (2009). Single-Myb-histone proteins from Arabidopsis thaliana: A quantitative study of telomere-binding specificity and kinetics. Biochemical Journal, 419(1), 221–230. https://doi.org/10.1042/BJ20082195 Hohenstatt, M. L., Mikulski, P., Komarynets, O., Klose, C., Kycia, I., Jeltsch, A., Farrona, S., & Schubert, D. (2018). PWWPDOMAIN INTERACTOR OF POLYCOMBS1 Interacts with Polycomb-Group Proteins and Histones and Regulates Arabidopsis Flowering and Development. The Plant Cell, 30(1), 117–133. https://doi.org/10.1105/tpc.17.00117 Holá, M., Kozák, J., Vágnerová, R., & Angelis, K. J. (2013). Genotoxin Induced Mutagenesis in the Model Plant Physcomitrella patens. BioMed Research International, 2013, e535049. https://doi.org/10.1155/2013/535049 Holt, S. E., Aisner, D. L., Baur, J., Tesmer, V. M., Dy, M., Ouellette, M., Trager, J. B., Morin, G. B., Toft, D. O., Shay, J. W., Wright, W. E., & White, M. A. (1999). Functional requirement of p23 and Hsp90 in telomerase complexes. Genes & Development, 13(7), 817–826. Hom, R. A., & Wuttke, D. S. (2017). Human CST Prefers G-Rich but Not Necessarily Telomeric Sequences. Biochemistry, 56(32), 4210–4218. https://doi.org/10.1021/acs.biochem.7b00584 Hong, J.-P., Byun, M. Y., An, K., Yang, S.-J., An, G., & Kim, W. T. (2010). OsKu70 Is Associated with Developmental Growth and Genome Stability in Rice. Plant Physiology, 152(1), 374–387. https://doi.org/10.1104/pp.109.150391 Hong, J.-P., Byun, M. Y., Koo, D.-H., An, K., Bang, J.-W., Chung, I. K., An, G., & Kim, W. T. (2007). Suppression of RICE TELOMERE BINDING PROTEIN1 Results in Severe and Gradual Developmental Defects Accompanied by Genome Instability in Rice. The Plant Cell, 19(6), 1770–1781. https://doi.org/10.1105/tpc.107.051953 57 Hrdličková, R., Nehyba, J., Liss, A. S., & Bose, H. R. (2006). Mechanism of Telomerase Activation by v-Rel and Its Contribution to Transformation. Journal of Virology, 80(1), 281–295. https://doi.org/10.1128/JVI.80.1.281- 295.2006 Hutchison, K. A., Stancato, L. F., Owens-Grillo, J. K., Johnson, J. L., Krishna, P., Toft, D. O., & Pratt, W. B. (1995). The 23kDa Acidic Protein in Reticulocyte Lysate Is the Weakly Bound Component of the hsp Foldosome That Is Required for Assembly of the Glucocorticoid Receptor into a Functional Heterocomplex with hsp90 (∗). Journal of Biological Chemistry, 270(32), 18841–18847. https://doi.org/10.1074/jbc.270.32.18841 Hwang, M. G., Chung, I. K., Kang, B. G., & Cho, M. H. (2001). Sequence-specific binding property of Arabidopsis thaliana telomeric DNA binding protein 1 (AtTBP1). FEBS Letters, 503(1), 35–40. https://doi.org/10.1016/S0014- 5793(01)02685-0 Chai, W., Ford, L. P., Lenertz, L., Wright, W. E., & Shay, J. W. (2002). Human Ku70/80 Associates Physically with Telomerase through Interaction with hTERT*. Journal of Biological Chemistry, 277(49), 47242–47247. https://doi.org/10.1074/jbc.M208542200 Chan, H., Wang, Y., & Feigon, J. (2017). Progress in Human and Tetrahymena Telomerase Structure Determination. Annual Review of Biophysics, 46(1), 199–225. https://doi.org/10.1146/annurev-biophys-062215-011140 Charbonnel, C., Rymarenko, O., Da Ines, O., Benyahya, F., White, C. I., Butter, F., & Amiard, S. (2018). The Linker Histone GH1-HMGA1 Is Involved in Telomere Stability and DNA Damage Repair. Plant Physiology, 177(1), 311–327. https://doi.org/10.1104/pp.17.01789 Chen, C. M., Wang, C. T., & Ho, C. H. (2001). A Plant Gene Encoding a Myb-like Protein That Binds Telomeric GGTTTAG Repeats in Vitro *. Journal of Biological Chemistry, 276(19), 16511–16519. https://doi.org/10.1074/jbc.M009659200 Chen, C.-M., Wang, C.-T., Kao, Y.-H., Chang, G.-D., Ho, C.-H., Lee, F., Hseu, M.-J. (2005). Functional redundancy of the duplex telomeric DNA-binding proteins in Arabidopsis. Janoušková, E., Nečasová, I., Pavloušková, J., Zimmermann, M., Hluchý, M., Marini, V., Nováková, M., & Hofr, C. (2015). Human Rap1 modulates TRF2 attraction to telomeric DNA. Nucleic Acids Research, 43(5), 2691–2700. https://doi.org/10.1093/nar/gkv097 Jenner, L. P., Peska, V., Fulnečková, J., & Sýkorová, E. (2022). Telomeres and Their Neighbors. Genes, 13(9), 1663. https://doi.org/10.3390/genes13091663 Jurečková, J. F., Sýkorová, E., Hafidh, S., Honys, D., Fajkus, J., & Fojtová, M. (2017). Tissue-specific expression of telomerase reverse transcriptase gene variants in Nicotiana tabacum. Planta, 245(3), 549–561. https://doi.org/10.1007/s00425-016-2624-1 Kang, S. S., Kwon, T., Kwon, D. Y., & Do, S. I. (1999). Akt Protein Kinase Enhances Human Telomerase Activity through Phosphorylation of Telomerase Reverse Transcriptase Subunit*. Journal of Biological Chemistry, 274(19), 13085– 13090. https://doi.org/10.1074/jbc.274.19.13085 Kannan, K., Nelson, A. D. L., & Shippen, D. E. (2008). Dyskerin Is a Component of the Arabidopsis Telomerase RNP Required for Telomere Maintenance. Molecular and Cellular Biology, 28(7), 2332–2341. https://doi.org/10.1128/MCB.01490-07 Kappei, D., Butter, F., Benda, C., Scheibe, M., Draškovič, I., Stevense, M., Novo, C. L., Basquin, C., Araki, M., Araki, K., Krastev, D. B., Kittler, R., Jessberger, R., Londoño-Vallejo, J. A., Mann, M., & Buchholz, F. (2013). HOT1 is a mammalian direct telomere repeat-binding protein contributing to telomerase recruitment. The EMBO Journal, 32(12), 1681–1701. https://doi.org/10.1038/emboj.2013.105 Karamysheva, Z. N., Surovtseva, Y. V., Vespa, L., Shakirov, E. V., & Shippen, D. E. (2004). A C-terminal Myb extension domain defines a novel family of double-strand telomeric DNA-binding proteins in Arabidopsis. The Journal of Biological Chemistry, 279(46), 47799–47807. https://doi.org/10.1074/jbc.M407938200 Kazda, A., Zellinger, B., Rössler, M., Derboven, E., Kusenda, B., & Riha, K. (2012). Chromosome end protection by bluntended telomeres. Genes & Development, 26(15), 1703–1713. https://doi.org/10.1101/gad.194944.112 Kim, J. H., Kim, W. T., & Chung, I. K. (1998). Rice proteins that bind single-stranded G-rich telomere DNA. Plant Molecular Biology, 36(5), 661–672. https://doi.org/10.1023/A:1005994719175 Ko, S., Jun, S.-H., Bae, H., Byun, J.-S., Han, W., Park, H., Yang, S. W., Park, S.-Y., Jeon, Y. H., Cheong, C., Kim, W. T., Lee, W., & Cho, H.-S. (2008). Structure of the DNA-binding domain of NgTRF1 reveals unique features of plant telomere-binding proteins. Nucleic Acids Research, 36(8), 2739–2755. https://doi.org/10.1093/nar/gkn030 Koonin, E. V. (2010). The origin and early evolution of eukaryotes in the light of phylogenomics. Genome Biology, 11(5), 209. https://doi.org/10.1186/gb-2010-11-5-209 Kotliński, M., Knizewski, L., Muszewska, A., Rutowicz, K., Lirski, M., Schmidt, A., Baroux, C., Ginalski, K., & Jerzmanowski, A. (2017). Phylogeny-Based Systematization of Arabidopsis Proteins with Histone H1 Globular Domain. Plant Physiology, 174(1), 27–34. https://doi.org/10.1104/pp.16.00214 Kováč, L.. Overview: bioenergetics between chemistry, genetics and physics (1987) Curr. Top. Bioenerg. 15, 331–372. 58 Kovařik, A., Fajkus, J., Koukalová, B., & Bezděk, M. (1996). Species-specific evolution of telomeric and rDNA repeats in the tobacco composite genome. Theoretical and Applied Genetics, 92(8), 1108–1111. https://doi.org/10.1007/BF00224057 Krause, K., Kilbienski, I., Mulisch, M., Rödiger, A., Schäfer, A., & Krupinska, K. (2005). DNA-binding proteins of the Whirly family in Arabidopsis thaliana are targeted to the organelles. FEBS Letters, 579(17), 3707–3712. https://doi.org/10.1016/j.febslet.2005.05.059 Kuchař, M., & Fajkus, J. (2004). Interactions of putative telomere-binding proteins in Arabidopsis thaliana: Identification of functional TRF2 homolog in plants. FEBS Letters, 578(3), 311–315. https://doi.org/10.1016/j.febslet.2004.11.021 Kupiec, M. (2014). Biology of telomeres: Lessons from budding yeast. FEMS Microbiology Reviews, 38(2), 144–171. https://doi.org/10.1111/1574-6976.12054 Kusová, A., Steinbachová, L., Přerovská, T., Drábková, L. Z., Paleček, J., Khan, A., Rigóová, G., Gadiou, Z., Jourdain, C., Stricker, T., Schubert, D., Honys, D., & Schrumpfová, P. P. (2023). Completing the TRB family: Newly characterized members show ancient evolutionary origins and distinct localization, yet similar interactions (s. 2022.11.23.517682). Plant Mol Biol . 2023 May;112(1-2):61-83 DOI: 10.1007/s11103-023-01348-2 Kwon, C., & Chung, I. K. (2004). Interaction of an Arabidopsis RNA-binding Protein with Plant Single-stranded Telomeric DNA Modulates Telomerase Activity*. Journal of Biological Chemistry, 279(13), 12812–12818. https://doi.org/10.1074/jbc.M312011200 Lamarche, B., Orazio, N., & Weitzman, M. (2010). The MRN complex in Double-Strand Break Repair and Telomere Maintenance. FEBS letters, 584, 3682–3695. https://doi.org/10.1016/j.febslet.2010.07.029 Lansdorp, P. M., Verwoerd, N. P., van de Rijke, F. M., Dragowska, V., Little, M.-T., Dirks, R. W., Raap, A. K., & Tanke, H. J. (1996). Heterogeneity in Telomere Length of Human Chromosomes. Human Molecular Genetics, 5(5), 685– 691. https://doi.org/10.1093/hmg/5.5.685 Lazzerini-Denchi, E., & Sfeir, A. (2016). Stop pulling my strings—What telomeres taught us about the DNA damage response. Nature Reviews Molecular Cell Biology, 17(6), Article 6. https://doi.org/10.1038/nrm.2016.43 Lee, W. K., & Cho, M. H. (2016). Telomere-binding protein regulates the chromosome ends through the interaction with histone deacetylases in Arabidopsis thaliana. Nucleic Acids Research, 44(10), 4610–4624. https://doi.org/10.1093/nar/gkw067 Lee, Y. W., & Kim, W. T. (2010). Tobacco GTBP1, a Homolog of Human Heterogeneous Nuclear Ribonucleoprotein, Protects Telomeres from Aberrant Homologous Recombination. The Plant Cell, 22(8), 2781–2795. https://doi.org/10.1105/tpc.110.076778 Leehy, K. A., Lee, J. R., Song, X., Renfrew, K. B., & Shippen, D. E. (2013). MERISTEM DISORGANIZATION1 Encodes TEN1, an Essential Telomere Protein That Modulates Telomerase Processivity in Arabidopsis. The Plant Cell, 25(4), 1343–1354. https://doi.org/10.1105/tpc.112.107425 Lei, M., Baumann, P., & Cech, T. R. (2002). Cooperative Binding of Single-Stranded Telomeric DNA by the Pot1 Protein of Schizosaccharomyces pombe. Biochemistry, 41(49), 14560–14568. https://doi.org/10.1021/bi026674z Lermontova, I., Schubert, V., Börnke, F., Macas, J., & Schubert, I. (2007). Arabidopsis CBF5 interacts with the H/ACA snoRNP assembly factor NAF1. Plant Molecular Biology, 65(5), 615–626. https://doi.org/10.1007/s11103-007- 9226-z Li, A. Y.-J., Lin, H. H., Kuo, C.-Y., Shih, H.-M., Wang, C. C. C., Yen, Y., & Ann, D. K. (2011). High-mobility group A2 protein modulates hTERT transcription to promote tumorigenesis. Molecular and Cellular Biology, 31(13), 2605–2617. https://doi.org/10.1128/MCB.05447-11 Liboz, T., Bardet, C., Le Van Thai, A., Axelos, M., & Lescure, B. (1990). The four members of the gene family encoding the Arabidopsis thaliana translation elongation factor EF-1α are actively transcribed. Plant Molecular Biology, 14(1), 107–110. https://doi.org/10.1007/BF00015660 Lildballe, D. L., Pedersen, D. S., Kalamajka, R., Emmersen, J., Houben, A., & Grasser, K. D. (2008). The Expression Level of the Chromatin-Associated HMGB1 Protein Influences Growth, Stress Tolerance, and Transcriptome in Arabidopsis. Journal of Molecular Biology, 384(1), 9–21. https://doi.org/10.1016/j.jmb.2008.09.014 Lim, C. J., Zaug, A. J., Kim, H. J., & Cech, T. R. (2017). Reconstitution of human shelterin complexes reveals unexpected stoichiometry and dual pathways to enhance telomerase processivity. Nature Communications, 8(1), Article 1. https://doi.org/10.1038/s41467-017-01313-w Liu, T., Li, S., Xia, C., & Xu, D. (2022). TERT promoter mutations and methylation for telomerase activation in urothelial carcinomas: New mechanistic insights and clinical significance. Frontiers in Immunology, 13, 1071390. https://doi.org/10.3389/fimmu.2022.1071390 Lu, F., Cui, X., Zhang, S., Liu, C., & Cao, X. (2010). JMJ14 is an H3K4 demethylase regulating flowering time in Arabidopsis. Cell Research, 20(3), Article 3. https://doi.org/10.1038/cr.2010.27 59 Lue, N. F. (2018). Evolving Linear Chromosomes and Telomeres: A C-Strand-Centric View. Trends in Biochemical Sciences, 43(5), 314–326. https://doi.org/10.1016/j.tibs.2018.02.008 Lugert, T., & Werr, W. (1994). A novel DNA-binding domain in theShrunken initiator-binding protein (IBP1). Plant Molecular Biology, 25(3), 493–506. https://doi.org/10.1007/BF00043877 Lundblad, V., & Szostak, J. W. (1989). A mutant with a defect in telomere elongation leads to senescence in yeast. Cell, 57(4), 633–643. https://doi.org/10.1016/0092-8674(89)90132-3 Maillet, G., White, C. I., & Gallego, M. E. (2006). Telomere-length regulation in inter-ecotype crosses of Arabidopsis. Plant Molecular Biology, 62(6), 859–866. https://doi.org/10.1007/s11103-006-9061-7 Majerová, E., Mandáková, T., Vu, G. T. H., Fajkus, J., Lysak, M. A., & Fojtová, M. (2014). Chromatin features of plant telomeric sequences at terminal vs. Internal positions. Frontiers in Plant Science, 5. https://www.frontiersin.org/articles/10.3389/fpls.2014.00593 Majerská, J., Sýkorová, E., & Fajkus, J. (2011). Non-telomeric activities of telomerase. Molecular BioSystems, 7(4), 1013–1023. https://doi.org/10.1039/C0MB00268B Makarov, V. L., Hirose, Y., & Langmore, J. P. (1997). Long G Tails at Both Ends of Human Chromosomes Suggest a C Strand Degradation Mechanism for Telomere Shortening. Cell, 88(5), 657–666. https://doi.org/10.1016/S0092- 8674(00)81908-X Mandáková, T., Joly, S., Krzywinski, M., Mummenhoff, K., & Lysak, M. A. (2010). Fast Diploidization in Close Mesopolyploid Relatives of Arabidopsis. The Plant Cell, 22(7), 2277–2290. https://doi.org/10.1105/tpc.110.074526 Mandáková, T., & Lysak, M. A. (2008). Chromosomal Phylogeny and Karyotype Evolution in x=7 Crucifer Species (Brassicaceae). The Plant Cell, 20(10), 2559–2570. https://doi.org/10.1105/tpc.108.062166 Mao, Y.-Q., & Houry, W. A. (2017). The Role of Pontin and Reptin in Cellular Physiology and Cancer Etiology. Frontiers in Molecular Biosciences, 4. https://www.frontiersin.org/articles/10.3389/fmolb.2017.00058 Marian, C. O., & Bass, H. W. (2005). The Terminal acidic SANT 1 (Tacs1) gene of maize is expressed in tissues containing meristems and encodes an acidic SANT domain similar to some chromatin-remodeling complex proteins. Biochimica et Biophysica Acta (BBA) - Gene Structure and Expression, 1727(2), 81–86. https://doi.org/10.1016/j.bbaexp.2004.12.010 Marian, C. O., Bordoli, S. J., Goltz, M., Santarella, R. A., Jackson, L. P., Danilevskaya, O., Beckstette, M., Meeley, R., & Bass, H. W. (2003). The maize Single myb histone 1 gene, Smh1, belongs to a novel gene family and encodes a protein that binds telomere DNA repeats in vitro. Plant Physiology, 133(3), 1336–1350. https://doi.org/10.1104/pp.103.026856 Mason, J. M., Randall, T. A., & Capkova Frydrychova, R. (2016). Telomerase lost? Chromosoma, 125(1), 65–73. https://doi.org/10.1007/s00412-015-0528-7 Matias, P. M., Gorynia, S., Donner, P., & Carrondo, M. A. (2006). Crystal Structure of the Human AAA+ Protein RuvBL1*. Journal of Biological Chemistry, 281(50), 38918–38929. https://doi.org/10.1074/jbc.M605625200 McClintock, B. (1942). The Fusion of Broken Ends of Chromosomes Following Nuclear Fusion. Proceedings of the National Academy of Sciences of the United States of America, 28(11), 458–463. Mikulski, P., Hohenstatt, M. L., Farrona, S., Smaczniak, C., Stahl, Y., Kalyanikrishna, Kaufmann, K., Angenent, G., & Schubert, D. (2019). The Chromatin-Associated Protein PWO1 Interacts with Plant Nuclear Lamin-like Components to Regulate Nuclear Size. The Plant Cell, 31(5), 1141–1154. https://doi.org/10.1105/tpc.18.00663 Min, J., Wright, W. E., & Shay, J. W. (2017). Alternative lengthening of telomeres can be maintained by preferential elongation of lagging strands. Nucleic Acids Research, 45(5), 2615–2628. https://doi.org/10.1093/nar/gkw1295 Moore, J. M. (2009). Investigating the DNA Binding Properties of the Initiator Binding Protein 2 (IBP2) in Maize (Zea Mays). https://diginole.lib.fsu.edu/islandora/object/fsu%3A180521/ Moriguchi, R., Kanahama, K., & Kanayama, Y. (2006). Characterization and expression analysis of the tomato telomerebinding protein LeTBP1. Plant Science, 171(1), 166–174. https://doi.org/10.1016/j.plantsci.2006.03.010 Moyzis, R. K., Buckingham, J. M., Cram, L. S., Dani, M., Deaven, L. L., Jones, M. D., Meyne, J., Ratliff, R. L., & Wu, J. R. (1988). A highly conserved repetitive DNA sequence, (TTAGGG)n, present at the telomeres of human chromosomes. Proceedings of the National Academy of Sciences, 85(18), 6622–6626. https://doi.org/10.1073/pnas.85.18.6622 Mozgová, I., Schrumpfová, P. P., Hofr, C., & Fajkus, J. (2008). Functional characterization of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. Phytochemistry, 69(9), 1814–1819. https://doi.org/10.1016/j.phytochem.2008.04.001 Muller, H.J. The remaking of chromosomes. Collect. Net. (1938), 13, 181–195. 60 Müller, F., Wicky, C., Spicher, A., & Tobler, H. (1991). New telomere formation after developmentally regulated chromosomal breakage during the process of chromatin diminution in ascaris lumbricoides. Cell, 67(4), 815–822. https://doi.org/10.1016/0092-8674(91)90076-B Natarajan, S., Begum, F., Gim, J., Wark, L., Henderson, D., Davie, J. R., Hombach-Klonisch, S., & Klonisch, T. (2016). High Mobility Group A2 protects cancer cells against telomere dysfunction. Oncotarget, 7(11), 12761–12782. https://doi.org/10.18632/oncotarget.6938 Nečasová, I., Janoušková, E., Klumpler, T., & Hofr, C. (2017). Basic domain of telomere guardian TRF2 reduces D-loop unwinding whereas Rap1 restores it. Nucleic Acids Research, 45(21), 12170–12180. https://doi.org/10.1093/nar/gkx812 Neumann, A. A., Watson, C. M., Noble, J. R., Pickett, H. A., Tam, P. P. L., & Reddel, R. R. (2013). Alternative lengthening of telomeres in normal mammalian somatic cells. Genes & Development, 27(1), 18–23. https://doi.org/10.1101/gad.205062.112 Nguyen, T. H. D. (2021). Structural biology of human telomerase: Progress and prospects. Biochemical Society Transactions, 49(5), 1927–1939. https://doi.org/10.1042/BST20200042 Nguyen, T. H. D., Tam, J., Wu, R. A., Greber, B. J., Toso, D., Nogales, E., & Collins, K. (2018). Cryo-EM structure of substrate-bound human telomerase holoenzyme. Nature, 557(7704), Article 7704. https://doi.org/10.1038/s41586-018-0062-x Nittis, T., Guittat, L., LeDuc, R. D., Dao, B., Duxin, J. P., Rohrs, H., Townsend, R. R., & Stewart, S. A. (2010). Revealing Novel Telomere Proteins Using in Vivo Cross-linking, Tandem Affinity Purification, and Label-free Quantitative LC-FTICR-MS*. Molecular & Cellular Proteomics, 9(6), 1144–1156. https://doi.org/10.1074/mcp.M900490- MCP200 Oganesian, L., & Karlseder, J. (2011). Mammalian 5′ C-Rich Telomeric Overhangs Are a Mark of RecombinationDependent Telomere Maintenance. Molecular Cell, 42(2), 224–236. https://doi.org/10.1016/j.molcel.2011.03.015 Ogita, N., Okushima, Y., Tokizawa, M., Yamamoto, Y. Y., Tanaka, M., Seki, M., Makita, Y., Matsui, M., OkamotoYoshiyama, K., Sakamoto, T., Kurata, T., Hiruma, K., Saijo, Y., Takahashi, N., & Umeda, M. (2018). Identifying the target genes of SUPPRESSOR OF GAMMA RESPONSE 1, a master transcription factor controlling DNA damage response in Arabidopsis. The Plant Journal, 94(3), 439–453. https://doi.org/10.1111/tpj.13866 Ogrocká, A., Sýkorová, E., Fajkus, J., & Fojtová, M. (2012). Developmental silencing of the AtTERT gene is associated with increased H3K27me3 loading and maintenance of its euchromatic environment. Journal of Experimental Botany, 63(11), 4233–4241. https://doi.org/10.1093/jxb/ers107 Oguchi, K., Tamura, K., & Takahashi, H. (2004). Characterization of Oryza sativa telomerase reverse transcriptase and possible role of its phosphorylation in the control of telomerase activity. Gene, 342(1), 57–66. https://doi.org/10.1016/j.gene.2004.07.011 Olovnikov, A. M. (1973). A theory of marginotomy: The incomplete copying of template margin in enzymic synthesis of polynucleotides and biological significance of the phenomenon. Journal of Theoretical Biology, 41(1), 181– 190. https://doi.org/10.1016/0022-5193(73)90198-7 Ozturk, N., Singh, I., Mehta, A., Braun, T., & Barreto, G. (2014). HMGA proteins as modulators of chromatin structure during transcriptional activation. Frontiers in Cell and Developmental Biology, 2. https://www.frontiersin.org/articles/10.3389/fcell.2014.00005 Palm, W., & de Lange, T. (2008). How Shelterin Protects Mammalian Telomeres. Annual Review of Genetics, 42(1), 301–334. https://doi.org/10.1146/annurev.genet.41.110306.130350 Palm, W., Hockezmezer, D., Kibe, T., &cde Lange, T.. (2009) Functional Dissection of Human and Mouse POT1 Proteins. Mol Cell Biol. 29(2): 471–482. doi: 10.1128/MCB.01352-08 Pecinka, A., Procházková Schrumpfová, P., Fischer, L., Dvořák Tomaštíková, E., & Mozgová, I. (2022). The Czech Plant Nucleus Workshop 2021. Biologia Plantarum, 66(1), 39–45. https://doi.org/10.32615/bp.2022.003 Pedersen, D. S., & Grasser, K. D. (2010). The role of chromosomal HMGB proteins in plants. Biochimica et Biophysica Acta (BBA) - Gene Regulatory Mechanisms, 1799(1), 171–174. https://doi.org/10.1016/j.bbagrm.2009.11.004 Pendle, A. F., Clark, G. P., Boon, R., Lewandowska, D., Lam, Y. W., Andersen, J., Mann, M., Lamond, A. I., Brown, J. W. S., & Shaw, P. J. (2005). Proteomic Analysis of the Arabidopsis Nucleolus Suggests Novel Nucleolar Functions. Molecular Biology of the Cell, 16(1), 260–269. https://doi.org/10.1091/mbc.e04-09-0791 Peska, V., Mátl, M., Mandáková, T., Vitales, D., Fajkus, P., Fajkus, J., & Garcia, S. (2020). Human-like telomeres in Zostera marina reveal a mode of transition from the plant to the human telomeric sequences. Journal of Experimental Botany, 71(19), 5786–5793. https://doi.org/10.1093/jxb/eraa293 Peška, V., Fajkus, P., Fojtová, M., Dvořáčková, M., Hapala, J., Dvořáček, V., Polanská, P., Leitch, A. R., Sýkorová, E., & Fajkus, J. (2015). Characterisation of an unusual telomere motif (TTTTTTAGGG)n in the plant Cestrum elegans 61 (Solanaceae), a species with a large genome. The Plant Journal: For Cell and Molecular Biology, 82(4), 644–654. https://doi.org/10.1111/tpj.12839 Peška, V., Schrumpfová, P. P., & Fajkus, J. (2011). Using the telobox to search for plant telomere binding proteins. Current Protein & Peptide Science, 12(2), 75–83. https://doi.org/10.2174/138920311795684968 Peška, V., Sýkorová, E., & Fajkus, J. (2009). Two faces of Solanaceae telomeres: A comparison between Nicotiana and Cestrum telomeres and telomere-binding proteins. Cytogenetic and Genome Research, 122(3–4), 380–387. https://doi.org/10.1159/000167826 Polanská, E., Dobšáková, Z., Dvořáčková, M., Fajkus, J., & Štros, M. (2012). HMGB1 gene knockout in mouse embryonic fibroblasts results in reduced telomerase activity and telomere dysfunction. Chromosoma, 121(4), 419–431. https://doi.org/10.1007/s00412-012-0373-x Pontvianne, F., Abou-Ellail, M., Douet, J., Comella, P., Matia, I., Chandrasekhara, C., DeBures, A., Blevins, T., Cooke, R., Medina, F. J., Tourmente, S., Pikaard, C. S., & Sáez-Vásquez, J. (2010). Nucleolin Is Required for DNA Methylation State and the Expression of rRNA Gene Variants in Arabidopsis thaliana. PLOS Genetics, 6(11), e1001225. https://doi.org/10.1371/journal.pgen.1001225 Pontvianne, F., Carpentier, M.-C., Durut, N., Pavlištová, V., Jaške, K., Schořová, Š., Parrinello, H., Rohmer, M., Pikaard, C. S., Fojtová, M., Fajkus, J., & Sáez-Vásquez, J. (2016). Identification of Nucleolus-Associated Chromatin Domains Reveals a Role for the Nucleolus in 3D Organization of the A. thaliana Genome. Cell Reports, 16(6), 1574–1587. https://doi.org/10.1016/j.celrep.2016.07.016 Price, C., Boltz, K. A., Chaiken, M. F., Stewart, J. A., Beilstein, M. A., & Shippen, D. E. (2010). Evolution of CST function in telomere maintenance. Cell Cycle, 9(16), 3177–3185. https://doi.org/10.4161/cc.9.16.12547 Price, C. M., & Cech, T. R. (1987). Telomeric DNA-protein interactions of Oxytricha macronuclear DNA. Genes & Development, 1(8), 783–793. https://doi.org/10.1101/gad.1.8.783 Qi, H., & Zakian, V. A. (2000). The Saccharomyces telomere-binding protein Cdc13p interacts with both the catalytic subunit of DNA polymerase alpha and the telomerase-associated est1 protein. Genes & Development, 14(14), 1777–1788. Qiao, X., Zhang, S., & Paterson, A. H. (2022). Pervasive genome duplications across the plant tree of life and their links to major evolutionary innovations and transitions. Computational and Structural Biotechnology Journal, 20, 3248–3256. https://doi.org/10.1016/j.csbj.2022.06.026 Raices, M., Verdun, R. E., Compton, S. A., Haggblom, C. I., Griffith, J. D., Dillin, A., & Karlseder, J. (2008). C. elegans Telomeres Contain G-Strand and C-Strand Overhangs that Are Bound by Distinct Proteins. Cell, 132(5), 745–757. https://doi.org/10.1016/j.cell.2007.12.039 Ramjeesingh, M., Huan, L. J., Garami, E., & Bear, C. E. (1999). Novel method for evaluation of the oligomeric structure of membrane proteins. Biochemical Journal, 342(Pt 1), 119–123. Recker, J., Knoll, A., & Puchta, H. (2014). The Arabidopsis thaliana Homolog of the Helicase RTEL1 Plays Multiple Roles in Preserving Genome Stability. The Plant Cell, 26(12), 4889–4902. https://doi.org/10.1105/tpc.114.132472 Reeves, R. (2015). High mobility group (HMG) proteins: Modulators of chromatin structure and DNA repair in mammalian cells. DNA Repair, 36, 122–136. https://doi.org/10.1016/j.dnarep.2015.09.015 Ren, S., Johnston, J. S., Shippen, D. E., & McKnight, T. D. (2004). TELOMERASE ACTIVATOR1 Induces Telomerase Activity and Potentiates Responses to Auxin in Arabidopsis. The Plant Cell, 16(11), 2910–2922. https://doi.org/10.1105/tpc.104.025072 Ren, S., Mandadi, K. K., Boedeker, A. L., Rathore, K. S., & McKnight, T. D. (2007). Regulation of Telomerase in Arabidopsis by BT2, an Apparent Target of TELOMERASE ACTIVATOR1. The Plant Cell, 19(1), 23–31. https://doi.org/10.1105/tpc.106.044321 Rice, C., & Skordalakes, E. (2016). Structure and function of the telomeric CST complex. Computational and Structural Biotechnology Journal, 14, 161–167. https://doi.org/10.1016/j.csbj.2016.04.002 Riha, K., Fajkus, J., Siroky, J., & Vyskot, B. (1998). Developmental Control of Telomere Lengths and Telomerase Activity in Plants. The Plant Cell, 10(10), 1691–1698. https://doi.org/10.1105/tpc.10.10.1691 Riha, K., McKnight, T. D., Fajkus, J., Vyskot, B., & Shippen, D. E. (2000). Analysis of the G-overhang structures on plant telomeres: Evidence for two distinct telomere architectures. The Plant Journal, 23(5), 633–641. https://doi.org/10.1046/j.1365-313x.2000.00831.x Riha, K., McKnight, T. D., Griffing, L. R., & Shippen, D. E. (2001). Living with Genome Instability: Plant Responses to Telomere Dysfunction. Science, 291(5509), 1797–1800. https://doi.org/10.1126/science.1057110 Riha, K., Watson, J. M., Parkey, J., & Shippen, D. E. (2002). Telomere length deregulation and enhanced sensitivity to genotoxic stress in Arabidopsis mutants deficient in Ku70. The EMBO Journal, 21(11), 2819–2826. https://doi.org/10.1093/emboj/21.11.2819 Richards, E. J., & Ausubel, F. M. (1988). Isolation of a higher eukaryotic telomere from Arabidopsis thaliana. Cell, 53(1), 127–136. https://doi.org/10.1016/0092-8674(88)90494-1 62 Rossignol, P., Collier, S., Bush, M., Shaw, P., & Doonan, J. H. (2007). Arabidopsis POT1A interacts with TERT-V(I8), an N-terminal splicing variant of telomerase. Journal of Cell Science, 120(20), 3678–3687. https://doi.org/10.1242/jcs.004119 Roy, A., Dutta, A., Roy, D., Ganguly, P., Ghosh, R., Kar, R. K., Bhunia, A., Mukhobadhyay, J., & Chaudhuri, S. (2016). Deciphering the role of the AT-rich interaction domain and the HMG-box domain of ARID-HMG proteins of Arabidopsis thaliana. Plant Molecular Biology, 92(3), 371–388. https://doi.org/10.1007/s11103-016-0519-y Růcková, E., Friml, J., Procházková Schrumpfová, P., & Fajkus, J. (2008). Role of alternative telomere lengthening unmasked in telomerase knock-out mutant plants. Plant Molecular Biology, 66(6), 637–646. https://doi.org/10.1007/s11103-008-9295-7 Ruiz-Herrera, A., Nergadze, S. G., Santagostino, M., & Giulotto, E. (2009). Telomeric repeats far from the ends: Mechanisms of origin and role in evolution. Cytogenetic and Genome Research, 122(3–4), 219–228. https://doi.org/10.1159/000167807 Safak, M., Gallia, G. L., & Khalili, K. (1999). Reciprocal interaction between two cellular proteins, Puralpha and YB-1, modulates transcriptional activity of JCVCY in glial cells. Molecular and Cellular Biology, 19(4), 2712–2723. https://doi.org/10.1128/MCB.19.4.2712 Sanders, J. L., & Newman, A. B. (2013). Telomere Length in Epidemiology: A Biomarker of Aging, Age-Related Disease, Both, or Neither? Epidemiologic Reviews, 35(1), 112–131. https://doi.org/10.1093/epirev/mxs008 Sauerwald, A., Sandin, S., Cristofari, G., Scheres, S. H. W., Lingner, J., & Rhodes, D. (2013). Structure of active dimeric human telomerase. Nature Structural & Molecular Biology, 20(4), Article 4. https://doi.org/10.1038/nsmb.2530 Seluanov, A., Chen, Z., Hine, C., Sasahara, T. H. C., Ribeiro, A. A. C. M., Catania, K. C., Presgraves, D. C., & Gorbunova, V. (2007). Telomerase activity coevolves with body mass not lifespan. Aging Cell, 6(1), 45–52. https://doi.org/10.1111/j.1474-9726.2006.00262.x Sfeir, A. (2012). Telomeres at a glance. Journal of Cell Science, 125(18), 4173–4178. https://doi.org/10.1242/jcs.106831 Shakirov, E. V., McKnight, T. D., & Shippen, D. E. (2009). POT1-independent single-strand telomeric DNA binding activities in Brassicaceae. The Plant Journal, 58(6), 1004–1015. https://doi.org/10.1111/j.1365- 313X.2009.03837.x Shakirov, E. V., Perroud, P.-F., Nelson, A. D., Cannell, M. E., Quatrano, R. S., & Shippen, D. E. (2010). Protection of Telomeres 1 Is Required for Telomere Integrity in the Moss Physcomitrella patens. The Plant Cell, 22(6), 1838– 1848. https://doi.org/10.1105/tpc.110.075846 Shakirov, E. V., & Shippen, D. E. (2004). Length Regulation and Dynamics of Individual Telomere Tracts in Wild-Type Arabidopsis. The Plant Cell, 16(8), 1959–1967. https://doi.org/10.1105/tpc.104.023093 Shakirov, E. V., Song, X., Joseph, J. A., & Shippen, D. E. (2009). POT1 proteins in green algae and land plants: DNAbinding properties and evidence of co-evolution with telomeric DNA. Nucleic Acids Research, 37(22), 7455–7467. https://doi.org/10.1093/nar/gkp785 Shakirov, E. V., Surovtseva, Y. V., Osbun, N., & Shippen, D. E. (2005). The Arabidopsis Pot1 and Pot2 Proteins Function in Telomere Length Homeostasis and Chromosome End Protection. Molecular and Cellular Biology, 25(17), 7725– 7733. https://doi.org/10.1128/MCB.25.17.7725-7733.2005 Shay, J. W., & Wright, W. E. (2000). Hayflick, his limit, and cellular ageing. Nature Reviews Molecular Cell Biology, 1(1), Article 1. https://doi.org/10.1038/35036093 Shay, J. W., & Wright, W. E. (2019). Telomeres and telomerase: Three decades of progress. Nature Reviews Genetics, 20(5), Article 5. https://doi.org/10.1038/s41576-019-0099-1 Shepelev, N., Dontsova, O., & Rubtsova, M. (2023). Post-Transcriptional and Post-Translational Modifications in Telomerase Biogenesis and Recruitment to Telomeres. International Journal of Molecular Sciences, 24(5), Article 5. https://doi.org/10.3390/ijms24055027 Scheibe, M., Arnoult, N., Kappei, D., Buchholz, F., Decottignies, A., Butter, F., & Mann, M. (2013). Quantitative interaction screen of telomeric repeat-containing RNA reveals novel TERRA regulators. Genome Research, 23(12), 2149–2157. https://doi.org/10.1101/gr.151878.112 Schmidt, J. C., & Cech, T. R. (2015). Human telomerase: Biogenesis, trafficking, recruitment, and activation. Genes & Development, 29(11), 1095–1105. https://doi.org/10.1101/gad.263863.115 Schmidt, J. C., Dalby, A. B., & Cech, T. R. (2014). Identification of human TERT elements necessary for telomerase recruitment to telomeres. eLife, 3, e03563. https://doi.org/10.7554/eLife.03563 Schořová, Š., Fajkus, J., & Schrumpfová, P. P. (2020). Optimized Detection of Protein-Protein and Protein-DNA Interactions, with Particular Application to Plant Telomeres. In R. Hancock (Ed.), The Nucleus (s. 139–167). Springer US. https://doi.org/10.1007/978-1-0716-0763-3_11 63 Schořová, Š., Fajkus, J., Záveská Drábková, L., Honys, D., & Schrumpfová, P. P. (2019). The plant Pontin and Reptin homologues, RuvBL1 and RuvBL2a, colocalize with TERT and TRB proteins in vivo, and participate in telomerase biogenesis. The Plant Journal, 98(2), 195–212. https://doi.org/10.1111/tpj.14306 Schrumpfová, P., Kuchař, M., Miková, G., Skříšovská, L., Kubičárová, T., & Fajkus, J. (2004). Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. Genome, 47(2), 316–324. https://doi.org/10.1139/g03-136 Schrumpfová, P. P., Kuchař, M., Paleček, J., & Fajkus, J. (2008). Mapping of interaction domains of putative telomerebinding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. FEBS Letters, 582(10), 1400–1406. https://doi.org/10.1016/j.febslet.2008.03.034 Schrumpfová, P. P., Fojtová, M., Mokroš, P., Grasser, K. D., & Fajkus, J. (2011). Role of HMGB proteins in chromatin dynamics and telomere maintenance in Arabidopsis thaliana. Current Protein & Peptide Science, 12(2), 105–111. https://doi.org/10.2174/138920311795684922 Schrumpfová, P., Vychodilová, I., Dvořáčková, M., Majerská, J., Dokládal, L., Schořová, Š., & Fajkus, J. (2014). Telomere repeat binding proteins are functional components of Arabidopsis telomeres and interact with telomerase. The Plant Journal, 77(5), 770–781. https://doi.org/10.1111/tpj.12428 Schrumpfová, P., Schořová, Š., & Fajkus, J. (2016a). Telomere- and Telomerase-Associated Proteins and Their Functions in the Plant Cell. Frontiers in Plant Science, 7. https://www.frontiersin.org/articles/10.3389/fpls.2016.00851 Schrumpfová, P. P., Vychodilová, I., Hapala, J., Schořová, Š., Dvořáček, V., & Fajkus, J. (2016b). Telomere binding protein TRB1 is associated with promoters of translation machinery genes in vivo. Plant Molecular Biology, 90(1), 189–206. https://doi.org/10.1007/s11103-015-0409-8 Schrumpfová, P. P., Majerská, J., Dokládal, L., Schořová, Š., Stejskal, K., Obořil, M., Honys, D., Kozáková, L., Polanská, P. S., & Sýkorová, E. (2017). Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo. Protoplasma, 254(4), 1547–1562. https://doi.org/10.1007/s00709-016-1042-3 Schrumpfová, P. P., Majerská, J., Dokládal, L., Schořová, Š., Stejskal, K., Obořil, M., Honys, D., Kozáková, L., Polanská, P. S., & Sýkorová, E. (2018). Correction to: Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo. Protoplasma, 255(2), 715–715. https://doi.org/10.1007/s00709-018-1224- 2 Schrumpfová, P., Fojtová, M., & Fajkus, J. (2019). Telomeres in Plants and Humans: Not So Different, Not So Similar. Cells, 8(1), Article 1. https://doi.org/10.3390/cells8010058 Schrumpfová, P. P., & Fajkus, J. (2020). Composition and Function of Telomerase-A Polymerase Associated with the Origin of Eukaryotes. Biomolecules, 10(10), 1425. https://doi.org/10.3390/biom10101425 Schwacke, R., Fischer, K., Ketelsen, B., Krupinska, K., & Krause, K. (2007). Comparative survey of plastid and mitochondrial targeting properties of transcription factors in Arabidopsis and rice. Molecular Genetics and Genomics, 277(6), 631–646. https://doi.org/10.1007/s00438-007-0214-4 Simonet, T., Zaragosi, L.-E., Philippe, C., Lebrigand, K., Schouteden, C., Augereau, A., Bauwens, S., Ye, J., Santagostino, M., Giulotto, E., Magdinier, F., Horard, B., Barbry, P., Waldmann, R., & Gilson, E. (2011). The human TTAGGG repeat factors 1 and 2 bind to a subset of interstitial telomeric sequences and satellite repeats. Cell Research, 21(7), Article 7. https://doi.org/10.1038/cr.2011.40 Smogorzewska, A., Karlseder, J., Holtgreve-Grez, H., Jauch, A., & de Lange, T. (2002). DNA Ligase IV-Dependent NHEJ of Deprotected Mammalian Telomeres in G1 and G2. Current Biology, 12(19), 1635–1644. https://doi.org/10.1016/S0960-9822(02)01179-X Song, J., Castillo-González, C., Ma, Z., & Shippen, D. E. (2021). Arabidopsis retains vertebrate-type telomerase accessory proteins via a plant-specific assembly. Nucleic Acids Research, 49(16), 9496–9507. https://doi.org/10.1093/nar/gkab699 Surovtseva, Y. V., Shakirov, E. V., Vespa, L., Osbun, N., Song, X., & Shippen, D. E. (2007). Arabidopsis POT1 associates with the telomerase RNP and is required for telomere maintenance. The EMBO Journal, 26(15), 3653–3661. https://doi.org/10.1038/sj.emboj.7601792 Sweetlove, L., & Gutierrez, C. (2019). The journey to the end of the chromosome: Delivering active telomerase to telomeres in plants. The Plant Journal, 98(2), 193–194. https://doi.org/10.1111/tpj.14328 Sýkorová, E., & Fajkus, J. (2009). Structure—Function relationships in telomerase genes. Biology of the Cell, 101(7), 375–406. https://doi.org/10.1042/BC20080205 Sýkorová, E., Fulnečková, J., Mokroš, P., Fajkus, J., Fojtová, M., & Peška, V. (2012). Three TERT genes in Nicotiana tabacum. Chromosome Research, 20(4), 381–394. https://doi.org/10.1007/s10577-012-9282-3 Sýkorová, E., Leitch, A. R., & Fajkus, J. (2006). Asparagales Telomerases which Synthesize the Human Type of Telomeres. Plant Molecular Biology, 60(5), 633–646. https://doi.org/10.1007/s11103-005-5091-9 64 Sýkorová, E., Lim, K. Y., Kunická, Z., Chase, M. w., Bennett, M. D., Fajkus, J., & Leitch, A. R. (2003). Telomere variability in the monocotyledonous plant order Asparagales. Proceedings of the Royal Society of London. Series B: Biological Sciences, 270(1527), 1893–1904. https://doi.org/10.1098/rspb.2003.2446 Tamura, K., Liu, H., & Takahashi, H. (1999). Auxin Induction of Cell Cycle Regulated Activity of Tobacco Telomerase*. Journal of Biological Chemistry, 274(30), 20997–21002. https://doi.org/10.1074/jbc.274.30.20997 Tan, L.-M., Zhang, C.-J., Hou, X.-M., Shao, C.-R., Lu, Y.-J., Zhou, J.-X., Li, Y.-Q., Li, L., Chen, S., & He, X.-J. (2018). The PEAT protein complexes are required for histone deacetylation and heterochromatin silencing. The EMBO Journal, 37(19), e98770. https://doi.org/10.15252/embj.201798770 Tani, A., & Murata, M. (2005). Alternative splicing of Pot1 (Protection of telomere)-like genes in Arabidopsis thaliana. Genes & Genetic Systems, 80(1), 41–48. https://doi.org/10.1266/ggs.80.41 Teano G., Concia L., Wolff L., Carron L., Biocanin I., Adamusová K., Fojtová M., Bourge M., Kramdi A., Colot V., Grossniklaus U., Bowler Ch., Baroux C., Carbone A., Probst A.V., Schrumpfová P.P., Fajkus J., Amiard S., Grob S., Bourbousse C., and Barneche F. 2023. Histone H1 protects telomeric repeats from H3K27me3 invasion in Arabidopsis. Cell Reports, 42(8):112894. doi: 10.1016/j.celrep.2023 To, T. K., Kim, J.-M., Matsui, A., Kurihara, Y., Morosawa, T., Ishida, J., Tanaka, M., Endo, T., Kakutani, T., Toyoda, T., Kimura, H., Yokoyama, S., Shinozaki, K., & Seki, M. (2011). Arabidopsis HDA6 Regulates Locus-Directed Heterochromatin Silencing in Cooperation with MET1. PLOS Genetics, 7(4), e1002055. https://doi.org/10.1371/journal.pgen.1002055 Tomaštíková, E., Yang, F., Mlynárová, K., Hafidh, S., Schořová, Š., Kusová, A., Pernisová, M., Přerovská, T., Klodová, B., Honys, D., Fajkus, J., Pecinka, A., & Schrumpfová, P. P. (b.r.). RUVBL proteins are involved in plant gametophyte development. The Plant Journal, n/a(n/a). https://doi.org/10.1111/tpj.16136 Tomáška, Ĺ., Nosek, J., Sepšiová, R., Červenák, F., Juríková, K., Procházková, K., Neboháčová, M., Willcox, S., & Griffith, J. D. (2018). Commentary: Single-stranded telomere-binding protein employs a dual rheostat for binding affinity and specificity that drives function. Frontiers in Genetics, 9, 742. https://doi.org/10.3389/fgene.2018.00742 Tran, T. D., Cao, H. X., Jovtchev, G., Neumann, P., Novák, P., Fojtová, M., Vu, G. T. H., Macas, J., Fajkus, J., Schubert, I., & Fuchs, J. (2015). Centromere and telomere sequence alterations reflect the rapid genome evolution within the carnivorous plant genus Genlisea. The Plant Journal, 84(6), 1087–1099. https://doi.org/10.1111/tpj.13058 Tremousaygue, D., Manevski, A., Bardet, C., Lescure, N., & Lescure, B. (1999). Plant interstitial telomere motifs participate in the control of gene expression in root meristems. The Plant Journal, 20(5), 553–561. https://doi.org/10.1046/j.1365-313X.1999.00627.x Tsuzuki, M., & Wierzbicki, A. T. (2018). Buried in PEAT—discovery of a new silencing complex with opposing activities. The EMBO Journal, 37(19), e100573. https://doi.org/10.15252/embj.2018100573 Turck, F., Roudier, F., Farrona, S., Martin-Magniette, M.-L., Guillaume, E., Buisine, N., Gagnot, S., Martienssen, R. A., Coupland, G., & Colot, V. (2007). Arabidopsis TFL2/LHP1 Specifically Associates with Genes Marked by Trimethylation of Histone H3 Lysine 27. PLoS Genetics, 3(6), e86. https://doi.org/10.1371/journal.pgen.0030086 Uchida, W., Matsunaga, S., Sugiyama, R., & Kawano, S. (2002). Interstitial telomere-like repeats in the Arabidopsis thaliana genome. Genes & Genetic Systems, 77(1), 63–67. https://doi.org/10.1266/ggs.77.63 Valuchova, S., Fulnecek, J., Prokop, Z., Stolt-Bergner, P., Janouskova, E., Hofr, C., & Riha, K. (2017). Protection of Arabidopsis Blunt-Ended Telomeres Is Mediated by a Physical Association with the Ku Heterodimer. The Plant Cell, 29(6), 1533–1545. https://doi.org/10.1105/tpc.17.00064 Vannier, J.-B., Sarek, G., & Boulton, S. J. (2014). RTEL1: Functions of a disease-associated helicase. Trends in Cell Biology, 24(7), 416–425. https://doi.org/10.1016/j.tcb.2014.01.004 Venteicher, A. S., Meng, Z., Mason, P. J., Veenstra, T. D., & Artandi, S. E. (2008). Identification of ATPases Pontin and Reptin as Telomerase Components Essential for Holoenzyme Assembly. Cell, 132(6), 945–957. https://doi.org/10.1016/j.cell.2008.01.019 Vicari, M. R., Bruschi, D. P., Cabral-de-Mello, D. C., & Nogaroto, V. (2022). Telomere organization and the interstitial telomeric sites involvement in insects and vertebrates chromosome evolution. Genetics and Molecular Biology, 45(3 Suppl 1), e20220071. https://doi.org/10.1590/1678-4685-GMB-2022-0071 Walden, N., German, D. A., Wolf, E. M., Kiefer, M., Rigault, P., Huang, X.-C., Kiefer, C., Schmickl, R., Franzke, A., Neuffer, B., Mummenhoff, K., & Koch, M. A. (2020). Nested whole-genome duplications coincide with diversification and high morphological disparity in Brassicaceae. Nature Communications, 11(1), Article 1. https://doi.org/10.1038/s41467-020-17605-7 Wang, F., Podell, E. R., Zaug, A. J., Yang, Y., Baciu, P., Cech, T. R., & Lei, M. (2007). The POT1–TPP1 telomere complex is a telomerase processivity factor. Nature, 445(7127), Article 7127. https://doi.org/10.1038/nature05454 Wang, M., Zhong, Z., Gallego-Bartolomé, J., Feng, S., Shih, Y.-H., Liu, M., Zhou, J., Richey, J. C., Ng, C., Jami-Alahmadi, Y., Wohlschlegel, J., Wu, K., & Jacobsen, S. E. (2023). Arabidopsis TRB proteins function in H3K4me3 65 demethylation by recruiting JMJ14. Nature Communications, 14(1), Article 1. https://doi.org/10.1038/s41467- 023-37263-9 Wang, Y., Ghosh, G., & Hendrickson, E. A. (2009). Ku86 represses lethal telomere deletion events in human somatic cells. Proceedings of the National Academy of Sciences, 106(30), 12430–12435. https://doi.org/10.1073/pnas.0903362106 Watson JM, Bulankova P, Riha K, Shippen DE, Vyskot B: Telomerase-independent cell survival in Arabidopsis thaliana. Plant J 43:662–674 (2005) Watson, J. M., Platzer, A., Kazda, A., Akimcheva, S., Valuchova, S., Nizhynska, V., Nordborg, M., & Riha, K. (2016). Germline replications and somatic mutation accumulation are independent of vegetative life span in Arabidopsis. Proceedings of the National Academy of Sciences, 113(43), 12226–12231. https://doi.org/10.1073/pnas.1609686113 Webb, C. J., & Zakian, V. A. (2016). Telomerase RNA is more than a DNA template. RNA Biology, 13(8), 683–689. https://doi.org/10.1080/15476286.2016.1191725 Wei, C., & Price, C. M. (2004). Cell Cycle Localization, Dimerization, and Binding Domain Architecture of the Telomere Protein cPot1. Molecular and Cellular Biology, 24(5), 2091–2102. https://doi.org/10.1128/MCB.24.5.2091- 2102.2004 Wellinger, R. J., & Zakian, V. A. (2012). Everything You Ever Wanted to Know About Saccharomyces cerevisiae Telomeres: Beginning to End. Genetics, 191(4), 1073–1105. https://doi.org/10.1534/genetics.111.137851 Wong, M. S., Wright, W. E., & Shay, J. W. (2014). Alternative splicing regulation of telomerase: A new paradigm? Trends in Genetics, 30(10), 430–438. https://doi.org/10.1016/j.tig.2014.07.006 Wood, A. M., Danielsen, J. M. R., Lucas, C. A., Rice, E. L., Scalzo, D., Shimi, T., Goldman, R. D., Smith, E. D., Le Beau, M. M., & Kosak, S. T. (2014). TRF2 and lamin A/C interact to facilitate the functional organization of chromosome ends. Nature Communications, 5(1), Article 1. https://doi.org/10.1038/ncomms6467 Wood, M. A., McMahon, S. B., & Cole, M. D. (2000). An ATPase/helicase complex is an essential cofactor for oncogenic transformation by c-Myc. Molecular Cell, 5(2), 321–330. https://doi.org/10.1016/s1097-2765(00)80427-x Wright, W. E., Tesmer, V. M., Huffman, K. E., Levene, S. D., & Shay, J. W. (1997). Normal human chromosomes have long G-rich telomeric overhangs at one end. Genes & Development, 11(21), 2801–2809. https://doi.org/10.1101/gad.11.21.2801 Wu, J., Liu, C., Liu, Z., Li, S., Li, D., Liu, S., Huang, X., Liu, S., & Yukawa, Y. (2019). Pol III-Dependent Cabbage BoNR8 Long ncRNA Affects Seed Germination and Growth in Arabidopsis. Plant and Cell Physiology, 60(2), 421–435. https://doi.org/10.1093/pcp/pcy220 Wu, J., Okada, T., Fukushima, T., Tsudzuki, T., Sugiura, M., & Yukawa, Y. (2012). A novel hypoxic stress-responsive long non-coding RNA transcribed by RNA polymerase III in Arabidopsis. RNA Biology, 9(3), 302–313. https://doi.org/10.4161/rna.19101 Wu, L., Multani, A. S., He, H., Cosme-Blanco, W., Deng, Y., Deng, J. M., Bachilo, O., Pathak, S., Tahara, H., Bailey, S. M., Deng, Y., Behringer, R. R., & Chang, S. (2006). Pot1 Deficiency Initiates DNA Damage Checkpoint Activation and Aberrant Homologous Recombination at Telomeres. Cell, 126(1), 49–62. https://doi.org/10.1016/j.cell.2006.05.037 Wu, R. A., Upton, H. E., Vogan, J. M., & Collins, K. (2017). Telomerase Mechanism of Telomere Synthesis. Annual Review of Biochemistry, 86(1), 439–460. https://doi.org/10.1146/annurev-biochem-061516-045019 Xi, L., & Cech, T. R. (2014). Inventory of telomerase components in human cells reveals multiple subpopulations of hTR and hTERT. Nucleic Acids Research, 42(13), 8565–8577. https://doi.org/10.1093/nar/gku560 Xin, H., Liu, D., Wan, M., Safari, A., Kim, H., Sun, W., O’Connor, M. S., & Songyang, Z. (2007). TPP1 is a homologue of ciliate TEBP-β and interacts with POT1 to recruit telomerase. Nature, 445(7127), Article 7127. https://doi.org/10.1038/nature05469 Yang, S. W., Jin, E., Chung, I. K., & Kim, W. T. (2002). Cell cycle-dependent regulation of telomerase activity by auxin, abscisic acid and protein phosphorylation in tobacco BY-2 suspension culture cells. The Plant Journal, 29(5), 617– 626. https://doi.org/10.1046/j.0960-7412.2001.01244.x Yang, S. W., Kim, D. H., Lee, J. J., Chun, Y. J., Lee, J.-H., Kim, Y. J., Chung, I. K., & Kim, W. T. (2003). Expression of the Telomeric Repeat Binding Factor Gene NgTRF1 Is Closely Coordinated with the Cell Division Program in Tobacco BY-2 Suspension Culture Cells*. Journal of Biological Chemistry, 278(24), 21395–21407. https://doi.org/10.1074/jbc.M209973200 Yang, S. W., Kim, S. K., & Kim, W. T. (2004). Perturbation of NgTRF1 Expression Induces Apoptosis-Like Cell Death in Tobacco BY-2 Cells and Implicates NgTRF1 in the Control of Telomere Length and Stability. The Plant Cell, 16(12), 3370–3385. https://doi.org/10.1105/tpc.104.026278 Ye, J., Lenain, C., Bauwens, S., Rizzo, A., Saint-Léger, A., Poulet, A., Benarroch, D., Magdinier, F., Morere, J., Amiard, S., Verhoeyen, E., Britton, S., Calsou, P., Salles, B., Bizard, A., Nadal, M., Salvati, E., Sabatier, L., Wu, Y., … Gilson, E. 66 (2010). TRF2 and Apollo Cooperate with Topoisomerase 2α to Protect Human Telomeres from Replicative Damage. Cell, 142(2), 230–242. https://doi.org/10.1016/j.cell.2010.05.032 Yoo, H. H., Kwon, C., Lee, M. M., & Chung, I. K. (2007). Single-stranded DNA binding factor AtWHY1 modulates telomere length homeostasis in Arabidopsis. The Plant Journal, 49(3), 442–451. https://doi.org/10.1111/j.1365- 313X.2006.02974.x Yu, E. Y., Kim, S. E., Kim, J. H., Ko, J. H., Cho, M. H., & Chung, I. K. (2000). Sequence-specific DNA recognition by the Myb-like domain of plant telomeric protein RTBP1. The Journal of Biological Chemistry, 275(31), 24208–24214. https://doi.org/10.1074/jbc.M003250200 Zachová, D., Fojtová, M., Dvořáčková, M., Mozgová, I., Lermontova, I., Peška, V., Schubert, I., Fajkus, J., & Sýkorová, E. (2013). Structure-function relationships during transgenic telomerase expression in Arabidopsis. Physiologia Plantarum, 149(1), 114–126. https://doi.org/10.1111/ppl.12021 Zellinger, B., Akimcheva, S., Puizina, J., Schirato, M., & Riha, K. (2007). Ku Suppresses Formation of Telomeric Circles and Alternative Telomere Lengthening in Arabidopsis. Molecular Cell, 27(1), 163–169. https://doi.org/10.1016/j.molcel.2007.05.025 Zentgraf, U. (1995). Telomere-binding proteins of Arabidopsis thaliana. Plant Molecular Biology, 27(3), 467–475. https://doi.org/10.1007/BF00019314 Zhang, J.-M., Yadav, T., Ouyang, J., Lan, L., & Zou, L. (2019). Alternative Lengthening of Telomeres through Two Distinct Break-Induced Replication Pathways. Cell Reports, 26(4), 955-968.e3. https://doi.org/10.1016/j.celrep.2018.12.102 Zhang, Q., Kim, N.-K., & Feigon, J. (2011). Architecture of human telomerase RNA. Proceedings of the National Academy of Sciences, 108(51), 20325–20332. https://doi.org/10.1073/pnas.1100279108 Zhang, Z., Sullivan, W., Felts, S. J., Prasad, B. D., Toft, D. O., & Krishna, P. (2010). Characterization of plant p23-like proteins for their co-chaperone activities. Cell Stress and Chaperones, 15(5), 703–715. https://doi.org/10.1007/s12192-010-0182-1 Zhou, Y., Hartwig, B., James, G. V., Schneeberger, K., & Turck, F. (2016). Complementary Activities of TELOMERE REPEAT BINDING Proteins and Polycomb Group Complexes in Transcriptional Regulation of Target Genes. The Plant Cell, 28(1), 87–101. https://doi.org/10.1105/tpc.15.00787 Zhou, Y., Wang, Y., Krause, K., Yang, T., Dongus, J. A., Zhang, Y., & Turck, F. (2018). Telobox motifs recruit CLF/SWN– PRC2 for H3K27me3 deposition via TRB factors in Arabidopsis. Nature Genetics, 50(5), Article 5. https://doi.org/10.1038/s41588-018-0109-9 Zhu, Q.-H., Ramm, K., Shivakkumar, R., Dennis, E. S., & Upadhyaya, N. M. (2004). The ANTHER INDEHISCENCE1 Gene Encoding a Single MYB Domain Protein Is Involved in Anther Development in Rice. Plant Physiology, 135(3), 1514– 1525. https://doi.org/10.1104/pp.104.041459 67 Supplements Relevant publications of the applicant arranged in chronological order. All publications referred in Supplements have been included in the thesis (references in bold and marked with letters of alphabet in the text). *Corresponding Author Supplement A Schrumpfová, P., Kuchar, M., Miková, G., Skrísovská, L., Kubicárová, T., Fajkus, J., 2004. Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. Genome 47, 316– 324 Supplement B Růčková, E., Friml, J., Schrumpfová, P.P., Fajkus, J., 2008. Role of alternative telomere lengthening unmasked in telomerase knock-out mutant plants. Plant Mol. Biol. 66, 637–646 Supplement C Schrumpfová, P.P., Kuchar, M., Palecek, J., Fajkus, J., 2008. Mapping of interaction domains of putative telomere-binding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. FEBS Lett. 582, 1400–1406. Supplement D Mozgová, I., Schrumpfová, P.P., Hofr, C., Fajkus, J., 2008. Functional characterization of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. Phytochemistry 69, 1814–1819 Supplement E Hofr, C., Sultesová, P., Zimmermann, M., Mozgová, I., Schrumpfová, P.P., Wimmerová, M., Fajkus, J., 2009. Single-Myb-histone proteins from Arabidopsis thaliana: a quantitative study of telomere-binding specificity and kinetics. Biochem. J. 419, 221–228 Supplement F Peška, V., Schrumpfová, P.P., Fajkus, J., 2011. Using the telobox to search for plant telomere binding proteins. Curr. Protein Pept. Sci. 12, 75–83 Supplement G Schrumpfová, P.P.*, Fojtová, M., Mokroš, P., Grasser, K.D., Fajkus, J., 2011. Role of HMGB proteins in chromatin dynamics and telomere maintenance in Arabidopsis thaliana. Curr. Protein Pept. Sci. 12, 105– 111 Supplement H Schrumpfová, P.P.*, Vychodilová, I., Dvořáčková, M., Majerská, J., Dokládal, L., Schořová, S., Fajkus, J., 2014. Telomere repeat binding proteins are functional components of Arabidopsis telomeres and interact with telomerase. Plant J. Cell Mol. Biol. 77, 770–781 Supplement I 68 Schrumpfová, P.P., Vychodilová, I., Hapala, J., Schořová, Š., Dvořáček, V., Fajkus, J., 2016. Telomere binding protein TRB1 is associated with promoters of translation machinery genes in vivo. Plant Mol.Biol. 90, 189– 206 Supplement J Schrumpfová, P.P.*, Schořová, Š., Fajkus, J., 2016. Telomere- and Telomerase-Associated Proteins and Their Functions in the Plant Cell. Front. Plant Sci. 7:851 Supplement K Schrumpfová, P.P., Majerská, J., Dokládal, L., Schořová, Š., Stejskal, K., Obořil, M., Honys, D., Kozáková, L., Polanská, P.S., Sýkorová, E., 2017. Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo. Protoplasma 254, 1547–1562 Supplement L Schrumpfová, P.P., Majerská, J., Dokládal, L., Schořová, Š., Stejskal, K., Obořil, M., Honys, D., Kozáková, L., Polanská, P.S., Sýkorová, E., 2018. Correction to: Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo. Protoplasma 255, 715 Supplement M Schrumpfová, P.P., Fojtová, M., Fajkus, J., 2019. Telomeres in Plants and Humans: Not So Different, Not So Similar. Cells 8 Supplement N Schořová, Š., Fajkus, J., Drábková, L.Z., Honys, D., Schrumpfová, P.P.*, 2019. The plant Pontin and Reptin homologues, RuvBL1 and RuvBL2a, colocalize with TERT and TRB proteins in vivo, and participate in telomerase biogenesis. Plant J. 98, 195–212 Supplement O Schrumpfová, P.P.* and Fajkus, J. 2020. Composition and Function of Telomerase - a polymerase associated with the origin of eukaryotes. Review. Biomolecules, 10(10):1425 Supplement P Pecinka A, Schrumpfová, P.P., Fischer L., Dvořák Tomaštíková E., and Mozgová I., 2022. The Czech Plant Nucleus Workshop 2021. Biologia Plantarum, 66: 39-45 Supplement Q Dvořák Tomaštíková E., Yang F., Mlynárová K., Hafid S., Schořová Š., Kusová A., Pernisová M., Přerovská M., Klodová B., Honys D., Fajkus J., Pecinka A., and Schrumpfova P.P.*, 2023 RUVBL proteins are involved in plant gametophyte development, The Plant Journal 114, 325–337 Supplement R Kusova A., Steinbachova L., Přerovská T., Záveská Drábková L., Paleček J., Khan A., Rigóová G., Gadoiu Z., Jourdain C., Stricker T., Schubert D., Honys D., and Schrumpfová P.P.*, 2023. Completing the TRB family: newly characterized members show ancient evolutionary origins and distinct localization, yet similar interactions Plant Molecular Biology 112:61–83 Supplement S Teano G., Concia L., Wolff L., Carron L., Biocanin I., Adamusová K., Fojtová M., Bourge M., Kramdi A., Colot V., Grossniklaus U., Bowler Ch., Baroux C., Carbone A., Probst A.V., Schrumpfová P.P., Fajkus J., Amiard S., Grob S., Bourbousse C., and Barneche F. 2023. Histone H1 protects telomeric repeats from H3K27me3 invasion in Arabidopsis. Cell Reports, 42(8):112894. 69 Supplement A Schrumpfová, P., Kuchar, M., Miková, G., Skrísovská, L., Kubicárová, T., Fajkus, J., 2004. Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. Genome 47, 316–324 P.P.S. performed most of the experiments (protein expression, EMSA, telomerase activity detection), evaluated data and participated in the ms writing and editing This journal did not provide open access, hence the article is not freely available. Supplement B Růčková, E., Friml, J., Schrumpfová, P.P., Fajkus, J., 2008. Role of alternative telomere lengthening unmasked in telomerase knock-out mutant plants. Plant Mol. Biol. 66, 637–646 P.P.S. was involved in the experimental part (plant cultivation and genotyping, telomere length measurement) and participated in the ms writing and editing Role of alternative telomere lengthening unmasked in telomerase knock-out mutant plants Eva Ru˚cˇkova´ Æ Jirˇı´ Friml Æ Petra Procha´zkova´ Schrumpfova´ Æ Jirˇı´ Fajkus Received: 23 November 2007 / Accepted: 14 January 2008 Ó Springer Science+Business Media B.V. 2008 Abstract Telomeres in many eukaryotes are maintained by telomerase in whose absence telomere shortening occurs. However, telomerase-deficient Arabidopsis thaliana mutants (Attert-/) show extremely low rates of telomere shortening per plant generation (250–500 bp), which does not correspond to the expected outcome of replicative telomere shortening resulting from ca. 1,000 meristem cell divisions per seed-to-seed generation. To investigate the influence of the number of cell divisions per seed-to-seed generation, Attert-/mutant plants were propagated from seeds coming either from the lower-most or the upper-most siliques (L- and U-plants) and the length of their telomeres were followed over several generations. The rate of telomere shortening was faster in U-plants, than in L-plants, as would be expected from their higher number of cell divisions per generation. However, this trend was observed only in telomeres whose initial length is relatively high and the differences decreased with progressive general telomere shortening over generations. But in generation 4, the L-plants frequently show a net telomere elongation, while the U-plants fail to do so. We propose that this is due to the activation of alternative telomere lengthening (ALT), a process which is activated in early embryonic development in both U- and L-plants, but is overridden in U-plants due to their higher number of cell divisions per generation. These data demonstrate what so far has only been speculated, that in the absence of telomerase, the number of cell divisions within one generation influences the control of telomere lengths. These results also reveal a fast and efficient activation of ALT mechanism(s) in response to the loss of telomerase activity and imply that ALT is probably involved also in normal plant development. Keywords Alternative telomere lengthening Á ALT Á Replicative telomere shortening Á Telomerase-deficient plants Introduction Incomplete replication of chromosome ends results in progressive telomere shortening unless a mechanism to elongate telomeres takes effect. It has been known for more than a decade that the common system of telomere maintenance in plants is provided by telomerase (Fajkus et al. 1996; Heller et al. 1996), although exemptions, in which a different type of telomeres and mechanism of their maintenance are in use have been described since about the same time in Allium (Pich et al. 1996; Sykorova et al. 2006) and later on in three Solanaceae genera (Sykorova et al. 2003). We hypothesized recently that alternative telomere lengthening (ALT) mechanisms are probably not restricted to species possessing ‘‘unusual’’ telomeres, but may be a normal part of plant development (Fajkus et al. 2005), an idea explored here. Arabidopsis thaliana knockout mutants in Telomerase Reverse Transcriptase (AtTERT) exhibit telomere shortening of 250–500 bp per generation E. Ru˚cˇkova´ Á J. Friml Á P. Procha´zkova´ Schrumpfova´ Á J. Fajkus (&) Department of Functional Genomics and Proteomics, Faculty of Science, Masaryk University, Building A2 – ILBIT, Kamenice 5, 625 00 Brno, Czech Republic e-mail: fajkus@sci.muni.cz E. Ru˚cˇkova´ Á J. Fajkus Institute of Biophysics ASCR, v.v.i., Kra´lovopolska´ 135, 61265 Brno, Czech Republic J. Friml VIB Department of Plant Systems Biology, Ghent University, Technologiepark 927, 9052 Gent, Belgium 123 Plant Mol Biol DOI 10.1007/s11103-008-9295-7 (Fitzgerald et al. 1999) and survive up to 10 generations with severe cytological and chromosomal abnormalities occurring after about eight generations (Riha et al. 2001; Siroky et al. 2003). The rate of the observed inter-generation shortening of telomeres is surprisingly low, given the high number of cell divisions per seed-to-seed generation. The high number results from the mode of plant development, which does not involve stem cell mobility and cell lines are not sequestrated for later use (as in the germ line of mammals). In plants an apical meristem consists of a small group of stem cells that generate a linear series of cells, which differentiate into an array of cell types that make a shoot and root. Flowers initiate from the shoot apical meristem in mature plants, which is organized in cell layers L1, L2 and L3, and divisions of those are roughly synchronized. L2 cell layer derivatives provide the mesodermal cells and the germ cells of pollen grains and ovules (Fletcher 2002, Grandjean et al. 2004). Consequently, meristem cells, which give rise to all tissues, including germ-line cells, undergo many divisions, calculated in A. thaliana to be approximately 1,000 divisions from seed to seed (Fajkus et al. 2005). When considering only 5–10 nucleotides lost per cell division (the average length of RNA primer for synthesis of Okazaki fragments) as the minimum plausible loss of telomeric DNA at each round of replication (under the very improbable scenario that the primer sits exactly at the 3’end of the parental DNA strand), then the number of cell divisions accounting for the observed telomere erosion per generation in Attert-/mutants would be only 25–50 cell divisions (as in mammals). But as already stated for A. thaliana the actual number of divisions is closer to 1,000. Thus we proposed that ALT system operates in telomerasedeficient plants and possibly also in normal plant development to partially compensate for replicative shortening of telomeres (Fajkus et al. 2005). To test this hypothesis, we compared telomere shortening in telomerase knock-out plants differing in the number of cell divisions per generation. We generated Attert-/- and Attert+/+ plants from heterozygous A. thaliana (Attert+/- ) line (SALK_061434.56.00.X) bearing T-insertion in Attert gene. In the obtained homozygous Attert-/plants, seeds were collected individually from the lowermost or uppermost siliques and seeds propagated. In subsequent generations, plants coming from the lowermost siliques were propagated again through seeds from the lowermost siliques, while plants coming from the uppermost siliques were again propagated through seeds from uppermost siliques. If our hypothesis is correct, the plant propagation scheme (Fig. 1) should result in additional cell divisions occurring in ‘‘upper silique’’ lines (U) compared with lower silique’’ lines (L). The number of additional cell divisions in U-plants compared with L-plants can be estimated as follows: the cell division rate after floral transition is 1–2 divisions per 24 h. That corresponds to 1–2 flower initiations for the same time, so the difference is 1 cell division between 2 consecutive flower initiations (Grandjean et al. 2004). Since ca. 30 siliques occur between the lowermost and the uppermost silique, we can expect about 30 additional cell divisions. Therefore in Attert-/mutants we might expect about 150–300 bp-shorter telomere lengths in the U-lines compared with L-lines. We show here that differential telomere shortening does occur between U- and L-lines but that the results are complicated through the activation of ALT. We reveal a differential rate of telomere shortening in different generations of mutants and telomere length oscillations, which, at least in Attert-/mutants, cannot be attributed to telomerase activity. Material and methods Plant material Homozygous mutant Attert-/plants (M) and corresponding Attert+/+ controls (Wt) were prepared from heterozygous A. thaliana line (Attert+/, H) bearing T-insertion in Attert gene (SALK_061434.56.00.X). Plants were initially cultivated on short day (8 h of light) and after 6 weeks at long day (16 h of light) in a greenhouse. Lines were derived from three heterozygous plant lineages (Ha, Hb1 and Hb2). Mutant (M) and control plants (Wt) were selected from their progeny and designated accordingly (Ma and Wta from Ha; Mb1, Mb2 and Wtb1 and Wtb2 from Hb). First generation plants (G1) were divided into lines propagated either from the lowermost or the uppermost siliques (L- and U-plants, respectively, see the schematic Fig. 1). These lines were cultured until the fourth generation (G4). Fig. 1 Schematic diagram of plant propagation system for the generation of plants derived from the heterozygote Attert+/- mutant Ha. The same scheme applies to the heterozygous mutants Hb1 and Hb2 and the nomenclature of the derived plants follow accordingly. For further details of the source materials see Materials and methods Plant Mol Biol 123 Genotyping DNA was extracted according to Edwards et al. (1991). A set of three primers was used for genotyping. Two primers were complementary to genomic DNA upstream and downstream of the T-insertion (tel+ and tel-, respectively) and produced a 876 bp product in wild-type. In mutants, a 702 bp product was synthesized using primer LBb1 (complementary to T-insertion) and tel-. Primer sequences were: tel+: 50 -CTg CTACTTTCAgCTTCAgC-30 , tel-: 50 -gCAAgAggATgCA TTgAAgTCCgg-30 , LBb1: 50 -gCgTggACCgCTTgCTgCA ACT-30 . The reaction mix (15 ll) contained 19 buffer (DyNAzyme II, Finnzymes), 0.25 mM dNTPs, 0.3 lM forward and 0.3 lM reverse primer, 0.3 U DyNAzyme II DNA polymerase (Finnzymes) and 5 ng of DNA. Initial denaturation (94°C/3 min) was followed by 35 thermal cycles (94°C/30 s, 56.5°C/30 s and 72°C/1 min), and a final extension (72°C/10 min). Induction of callus cultures Seeds were sterilized by shaking for 15 min in 50 mg ml-1 Ca(OCl)2 solution, then rinsed three times with sterile water with 0.01% Triton X-100 and placed on solid MS medium supplemented with 20 g l-1 sucrose. Plants were genotyped and leaves from mutant, wild-type and heterozygous plants were harvested, cut and cultivated on solid MS medium (Duchefa M0231) supplemented with 1 mg/l 2,4-D (2,4-dichlorophenoxyacetic acid), 0.2 mg/l kinetin and 20 g/l sucrose. Calli were maintained on this medium and transferred to fresh medium every 4 weeks. Detection of telomerase activity Protein extracts from calli were prepared as described in Fitzgerald et al. (1996). These extracts were then tested for telomerase activity by plant telomere repeat amplification protocol (TRAP) as described by Fajkus et al. (1998). Determination of telomere length DNA was extracted from three rosette leaves according to Dellaporta et al. (1983). Primer extension telomere repeat amplification (PETRA) analysis was performed with an equivalent amount of DNA as described in Heacock et al. (2004) and Watson and Shippen (2007). Individual telomeres were designated according to Heacock et al. (2004) with a number identifying a chromosome, and R or L letter indicating a chromosome arm, where R corresponds to South and L to North (Arabidopsis Genome Initiative 2000). To measure telomere length, signals of PETRA products were analyzed by TotalLab using a 1-kb DNA ladder (Fermentas) as standard. The distance of the PETRA primer to the telomere was subtracted from the total length of PETRA product to give the actual length of the telomere tract. The average telomere lengths were visualized using Southern hybridization of terminal restriction fragments (TRF) (Fajkus et al. 1995) produced by digestion with MseI restriction endonuclease. Both PETRA and TRF products were detected using telomeric oligonucleotide (50 -GGTT TAGGGTTTAGGGTTTAGGGTTTAG-30 ) end-labelled with [c-32 P]ATP using polynucleotide kinase (NEB). Results Arabidopsis tert mutants do not possess residual telomerase activity To make sure that A. thaliana Attert-/mutants do not possess any residual telomerase activity, tissue cultures derived from original Attert+/plants and their mutant (Attert-/) and Wt (Attert+/+ ) progeny were assayed for telomerase activity by TRAP assay. The results (Fig. 2) show the absence of telomerase activity in extracts obtained from Attert-/mutant, while both Attert+/- and Attert+/+ cultures are telomerase-positive. Telomere lengthening upon transition from Attert+/(H) to Attert+/+ (Wt) state To evaluate telomere length changes between mutant plants propagated via upper and lower silique seeds (undergoing a different number of cell divisions per seedto-seed generation) and to distinguish them from natural variations in telomere lengths in telomerase-positive plants, control Wt plants were generated from the same original heterozygous plants as Attert-/mutants. Wt plants were propagated according to the same schematic protocol as U- and L-lines of Attert-/mutants (Fig. 1) to reveal any potential stochastic changes in telomere lengths. Examples of primary PETRA and TRF results are shown in Fig. 3. The results were repeated in three independent lineages (Ha, Hb1 and Hb2). Measurement of PETRA fragment sizes revealed increases in both 2R and 3L telomeres in most first generation Attert+/+ plants within a range of 120–380 bp (Figs. 4, 5). The exception is the 3L telomere in Wta plant where telomere lengths show a mild (90 bp) shortening (Fig. 5). Telomere dynamics in the following generations of Wt plants displayed changes in both directions, but overall most plants showed a net increase in telomere lengths in G4 compared to their length in the Plant Mol Biol 123 original heterozygotes. This observation suggests that the mutation of a single Attert allele acts via haploinsufficiency, as in human (Hauguel and Bunz 2003; Zhang et al. 2003) and mouse tert genes (Erdmann et al. 2004). No substantial differences in telomere lengths were observed between U- and L-lines of corresponding Attert+/+ plants, revealing that telomerase in meristem cells maintains telomere stability during plant development (Fajkus et al. 1998; Riha et al. 1998) regardless of number of cell divisions. Telomere dynamics in Attert-/- plants The key question of this study was to analyze if there is a relationship between telomere shortening per plant generation and the number of meristem cell divisions between generations. To address this question, telomere lengths were measured during propagation of Attert-/mutants as U- and L-lines. Only the first few generations of Attert-/mutants were analyzed (G1–G4) to avoid accumulated cytogenetic abnormalities expected from the 6th generation (Riha et al. 2001). Results of telomere analysis (Figs. 4, 5) revealed: (i) Telomere length differences between consecutive plant generations can be highly variable ranging from tens of bp to 800 bp. There is some evidence that large changes in telomere length ([500 bp) in one generation is followed by smaller changes in the subsequent generation. For example, in 2R telomeres in mutant Ma-derived U-plant lines, large changes occur between G1 and G2 (Figs. 4, 5) and subsequently are less dramatic. In 3L telomeres of all the Maderived U-lines, substantial telomere shortening occurs between G3 and G4, while in earlier generations there are only moderate changes (Fig. 5). (ii) Telomeres 2R and 3L in the same plants behave relatively independently. Compare e.g., a 150 bp shortening between G1 and G2 generations in the 2R telomere in L-plants derived from the Mb1 mutant with ca. 800 bp shortening of 3L telomeres in the same plants (Figs. 4–6). (iii) Two patterns were observed in the dynamics of telomere shortening in Attert-/mutants. The first is represented by 2R telomeres of plants derived from Ma and Mb1 mutants. These results reveal, as predicted, a faster rate of telomere shortening per generation in plants propagated through seeds from the upper siliques (U-lines) than in those from the lower siliques (L-lines, Figs. 4, 7A). The second is observed in 3L telomeres of all lines and of 2R telomeres of Mb2-derived lines. Here the rate of Fig. 2 Results of TRAP assay in calli derived from the original heterozygous Attert+/plant (H), and its progeny—mutant Attert-/(M) and wild type Attert+/+ (Wt). A 50 bp marker (GeneRuler, Fermentas) is used as a marker (m). Telomerase extract from standard A. thaliana wt seedlings was used as positive control (+), and the extraction buffer served as a negative control (-) Fig. 3 Example of PETRA and TRF results. One of the Attert+/plants (Ha) and two generations of Attert-/(Ma) and Attert+/+ (Wta) plants derived from the Ha plant, were propagated as L- and U-lines (as indicated) and assayed by PETRA with a primer specific for the subtelomere 2R. Apart from stronger bands that correspond to main products of PETRA there are also weaker bands of a higher mobility. These weaker bands in a given lane correspond to the stronger bands in both their number and mutual position. The result labelled as TRF shows the hybridization pattern of terminal restriction fragments generated with restriction enzyme MseI in Ha plant. A 1 kb GeneRuler (Fermentas) has been used as marker (m) Plant Mol Biol 123 telomere shortening is not substantially different between U- and L-lines (Figs. 4, 5, 7B). If it is assumed that ALT becomes activated in response to telomere shortening, then the mechanism is more active in 3L telomeres in earlier generations compared to 2R telomeres, and acts to override any losses incurred through an increased number of cell divisions between U- and L-plant generations. (iv) A difference between U- and L-lines is apparent from the evaluation of relative telomere length changes between generations (Fig. 6). While the overall telomere lengthening is occasionally observed between G2 and G3 generations in both U- and L-lines, it is much more frequent between G3 and G4 generations—but only in L-lines, whereas it is entirely absent in any of the U-line plants between the same generations. This remarkable difference suggests that ALT processes may be more frequent in gametogenesis or early developmental stages. In L-plants, such elongation can sometimes be even higher than replicative shortening corresponding to formation of lower-silique seeds, while in U-plants the greater number of cell divisions overrides the efficiency of ALT. Fig. 4 Telomere lengths in 2R chromosome arms in wild-type (Wta, Wtb1, Wtb2) and mutant (Ma, Mb1, Mb2) plants. Each graph consists of consecutive results from the original heterozygous Attert+/- plant (Ha or Hb), and four subsequent generations (G1–G4) of the individual U- or L-lines coming from the given first-generation plant (Wta, Wtb1, Wtb2 and Ma, Mb1, Mb2). Overlapping points cannot be seen separately Fig. 5 Telomere lengths in 3L chromosome arms in wild-type (Wta, Wtb1, Wtb2) and mutant (Ma, Mb1, Mb2) plants. Each graph consists of consecutive results from the original heterozygous Attert+/- plant (Ha or Hb), and four subsequent generations (G1–G4) of the individual U- or L-lines coming from the given first-generation plant (Wta, Wtb1, Wtb2 and Ma, Mb1, Mb2) Plant Mol Biol 123 Early onset of phenotype changes in Attert mutants The first phenotypic abnormalities appeared in some plants from the U- and L-lines in the 4th generation of Attert-/mutants, primarily from Mb1 and Mb2 plants. Some plants showed mild abnormalities, leaves were asymmetric and lobed. Other plants had more serious abnormalities both in leaf morphology and shoot structure—stems were split and in some cases appeared to have lost the apical dominance; leaves were smaller than wild-type, had an irregular shape, a rough surface and were more plentiful than in wild-type plants (Fig. 8); siliques were smaller with fewer seeds, or were completely sterile. The most severe phenotype arose in a G4 plant from U-line derived from the Mb1 mutant. In that plant there were many abnormalities including leaves with a distinct trichomes, a small stem and sterile siliques (Fig. 8A–C). 2R and 3L telomeres in this G4 plant were 1340 and 770 bp long, respectively, i.e. similar to the other G4 mutant plants included in Figs. 4–7. For other plants, the occurrence of abnormal phenotypes did not differ substantially between mutants of the U- and L-lines. The phenotypic abnormalities observed correspond well to those described by Riha et al. (2001), but in our experiments some plants had severe phenotypes in the 4th generation rather than the 8th generation described in that earlier work. Fig. 6 Summary of relative telomere shortening in each generation of all three mutant lines (Ma, Mb1 and Mb2). Telomere lengths in each generation were subtracted from lengths in the previous generation, therefore values are positive in case of shortening and negative in elongation events Fig. 7 Graphs showing two different trends observed in rates of telomere shortening between plants from U- and L-lines. Telomere lengths of all plants in either U- or L-line were averaged in a given generation. (A) (2R telomeres in Ma lines) shows example of the case in which the rate of telomere shortening is faster in the U-line then in L-line. (B) (2R telomeres in Mb2 lines) exemplifies the similar rate of telomere shortening in U- and L-lines. The former course was observed in 2R telomeres in Ma and Mb1 lines, while the latter occurred in 2R telomeres in Mb2 line and 3L telomeres of all lines Plant Mol Biol 123 Discussion Involvement of telomerase-independent processes in plant telomere dynamics Our analysis of telomere dynamics in Arabidopsis Attert-/mutants challenges current perceptions of the relative contributions of telomerase activity, ALT and telomere rapid deletion (TRD) in plant telomere dynamics. Watson and Shippen (2007) showed that elongated telomeres in A. thaliana Ku70 mutants shorten to the length typical for wt plants after three generations when restored with wild-type Ku70. This corresponds to an average loss of 2.3 ± 0.8 kb of telomeric DNA per generation, which is interpreted as the result of TRD, as it exceeds the previously reported rate of telomere shortening per generation in tert mutants. The rate of telomere shortening per generation observed in this work, as well as in our study was not constant (Fig. 7), but decreased as telomeres approached the length of about 2 kb. In addition to the reported TRD, the opposing process—ALT— was detected in ku70 tert double mutants with elongated telomeres (Watson and Shippen 2007). Although it is possible that both TRD and ALT are actively involved in telomere length regulation in A. thaliana, according to our previous estimates and calculations (Fajkus et al. 2005), the observed rate of telomere shortening in Attert-/mutants is about ten times less than the expected replicative loss resulting from number of cell divisions that occurs between generations. As telomeres approach a critical length, the frequency of ALT events may increase, which results in more substantial telomere length changes in both directions. This hypothesis might also explain why 3L telomeres respond similarly between U- and L-lines. The initial length of 3L telomeres is about 500 bp shorter than 2R telomeres. Perhaps because the telomeres are shorter, stochastic ALT processes are active in both lines in earlier generations in 3L telomeres. Therefore we conclude that the apparent slower rates of telomere attrition in G3–G4 generations are presumably due to the up-regulation of ALT by shortened telomeres themselves, or by their changed nucleoprotein structure. Is ALT restricted to a specific developmental stage? Frequently telomere elongation was observed between G3 and G4 in L-plants, while they are entirely absent in U-plants. This remarkable difference suggests that ALT processes may be more frequent in gametogenesis or early developmental stages. In L-plants, such elongation can sometimes be even greater than subsequent replicative shortening corresponding to formation of lower-silique seeds, while in U-plants the additional 30 divisions counteract length gains produced by ALT at a particular time in development. The activity of ALT in early embryonic Fig. 8 Examples of mutant phenotypes. Different life stages of the G4 plant from U-line, which had the most serious phenotype defects (A–C). Leaves were asymmetric and rough, had conspicuous trichomes, stem was short and thin and the plant was sterile. A detailed picture of a split stem is given in panel D Plant Mol Biol 123 development is supported also by recent observations of Liu et al. (2007) in mammalian oocyte cells. They show that oocytes have shorter telomeres than somatic cells and lack telomerase activity, but their telomeres lengthen remarkably during early cleavage cycles following fertilization through a recombination-based mechanism. From the blastocyst stage onwards, telomerase becomes activated and maintains the telomere length established by this alternative mechanism. The involvement of ALT in normal development thus gains support in such divergent organisms as mammalians and plants, and this conservation suggests its key importance. The observation of the overall telomere elongation events in L-plants also poses an interesting question: why is the ALT process not able to compensate completely for the lack of telomerase in Attert-/mutants? A possible explanation is provided by ALT itself. Studies on different model systems show that ALT involves homologous recombination (HR) and HR-dependent DNA replication (reviewed in Cesare and Reddel 2008). Experimental evidence supports a ‘‘roll and spread’’ model of ALT, in which a 30 telomeric overhang invades either the duplex region of its own telomere (forming the t-loop), or an extrachromosomal telomeric circle (a product of t-loop junction resolution), and is extended with DNA polymerase using either of the above templates. The extended telomere can then spread to other chromosome termini via HR. This scenario is supported by the presence of both types of candidate templates (t-loops and t-circles) for the initial phase of ALT in plants (Cesare et al. 2003; Zellinger et al. 2007). Moreover, it was shown recently that Ku protein suppresses formation of t-circles and ALT lengthening in A. thaliana (Zellinger et al. 2007). If initial telomere elongation events are limited to the early embryogenesis or gametogenesis, as suggested above, and the ‘‘reservoir’’ of elongated telomere sequence available for the individual plant generation is finite, then the extension potential will depend on the initial length of telomere. In early generations, when the initial telomere length is relatively long, the ALT is able to compensate for the replicative telomere shortening almost completely (compare the expected telomere loss of several kb per generation in Attert-/- mutants with the observed average rate of telomere shortening of ca. 250–500 bp per generation (Fitzgerald et al. 1999)). The slow but progressive telomere shortening, nevertheless, gradually decreases the efficiency of ALT, and the observed massive increase in genome instability observed since the 6th generation of Attert mutants (Riha et al. 2001) may reflect this. Moreover, the increase of recombination frequency due to the excessive activation of ALT itself can contribute to an increase in genome instability (Jeyapalan et al. 2005). Immortal strand hypothesis resurrected in animal cells—does it live in plant cells too? The remarkably low telomere-shortening rate of Attert-/mutants could be also explained without invoking ALTmechanisms. The ‘‘immortal strand hypothesis’’ proposed decades ago (Cairns 1975) suggests stem cells might limit acquired mutations that give rise to cancer through the directed inheritance of parental DNA strands. Though largely disregarded, this hypothesis of the template DNA strand co-segregation in dividing stem cells and their progeny has implications in telomere biology. Recent results (Conboy et al. 2007) provide experimental support to this hypothesis. These authors used sequential pulses of three different halogenated thymidine analogs and analyzed stem cell progeny during induced regeneration in vivo. They observed extraordinarily high frequencies of segregation of older and younger template strands during a period of proliferative expansion of muscle stem cells. Template strand co-segregation was strongly associated with asymmetric cell divisions yielding daughters with divergent fates. Daughter cells inheriting the older templates retained the more immature phenotype, whereas daughters inheriting the newer templates acquired a more differentiated phenotype (Conboy et al. 2007). It has yet to be shown if this behaviour is also present in plant meristem cells, but the idea is certainly worth considering and testing experimentally. While the mechanism could contribute to the low telomere shortening rate, it does not provide explanation for the overall telomere lengthening observed namely between G3 and G4 generations of L-line plants. The onset of abnormal phenotype preceeds critical telomere shortening The onset ofabnormalphenotype effects was observed atleast two generations earlier in this work than in the previous study (Riha et al. 2001). The severity of phenotype changes, however, did not show any direct relationship to telomere lengths ortheoccurrenceofplantsineitherU-orL-lines.Inparticular, the severely affected G4 mutant plant did not show markedly shorter 2R or 3L telomeres than the other mutants of the same generation. Moreover, we also measured other telomeres in this particular plant (1R = 840 bp, 1L = 820 bp, 4R = 1,170 bp, 5R = 850 bp, 5L = 970 bp) and all of these ranged well above the previously published minimum functional telomere length in Arabidopsis (300–400 bp), which was based on measurement of telomere fusion sites (Heacock et al. 2004). Potentially therefore the phenotype characteristics may not be due to reduced telomere lengths directly, but perhaps a resultof ALT-associated abundant levelof telomere recombination, leading to genome instability. Plant Mol Biol 123 Possible origins of multiple products in PETRA assays The presence of two or multiple signals coming from a given chromosome arm using the PETRA technique is usually attributed to different telomere lengths at homologous chromosome arms, or to the hypothetical TRD events in a subpopulation of telomeres (Watson and Shippen 2007). Besides these bands of comparable intensity, we observed also some weaker bands of higher mobility, which corresponded to the stronger bands in the same lane in their number and mutual positions (see Fig. 3). Interestingly, the length differences between weaker and stronger products decreased proportionally with the size of the stronger band (see Fig. 9). The weaker band occasionally disappears in association with the shortest telomeres. We propose that these bands arise through secondary annealing sites of the telomeric PETRA primer to the region of the displacement loop (D-loop) at the site of G-overhang invasion into the double-stranded telomere region (see Fig. 10). Since the annealing occurs under native conditions, at least a fraction of t-loops may be preserved at the initial phase of PETRA. Subsequent convergence of the sizes of the two products could then be explained by tightening of the t-loop which is, however critically limited by bendability of the chromatin fibre (Fajkus and Trifonov 2001). Continued telomere shortening below the critical lengths impedes further t-loop formation (Forsyth et al. 2002), thus leading to a single telomere product in PETRA. In conclusion, our results suggest that telomere dynamics in Attert-/knock-out mutants can be explained solely by involvement of ALT, without a necessity to presume TRD events. The ALT events appear to be timelimited, probably to gametogenesis or early embryonic development. The fast and efficient activation of ALT in response to the loss of telomerase activity, and its probable developmental regulation imply that ALT may be a normal part of plant development. Acknowledgements We thank Professor Andrew R. Leitch for insightful comments and the manuscript revision, Eva Sy´korova´ and Dagmar Zachova´ for helpful advice for cultivation of plant tissue cultures. This work was supported by the Czech Ministry of Education (MSM0021622415, LC06004), Czech Academy of Sciences (AVOZ50040507) and grants from GACR (521/05/0055) and GA ASCR (IAA600040505 and IAA601630703). References Arabidopsis Genome Initiative (2000) Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 408:796–815 Cairns J (1975) Mutation selection and the natural history of cancer. Nature 255:197–200 Cesare AJ, Reddel RR (2008) Alternative lengthening of telomeres in mammalian cells. In: Nosek J, Tomaska L (eds) Origin and evolution of telomeres. Austin, Landes Bioscience (in press) Fig. 9 By-products of PETRA. Besides major products of PETRA, we observed also weaker and faster-migrating bands, which correspond to bands of the major products both in their number and lengths ratio (compare to Fig. 3). The length differences between the main products and by-products in a given lane obtained from all PETRA experiments were plotted against the lengths of main products. These differences are proportional to the lengths of corresponding main products as is displayed in the graph. The graph includes data obtained in both 2R and 3L telomeres in all plants till G3 Fig. 10 Schematic picture of a possible origin of the shorter and weaker bands accompanying the main products of PETRA (see also Figs. 3 and 9). These by-products may originate from a secondary annealing of telomeric PETRA primer (dashed arrow) in the displacement loop in a fraction of t-loops surviving DNA isolation (A). Telomere shortening results in the corresponding tightening of the t-loop (B) which is limited by the bendability of telomeric chromatin fibre. Shortening below the critical limit results in the complete opening of t-loop (C) and the loss of the by-products of PETRA Plant Mol Biol 123 Cesare AJ, Quinney N, Willcox S, Subramanian D, Griffith JD (2003) Telomere looping in P-sativum (common garden pea). Plant J 36:271–279 Conboy MJ, Karasov AO, Rando TA (2007) High incidence of nonrandom template strand segregation and asymmetric fate determination in dividing stem cells and their progeny. PLoS Biol 5:e102 Dellaporta SL, Wood J, Hicks JB (1983) A plant DNA minipreparation: version II. Plant Mol Biol Rep 1:19–21 Edwards K, Johnstone C, Thompson C (1991) A simple and rapid method for the preparation of plant genomic DNA for PCR analysis. Nucleic Acids Res 19:1349 Erdmann N, Liu Y, Harrington L (2004) Distinct dosage requirements for the maintenance of long and short telomeres in mTert heterozygous mice. Proc Natl Acad Sci USA 101:6080–6085 Fajkus J, Trifonov EN (2001) Columnar packing of telomeric nucleosomes. Biochem Biophys Res Commun 280:961–963 Fajkus J, Kovarik A, Kralovics R, Bezdek M (1995) Organization of telomeric and subtelomeric chromatin in the higher-plant Nicotiana-tabacum. Mol Gen Genet 247:633–638 Fajkus J, Kovarik A, Kralovics R (1996) Telomerase activity in plant cells. FEBS Lett 391:307–309 Fajkus J, Fulneckova J, Hulanova M, Berkova K, Riha K, Matyasek R (1998) Plant cells express telomerase activity upon transfer to callus culture, without extensively changing telomere lengths. Mol Gen Genet 260:470–474 Fajkus J, Sykorova E, Leitch AR (2005) Telomeres in evolution and evolution of telomeres. Chromosome Res 13:469–479 Fitzgerald MS, McKnight TD, Shippen DE (1996) Characterization and developmental patterns of telomerase expression in plants. Proc Natl Acad Sci USA 93:14422–14427 Fitzgerald MS, Riha K, Gao F, Ren S, McKnight TD, Shippen DE (1999) Disruption of the telomerase catalytic subunit gene from Arabidopsis inactivates telomerase and leads to a slow loss of telomeric DNA. Proc Natl Acad Sci USA 96:14813–14818 Fletcher JC (2002) Coordination of cell proliferation and cell fate decisions in the angiosperm shoot apical meristem. Bioessays 24:27–37 Forsyth NR, Wright WE, Shay JW (2002) Telomerase and differentiation in multicellular organisms: turn it off, turn it on, and turn it off again. Differentiation 69:188–197 Grandjean O, Vernoux T, Laufs P, Belcram K, Mizukami Y, Traas J (2004) In vivo analysis of cell division, cell growth, and differentiation at the shoot apical meristem in Arabidopsis. Plant Cell 16:74–87 Hauguel T, Bunz F (2003) Haploinsufficiency of hTERT leads to telomere dysfunction and radiosensitivity in human cancer cells. Cancer Biol Ther 2:679–684 Heacock M, Spangler E, Riha K, Puizina J, Shippen DE (2004) Molecular analysis of telomere fusions in Arabidopsis: multiple pathways for chromosome end-joining. Embo J 23:2304–2313 Heller K, Kilian A, Piatyszek MA, Kleinhofs A (1996) Telomerase activity in plant extracts. Mol Gen Genet 252:342–345 Jeyapalan JN, Varley H, Foxon JL, Pollock RE, Jeffreys AJ, Henson JD, Reddel RR, Royle NJ (2005) Activation of the ALT pathway for telomere maintenance can affect other sequences in the human genome. Hum Mol Genet 14:1785–1794 Liu L, Bailey SM, Okuka M, Munoz P, Li C, Zhou L, Wu C, Czerwiec E, Sandler L, Seyfang A et al (2007) Telomere lengthening early in development. Nat Cell Biol 9:1436–1441 Pich U, Fuchs J, Schubert I (1996) How do Alliaceae stabilize their chromosome ends in the absence of TTTAGGG sequences? Chromosome Res 4:207–213 Riha K, Fajkus J, Siroky J, Vyskot B (1998) Developmental control of telomere lengths and telomerase activity in plants. Plant Cell 10:1691–1698 Riha K, McKnight TD, Griffing LR, Shippen DE (2001) Living with genome instability: plant responses to telomere dysfunction. Science 291:1797–1800 Siroky J, Zluvova J, Riha K, Shippen DE, Vyskot B (2003) Rearrangements of ribosomal DNA clusters in late generation telomerase-deficient Arabidopsis. Chromosoma 112:116–123 Sykorova E, Lim KY, Chase MW, Knapp S, Leitch IJ, Leitch AR, Fajkus J (2003) The absence of Arabidopsis-type telomeres in Cestrum and closely related genera Vestia and Sessea (Solanaceae): first evidence from eudicots. Plant J 34:283–291 Sykorova E, Fajkus J, Meznikova M, Lim KY, Neplechova K, Blattner FR, Chase MW, Leitch AR (2006) Minisatellite telomeres occur in the family Alliaceae but are lost in Allium. Am J Bot 93:814–823 Watson JM, Shippen DE (2007) Telomere rapid deletion regulates telomere length in Arabidopsis thaliana. Mol Cell Biol 27: 1706–1715 Zellinger B, Akimcheva S, Puizina J, Schirato M, Riha K (2007) Ku suppresses formation of telomeric circles and alternative telomere lengthening in Arabidopsis. Mol Cell 27:163–169 Zhang A, Zheng C, Hou M, Lindvall C, Li KJ, Erlandsson F, Bjorkholm M, Gruber A, Blennow E, Xu D (2003) Deletion of the telomerase reverse transcriptase gene and haploinsufficiency of telomere maintenance in Cri du chat syndrome. Am J Hum Genet 72:940–948 Plant Mol Biol 123 Supplement C Schrumpfová, P.P., Kuchar, M., Palecek, J., Fajkus, J., 2008. Mapping of interaction domains of putative telomere-binding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. FEBS Lett. 582, 1400–1406 P.P.S. was significantly involved in the experimental part (more then 50 % of experiments) and participated in the ms writing and editing This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright Author's personal copy Mapping of interaction domains of putative telomere-binding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana Petra Procha´zkova´ Schrumpfova´a,1 , Milan Kucharˇa,1 , Jan Palecˇeka , Jirˇı´ Fajkusa,b,* a Department of Functional Genomics and Proteomics, Institute of Experimental Biology, Faculty of Science, Masaryk University, Kamenice 5, CZ-62500 Brno, Czech Republic b Laboratory of DNA-Molecular Complexes, Institute of Biophysics, Czech Academy of Sciences, Kra´lovopolska´ 135, CZ-61265 Brno, Czech Republic Received 12 December 2007; revised 19 February 2008; accepted 2 March 2008 Available online 1 April 2008 Edited by Gianni Cesareni Abstract We previously searched for interactions between plant telomere-binding proteins and found that AtTRB1, from the single-myb-histone (Smh) family, interacts with the Arabidopsis POT1-like-protein, AtPOT1b, involved in telomere capping. Here we identify domains responsible for that interaction. We also map domains in AtTRB1 responsible for interactions with other Smh-family-members. Our results show that the N-terminal OB-fold-domain of AtPOT1b mediates the interaction with AtTRB1. This domain is characteristic for POT1- proteins and is involved with binding the G-rich-strand of telomeric DNA. AtPOT1b also interacts with AtTRB2 and AtTRB3. The central histone-globular-domain of AtTRB1 is involved with binding to AtTRB2 and 3, as well as to AtPOT1b. AtTRB1-heterodimers with other Smh-family-members are more stable than AtTRB1-homodimers. Our results reveal interaction networks of plant telomeres. Structured summary: MINT-6440051: AtTRB1 (uniprotkb:Q8VWK4) physically interacts (MI:0218) with AtTRB1 (uniprotkb:Q8VWK4) by two-hybrid (MI:0018) MINT-6440068: AtTRB2 (uniprotkb:Q8VX38) physically interacts (MI:0218) with AtTRB1 (uniprotkb:Q8VWK4) by two-hybrid (MI:0018) MINT-6440083: AtTRB3 (uniprotkb:Q9M2X3) physically interacts (MI:0218) with AtTRB1 (uniprotkb:Q8VWK4) by two-hybrid (MI:0018) MINT-6440099: AtPOT1b (uniprotkb:Q6Q835) physically interacts (MI:0218) with AtTRB1 (uniprotkb:Q8VWK4) by two-hybrid (MI:0018) MINT-6440119: AtPOT1b (uniprotkb:Q6Q835) physically interacts (MI:0218) with AtTRB2 (uniprotkb:Q8VX38) by two-hybrid (MI:0018) MINT-6440138: AtPOT1b (uniprotkb:Q6Q835) physically interacts (MI:0218) with AtTRB3 (uniprotkb:Q9M2X3) by two-hybrid (MI:0018) MINT-6440216: AtPOT1b (uniprotkb:Q6Q835) physically interacts (MI:0218) with AtTRB1 (uniprotkb:Q8VWK4) by coimmunoprecipitation (MI:0019) MINT-6440157: AtTRB2 (uniprotkb:Q8VX38) physically interacts (MI:0218) with AtTRB1 (uniprotkb:Q8VWK4) by coimmunoprecipitation (MI:0019) MINT-6440177: AtTRB3 (uniprotkb:Q9M2X3) physically interacts (MI:0218) with AtTRB1 (uniprotkb:Q8VWK4) by coimmunoprecipitation (MI:0019) Ó 2008 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved. Keywords: Plant; Telomere; Protein–protein interaction; AtTRB1; AtTRB2; AtTRB3; AtPOT1b 1. Introduction Telomere proteins play a role in the protection and maintenance of chromosome ends. In human cells, the minimal functional set of proteins participating in telomere protection is collectively called ‘‘shelterin’’ [1]. Shelterin consists of three proteins (TRF1, TRF2 and POT1) that directly recognize telomeric DNA and are interconnected by at least three other proteins (TIN2, TPP1 and Rap1), forming a telomere-specific protective cap. Similar complexes are also likely to exist in plants and these are particularly attractive to study due to the telomerase-competent status (i.e., reversible telomerase activity regulation) of plant somatic cells [2,3]. A number of putative plant telomeric proteins have been found by homology searches of DNA and protein sequence databases and tested for their affinity to telomeric DNA sequences in vitro (reviewed in [4]). There is however very little data relevant to their telomeric function. Of the putative ‘‘plant shelterin’’ components, functional data relevant to telomere homeostasis is available for two Arabidopsis thaliana POT1-like proteins, AtPOT1a and AtPOT1b. These proteins contain the oligonucleotide-binding (OB) fold domain which binds to the G-rich strand of telomeric DNA but their overall sequence similarity is low (49%). The functions of AtPOT1a and AtPOT1b proteins are different: AtPOT1a functions mainly in telomerase regulation, while AtPOT1b contributes to chromosome endprotection and genome stability [5–9]. Recently, another Arabidopsis protein, AtTBP1, has been shown to be involved in telomere length regulation [10]. This * Corresponding author. Address: Department of Functional Genomics and Proteomics, Institute of Experimental Biology, Faculty of Science, Masaryk University, Kamenice 5, CZ-62500 Brno, Czech Republic. Fax: +420 549492654. E-mail address: fajkus@sci.muni.cz (J. Fajkus). 1 First two authors contributed equally to this work. 0014-5793/$34.00 Ó 2008 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved. doi:10.1016/j.febslet.2008.03.034 FEBS Letters 582 (2008) 1400–1406 Author's personal copy protein binds double-stranded telomeric DNA in vitro via a characteristic Myb-like domain, referred to as a telobox, located at its C-terminus [11,12]. To identify other components of ‘‘plant shelterin’’, we analyzed a number of putative A. thaliana telomere proteins for their mutual interactions. We previously found that AtTRP1, the Arabidopsis myb-like protein bearing a C-terminal telobox, interacts with AtKu70 [5], which itself plays a role in plant telomere homeostasis [13,14]. Furthermore AtTRP1 may be a functional homolog of mammalian TRF2 [5]. In addition, an Arabidopsis POT1-like protein, AtPOT1b, interacts with AtTRB1, a protein from the single myb histone (Smh) family [5]. The Smh family is characterised by a unique triple motif structure containing a N-terminal myb-like domain, a central GH1/GH5 histone globular domain and a C-terminal coiled-coil domain [15]. Proteins of this family in Arabidopsis show specific binding to telomeric DNA and can form homo- and heteromeric protein–protein complexes [16]. The abundance of candidate telomere proteins in plants, arising from numerous paralogs of telomere-binding protein and plant-specific proteins, coupled with an apparent absence of some constitutive animals and fungi shelterin components, makes imperative analyses of interactions between the candidate plant telomere proteins. Using a combination of the yeast two-hybrid system (Y2H) and co-immunoprecipitation (CoIP), we characterise here the protein domains involved in interactions between AtTRB1 and AtPOT1b, as well as domains engaged in the formation of homomeric and heteromeric complexes of AtTRB proteins. 2. Materials and methods 2.1. Cloning of full length proteins and their deletion variants for twohybrid assay An overview of cloned candidate telomeric DNA-binding proteins is given in Table 1. cDNA sequences of AtTRB1, AtTRB2, AtTRB3 and AtPOT1b have been cloned as described previously [16]. To localize the interaction domains the deletion forms of AtTRB1 and AtPOT1b were generated by PCR and cloned into the vector pGBKT7 or pGADT7, respectively. Sequence-specific primers with restriction sites were used for cloning individual cDNAs as shown in Table 2. To localize the interaction domain(s) in AtTRB1, cDNA fragments were cloned in pGADT7 and denominated according to primers used (for example, the fragment F1R1 was generated using TRB1 F1 as forward and TRB1 R1 as reverse primers – see Figs. 1A and 2B). Similarly, AtPOT1b fragments were generated to localize the region of AtPOT1b responsible for interaction with AtTRB1 (see Fig. 2C). Prior to two-hybrid screening, cloned constructs were checked for the correct reading frame and absence of mutations by DNA sequencing on an ABI PRISM 310 sequencer (Perkin–Elmer). 2.2. Yeast two-hybrid (Y2H) system Two strains of Saccharomyces cerevisiae, PJ69-4a and PJ69-4a were used [17]. Protein AtPOT1b, its deletion variants and AtTRB2, AtTRB3 were expressed from the yeast vector pGBKT7 in strain PJ69- 4a, and AtTRB1 and its fragments from vector pGADT7 in strain PJ69-4a. This division enabled proper combining of the proteins and their deletion variants in interaction assays. Both strains, identical except for the mating type, were mixed on Petri-dishes with YPD medium (1.1% yeast extract, 2.2% bacteriological peptone, 2% glucose and 2% agar) to fuse yeast haploid cells of different strains, and incubated at 30 °C for 8–10 h. The diploid cells were printed by velvet stamp onto control -Leu,-Trp selective plates (0.67% yeast nitrogen base, 2% glucose, 0.12% amino acid mixture without Leu and Trp, 2% agar, pH adjusted by NaOH to 6.8) and then onto -Ade selective plates to test the interaction (0.67% yeast nitrogen base, 2% glucose, 0.12% amino acid mixture without Ade, 2% agar, pH adjusted by NaOH to 6.8) and were incubated at 30 °C for a few days until colonies had grown. Alternatively, PJ69-4a cells were cotransformed with both pGBKT7 and pGADT7 plasmids and grown on -Leu,-Trp plates. Colonies were inoculated into YPD liquid medium and incubated at 30 °C overnight. Ten-times diluted aliquots were dropped onto both -Ade and -His plates. For a semi-quantitative test, 5 ll aliquots were dropped onto selective -His plates containing increasing concentrations of 3-aminotriazol (3-AT). As the 3-AT inhibits His3 activity, the ability of yeast cells to grow on higher concentrations correlates with the higher binding affinity of the hybrid proteins. To verify our results we also used the yeast strain MaV203, where the His3-reporter gene is under a less tightly controlled promoter (Invitrogen). The drop test was executed in the similar way as with the PJ69 strain. 2.3. In vitro translation and co-immunoprecipitation Proteins were co-expressed from the same constructs as were used in Y2H system with an hemagglutinin tag (pGADT) or a myc-tag (pGBKT) by use of the TNT Quick Coupled Transcription/Translation System (Promega) in 15–25 ll of each reaction according to the manufacturerÕs instruction. For Myc pull-down experiments, 15– 25 ll of in vitro-expressed proteins in total volume of 100 ll of HEPES buffer (25 mM HEPES, 150 mM KCl, 50 mM NaCl, 3 mM MgCl2, 10% glycerol, 0.1% NP-40, 1 mM DTT, 2 mM PMSF, 2 lg/ll leupeptine, 1 lg/ll pepstatine) were mixed with 1 lg anti-Myc-tag polyclonal antibody (Abcam) and incubated overnight at 4 °C (Input fraction). 10 ll of Protein G magnetic particles (Dynabeads, Invitrogen-Dynal) were then added, and the mixture was incubated for 1 h/4 °C (UnTable 1 Overview of cloned proteins Protein Protein group Characteristic domain GenBank accession number Reference AtTRB1 dsDNA binding N-terminal Myb domain AAL73123 [16] AtTRB2 Proteins AAL73441 [17] AtTRB3 NP_190554 AtPOT1b ssDNA binding proteins Pot1 domain NP_196249 [5,9,23] Table 2 Complete list of primers used for cloning Primer Restriction site Sequence of primer (50 fi 30 ) POT1b F BamHI ATGGATCCTAATGGAGGAGGAGAGAAGAG POT1b F1 BamHI ATGGATCCTAAAGATTGTGCTGATTAACC POT1b F2 BamHI TAGGATCCACTTCTTATCGAATCTGAGAG POT1b F3 BamHI TTGGATCCTTAAGTCAGAAAGGCTTC POT1b R XhoI ATTCTCGAGTCATGAAGCATTGATCCAAG POT1b R1 XhoI TTACTCGAGCCCTTCATCAGCATATAGAG POT1b R2 XhoI TTACTCGAGCCTGTGATTTCAGAATGTG POT1b R3 XhoI TTACTCGAGGGTTGAAGACAGTGAATG POT1b R4 XhoI TTACTCGAGATCTTCAAACTTGTACGTG POT1b R5 XhoI CTTCTCGAGGGTTAATCAGCACAATCTTTA TRB1 F BamHI ATGGATCCGAATGGGTGCTCCTAAGCAG TRB1 F1 BamHI CGGGATCCAAGATGCGACCTCTGGACTCC TRB1 F2 BamHI GAGGATCCAAGGTCTGGGGGTGTTTGGA TRB1 F01 BamHI CGGGATCCTAGTCATGGCAAATGGCTGG TRB1 R XhoI TGGCTCGAGAGGCACGGATCATCATTTTG TRB1 R1 BamHI TCGGATCCTCCAAACACCCCCAGACC TRB1 R2 BamHI GAGGATCCGGAGTCCAGAGGTCGCATC TRB1 R12 BamHI CAGGATCCGCGTTTGAAGTCTGGTGGAG P.P. Schrumpfova´ et al. / FEBS Letters 582 (2008) 1400–1406 1401 Author's personal copy bound fraction). The beads were washed five times with HEPES buffer and then incubated with 10 ll of SDS-loading buffer for 10 min/85 °C to elute bound proteins (Bound fraction). Input, unbound, and bound fractions were separated by 12% SDS–PAGE and analyzed by STORM860 (GE Healthcare). 3. Results 3.1. Interactions between AtTRB proteins We have shown previously that the telomere-binding proteins AtTRB1, AtTRB2, and AtTRB3 form homodimeric and heterodimeric complexes in Y2H assays [5,16]. These proteins are similar to each other at the level of amino acid sequence and belong to the same family of Smh proteins [15]. In contrast to other myb-like telomere-binding proteins, in Smh proteins the myb-like domain is N-terminal. The myb-like domain is followed by GH1/GH5 histone globular domain and a C-terminal coiled-coil domain (Fig. 1A). Interactions of AtTRB proteins were expected to be mediated by the C-terminal coiled-coil domain because this domain supports protein oligomerization [18]. Therefore, we designed two deletion mutants AtTRB1 F1R (aa position 101–300) and AtTRB1 F2R (196–300) that comprised the C-terminus. Each construct was transformed into PJ69-4a two-hybrid strain and crossed with PJ69-4a containing full-length (FL) AtTRB constructs. Only AtTRB1 F1R construct containing both coiled-coil and Fig. 1. Histone GH1/GH5 globular domain of AtTRB1 binds to AtTRB1, 2 and 3 protein. (A) The AtTRB1 protein contains a myb-like domain (myb-like), followed by GH1/GH5 histone globular domain and C-terminal coiled-coil domain (coiled-coil). Full-length (FL) AtTRB1, F1R (aa 101– 300), FR1 (aa 1–201), F01R1 (aa 58–201), FR12 (aa 1–159), F01R12 (aa 58–159), F2R (aa 196–300), FR2 (1–106), F1R1 (aa 101–201) fragments (in PJ69-4a yeast strain) are combined with AtTRB1 FL, AtTRB2 FL, AtTRB3 FL two-hybrid constructs (in PJ69-4a yeast strain) and the diploid cells are tested on -Ade plates (-Ade) for protein–protein interactions (top and middle panel). Only fragments containing the GH1/GH5 histone globular domain interact with all three AtTRB proteins. The PJ69-4 cells containing the F1R1 construct are also tested on -His plates with 2 mM concentration of 3-AT (-His/2 mM). Only the interactions of F1R1 with AtTRB2 and AtTRB3 are detectable (bottom panel). Empty pGBKT7 (right panel) and pGADT7 vectors are used as negative controls. (B) AtTRB2 and AtTRB3 proteins are able to pull-down F1R1 fragment. The TNT expressed full-length AtTRB1 FL (lanes 1–3), AtTRB2 FL (lanes 4–6) and/or AtTRB3 FL (lanes 7–9) proteins were mixed with AtTRB1 GH1/GH5 fragment (F1R1, lanes 1–12) and incubated with anti-myc antibody overnight. Then protein G magnetic beads were added and proteins were immunoprecipitated for 1 h. In the control experiment, the F1R1 fragment was incubated with antibody and beads in the absence of partner protein (lanes 10–12). Input (I), unbound (U), and bound (B) fractions were collected and run in SDS–12% PAGE gels. 1402 P.P. Schrumpfova´ et al. / FEBS Letters 582 (2008) 1400–1406 Author's personal copy GH1/GH5 histone globular domain supported the growth of PJ69-4 diploid strain on -Ade plates (Fig. 1A, top panel), suggesting that the interactions are not mediated by the coiled-coil region. Instead, interactions between all three AtTRB proteins and fragments containing the GH1/GH5 histone globular domain were observed, in particular AtTRB1 FR1 (1–201), AtTRB1 F01R1 (58–201) and AtTRB1 F1R1 (101–201). Fragments containing truncated or completely deleted GH1/ GH5, as in the AtTRB1 FR2 (1–106), AtTRB1 F01R12 (58– 159) and AtTRB1 FR12 (1–159) constructs, lost the interaction with AtTRB1, as well as with the other AtTRB proteins. Thus, the shortest AtTRB1 fragment displaying the interaction with AtTRB proteins is F1R1 (101–201), which contains a GH1/GH5 domain and short flanking regions (Fig. 1A, middle panel). The PJ69-4 cells containing the F1R1 construct were also grown on -His plates with increasing concentrations of 3-amino-1,2,3-triazol (3-AT) to compare binding affinities to Fig. 2. Histone GH1/GH5 globular domain of AtTRB1 binds to the N-terminus of the AtPOT1b protein. (A) Full-length AtPOT1b FL interacts with AtTRB1 FL, AtTRB2 FL and AtTRB3 FL in two-hybrid assay when scored for growth on -His plates (for Y2H details see Fig. 1A). (B) Yeast two-hybrid cells containing FR1 (aa 1–201), F01R1 (aa 58–201), FR12 (aa 1–159) fragments are combined with AtPOT1b FL and tested on -Ade plates for protein–protein interactions. Only fragments containing the GH1/GH5 histone globular domain interact with the AtPOT1b protein. (C) In the co-immunoprecipitation assay, the TNT expressed AtTRB1 FR1 fragment (lanes 1–6) was mixed with full-length AtPOT1b FL protein (lanes 4– 6) and incubated with anti-myc antibody (same conditions as in Fig. 1B). In the control experiment, the FR1 fragment was incubated with antibody and beads in the absence of the partner protein (lanes 1–3). (D) The PJ69-4a cells containing AtTRB1 fragment F1R1 (aa 101–201) were cotransformed with full-length (FL) and/or following fragments F2R (aa 135–454), FR4 (aa 1–150), FR5 (aa 1–90) of AtPOT1b. Transformants containing FL, FR4 and FR5 fragments grow on -Ade plate (first column), however, a weak self-activation can be seen with FR4 fragment on a control plate (second column). When these cells were grown on -His plates with increasing concentrations of 3-AT the addition of 3-AT to 5 mM concentration abolishes the self-activation of FR4 (fourth column) while keeping its specific interaction with the F1R1 fragment of AtTRB1 (third column). These results suggest that the AtTRB1-AtPOT1b interaction is mediated by the binding of the GH1/GH5 domain of the AtTRB protein to the N-terminus (bearing the OB-fold domain) of the AtPOT1b protein. P.P. Schrumpfova´ et al. / FEBS Letters 582 (2008) 1400–1406 1403 Author's personal copy different AtTRB proteins (data not shown). At 2 mM 3-AT only the interactions of F1R1 with AtTRB2 and AtTRB3 were detected (Fig. 1A, bottom panel). To verify these results we used another two-hybrid strain (MaV203 strain has His3-reporter gene under a less tightly controlled promoter). When the full-length AtTRB1 or its FR1 fragment (covering both myb-like and GH1/GH5 domains) was used in the His-reporter assay, interactions with all three AtTRB proteins were positive up to 5 mM 3-AT. A further increase of 3-AT to 20 mM resulted in a loss of interaction with AtTRB1, while keeping interactions with AtTRB2 and 3 (Supplementary Fig. S1). Altogether, these data suggest that heterotypic complexes of AtTRB1 are more stable than homotypic ones. To test the above Y2H results by an independent approach, co-immunoprecipitation (CoIP) assays were performed with the above AtTRB1 fragments. In particular, interactions were assayed between F1R1 and the three AtTRB proteins. Fig. 1B shows that the myc-tagged AtTRB2 and AtTRB3 proteins are able to pull-down the F1R1 fragment while the full-length AtTRB1 is not. These results confirm a low affinity of the GH1/GH5 domain of AtTRB1 to full-length AtTRB1 (insufficient to provide a positive result in CoIP) and a higher affinity to both AtTRB2 and AtTRB3. 3.2. AtTRB1 interaction with AtPOT1b The protein AtPOT1b is thought important for ‘‘chromosome capping’’ in A. thaliana and interactions previously detected with AtTRB1 may be of functional significance [5,6]. Fig. 2A shows that the interaction of AtPOT1b is not limited to AtTRB1, but also occurs with AtTRB2 and AtTRB3 proteins (Fig. 2A). For mapping the interaction between the AtTRB1 and AtPOT1b the same AtTRB1 fragments in PJ69-4a cells (see Fig. 1A) were crossed with PJ69-4a cells containing full-length AtPOT1b. Only the diploid cells with AtTRB1 fragments containing the GH1/GH5 domain grow on -Ade plates (Fig. 2B and D; data not shown). These results suggest a role of the GH1/GH5 domain in binding to AtPOT1b. CoIP assays were performed using the FR1 fragment of AtTRB1 and the full-length AtPOT1b. The FR1 fragment co-precipitated with the myc-tagged AtPOT1b protein (Fig. 2C). This positive result confirms the above findings obtained by Y2H. In the case of AtPOT1b the following fragments were generated and cloned into pGBKT7: POT1b F2R (135–454), POT1b FR4 (1–150) and POT1b FR5 (1–90). The PJ69-4a cells containing AtTRB1 fragment F1R1 were cotransformed with Fig. 3. Schematic depiction of the principal telomere components and their interactions in humans and plants. In humans (A), the complex of ubiquitously present telomere-associated proteins, termed as shelterin [1], consists of two components that can bind telomeric dsDNA (TRF1, TRF2), and recruit the shelterin components TIN2, TPP1 and Rap1. The sixth partner in shelterin is the single-stranded TTAGGG repeat-binding protein, POT1. In addition to its binding to the G-strand of telomeric DNA, it can associate with telomeres also through its interaction with TPP1. Transitions between these states are illustrated with arrows. Examples of functionally important interactions (with Ku-proteins and telomerase) are also shown. In plants (B), a number of TRF-like proteins, exemplified here by AtTRP1, have been identified which are able to form homodimers and bind telomeric dsDNA with their C-terminal telobox Myb-like domain [11,12]. Analogous to TRF2, AtTRP1 is able to interact with AtKu70 [5], providing thus (in similarity to other organisms) association of Ku-heterodimer with telomeres via protein–protein interaction, in addition to its possible direct DNA-binding. Besides the TRF-like proteins, plant possess the SMH family of proteins which are characterised by the N-terminal myb-like domain, a central GH1/GH5 histone globular domain and a C-terminal coiled-coil domain [15]. These proteins form both homomeric and heteromeric complexes among each other using their central GH1/GH5 domain. This domain can also bind DNA in a sequence-non-specific manner (not shown), the interaction possibly important to avoid protein aggregation or telomere chromatin folding [23]. In addition, SMH proteins can interact (using the same GH1/GH5 domain) with AtPOT1b [5], one of the POT1-like proteins in Arabidopsis thaliana, which participates in telomere end-protection [6]. It is noteworthy that AtPOT1b uses the same domain (N-terminal OB-fold) for interaction with both telomeric ssDNA, and AtTRB proteins. Transitions of AtPOT1b between its DNA- and SMH-associated state (arrows) may be important for AtPOT1b recruitment to telomeres and for its protective function. The other AtPOT1 paralog, AtPOT1a, possibly functions in telomerase regulation and recruitment [7,8,22]. The presence of two functionally divergent POT1-like proteins in Arabidopsis is similar to the situation in mice [24]. Only linear telomere conformation is shown and nucleosomes are not depicted for simplicity. 1404 P.P. Schrumpfova´ et al. / FEBS Letters 582 (2008) 1400–1406 Author's personal copy full-length (FL) and/or fragments of AtPOT1b. Only transformants containing FL, FR4 and FR5 fragments grew on -Ade plates (Fig. 2D), however, a weak self-activation can be seen with fragment FR4 on the control plate. When these cells were grown on -His plates with increasing concentrations of 3-AT the addition of 3-AT to 5 mM concentration abolished the self-activation of FR4 whilst keeping its specific interaction with the F1R1 fragment of AtTRB1 (Fig. 2D, right panel). These results suggest that the AtTRB1-AtPOT1b interaction is mediated by the binding of GH1/GH5 domain of the AtTRB protein to the N-terminus (bearing the OB-fold domain) of the AtPOT1b protein. 4. Discussion Our results confirm previously published interactions between AtTRB proteins [16] and interactions between AtTRB1 and AtPOT1b [5]. They also provide a detailed map of those interactions. Interactions between AtPOT1b and other members of the Smh family (AtTRB2 and 3) are newly reported here. The function of AtPOT1b in chromosome protection and genome stability [6], highlights the importance of its interactions. AtTRB proteins, the only interaction partners of AtPOT1b identified so far, are representatives of the plant-specific Smh family of telomere-binding proteins. They have an N-terminal myb-like domain (instead of a usual C-terminal position in TRF-like proteins), and a central GH1/GH5 domain. The results of the assays show that AtTRB1 uses the central GH1/GH5 histone globular domain for interaction with AtTRB proteins and AtPOT1b. The GH1 and GH5 sub-domains are members of the Ôwinged helixÕ class of DNA-binding domains, although in contrast to other members of the family, they contain a distinct, additional cluster of positively charged amino acids. These residues form a second DNA-binding surface on the opposite side of the protein to the primary DNAbinding site [19]. Besides the ability to bind DNA, the GH5 domain is able to self-associate [20], which mediates AtTRB1 selfinteraction and interactions with AtTRB2, AtTRB3 and AtPOT1b. Possibly, the weak interactions of the GH1/GH5 histone globular domain in AtTRB1 are of a similar nature to GH5 hydrophobic protein–protein interactions described previously [20]. The GH1/GH5 histone globular domain can also bind DNA in a sequence-non-specific manner, while their Nterminal Myb-like domain [16] provides the sequence-specific binding of AtTRB to telomeres (Mozgova et al., submitted for publication). The use of the same domain for interacting with different proteins may be of functional importance. It is not only AtTRB1 which uses the same region to interact with all AtTRB proteins and AtPOT1b, but this is also observed for the N-terminus of AtPOT1b, which bears the OB-fold domain. This domain is thought to be involved in binding to telomere DNA and interacting with AtTRB proteins. These overlapping functions may be part of a regulatory mechanism, similar to that provided by binding properties of the components of the mammalian shelterin complex (Fig. 3). In the latter complex, there is a dynamic balance between POT1 bound directly to the telomere (to telomeric DNA) and via protein–protein interactions [21]. In plant shelterin, differences in expression levels of the proteins (see Supplementary figure S2) may also participate in modulation of telomere metabolism in a tissueand developmental stage-specific manner. The observed differential tendency of AtTRB1 to form homomeric and heteromeric complexes with AtTRB proteins, plus the ability of AtPOT1b to form complexes with all three tested AtTRB proteins and the presence of two functionally divergent AtPOT1 proteins in A. thaliana [6–8,22] suggest that plant shelterins are highly complex and carry features not found in animal and fungal shelterin. Acknowledgements: We thank Lumı´r Krejcˇı´ for suggestions regarding the two-hybrid techniques and Andrew Leitch (Queen Mary University, London) for critical reading of the manuscript. This work was supported by the Grant Agency of the Czech Republic (projects 521/ 05/0055 and 521/08P452), Czech Ministry of Education (LC06004) and the institutional support (MSM0021622415 and AVOZ50040507). Appendix A. Supplementary data Supplementary data associated with this article can be found, in the online version, at doi:10.1016/j.febslet. 2008.03.034. References [1] de Lange, T. (2005) Shelterin: the protein complex that shapes and safeguards human telomeres. Genes Dev. 19, 2100–2110. [2] Fajkus, J., Fulneckova, J., Hulanova, M., Berkova, K., Riha, K. and Matyasek, R. (1998) Plant cells express telomerase activity upon transfer to callus culture, without extensively changing telomere lengths. Mol. Gen. Genet. 260, 470–474. [3] Riha, K., Fajkus, J., Siroky, J. and Vyskot, B. (1998) Developmental control of telomere lengths and telomerase activity in plants. Plant Cell 10, 1691–1698. [4] Kuchar, M. (2006) Plant telomere-binding proteins. Biol Plantarum 50, 1–7. [5] Kuchar, M. and Fajkus, J. (2004) Interactions of putative telomere-binding proteins in Arabidopsis thaliana: identification of functional TRF2 homolog in plants. FEBS Lett. 578, 311–315. [6] Shakirov, E.V., Surovtseva, Y.V., Osbun, N. and Shippen, D.E. (2005) The Arabidopsis Pot1 and Pot2 proteins function in telomere length homeostasis and chromosome end protection. Mol. Cell. Biol. 25, 7725–7733. [7] Rossignol, P., Collier, S., Bush, M., Shaw, P. and Doonan, J.H. (2007) Arabidopsis POT1A interacts with TERT-V(I8), an Nterminal splicing variant of telomerase. J. Cell. Sci. 120, 3678– 3687. [8] Surovtseva, Y.V., Shakirov, E.V., Vespa, L., Osbun, N., Song, X. and Shippen, D.E. (2007) Arabidopsis POT1 associates with the telomerase RNP and is required for telomere maintenance. EMBO J. 26, 3653–3661. [9] Baumann, P., Podell, E. and Cech, T.R. (2002) Human Pot1 (protection of telomeres) protein: cytolocalization, gene structure, and alternative splicing. Mol. Cell Biol. 22, 8079–8087. [10] Hwang, M.G. and Cho, M.H. (2007) Arabidopsis thaliana telomeric DNA-binding protein 1 is required for telomere length homeostasis and its Myb-extension domain stabilizes plant telomeric DNA binding. Nucleic Acids Res. 35, 1333–1342. [11] Hwang, M.G., Chung, I.K., Kang, B.G. and Cho, M.H. (2001) Sequence-specific binding property of Arabidopsis thaliana telomeric DNA binding protein 1 (AtTBP1). FEBS Lett. 503, 35–40. [12] Bilaud, T., Koering, C.E., Binet-Brasselet, E., Ancelin, K., Pollice, A., Gasser, S.M. and Gilson, E. (1996) The telobox, a Myb-related telomeric DNA binding motif found in proteins from yeast, plants and human. Nucleic Acids Res. 24, 1294–1303. [13] Bundock, P., van Attikum, H. and Hooykaas, P. (2002) Increased telomere length and hypersensitivity to DNA damaging agents in an Arabidopsis KU70 mutant. Nucleic Acids Res. 30, 3395–3400. [14] Riha, K., Watson, J.M., Parkey, J. and Shippen, D.E. (2002) Telomere length deregulation and enhanced sensitivity to genoP.P. Schrumpfova´ et al. / FEBS Letters 582 (2008) 1400–1406 1405 Author's personal copy toxic stress in Arabidopsis mutants deficient in Ku70. EMBO J. 21, 2819–2826. [15] Marian, C.O. et al. (2003) The maize Single myb histone 1 gene, Smh1, belongs to a novel gene family and encodes a protein that binds telomere DNA repeats in vitro. Plant Physiol. 133, 1336–1350. [16] Schrumpfova, P., Kuchar, M., Mikova, G., Skrisovska, L., Kubicarova, T. and Fajkus, J. (2004) Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. Genome 47, 316–324. [17] James, P., Halladay, J. and Craig, E.A. (1996) Genomic libraries and a host strain designed for highly efficient two-hybrid selection in yeast. Genetics 144, 1425–1436. [18] Lupas, A., Van Dyke, M. and Stock, J. (1991) Predicting coiled coils from protein sequences. Science 252, 1162–1164. [19] Ramakrishnan, V. (1997) Histone structure and the organization of the nucleosome. Annu. Rev. Biophys. Biomol. Struct. 26, 83–112. [20] Maman, J.D., Yager, T.D. and Allan, J. (1994) Self-association of the globular domain of histone H5. Biochemistry 33, 1300–1310. [21] Liu, D., Safari, A., OÕConnor, M.S., Chan, D.W., Laegeler, A., Qin, J. and Songyang, Z. (2004) PTOP interacts with POT1 and regulates its localization to telomeres. Nat. Cell Biol. 6, 673–680. [22] Rotkova, G., Sykorova, E. and Fajkus, J. (2008) Protect and regulate: recent findings on plant POT1-like proteins. Biol. Plantarum. 52. [23] Mozgova, I, Prochazkova Schrumpfova, P., Hofr, C. and Fajkus, J. (2008) Functional characterization of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. Phytochemistry 69. [24] Hockemeyer, D., Daniels, J.P., Takai, H. and de Lange, T. (2006) Recent expansion of the telomeric complex in rodents: two distinct POT1 proteins protect mouse telomeres. Cell 126, 63–77. 1406 P.P. Schrumpfova´ et al. / FEBS Letters 582 (2008) 1400–1406 Supporting Information Supplementary Figure S1. AtTRB1 forms preferentially heterodimers with AtTRB2 and 3. Yeast two‐ hybrid strain MaV203 (with His3‐reporter gene under less tightly controlled promoter) was cotransformed with either full‐length AtTRB1 or FR1 fragment and with either of AtTRB FL constructs (for further details see Fig. 1). Drop test was performed with ‐His plates containing 5 mM and/or 20 mM 3‐AT. At the 20 mM 3‐AT concentration only AtTRB2 and AtTRB3 transformants grew suggesting that heterotypic complexes of AtTRB1 are more stable than homotypic ones. Supplementary Figure S2. Gene expression in various organs of A. thaliana. A. Gene expression of AtTRB1, 2 and 3 proteins (=SMH). Original data from chip expression database AtGenExpress were graphically transformed in Arabidopsis Gene Family Profiler (http://agfp.ueb.cas.cz) and modified for this article. B. Gene expression data for AtPOT1b protein are not available in any of general chip database. A schematic table of AtPOT1b expression is based on RT‐PCR data published in [6]. 105 Supplement D Mozgová, I., Schrumpfová, P.P., Hofr, C., Fajkus, J., 2008. Functional characterization of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. Phytochemistry 69, 1814–1819 P.P.S. participated in the experiments design, evaluation of data and the ms editing This journal did not provide open access, hence the article is not freely available. Supplement E Hofr, C., Sultesová, P., Zimmermann, M., Mozgová, I., Schrumpfová, P.P., Wimmerová, M., Fajkus, J., 2009. Single-Myb-histone proteins from Arabidopsis thaliana: a quantitative study of telomere-binding specificity and kinetics. Biochem. J. 419, 221–228 P.P.S. was involved in the experimental part (protein cloning, expression and purification) Biochem. J. (2009) 419, 221–228 (Printed in Great Britain) doi:10.1042/BJ20082195 221 Single-Myb-histone proteins from Arabidopsis thaliana: a quantitative study of telomere-binding specificity and kinetics Ctirad HOFR*1 , Pavla ˇSULTESOV´A*, Michal ZIMMERMANN*, Iva MOZGOV´A*, Petra PROCH´AZKOV´A SCHRUMPFOV´A*, Michaela WIMMEROV´A† and Jiˇr´ı FAJKUS*‡1 *Department of Functional Genomics and Proteomics, Institute of Experimental Biology, Faculty of Science, Masaryk University, CZ-62500 Brno, Czech Republic, †National Centre for Biomolecular Research and Department of Biochemistry, Faculty of Science, Masaryk University, CZ-61137 Brno, Czech Republic, and ‡Laboratory of DNA–Molecular Complexes, Institute of Biophysics, Czech Academy of Sciences, CZ-61265 Brno, Czech Republic Proteins that bind telomeric DNA modulate the structure of chromosome ends and control telomere function and maintenance. It has been shown that AtTRB (Arabidopsis thaliana telomererepeat-binding factor) proteins from the SMH (single-Mybhistone) family selectively bind double-stranded telomeric DNA and interact with the telomeric protein AtPOT1b (A. thaliana protection of telomeres 1b), which is involved in telomere capping. In the present study, we performed the first quantitative DNA-binding study of this plant-specific family of proteins. Interactions of full-length proteins AtTRB1 and AtTRB3 with telomeric DNA were analysed by electrophoretic mobility-shift assay, fluorescence anisotropy and surface plasmon resonance to reveal their binding stoichiometry and kinetics. Kinetic analyses at different salt conditions enabled us to estimate the electrostatic component of binding and explain different affinities of the two proteins to telomeric DNA. On the basis of available data, a putative model explaining the binding stoichiometry and the protein arrangement on telomeric DNA is presented. Key words: Arabidopsis thaliana, fluorescence anisotropy, kinetics, single-Myb-histone protein (SMH protein), surface plasmon resonance, telomere protein–DNA interaction. INTRODUCTION Telomeres are nucleoprotein complexes consisting of repetitive DNA sequences, general chromatin proteins and telomere-specific proteins. Tandem repeats of telomeric DNA are short T- and G-rich sequences, such as d(GGGTTA) in humans and d(GGGTTTA) in the majority of plants. Telomeres form protective capping structures at the ends of chromosomes [1]. These structures are essential for cell viability as they prevent chromosomes from unwanted end-to-end joining and recognition of chromosome tips as unrepaired double-strand breaks by the repair system of the cell. Changes in telomere structure and function induce chromosomal abnormalities and are directly connected with human aging and cancer [2]. Telomeres are usually maintained by telomerase, a ribonucleoprotein enzyme that adds telomeric repeats to the 3 -overhang of the G-rich DNA strand. The action of telomerase is regulated by its expression and by numerous proteins that control telomerase access to telomeres and organize telomeres into specific capping structures, such as telomeric loops that were observed in a number of organisms, including humans and plants [3,4]. Three DNA-binding proteins have been found to be responsible for specific recognition and direct interactions with the telomeric repeat sequence in humans. Two of them, TRF1 and TRF2 (where TRF is telomeric repeat-binding factor), described as negative regulators of telomere length [5], show substantial structural similarity and bind double-stranded telomeric DNA. The third protein, POT1 (protection of telomeres 1), binds the G-rich strand of telomeric DNA, participates in chromosome capping and is able to control telomere extension by telomerase, both positively and negatively [6,7]. The human TRFs and their homologues in other organisms possess a well conserved DNA-binding structural motif similar to the c-Myb-family of transcriptional activators [8]. The Myb domain of TRFs is C-terminally positioned and consists of three helices connected in a helix–turn–helix manner. The third helix contains a conserved amino acid sequence called a ‘telobox’, which has been shown to be important for recognition of telomeric double-stranded DNA [8]. Numerous TRF-like proteins have been identified in plants (reviewed in [9]), and, in a few cases, the influence of these proteins on telomere length homoeostasis has been demonstrated [10,11]. Interestingly, besides the TRF-like proteins, a plantspecific family of other telobox proteins has been described [12]. This group of proteins, termed the SMH (single-Myb-histone) family, is characterized by a triple-domain structure consisting of an N-terminal Myb domain, central globular histone H1/5 domain, and a C-terminal coiled-coil domain. In Arabidopsis thaliana (At), five SMH proteins were identified (AtTRB1–AtTRB5, where TRB is telomere-repeat-binding factor) [12], and three of them have been characterized [13,14]. These proteins show not only specific interactions with telomeric DNA, but also a number of protein–protein interactions functionally related to telomeres. In addition to their ability to form homodimers (similarly to TRFs), they can also form heterodimers and both homo- and heterotypic multimers [13–15] via their H1/5 histone domain. They also interact (using the same H1/5 domain) with one of the POT1 proteins in A. thaliana, AtPOT1b [15,16], which participates in telomere capping [17]. The emerging complexity of interactions of AtTRBs urges more detailed and quantitative studies of their DNA–protein and protein–protein interactions to reveal principles of their regulatory role. So far, only structural data for the Myb DNA-binding Abbreviations used: At, Arabidopsis thaliana; EMSA, electrophoretic mobility-shift assay; FA, fluorescence anisotropy; LB, Luria–Bertani; POT1, protection of telomeres 1; RedX, Rhodamine Red-X; RT, reverse transcription; SMH, single-Myb-histone; SPR, surface plasmon resonance; TRB, telomererepeat-binding factor; TRF, telomeric repeat-binding factor. 1 Correspondence may be addressed to either of these authors (email hofr@sci.muni.cz or fajkus@sci.muni.cz). c The Authors Journal compilation c 2009 Biochemical Society www.biochemj.org BiochemicalJournal 222 C. Hofr and others domain are available [18]. Similarly, kinetic studies are limited to the interaction of a Myb-domain-bearing fragment with a short telomeric DNA oligonucleotide (13 bp) [18], and a nonequilibrium technique was used to describe binding kinetics of TRFs in rice [19]. In order to describe binding interactions more thoroughly, associations of the full-length proteins with telomeric DNA need to be evaluated. The equilibrium binding kinetics of the full-length proteins can be studied by quantitative biophysical approaches. The binding of proteins to fluorescently labelled DNA may be monitored by FA (fluorescence anisotropy). This method gives well-resolved binding isotherms at different buffer conditions and therefore reliable kinetic and energetic parameters of binding. If the solution contains only free DNA molecules, FA is relatively low, owing to the fast rotational rearrangement of DNA molecules. After the binding of protein to DNA, a bulky slower-rotating protein– DNA complex is formed and the anisotropy is increased. Thus the anisotropy change of fluorescently labelled DNA duplexes, after each addition of protein into solution, describes the extent of protein–DNA binding [20,21]. In the present paper, we report a detailed study to reveal stoichiometry and kinetics of AtTRB1 and AtTRB3 binding to telomeric DNA. Proteins AtTRB1 and AtTRB3 have been chosen for these functional assays because they showed the highest structural stability within the AtTRB family of proteins. Interactions of full-length proteins with telomeric DNA are analysed by a combination of EMSA (electrophoretic mobilityshift assay) and quantitative biophysical methods employing FA and SPR (surface plasmon resonance). Kinetic analyses at different salt conditions enable us to estimate the electrostatic component of binding and explain different affinities of the two AtTRBs to telomeric and non-telomeric DNA. The kinetic measurements also contribute to the estimation of the length of double-stranded DNA for proper protein binding. On the basis of these data, a speculative model for binding stoichiometry and protein arrangement on telomeric DNA is presented. EXPERIMENTAL Cloning, expression and purification of AtTRB1 and AtTRB3 The cDNA sequence of AtTRB1 (locus At1g49950) was obtained by RT (reverse transcription)–PCR from total RNA as described previously [16]. AtTRB1 has been cloned into pET15b vector (Novagen) and expressed as a His-tagged fusion protein in Escherichia coli C41(DE3) cells [14]. The cells were grown on LB (Luria–Bertani) medium with ampicillin (100 μg/ml) at 37◦ C overnight. The next day, cells were diluted 20-fold into ZYM 5052 complex autoinducing medium containing ampicilin [22]. The cells were incubated at 37◦ C for 5 h. Then the temperature was set to 20◦ C, and the incubation continued overnight. The cDNA sequence of AtTRB3 (locus At3g49850) was obtained by RT–PCR from total RNA as described previously [13]. AtTRB3 has been cloned into pET30a(+) vector (Novagen) and expressed as a His-tagged fusion protein in E. coli BL21(DE3)pLysS cells. The cells were grown on LB medium with kanamycin (50 μg/ml) at 37◦ C for 4 h. At a D600 of 0.6, the overexpression of AtTRB3 was induced by the addition of IPTG (isopropyl β-D-thiogalactoside) to a concentration of 1 mM. After lowering the incubation temperature to 25◦ C, the growth continued for an additional 3 h. The following extraction and purification steps were the same for both recombinant proteins. After harvesting by centrifugation at 8000 g for 8 min, the pellet was dissolved in buffer containing 50 mM sodium phosphate (pH 8.0) with 300 mM NaCl and Figure 1 Proteins and oligonucleotide duplexes used for binding studies (a) Organization of the AtTRB1 and AtTRB3 polypeptide chains. The localization of the Myb domain, histone-like H1/5 domain and coiled-coil domain is shown together with numbers denoting their positions in the sequence. (b) Base sequence of telomeric oligonucleotide duplex R4 and R2 along with non-telomeric duplex N. RedX denotes fluorescent label Rhodamine RedX. The nucleotides of putative Myb-domain-binding sites are shaded grey [18]. 10 mM imidazole and was sonicated for 5 min. The sonicated cell extract was cleared by centrifugation at 14000 rev./min for 1 h at 4◦ C using a Beckman JA 14 rotor and subsequent filtration (0.45 μm filter). Affinity purification was performed on a column filled with a TALON® metal-affinity resin (BD Biosciences). Protein was eluted at 80 mM imidazole. The eluent was loaded on to a heparin HiTrapTM column (GE Healthcare). A concentration gradient of NaCl from 0.4 to 1 M NaCl was used for protein elution. The fractions containing pure protein were concentrated, and buffer-exchanged usually into 50 mM sodium phosphate (pH 7.5) with 100 mM NaCl by ultrafitration (Amicon 10K, Millipore) or by extensive dialysis. A typical yield was 1 mg of purified protein per litre of bacterial culture. The concentration of purified protein was determined using the Bradford assay [23]. DNA substrates Oligodeoxynucleotides were synthesized and HPLC-purified by Core Laboratory at Masaryk University. One of the strands in the duplexes was synthesized with the 3 -end C6 aminoalkyl linker and labelled with RedX (Rhodamine Red-X) (Molecular Probes) using the protocol provided by the manufacturer. The duplexes comprising four and two telomeric repeats were denoted as R4 and R2 respectively. The DNA duplex with non-telomeric sequence was denoted as N. The molar absorption coefficients of the single strands were estimated with the employment of phosphate assay [24]. Molar absorption coefficients were 281000 (RedX-labelled strand in R4), 278000 (complementary strand in R4), 148000 (RedX-labelled strand in R2), 140000 (complementary strand in R2), 284000 (RedX-labelled strand in N) and 265000 M−1 · cm−1 (complementary strand in R2) for DNA oligonucleotides shown in Figure 1. EMSA Protein–DNA-binding reactions were performed in 10 μl volumes containing the same amount of labelled DNA duplex (30 pmol) and various concentrations of protein (0–180 pmol) in 50 mM sodium phosphate (pH 7.0) with 200 mM NaCl. c The Authors Journal compilation c 2009 Biochemical Society DNA-binding kinetics of SMH proteins 223 Reaction mixtures were incubated for 10 min at 25◦ C. Protein– DNA complexes were resolved on horizontal 7.5% (w/v) acrylamide/0.3% bisacrylamide gels, as described in [25]. The electrophoresis proceeded at 1.5 V/cm for 30 min and for an additional 90 min at 3 V/cm. Gels were analysed with a LAS 3000 imaging system (Fujifilm). After the fluorescence imaging, Coomassie Blue staining of the gel was performed to reveal protein-containing bands in the gel. Fluorescence anisotropy Florescence anisotropy was measured on a FluoroMax-4 spectrofluorimeter (Horiba) with an L-format set up under control of an Origin-based FluorEssence software (version 2.1.6). Excitation and emission wavelengths were 572 and 591 nm respectively, with the same excitation and emission bandpath, 8 nm. The integration time was 3s. For each anisotropy value, five measurements were averaged. The titration experiments were carried out in a 10 mm × 4 mm quartz-glass cuvette with a magnetic bar stirrer. All measurements were conducted at 25◦ C in 50 mM sodium phosphate buffer (pH 7.5) containing 100 mM NaCl if not stated otherwise. To 1500 μl of DNA solution (20 nM) in the buffer, protein solution was added stepwise. The decrease in DNA concentration during the titration was taken into account in the analysis of the data. A control titration of protein to RedX solution (without DNA) has been performed to confirm that there was no interaction between RedX and protein. Dissociation constants of protein binding were evaluated by fitting of dilution-corrected binding isotherms using programs SigmaPlot 8 (Systat Software) and DynaFit3 (version 3.28) [26]. Analysis of the binding of protein to DNA duplexes was performed with the assumption of a non-co-operative binding mode. The association constants were calculated as reciprocal values of dissociation constants (Ka = 1/Kd). The association constants provided the free energies of association. Electrostatic component of binding In order to determine the contribution of electrostatic interactions upon binding of DNA with protein, the equilibrium binding constant was measured at different concentrations of NaCl (see Figure 4 and Table 2). The electrostatic component of binding originates from the formation of ion pairs between the cationic amino acid residues of the protein and the negatively charged DNA. The number of ion pairs formed upon protein–DNA binding and corresponding electrostatic contribution to overall binding affinity (Ka) could be derived from the dependence of the binding constant on salt concentration according to the eqn (1): log Ka = log Ka nel − Zϕ · log [NaCl] (1) where Z is the number of DNA phosphates that interact with the protein, ϕ is the number of Na+ cations per phosphate released upon protein binding. For B–DNA duplexes of 24 bp and shorter, the value for ϕ is approx. 0.64 [27]. The right-hand side of the equation divides overall binding affinity into the non-electrostatic part described by Ka nel and a salt-dependent electrostatic part [28,29]. When the linear dependence of log Ka is extrapolated to the salt concentration of 1 M, the electrostatic term in eqn (1) can be removed: log Ka = log Ka nel , i.e. the binding affinity is given only by non-electrostatic interactions. Similarly to the binding affinity, the overall binding energy defined as Ga = −2.3RT · log Ka could be divided into electrostatic and non-electrostatic terms Ga = Ga nel + Ga el . The electrostatic term Ga el disappears when the salt concentration approaches 1 M and the overall energy of binding is given only by the non-electrostatic term, Ga = Ga nel = −2.3RT · log Ka nel . Surface plasmon resonance Sensorgrams were recorded on a Biacore 3000 instrument (GE Healthcare) using CM5 chips. More details are available in the Supplementary Online Data at http://www.BiochemJ.org/bj/419/ bj4190221add.htm. RESULTS Stoichiometry of protein–DNA complexes In order to estimate the binding ratio of AtTRBs and DNA, oligonucleotide substrates containing two or one putative binding sites were designed. The telomeric duplex R4 covers the length of four plant telomeric repeats and comprises two putative Mybdomain-binding sites. The shorter duplex, double-stranded DNA fragment R2, consists of two telomeric repeats and contains one Myb-domain-binding site. For comparative purposes, oligonucleotide duplex N, as a representative of non-telomeric DNA, was used in the present study (Figure 1). Both AtTRB1 and AtTRB3 bind telomeric DNA with the stoichiometry of one protein monomer per one telomeric repeat The binding stoichiometry was analysed by EMSA with samples containing a variable protein/DNA ratio. Figure 2 shows fluorescently visualized bands indicating the mobility of free and protein-bound DNA duplexes in non-denaturating acrylamide gels. Increasing the concentration of protein shifted the free labelled DNA duplex to a new position corresponding to a protein–DNA complex. The band corresponding to the free duplex R4 disappeared when the AtTRB1/R4 ratio was 4:1 (Figure 2a). Similarly, the complete binding of AtTRB3 to substrate R4 was observed at the same protein/DNA ratio (Figure 2b). Both AtTRB1 and AtTRB3 bind telomeric DNA with the stoichiometry of one protein monomer per one telomeric repeat. In order to characterize interaction stoichiometry of AtTRBs with telomeric DNA further, proteins were allowed to interact with the shorter substrate R2 bearing two telomeric repetitions (Figure 2c). The results of EMSA with R2 demonstrate that a 2-fold decrease in the length of DNA reduces the protein/DNA binding ratio proportionally. These results confirmed that the stoichiometry of binding is one monomer of AtTRB1 or AtTRB3 per one telomeric repeat. If we consider binding of protein in dimeric form, as was shown in our recent study [14], then two protein dimers bind one R4 substrate (four telomeric repeats) or, in other words, one dimer of AtTRB binds the fragment R2 (two telomeric repeats). On the basis of these data, we could rephrase our initial statement regarding stoichiometry to the following form: one dimer of AtTRB binds the region of two telomeric repeats. AtTRB1 shows the same binding stoichiometry for telomeric and non-telomeric DNA sequences, whereas AtTRB3 exhibits different binding capacities for telomeric and non-telomeric DNA sequences The effect of DNA sequence on binding ability of AtTRB1 and AtTRB3 was analysed by comparing the protein/DNA ratio needed for complete saturation of telomeric R4 and non-telomeric N substrate. In this respect, AtTRB1 behaves similarly in both cases; the binding stoichiometry of AtTRB1 remained the same, as demonstrated in Figure 2(a). c The Authors Journal compilation c 2009 Biochemical Society 224 C. Hofr and others Figure 2 Non-denaturing EMSA (a) AtTRB1 binding to fluorescently labelled oligonucleotide R4 with telomeric sequence and oligonucleotide N with non-telomeric sequence. (b) AtTRB3 binding to DNA oligonucleotides R4 and N. (c) AtTRB1 or AtTRB3 binding to oligonucleotide R2 with the sequence of two telomeric repetitions. The DNA oligonucleotides and AtTRBs were incubated with increasing amounts of protein. The numbers under electrophoretic lanes denote the stoichiometric protein/DNA ratio. The protein/DNA ratio corresponding to binding saturation is indicated with a grey line. In contrast, AtTRB3 exhibits a markedly stronger dependence of the binding ability on DNA sequence that was manifested by a shift in ratio needed for saturation of the non-telomeric substrate N. The protein/DNA ratio was shifted to the higher values (>5:1) in the case of duplex N than was the ratio for the telomeric duplex R4 (Figure 2b). The difference in DNA-sequence-dependent saturation might be a result of different binding kinetics of AtTRB1 and AtTRB3. To assess this possibility, direct kinetic measurements were performed using FA. Binding kinetics The binding affinity of AtTRB variants to double-stranded DNA was analysed further by FA measurements. In these measurements, protein aliquots were added to the solution of labelled DNA duplex, and an increase of FA was observed. The equilibrium dissociation constants (Kd) obtained by analyses of anisotropy curves for binding are listed in Table 1. AtTRB1 and AtTRB3 bind telomeric DNA with high affinity and specificity The binding affinity of AtTRB1 to telomeric DNA is significantly higher in comparison with the binding to non-telomeric DNA. The titration curves obtained for AtTRB1 binding to DNA substrates Table 1 Dissociation and association constants for binding of AtTRB1 and AtTRB3 to DNA Values are means + − S.E.M. for three independent experiments in 50 mM sodium phosphate (pH 7.5) and 100 mM NaCl measured at 25◦C. R4 N R2 Ka Ka Ka Protein Kd (nM) (10−6 M−1 ) Kd (nM) (10−6 M−1 ) Kd (nM) (10−6 M−1 ) AtTRB1 90 + − 20 11.0 1200 + − 300 0.83 210 + − 30 4.8 AtTRB3 400 + − 60 2.5 2900 + − 300 0.35 800 + − 100 1.3 Figure 3 Binding of AtTRB1 and AtTRB3 to DNA duplexes (a) FA measurements of binding of AtTRB1 to telomeric duplex R4 or non-telomeric duplex N. The binding at 20 nM DNA occurred in buffer containing 50 mM sodium phosphate (pH 7.5) and 100 mM NaCl. (b) FA measurements of binding of AtTRB3 to R4 or N duplex. Binding conditions were the same as in (a). (c) Binding isotherms of AtTRB1 and AtTRB3 with telomeric duplex R2 measured by FA. Binding conditions were the same as in (a). R4 and N are shown in Figure 3(a). As expected, AtTRB1 shows significantly higher binding affinity to telomeric R4 than to the non-telomeric N DNA substrate. This can be clearly seen from the steeper rise of the curve corresponding to binding telomeric DNA. The evaluation of binding curves revealed Kd values of 90 and 1200 nM for R4 and N substrate respectively (Table 1). Comparison of dissociation constants thus demonstrates more than 13-fold higher affinity and binding specificity of AtTRB1 to DNA bearing telomeric sequences. The binding affinity of AtTRB3 to telomeric sequence is higher in comparison with the binding to non-telomeric sequence, but the difference is less pronounced than in case of AtTRB1. AtTRB3 was allowed to bind either the telomeric substrate R4 or the nontelomeric duplex N (Figure 3b). The Kd values for the binding of AtTRB3 to R4 and N were 400 and 2900 nM respectively. c The Authors Journal compilation c 2009 Biochemical Society DNA-binding kinetics of SMH proteins 225 Figure 4 Dependence of the association constants for binding of AtTRB1 and AtTRB3 to substrate R4 on NaCl concentration The inset shows the electrostatic and non-electrostatic components of the free energy of association of AtTRB1 or AtTRB3 with substrate R4. AtTRB3 shows more than 7-fold higher affinity to telomeric DNA duplex than to non-telomeric DNA. The absolute value of the dissociation constant was verified by SPR In order to confirm the absolute values of binding constants obtained using FA, a reverse-order experiment was performed using SPR. In this experiment, AtTRB3 was immobilized on the chip surface, and duplex R4 was allowed to bind. The reverse arrangement of the SPR experiment changes interaction stoichiometry (one DNA duplex interacts with one immobilized protein, whereas four protein monomers bind one DNA duplex during FA measurements). This had been considered when the equilibrium binding constant was evaluated. The output of the non-linear fitting of SPR curves for different concentrations of DNA produces a Kd of 1700 nM, which agrees with the value determined previously with a factor of 2 at a similar salt concentration (see the Supplementary Online Data). AtTRB1 and AtTRB3 show reduced binding affinity to R2 when compared with binding affinity to R4 When the length of DNA duplex is shortened from four to two telomeric repeats, the binding affinity decreases to the level of binding affinity recorded for the non-telomeric DNA. Even though there is one putative binding site present on the duplex R2, the binding affinity of AtTRB1 is quite low and is characterized by a Kd similar to that obtained for binding to duplex N. The shortening of telomeric DNA substrate has a similar effect on binding affinity of AtTRB3 (Table 1). The length reduction of telomeric DNA substrate thus results in a substantial fall in the binding affinity of both AtTRB1 and AtTRB3. Electrostatic contribution to binding affinity The binding of AtTRB1 or AtTRB3 to duplex R4 containing two putative binding sites induces the formation of four or three ion pairs respectively. Binding of both proteins to the substrate R4 was measured at different concentrations of NaCl. The change of binding parameters is set out in the double-log-plot of the association constants against salt concentration (Figure 4 and Table 2). From the slope, the parameter Z was calculated. Z denotes the number of newly formed ionic bonds between protein and DNA. This number is 4 (after rounding) for binding of Table 2 Salt-concentration-dependence of association constants for binding of AtTRB1 and AtTRB3 to R4 Values are means + − S.E.M. Protein [NaCl] (mM) log Ka δlog Ka/δlog [NaCl] log Ka nel Z AtTRB1 100 7.04 2.8 + − 0.2 4.31 + − 0.2 4.4 119 7.08 141 6.67 167 6.54 200 6.25 AtTRB3 100 6.41 2.0 + − 0.1 4.4 + − 0.1 3.2 119 6.31 141 6.06 167 6.00 200 5.81 AtTRB1, and 3 for AtTRB3. Thus approx. four ion pairs are formed upon binding of AtTRB1 and approx. three ion pairs upon binding of AtTRB3 to the telomeric DNA. The binding energy is provided mainly by a non-electrostatic component in the case of both AtTRBs Further evaluation of the salt-dependent binding constant was performed to obtain the non-electrostatic contribution to the binding affinity. The electrostatic and non-electrostatic components of the binding energy for AtTRB1 or AtTRB3 to R4 are shown in the inset of Figure 4. It is notable that the nonelectrostatic components of binding energy Ga nel for the two proteins are identical within error range with magnitudes of 25 kJ · mol−1 for both AtTRB1 and AtTRB3. If this value is compared with the values of the overall binding energy 40 kJ · mol−1 for AtTRB1 and 37 kJ · mol−1 for AtTRB3, it can be concluded that the non-electrostatic interactions contribute to the total energy of binding by approx. 60% for AtTRB1 and by approx. 70% for AtTRB3. Hence, it is apparent that the major part of the binding energy originates from the non-electrostatic interactions. The greater electrostatic component is responsible for a more favourable overall binding energy of AtTRB1 compared with AtTRB3 Further inspection of calculated energetic data allowed us to identify the main reason for different binding affinities between these similar proteins. It is demonstrated that the kinetics of protein–DNA interactions are different because of the electrostatic term of the binding energy (inset in Figure 4). In other words, the difference in the total binding energy for AtTRB1 and AtTRB3 is entirely given by the change in the electrostatic component of binding. DISCUSSION Kinetics and stoichiometry of binding The present study shows that binding of AtTRB1 and AtTRB3 with the telomeric DNA proceeds with the stoichiometry of one protein monomer per one telomeric repeat. A higher protein/DNA ratio was observed only in case of AtTRB3 binding to nontelomeric DNA (Figure 2b). The shift in the ratio can be explained by the observed lower affinity of AtTRB3 for non-telomeric DNA. The decrease in binding affinity with the change from telomeric to non-telomeric sequence was confirmed also by our kinetic measurements (Table 1). All recently characterized AtTRBs form tightly bound homo- and hetero-dimers and multimers [14,15]. Relatively strong mutual interactions of AtTRBs were also c The Authors Journal compilation c 2009 Biochemical Society 226 C. Hofr and others verified independently using SPR (results not shown) and their dimerization ability was demonstrated by gel chromatography (see the Supplementary Online Data). Therefore the feasibility of protein dimerization and stoichiometric data of the present study support the assumption that AtTRBs bind to DNA in dimeric form. In this respect, the AtTRBs behave similarly to human TRF1 and TRF2 [30–32], with the exception that TRFs do not form heterodimers. Surprisingly, the affinity of AtTRB1 to telomeric substrate R4 is 4-fold higher than that of AtTRB3, although AtTRB1 and AtTRB3 are relatively similar in their primary sequences. Interestingly, it has been found that Kd values observed in the present study for AtTRB1 and AtTRB3 correspond very well to Kd values obtained for the DNA-binding domain of human TRF1 and TRF2 when interacting with telomeric DNA [33]. Moreover, similarly to AtTRB1 and AtTRB3, human TRF1 binds telomeric DNA with a 4-fold higher affinity than that of TRF2. In order to explain potential reasons for the different binding manner of AtTRB1 and AtTRB3, we compared our findings with available equilibrium kinetic data for the binding of the Myb domain. The Kd obtained for the binding of the Myb domain alone to telomeric DNA from NMR studies was in the range of 1 μM [18]. If we compare this value measured at physiological salt concentration with the values for the binding of full-length proteins measured in the present study at a corresponding NaCl concentration, the magnitude of Kd for AtTRB3 is slightly lower at 0.9 μM (see Table 1), and the Kd for AtTRB1 is significantly lower (0.2 μM). Both full-length proteins showed higher binding capacities than that reported for a Myb domain alone. Since the Myb domain sequence is highly conserved between AtTRB1 and AtTRB3, the higher binding affinity of AtTRB1 should originate from another part of the protein. The domain that may contribute to the tuning of binding affinity of AtTRBs to DNA is the H1/5 domain [13], as supported by our recent findings [14]. The conservation of the H1/5 domain between AtTRB1 and AtTRB3 is lower than that of the Myb domain and differs in a way that might allow the corresponding protein region to adopt a structure with a different net charge on the surface. The surface net charge is important for a long-range non-specific electrostatic attraction among proteins and DNA, whereas non-electrostatic interactions that are important for specific recognition of a DNA sequence comprise hydrogen bonds between outer groups of DNA and polar residues of the protein [18]. Electrostatic component of binding Proteins controlling and regulating nucleic acid structure and function usually show both sequence-non-specific binding to DNA and a higher-affinity binding of their specific physiological DNA target. In general, protein–DNA binding takes place in two steps. In the first step, a non-specific, mainly electrostatic, binding to the phosphate backbone occurs; in the second step, the protein explores the DNA surface for specific non-electrostatic interactions such as hydrogen bonds [34]. Different contributions of electrostatic and non-electrostatic interactions to binding were observed for different classes of DNA-binding proteins. For example, telomere-binding protein α from Oxytricha nova induces the formation of two ion pairs upon binding to DNA, and the electrostatic contribution to the free energy of binding is approx. 15% [25]. On the other hand, proteins containing a strongly positively charged scissor-grip motif for DNA recognition induce the formation of six ion pairs with the electrostatic contribution to the total free energy of binding being 45% [29]. The different contribution of electrostatic attraction for binding of AtTRB1 and AtTRB3 was observed. We estimated the number of four and three ion pairs upon AtTRB1 or AtTRB3 binding to R4 and the corresponding electrostatic contribution to the total free energy of binding at 40 and 30% respectively. This correlates well with data available for electrostatic interactions of other DNA-binding proteins. The DNA-binding event of AtTRBs is driven mainly by non-electrostatic interactions. On the whole, our results show that AtTRBs bind telomeric DNA primarily in a sequence-specific manner that is essential for the recognition of binding sites within telomeric DNA. Kinetic data contribute to understanding of nucleoprotein complex arrangement Analyses of our kinetic data together with available structural data may be also used to elucidate the arrangement of nucleoprotein complexes of AtTRBs with telomeric DNA. In general, one might suppose that the same binding preferences to telomeric DNA are given primarily by the occurrence of the recognition sequence in DNA. For this reason, one would also expect the same binding kinetics for the telomeric DNA with one or two putative binding sites under the consideration of a nonco-operative independent binding. As follows from the previous assumptions, the duplex R2, containing one binding site, should have reached the saturation of binding sites faster (Kd would be lower) when compared with that for duplex R4, with two binding sites. However, our data show the opposite. The binding affinity of both examined proteins to duplex R2 is lower (Kd is shifted to higher values) than in the case of binding to R4. Our quantitative kinetics results confirmed a previously reported decrease in binding affinity of AtTRBs with the shortening of telomeric DNA substrate [13]. Moreover, the lower affinity to DNA containing only one putative binding site might be an indication of an insufficient space for the binding of an active protein. Importantly, it has been shown that the minimum length of DNA for Myb domain binding is approx. 13 bp [18]. If AtTRBs interacted with the DNA exclusively through the Myb domain and binding sites were positioned suitably within the sequence, the length of R2 duplex (14 bp) should have been sufficient for proper binding without a change in binding affinity. Since a significant fall in binding affinity was observed, the kinetics data suggest that there is also another domain taking part in the interaction. As a result, the binding affinity of AtTRBs to the 14 bp long and 28 bp long DNA duplex differs substantially. In our recent results, the H1/5 domain promotes interaction with DNA [14]. Presumably, the short length might prevent the H1/5 domain from properly interacting with the DNA. Hence, the constrained binding without H1/5 domain might be the main reason for the reduction of the overall binding affinity to substrate R2. Although the picture of a molecular mechanism controlling telomerase activity is far from complete, it is important to consider how the protein-binding events measured in the present study relate to structural arrangements and subsequent interactions essential for the biology of telomeres. If we take into account the kinetic data and the dimerization ability of AtTRBs, a speculative protein arrangement on telomeric DNA could be considered (Figure 5). The model of binding arrangement considers that the protein monomers form a dimer that binds two adjacent binding sites simultaneously. This type of interaction mode is quite common in the sequence-specific binding of proteins that take part in regulatory mechanisms [35]. This model, where two recognition sites on DNA are bound by one protein dimer, might explain well the fall of binding activity when the DNA substrate is c The Authors Journal compilation c 2009 Biochemical Society DNA-binding kinetics of SMH proteins 227 Figure 5 Speculative model of interaction of AtTRBs with telomeric DNA Both homo- and hetero-dimers of AtTRB may participate in the interaction with telomeric DNA. shortened from 28 to 14 bp as was observed for binding to the R4 and R2 duplex respectively. The introduced model is supported by stoichiometric and kinetic data presented here and it is also in accordance with our previous study demonstrating weaker binding to DNA containing fewer telomeric repeats [13]. The binding arrangement shown in Figure 5 also takes into consideration the multimerization ability of the H1/5 domain that could promote the arrangement of protein monomers in the DNA region between the binding sites. Moreover, the formation of homo- and hetero-multimers of SMH proteins and their ability to interact with other proteins (e.g. AtPOT1b [15,16]) contribute to a network of protein interactions that could be employed in the organization of telomere to form highly ordered chromatin structures, such as t-loops, in a similar way to human TRFs [31,32]. Thus, on the basis of results of the present study and the data available, we suggest that interactions of the two AtTRBs with telomeric DNA occur simultaneously with two binding sites. Therefore the minimal length of duplex DNA required for the proper binding of full-length AtTRB1 and AtTRB3 should harbour at least two putative binding sites that are bound by two dimers of AtTRBs. Consequently, SMH proteins are able to distinguish between short (<10 bp) telomere-like sequences that are dispersed throughout the genome, e.g. in promoter regions [36], and longer tracts of telomere repeats occurring in telomeres. There is still a considerable lack of general knowledge of intracellular arrangement, molecular crowding effects, association mechanisms and kinetics of protein–DNA-binding events in a living cell. Nevertheless, we can draw a speculative view of the in vivo consequences of our in vitro data, if we consider that the behaviour of a protein would not be markedly changed in the cellular environment. The access of AtTRBs to their telomeric target sites is restricted in both spatial and temporal ways by chromatin structure: the telomeric heterochromatin structure provides low accessibility upon its tight condensation, and thus the binding of specific proteins to DNA may occur preferentially in a short time slot between DNA replication and chromatin condensation [37]. AtTRBs might be first recruited by a weak non-specific binding to multiple chromosome regions. Then, once the specific target sites become accessible, highly specific binding occurs. On the other hand, the AtTRB molecules which are bound only by a highly dynamic non-specific interaction (in non-telomeric regions) can be easily displaced by other proteins binding with a higher affinity. Thus AtTRBs at non-telomeric sites do not impede other functional DNA–protein interactions. In this way, non-specific binding could serve as a tool for increasing the local concentration of the proteins on DNA [34]. Accumulation of SMH proteins on DNA via non-specific electrostatic interactions may be important for their immediate availability for functional and specific binding to their telomere target sites. Although further details of the binding interactions of proteins and their biological significance have yet to be determined, these results demonstrate the advantage of the approach employed in the present study by using a complete protein for in vitro studies rather than the commonly used Myb-domain-bearing fragment. Our data imply that AtTRB1 and AtTRB3 are telomere-specific proteins that bind telomeric DNA with distinct kinetics given by differences in their electrostatic interactions with DNA. To our knowledge, this is the first quantitative study of the plant-specific SMH family of proteins. The present paper demonstrates that the detailed quantification of protein–DNA interactions may provide new insights into the structural dynamics of telomeres. ACKNOWLEDGEMENTS We are grateful to M. Chester for the critical reading of the manuscript. FUNDING This work was supported by the Grant Agency of the Czech Republic [grant numbers 521/ 08/P452 and 204/08/H054], the Czech Ministry of Education [grant number LC06004], the Grant Agency of the Czech Academy of Sciences [grant number IAA500040801] and the institutional support [grant numbers MSM0021622415, MSM0021622413, AV0Z50040702 and AV0Z50040507]. REFERENCES 1 Zakian, V. A. (1995) Telomeres: beginning to understand the end. Science 270, 1601–1607 2 Smogorzewska, A. and de Lange, T. (2004) Regulation of telomerase by telomeric proteins. Annu. Rev. Biochem. 73, 177–208 3 Griffith, J. D., Comeau, L., Rosenfield, S., Stansel, R. M., Bianchi, A., Moss, H. and de Lange, T. (1999) Mammalian telomeres end in a large duplex loop. Cell 97, 503–514 4 Cesare, A. J., Quinney, N., Willcox, S., Subramanian, D. and Griffith, J. D. (2003) Telomere looping in P. sativum (common garden pea). Plant J. 36, 271–279 5 Smogorzewska, A., van Steensel, B., Bianchi, A., Oelmann, S., Schaefer, M. R., Schnapp, G. and de Lange, T. (2000) Control of human telomere length by TRF1 and TRF2. Mol. Cell. Biol. 20, 1659–1668 6 Baumann, P. and Cech, T. R. (2001) Pot1, the putative telomere end-binding protein in fission yeast and humans. Science 292, 1171–1175 7 Hockemeyer, D., Sfeir, A. J., Shay, J. W., Wright, W. E. and de Lange, T. (2005) POT1 protects telomeres from a transient DNA damage response and determines how human chromosomes end. EMBO J. 24, 2667–2678 8 Bilaud, T., Koering, C. E., Binet-Brasselet, E., Ancelin, K., Pollice, A., Gasser, S. M. and Gilson, E. (1996) The telobox, a Myb-related telomeric DNA binding motif found in proteins from yeast, plants and human. Nucleic Acids Res. 24, 1294–1303 9 Kuchar, M. (2006) Plant telomere-binding proteins. Biol. Plant. 50, 1–7 10 Yang, S. W., Kim, S. K. and Kim, W. T. (2004) Perturbation of NgTRF1 expression induces apoptosis-like cell death in tobacco BY-2 cells and implicates NgTRF1 in the control of telomere length and stability. Plant Cell 16, 3370–3385 11 Hwang, M. G. and Cho, M. H. (2007) Arabidopsis thaliana telomeric DNA-binding protein 1 is required for telomere length homeostasis and its Myb-extension domain stabilizes plant telomeric DNA binding. Nucleic Acids Res. 35, 1333–1342 12 Marian, C. O., Bordoli, S. J., Goltz, M., Santarella, R. A., Jackson, L. P., Danilevskaya, O., Beckstette, M., Meeley, R. and Bass, H. W. (2003) The maize single myb histone 1 gene, Smh1, belongs to a novel gene family and encodes a protein that binds telomere DNA repeats in vitro. Plant Physiol. 133, 1336–1350 13 Schrumpfova, P., Kuchar, M., Mikova, G., Skrisovska, L., Kubicarova, T. and Fajkus, J. (2004) Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. Genome 47, 316–324 14 Mozgova, I., Prochazkova Schrumpfova, P., Hofr, C. and Fajkus, J. (2008) Functional characterisation of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. Phytochemistry 69, 1814–1819 c The Authors Journal compilation c 2009 Biochemical Society 228 C. Hofr and others 15 Prochazkova Schrumpfova, P., Kuchar, M., Palecek, J. and Fajkus, J. (2008) Mapping of interaction domains of putative telomere-binding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. FEBS Lett. 582, 1400–1406 16 Kuchar, M. and Fajkus, J. (2004) Interactions of putative telomere-binding proteins in Arabidopsis thaliana: identification of functional TRF2 homolog in plants. FEBS Lett. 578, 311–315 17 Shakirov, E. V., Surovtseva, Y. V., Osbun, N. and Shippen, D. E. (2005) The Arabidopsis Pot1 and Pot2 proteins function in telomere length homeostasis and chromosome end protection. Mol. Cell. Biol. 25, 7725–7733 18 Sue, S. C., Hsiao, H. H., Chung, B. C., Cheng, Y. H., Hsueh, K. L., Chen, C. M., Ho, C. H. and Huang, T. H. (2006) Solution structure of the Arabidopsis thaliana telomeric repeat-binding protein DNA binding domain: a new fold with an additional C-terminal helix. J. Mol. Biol. 356, 72–85 19 Byun, M. Y., Hong, J. P. and Kim, W. T. (2008) Identification and characterization of three telomere repeat-binding factors in rice. Biochem. Biophys. Res. Commun. 372, 85–90 20 Heyduk, T. and Lee, J. C. (1990) Application of fluorescence energy transfer and polarization to monitor Escherichia coli cAMP receptor protein and lac promoter interaction. Proc. Natl. Acad. Sci. U.S.A. 87, 1744–1748 21 LeTilly, V. and Royer, C. A. (1993) Fluorescence anisotropy assays implicate protein-protein interactions in regulating trp repressor DNA binding. Biochemistry 32, 7753–7758 22 Studier, F. W. (2005) Protein production by auto-induction in high density shaking cultures. Protein Expression Purif. 41, 207–234 23 Bradford, M. M. (1976) A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 72, 248–254 24 Murphy, J. H. and Trapane, T. L. (1996) Concentration and extinction coefficient determination for oligonucleotides and analogs using a general phosphate analysis. Anal. Biochem. 240, 273–282 25 Buczek, P. and Horvath, M. P. (2006) Thermodynamic characterization of binding Oxytricha nova single strand telomere DNA with the α protein N-terminal domain. J. Mol. Biol. 359, 1217–1234 26 Kuzmic, P. (1996) Program DYNAFIT for the analysis of enzyme kinetic data: application to HIV proteinase. Anal. Biochem. 237, 260–273 27 Olmsted, M. C., Bond, J. P., Anderson, C. F. and Record, Jr, M. T. (1995) Grand canonical Monte Carlo molecular and thermodynamic predictions of ion effects on binding of an oligocation (L8+) to the center of DNA oligomers. Biophys. J. 68, 634–647 28 Record, Jr, M. T., Zhang, W. and Anderson, C. F. (1998) Analysis of effects of salts and uncharged solutes on protein and nucleic acid equilibria and processes: a practical guide to recognizing and interpreting polyelectrolyte effects, Hofmeister effects, and osmotic effects of salts. Adv. Protein Chem. 51, 281–353 29 Dragan, A. I., Liu, Y., Makeyeva, E. N. and Privalov, P. L. (2004) DNA-binding domain of GCN4 induces bending of both the ATF/CREB and AP-1 binding sites of DNA. Nucleic Acids Res. 32, 5192–5197 30 Bianchi, A., Smith, S., Chong, L., Elias, P. and de Lange, T. (1997) TRF1 is a dimer and bends telomeric DNA. EMBO J. 16, 1785–1794 31 Fairall, L., Chapman, L., Moss, H., de Lange, T. and Rhodes, D. (2001) Structure of the TRFH dimerization domain of the human telomeric proteins TRF1 and TRF2. Mol. Cell 8, 351–361 32 Khan, S. J., Yanez, G., Seldeen, K., Wang, H., Lindsay, S. M. and Fletcher, T. M. (2007) Interactions of TRF2 with model telomeric ends. Biochem. Biophys. Res. Commun. 363, 44–50 33 Hanaoka, S., Nagadoi, A. and Nishimura, Y. (2005) Comparison between TRF2 and TRF1 of their telomeric DNA-bound structures and DNA-binding activities. Protein Sci. 14, 119–130 34 Revzin, A. (1990) The Biology of Nonspecific DNA–Protein Interactions, CRC Press, Boca Raton 35 von Hippel, P. H. (2007) From “simple” DNA–protein interactions to the macromolecular machines of gene expression. Annu. Rev. Biophys. Biomol. Struct. 36, 79–105 36 Regad, F., Lebas, M. and Lescure, B. (1994) Interstitial telomeric repeats within the Arabidopsis thaliana genome. J. Mol. Biol. 239, 163–169 37 Fajkus, J. and Trifonov, E. N. (2001) Columnar packing of telomeric nucleosomes. Biochem. Biophys. Res. Commun. 280, 961–963 Received 7 November 2008/18 December 2008; accepted 22 December 2008 Published as BJ Immediate Publication 22 December 2008, doi:10.1042/BJ20082195 c The Authors Journal compilation c 2009 Biochemical Society Biochem. J. (2009) 419, 221–228 (Printed in Great Britain) doi:10.1042/BJ20082195 SUPPLEMENTARY ONLINE DATA Single-Myb-histone proteins from Arabidopsis thaliana: a quantitative study of telomere-binding specificity and kinetics Ctirad HOFR*1 , Pavla ˇSULTESOV´A*, Michal ZIMMERMANN*, Iva MOZGOV´A*, Petra PROCH´AZKOV´A SCHRUMPFOV´A*, Michaela WIMMEROV´A† and Jiˇr´ı FAJKUS*‡1 *Department of Functional Genomics and Proteomics, Institute of Experimental Biology, Faculty of Science, Masaryk University, CZ-62500 Brno, Czech Republic, †National Centre for Biomolecular Research and Department of Biochemistry, Faculty of Science, Masaryk University, CZ-61137 Brno, Czech Republic, and ‡Laboratory of DNA–Molecular Complexes, Institute of Biophysics, Czech Academy of Sciences, CZ-61265 Brno, Czech Republic EXPERIMENTAL Gel-filtration chromatography The molecular masses of the protein in monomeric and dimeric forms were estimated by size-exclusion gel-filtration chromatography through a Superdex 200 10/30 GL column (GE Healthcare), using a gel-filtration standard (Bio-Rad Laboratories) in a buffer containing 50 mM sodium phosphate (pH 7.5) and 300 mM NaCl. The molecular masses of proteins were estimated from a linear fit to the log Mr against elution volume plot generated with the protein standards. Supplementary Figure S1 shows the chromatograms. Surface plasmon resonance All SPR experiments were performed with a Biacore 3000 instrument (GE Healthcare) at 25◦ C using TBST (Tris-buffered saline with Tween 20: 10 mM Tris/HCl, pH 7.5, 150 mM NaCl, containing 0.005% Tween 20) and a flow rate of 5 μl/min. AtTRB3 was immobilized on the research-grade CM5 sensor chip in a buffer containing 10 mM Hepes, 150 mM NaCl (pH 7.5) and 0.005% Tween 20. Sensorgrams were run in the automatic subtraction mode using FC (flowcell) 1 as an unmodified reference. Data were collected for FC 2, FC 3 and FC 4, which contained various amounts of AtTRB3. Injections of DNA were made using the ‘quickinject’ injection mode, going from lowest to highest concentration samples, with a 5 min contact time and a 1200 s dissociation phase in all cases. Regeneration was achieved using several (two to five) 1 min pulses of 50 mM NaOH. All sensorgrams were obtained at 25◦ C. Data were analysed by equilibrium analysis in addition to the kinetic analysis. The equilibrium response was plotted against the concentration of DNA and fitted to: R = Ka[DNA]Rmax(Ka[DNA] + 1) where R is the equilibrium response at a specific concentration of DNA substrate, Rmax is the response at saturation of the DNA substrate on the chip, Ka is the equilibrium association constant, which is the reciprocal of the dissociation constant Kd (Ka = 1/Kd). When assuming a non-co-operative binding model, the apparent Kd from SPR experiments should be divided by 4 to resemble Figure S1 Size-exclusion chromatograms of protein AtTRB3 (continuous line) and molecular-mass standard (broken line) Abs, absorbance; MW standard, molecular-mass standard. The numbers next to the arrows indicate determined molecular-mass values of monomeric and dimeric protein forms. different binding stoichiometry of FA and SPR experiments. The output of the non-linear fitting of SPR curves for different concentrations of DNA produces a Kd of 6.8 μM, which, divided by 4, gives 1.7 μM. This value agrees well with the value of Kd determined from FA measurements considering different buffer conditions. Supplementary Figure S2 shows the sensorgrams and the response curve. Purification of AtTRBs AtTRB1 and AtTRB3 were expressed in soluble forms in cytoplasm of E. coli. The purification strategy consisted of two affinity steps. A capture step by IMAC (immobilized metalion-affinity chromatography) was followed by a purification step using HAC (heparin-affinity chromatography). To confirm final purity, collected fractions were separated by SDS/PAGE (0.1% SDS, 10% acrylamide). Supplementary Figure S3 shows the gelpurified proteins. 1 Correspondence may be addressed to either of these authors (email hofr@sci.muni.cz or fajkus@sci.muni.cz). c The Authors Journal compilation c 2009 Biochemical Society C. Hofr and others Figure S2 Binding of telomeric duplex R4 to immobilized AtTRB3 Response signals from the saturated region of the sensorgram have been used to calculate equilibrium dissociation constant Kd. RU, response units. Figure S3 Analysis of purification steps using SDS/PAGE Lane 1, collected fractions containing AtTRB3 after IMAC (immobilized metal-ion-affinity chromatography) and subsequent HAC (heparin-affinity chromatography) (10 μg); lane 2, collected fractions containing AtTRB3 after IMAC (15 μg); lanes M, molecular-mass markers (sizes are indicated in kDa); lane 5, collected fractions containing AtTRB1 after IMAC and subsequent HAC (3 μg); lane 6, clarified cytoplasmic extract with expressed AtTRB1 (35 μg). The proteins were stained with Coomassie Brilliant Blue. Received 7 November 2008/18 December 2008; accepted 22 December 2008 Published as BJ Immediate Publication 22 December 2008, doi:10.1042/BJ20082195 c The Authors Journal compilation c 2009 Biochemical Society Supplement F Peška, V., Schrumpfová, P.P., Fajkus, J., 2011. Using the telobox to search for plant telomere binding proteins. Curr. Protein Pept. Sci. 12, 75–83 P.P.S. participated in the ms writing and editing This journal did not provide open access, hence the article is not freely available. Supplement G Schrumpfová, P.P.*, Fojtová, M., Mokroš, P., Grasser, K.D., Fajkus, J., 2011. Role of HMGB proteins in chromatin dynamics and telomere maintenance in Arabidopsis thaliana. Curr. Protein Pept. Sci. 12, 105–111 P.P.S. was significantly involved in the experimental, data evaluation, and in the ms writing and editing This journal did not provide open access, hence the article is not freely available. Supplement H Schrumpfová, P.P.*, Vychodilová, I., Dvořáčková, M., Majerská, J., Dokládal, L., Schořová, S., Fajkus, J., 2014. Telomere repeat binding proteins are functional components of Arabidopsis telomeres and interact with telomerase. Plant J. Cell Mol. Biol. 77, 770–781. P.P.S. participated in the design of experiments, significantly involved in the experimental part, data evaluation, and in the ms writing and editing Telomere repeat binding proteins are functional components of Arabidopsis telomeres and interact with telomerase Petra Prochazkova Schrumpfova1,2, *, Ivona Vychodilova1,2 , Martina Dvorackova1,3 , Jana Majerska1,† , Ladislav Dokladal1,3 , Sarka Schorova1,2 and Jirı Fajkus1,2,3, * 1 Mendel Centre for Plant Genomics and Proteomics, Central European Institute of Technology, Masaryk University, Kamenice 5, Brno, CZ 62500, Czech Republic, 2 Functional Genomics and Proteomics, CEITEC National Centre for Biomolecular Research, Faculty of Science, Masaryk University, Kamenice 5, Brno, CZ 62500, Czech Republic, and 3 Institute of Biophysics, Academy of Sciences of the Czech Republic, v.v.i, Kralovopolska 135, Brno, CZ 61265, Czech Republic Received 11 October 2013; revised 6 December 2013; accepted 23 December 2013; published online 8 January 2014. *For correspondence (e-mails fajkus@sci.muni.cz or schpetra@centrum.cz). † Present address: Swiss Institute for Experimental Cancer Research, Ecole Polytechnique Federale de Lausanne, Station 19, 1015, Lausanne, Switzerland. SUMMARY Although telomere-binding proteins constitute an essential part of telomeres, in vivo data indicating the existence of a structure similar to mammalian shelterin complex in plants are limited. Partial characterization of a number of candidate proteins has not identified true components of plant shelterin or elucidated their functional mechanisms. Telomere repeat binding (TRB) proteins from Arabidopsis thaliana bind plant telomeric repeats through a Myb domain of the telobox type in vitro, and have been shown to interact with POT1b (Protection of telomeres 1). Here we demonstrate co-localization of TRB1 protein with telomeres in situ using fluorescence microscopy, as well as in vivo interaction using chromatin immunoprecipitation. Classification of the TRB1 protein as a component of plant telomeres is further confirmed by the observation of shortening of telomeres in knockout mutants of the trb1 gene. Moreover, TRB proteins physically interact with plant telomerase catalytic subunits. These findings integrate TRB proteins into the telomeric interactome of A. thaliana. Keywords: telomerase, telomere, telomere repeat binding (TRB), Arabidopsis thaliana, telomere protein interaction, plant shelterin. INTRODUCTION Telomeres, nucleoprotein structures that form and protect the ends of chromosomes, have been the subject of intense studies for about three decades, starting with a description of the telomere DNA component (Blackburn and Gall, 1978) and the most common system of telomere maintenance by the ribonucleoprotein complex of telomerase (Greider and Blackburn, 1985, 1989). Proteins essential for telomere functions have been described in detail in yeasts and vertebrates. Among protein components of telomeres, the most important is indisputably telomerase itself, but other proteins are necessary to perform other functions of telomeres, such as inhibiting the DNA damage response at telomeres (de Lange, 2009), recruiting telomerase to chromosome ends (Nandakumar et al., 2012), or facilitating telomere replication (Sfeir et al., 2009). Current evidence suggests that these components assemble into two distinct complexes known as shelterin (de Lange, 2005) and CST (composed of CTC1/STN1/TEN1 proteins) complexes (Surovtseva et al., 2009). Human shelterin consists of six core components: telomeric repeat-binding factor 1 (TRF1), telomeric repeatbinding factor 2 (TRF2), represor/activator protein 1 (RAP1), TRF1-interacting protein (TIN2), TINT1/PIP1/PTOP1 (TPP1), protection of telomeres 1 (POT1). TRF1 and TRF2 anchor the complex to double-stranded telomeric DNA using a specific Myb-like motif termed a telobox (Bilaud et al., 1996), and recruit two other shelterin components, RAP1 and TIN2, to the telomeres. TIN2 further interacts with TPP1 protein, which binds the final shelterin component, POT1. POT1 also binds the G–rich strand of telomeric DNA from either the single-stranded G–overhang or displacement loop (D–loop). In this way, shelterin may bridge the double- and single-stranded parts of telomeric DNA. © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd. This is an open access article under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs License, which permits use and distribution in any medium, provided the original work is properly cited, the use is non-commercial and no modifications or adaptations are made. 770 The Plant Journal (2014) 77, 770–781 doi: 10.1111/tpj.12428 The CST complex, consisting of three components (Cdc13, Stn1 and Ten1), was originally described in yeast (Gao et al., 2007) as a telomere-specific replication protein A-like complex that protects single-stranded chromosome termini and regulates telomere replication. Subsequent studies have shown that a CST-like complex also exists in plants and humans and contributes to telomere protection and replication (Surovtseva et al., 2009; Price et al., 2010). According to recent studies, both complexes participate in telomere capping, telomerase regulation and 3′ overhang formation (Giraud-Panis et al., 2010; Pinto et al., 2011; Chen et al., 2012; Wu et al., 2012). In contrast to the CST complex, no functional and structural equivalent of shelterin has been found in plants. Although many putative shelterin-like protein components have been found in plants (Peska et al., 2011), including those bearing a telobox Myb-like domain at their C–terminus (Hwang et al., 2001, 2005; Karamysheva et al., 2004) or N–terminus (Marian et al., 2003; Schrumpfova et al., 2004), as well as POT1 homologues (Baumann et al., 2002; Kuchar and Fajkus, 2004; Shakirov et al., 2005; Tani and Murata, 2005; Peska et al., 2008), none of these have been shown to specifically associate with telomeres in situ or in vivo. Molecular components responsible for reversible telomerase regulation in plant cells (Fajkus et al., 1998; Riha et al., 1998) are an attractive target for possible biomedical applications of telomere biology, and are sought primarily at the levels of protein components of plant telomeres, and regulation of the basic telomerase subunits TERT (telomerase reverse transcriptase) and TER (telomerase RNA). In this study, we investigated the interactions and roles of Single myb histone (Smh) proteins at plant telomeres. Five members of the Smh family are encoded by the A. thaliana genome (TRB1–5). These proteins are specific to plants, and consist of an N–terminal Myb-like domain of the telobox type, which is responsible for specific recognition of double/single-stranded telomeric DNA (Schrumpfova et al., 2004; Hofr et al., 2009), a central histone-like domain, which is involved in non-specific DNA–protein interactions and mediates protein–protein interactions, including formation of homo- and heteromeric complexes of TRB proteins (Mozgova et al., 2008), and a C–terminal coiled-coil domain to which no specific function has yet been attributed. We previously reported that TRB proteins interact via their histone-like domain with POT1b, an A. thaliana homologue of the G–overhang binding protein POT1 (Kuchar and Fajkus, 2004; Schrumpfova et al., 2008; Rotkova et al., 2009). In addition, POT1b also associates with an alternative telomerase nucleoprotein complex in Arabidopsis (Surovtseva et al., 2007; Cifuentes-Rojas et al., 2012). We have previously shown that TRB1 is localized in the nucleus and nucleolus in vivo and shows highly dynamic association with chromatin (Dvorackova et al., 2010). Together, these findings indicate that TRB proteins are promising candidates for plant shelterin-like components. Here we demonstrate that TRB proteins act as components of a plant telomere-protection complex. Microscopic and chromatin immunoprecipitation techniques showed co-localization of TRB1 with telomeric tracts in vivo and physical interaction of TRB proteins with the N–terminal part of the catalytic subunit of telomerase. In addition, loss of TRB1 protein leads to telomere shortening. RESULTS TRB1 co-localizes with telomeres Although a possible association of GFP–TRB1 (35Spro: GFP-TRB1) with the telomere was suggested previously (Dvorackova et al., 2010), whether the nuclear speckles are directly associated with telomeres remained to be deter- mined. Here, we took advantage of the well-established protocol of Nicotiana benthamiana leaf infiltration and the fact that N. benthamiana has longer telomeres that are easier to visualize compared to Arabidopsis. As shown in Figure 1(c), the localization of transiently transformed TRB1 in N. benthamiana leaf is similar to that observed in Arabidopsis cell cultures, as was shown by Dvorackova et al. (2010), labelling the whole nucleus, with strong nucleolar signal and relatively strong nuclear speckles. Nuclei from transformed leaves were isolated and used for telomere peptide nucleic acid FISH. Fluorescence from GFP–TRB1 remained very bright during the isolation procedure; however, a gentle denaturation step was necessary during the FISH protocol to preserve the integrity of the GFP signal. These FISH results showed that telomeres colocalize or associate with TRB1 speckles in 59% and 31% of cases, respectively, with 90% association overall (Figure 1 and Table S1). Telomeric signals sometimes appeared as double dots connected to the TRB1 foci (Figure 1d, images 1, 2 and 3), but in other cases co-localize directly with TRB1 (Figure 1d, images 4 and 5). These results provide in situ evidence of telomere occupancy by TRB1. TRB1 is associated with telomeric sequence in vivo The observed co-localization of TRB1 with telomeric tracts, together with our previous detailed analyses of TRB1 binding to telomeric DNA in vitro (Schrumpfova et al., 2004; Hofr et al., 2009), suggest the possibility that TRB1 protein directly recognizes telomeric repeats and belongs to the core components that shelter telomeres. We used a chromatin immunoprecipitation assay to isolate DNA sequences associated with TRB1 protein. As source material, we used formaldehyde cross-linked seedlings of Arabidopsis plants stably transformed with a TRB1–GFP construct driven by the native promoter (TRB1pro:TRB1GFP) (Dvorackova et al., 2010). Despite using the native © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 TRB interactions with telomeres and telomerase 771 promoter, enhanced levels of TRB1–GFP protein were observed (see below). TRB1–GFP protein was immunoprecipitated from purified nuclei using GFP-Trap A matrix, which contains a single variable antibody domain that recognizes GFP. Non-specific binding of TRB1-GFP to the GFP-Trap A matrix was excluded by precise detection of GFP in all fractions (input, bound, unbound, wash, elution). We have shown that TRB1, but not TRB1-GFP is washed out (Figure S1). DNA co-purifying with TRB1–GFP was dotblotted onto nylon membranes, and visualized by hybridization with radioactively labelled telomeric probe. Figure 2 shows that TRB1 protein is indeed associated with telomeric sequence in vivo, as telomeric sequence was repeatedly detected in TRB1–GFP but not wild-type samples. To demonstrate that the observed enrichment is indeed due to sequence-specific association and not due to the high copy number of the telomeric DNA, we hybridized DNA co-purified with TRB–GFP with a centromeric probe. As our previous results (Dvorackova et al., 2010) showed localization of TRB1 protein in the nucleus and nucleolus, another candidate sequence investigated for association with TRB1 was ribosomal DNA (rDNA). Only negligible enrichment of the centromeric or 18S rDNA probe, in contrast to significant enrichment of the telomeric probe, was observed when comparing each wild-type to TRB–GFP sample (Figure S2). Analysis of TRB1 expression in trb1 mutant, wild-type and TRB1pro:TRB1-GFP-transformed plants To examine the role of TRB1 in planta, we analysed T–DNA insertion line SALK_025147 (ecotype Col–0). Three parallel wild-type (wild-type) and homozygous trb1 (trb1À/À) lines (A, B and C) were derived from three independent heterozygous plants (see Figure 4a). The homozygosity of each parallel wild-type and trb1 mutant plant line was determined by PCR (Figure 3b). The T–DNA insertion is located in the second intron (Figure 3a), and the absence of trb1 transcript was confirmed by RT–PCR (Figure 3c). TRB proteins consist of three domains: Myb-like, histone-like and a coiled-coil domain (Figure 3a). As no antibody recognizing either TRB proteins or the plant Myb (a) (c) (d) (b) Figure 1. Co-localization of TRB protein with telomeric probe. Nuclei isolated from N. benthamiana were transformed with 35Spro:GFP-TRB1 construct and hybridized with telomeric peptide nucleic acid (PNA) Cy3-labelled probe. (a) Co-localization between GFP–TRB1 nuclear speckles (green) and telomeric PNA probe (red) is detectable in most of the foci. (b) Control experiment without telomeric probe showing very little background present in the red channel. (c) Confocal image of GFP–TRB1 expression in an N. benthamiana leaf without any further sample processing. (d) Details of co-localizing speckles; ImageJ (http://rsbweb.nih.gov/ij/) was used to create intensity plots for red and green channels. © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 772 Petra Prochazkova Schrumpfova et al. domain of the telobox type is commercially available, we developed specific mouse monoclonal antibodies in our laboratory. Two of them were used in this study: 1.2 (specific to TRB1) and 5.2 (specific for the conservative part of the Myb domain; this also recognizes other TRB proteins). The location of antibody recognition sites within the structure of TRB1 as determined by ELISA (Figure S3) is shown in Figure 3(a). Although the conservative part of the telobox Myb domain is also present in Arabidopsis TRFlike family (TRFL) proteins (Karamysheva et al., 2004), these proteins are not recognized by the 1.2 or 5.2 antibodies. The anti-TRB 1.2 or 5.2 antibodies were unable to detect in vitro expressed TRFL2 or 9 or TRP1 (telomeric repeat binding protein 1) from the TRFL family (Figure S4, constructs kindly provided by D.E. Shippen, Department of Biochemistry,Texas A&M University,TX, USA). Antibody 1.2 was used to detect native TRB1 protein in Arabidopsis plant protein extracts. The natural level of TRB1 protein was clearly observed on Western blots of wild-type plants (Figure 3d), but no TRB1 protein was observed for extracts from trb1 mutant plants (Figure 3d). In addition, plant lines stably transformed with TRB1–GFP construct under the control of native promoter showed a distinct abundance of TRB1–GFP protein compared to the native TRB1 protein. Various expression levels of native TRB1 protein and TRB1–GFP were also apparent after immunolocalization in vivo (Figure 3e), in which TRB1 protein is visualized using either anti-TRB1 1.2 antibody or anti-GFP antibody. We tested both antibodies by indirect immunofluorescence on trb1 mutant and GFP–TRB1-expressing plants. These experiments showed evenly distributed nuclear and nucleolar signals for both 1.2 and 5.2 antibodies. Antibody 1.2 did not detect any signal in trb1À/À plants, but antibody 5.2 recognizes some epitopes in trb1À/À (Figure S5). However, the generated antibodies do not appear to be of sufficient quality for more demanding immunolocalization or ChIP experiments (as concluded from further testing). Telomere shortening in trb1 null mutant plants Derivation of independent wild-type and trb1À/À plant lines from three heterozygous progeny (Figure 4a) provided reliable material for phenotypic studies of the trb1 null mutation effect. All six homozygous plant lines were propagated for five generations. Obvious shortening of telomeres was observed by terminal restriction fragment (TRF) analysis in all three trb1 mutant lines analysed in the fifth generation compared to their segregated wild-type siblings. Hybridization with a radioactively labelled telomeric probe (Figure 4b) revealed truncation of telomeric tracts in trb1 lines by approximately 10–20% (Figure 4c). The graph represents evaluation in the three biological replicates. Observations in earlier generations of trb1 lines (Figure S6) show mild but progressive shortening that continues through the generations. Despite clear and reproducible telomere shortening in trb1, no significant morphological differences were observed in rosette diameter, leaf number, flowering and seed set when analysing soil-grown wild-type and trb1À/Àplants. TRB proteins interact with telomerase in planta Our previous finding that TRB1 protein interacts with POT1b and evidence presented here showing that TRB1 co-localizes with telomeric repeats and is involved in regulation of telomere maintenance suggest its possible association with telomerase (Kuchar and Fajkus, 2004; Schrumpfova et al., 2008). We therefore tested the possibility of direct interaction between TRB1 and TERT, as well as the influence of TRB1 on telomerase activity in vitro. As TERT is a high-molecular-weight protein (approximately 130 kDa), we used TERT fragments containing N– terminal domains associated with distinct telomeric functions (Sykorova and Fajkus, 2009) to detect a possible direct interaction between TERT and TRB proteins (Figure 5a). We tested their ability to interact using a GAL4 based yeast two-hybrid system, in which interactions take place inside the nucleus. As shown in Figure 5(b), strong interaction between TRB1 and the TERT 1-271 fragment was observed on histidine-deficient plates. This interaction was confirmed under stringent adenine selection. Clear interactions between TRB3 and TERT 1-271 and a weak interaction between TRB2 and TERT 1-271 were also observed under histidine selection. Further testing TRB1 Figure 2. TRB1 proteins are associated with telomeric sequence in vivo. DNA cross-linked with TRB1 protein was isolated by ChIP analysis using GFP-Trap A matrix from wild-type (Wt) and TRB1pro:TRB1-GFP plants. Hybridization of isolated DNA with radioactively labelled telomeric oligonucleotide (CCCTAAA)4 in three biologically and technically replicated experiments confirmed the hypothesis that TRB1 protein is associated with telomeric sequence in vivo. As a control, telomeric oligonucleotide (TTTAGGG)4 was dot-blotted on the same membrane and visualized together with immunoprecipitated DNA. © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 TRB interactions with telomeres and telomerase 773 and TRB2 with a longer fragment of TERT (amino acids 1- 582) confirmed these interactions. To test whether the interactions observed in a yeast-two hybrid system are reproducible in the plant cell, we used a bimolecular fluorescence complementation assay (BiFC). Arabidopsis protoplasts were transfected with plasmids encoding nYFP-tagged TRB constructs and cYFP-tagged TERT fragments, and a clear intra-nuclear interaction was observed (Figure 5c and Figure S7). The TERT fragments used in BiFC (TERT 1-271 and TERT 229-582) overlap with the fragments tested in the yeast two-hybrid system. The interaction was further verified by co-immunoprecipitation experiments in which proteins were expressed in rabbit reticulocyte lysate from the same vectors used in yeast two-hybrid system. As shown in Figure S8, clear interactions between TRB1 and all three TERT fragments (1-271, 229-582 and 1-582) were observed. Obvious interactions were also detected between TRB3 and TERT 1-271 or TERT 229-582, but only weak interactions were observed between TRB2 and TERT fragments. The generally weaker interactions of TRB2 or TRB3 proteins with TERT fragments in comparison to the corresponding interactions of TRB1 (a) (b) (d) (e) (c) Figure 3. Expression analysis of TRB1 protein in mutant (trb1À/À), wild-type (Wt) and transformed TRB1pro:TRB1-GFP plants. (a) Schematic illustration of specific primers and T–DNA insertion location within the trb1 gene. The domain location and antibody recognition sites for two specific antibodies developed in our laboratory are shown below. (b) Three individual plant lines (A, B and C) were derived from heterozygous progenitors (as shown in Figure 4a). Example of PCR analysis of genomic DNA isolated from Wt plants (primers P3 + P2) and mutant (trb1À/À) plants (primers P1 + P2) of line B. (c) RT–PCR of RNA isolated from Wt and mutant (trb1À/À) plants of line B using primers P4 + P5. (d) Immunodetection by Western blot analysis of TRB1 protein in Wt and mutant (trb1À/À) plants of line B and TRB1pro:TRB1-GFP plant nuclear extracts using specific antibody recognizing the Myb-like domain of TRB1 (anti-TRB1 1.2). The level of native TRB1 protein is lower compared to the TRB1–GFP fusion protein construct expressed under the control of the native promoter. (e) Immunolocalization of TRB1 protein using anti-GFP and anti-TRB1 antibody. The level of native TRB1 protein in the wild-type is very low. The selected plant line does not show GFP labelling in all cells, so some nuclei contain a wild-type level of TRB1 and others show higher expression due to TRB1–GFP. Thus the intensity of signal may be clearly measured as Wt and over-expressing nuclei are present together on one slide and may be clearly distinguished using specific anti-TRB1 protein antibody 1.2 and anti-GFP antibody. © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 774 Petra Prochazkova Schrumpfova et al. were due to lower expression of TRB2 and TRB3 proteins in rabbit reticulocyte lysate. To determine whether interaction between TRB proteins and TERT directly influences telomerase activity, we used a telomere repeat amplification protocol (TRAP). In extracts from trb1À/À plants, no changes in telomerase activity or processivity were observed. Correspondingly, no variations in telomerase activity were detected in transformed plants (TRB1pro:TRB1-GFP) expressing higher levels of protein (Figure S9). This observation is in agreement with our previous experiments in which Escherichia coli-expressed and purified TRB2 and TRB3 proteins were added to the TRAP assay (Schrumpfova et al., 2004). DISCUSSION The composition of plant shelterin-like complex has long remained elusive due to the high number of candidate proteins with apparently redundant functions (Peska et al., 2011). These obstacles and lack of convincing evidence raised doubts over the existence of such a complex, and its functions have been mostly attributed to the previously described CST complex, which is conserved throughout eukaryotes (Nelson and Shippen, 2012b). However, absence of evidence is not evidence of absence. Our present data suggest the existence of a telomere protein complex that includes plant-specific Smh proteins (termed TRB proteins in Arabidopsis). These proteins interact directly with the catalytic subunit of telomerase: TRB1 protein co-localizes with telomeres, specifically binds telomeric DNA in vitro and in vivo, and TRB1 loss results in telomere shortening. Moreover, TRB proteins also interact with POT1b, a POT1-like orthologue in A. thaliana (Kuchar and Fajkus, 2004; Schrumpfova et al., 2008). In previous studies, we considered in detail the localization of TRB1 protein, showing that, similar to TRB2 and TRB3, this is a nuclear factor with markedly increased nucleolar labelling and speckles present in the nucleus, especially in Arabidopsis cell cultures transiently transformed with GFP–TRB1 (Dvorackova et al., 2010). We have previously speculated on the telomeric association of GFP– TRB1 speckles, but the low expression of GFP–TRB1 in stably transformed Arabidopsis plants/cultures and the short size of Arabidopsis telomeres impeded its direct demonstration (Dvorackova et al., 2010). In this study, we used a plant system with longer telomeres and sufficient expression of TRB1–GFP protein, and clearly showed that TRB1 co-localizes with telomeres in plant leaves. Close linkage between TRB1 protein and the telomere was further supported by the finding that plant telomeric sequence may be isolated directly from plant seedlings together with TRB1–GFP using the anti-GFP immunoprecipitation system. Specific anti-TRB1 antibodies were also developed and successfully used for detection of TRB1 alone or all TRB proteins in the whole-protein extract by Western blot or ELISA procedures. Using these antibodies, clear nuclear and nucleolar localization of TRB1 protein was demonstrated. Although TRB1 protein need not associate exclusively with telomeres in vivo, the preferential association of TRB1 with the telomeric tracts as described here is in agreement with previous observations using independent approaches (Mozgova et al., 2008; Hofr et al., 2009; Dvorackova et al., 2010). The obvious association of TRB1 with the nucleolus, which contains sub-telomeric clusters of rDNA, may be due to the fact that nucleoli associate with telomeres and telomerase at the cellular level: telomerase assembly occurs in nucleoli in a number of model organisms including plants (Lo et al., 2006; Brown and Shaw, 2008; Kannan et al., 2008), and nucleolus-associated telomere clustering and pairing precede meiotic chromosome synapsis in Arabidopsis thaliana (Armstrong et al., 2001). The key finding of this work is that TRB proteins interact with the N–terminal part of TERT. This part contains the telomerase-specific motifs TEN (telomerase essential N– terminal domain) and TRBD (N–terminal RNA-binding domain). The most conserved motif, the T–motif, with a high-affinity binding site for the TER subunit, is included in the TRBD domain (Lai et al., 2001). Several distinct functions have been proposed for the TEN domain: e.g. as an (a) (b) (c) Figure 4. The telomeres are shortened in all three individually derived trb1À/À mutant plant lines. (a) Derivation of three independent plant lines (A, B, C) that were propagated for five generations (G5). (b) Terminal restriction fragment analysis, showing telomere shortening in the trb1À/À mutant line compared with the wild-type control in the fifth generation. (c) Difference in mutant trb1À/À and wild-type telomere lengths in three independent plant lines. Error bars represent standard deviation. © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 TRB interactions with telomeres and telomerase 775 anchor during template translocation (Lue, 2005; Wyatt et al., 2007; Sealey et al., 2010), involvement in positioning the 3′ end of a telomeric DNA primer in the active site during nucleotide addition (Jurczyluk et al., 2011), putative mitochondrial localization (Santos et al., 2004), and, last but not least, involvement in protein–protein interactions (Sealey et al., 2011). Hence, the positioning of the region involved in interaction between TERT and TRB proteins in the N–terminal part of telomerase is not surprising. Identification of TRB proteins as the interaction partner of TERT is also supported by the observation that TRB1 protein is present in a group of proteins that were co-purified with the N–terminal part of TERT using tandem affinity purification (P.P.S., J.M., L.D., E.S and J.F., unpublished results). The observation of TRB/telomerase interaction, together with the previously detected interaction between TRB and POT1b (Kuchar and Fajkus, 2004; Schrumpfova et al., 2008), suggest that TRB proteins are part of the telomeric interactome of A. thaliana. Interaction of POT1b protein with the TRB1 protein is mediated by the central TRB histone-like domain (Schrumpfova et al., 2008), but it is not yet clear how the interaction between telomerase and TRB is mediated. Determination of whether it occurs through the same histone-like domain or the N–terminal Myb domain or C–terminal coiled-coil domain would help to (a) (b) (c) Figure 5. TRB proteins interact with plant telomerase (TERT). (a) Schematic depiction of the catalytic subunit of telomerase (TERT) showing evolutionarily conserved motifs. N-terminal fragments containing the telomerasespecific motifs TEN (telomerase essential N–terminal domain) and TRBD (N–terminal RNA-binding domain) were used in protein–protein interaction analysis (amino acid numbering is shown). (b) Yeast two-hybrid system was used to assess interaction of TRB proteins with N–terminal TERT fragments. Two sets of plasmids carrying the indicated segments of TERT fused to either the GAL4 DNA-binding domain (BD) or the GAL4 activation domain (AD) were constructed and introduced into yeast strain PJ69– 4a carrying reporter genes His3 and Ade2. Although weak interactions often fail to rescue growth under stringent adenine selection, plausible TRB–TERT interactions were observed on histidine-deficient plates. Co-transformation with an empty vector (AD/BD/vector) served as a negative control. (c) Bimolecular fluorescence complementation confirmed the interaction of TRB proteins with TERT fragments. Arabidopsis leaf protoplasts were co-transfected with 10 lg each of plasmids encoding nEYFP-tagged TRB clones, cEYFP-tagged TERT fragments or Gaut10 (as negative control) and mRFP-VirD2NLS (to label cell nuclei and to determine transfection efficiency). The cells were imaged by epifluorescence microscopy after overnight incubation. Clear nuclear interactions of TRB proteins with TERT fragments are observed on the protoplast images: YFP fluorescence (yellow), mRFP fluorescence (red), chloroplast autofluorescence (green pseudocolor); chloroplast autofluorescence is also visible in the YFP channel (indicated by arrows). Scale bars = 7 lm. © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 776 Petra Prochazkova Schrumpfova et al. determine mutual exclusion or co-existence of TERT and POT1b association with TRB proteins. Importantly, POT1b is also an interaction partner of TER2, an alternative telomerase RNA subunit in Arabidopsis (Cifuentes-Rojas et al., 2011, 2012). Together with TERT, dyskerin and Ku, these components form telomerase ribonucleoprotein complex that may participate in telomerase regulation, the DNA damage response and telomere protection, but do not substantially contribute to telomere maintenance (Cifuentes-Rojas et al., 2011, 2012). Therefore, the observation that TRB1 interacts with TERT but its loss or increased expression does not change telomerase activity is not surprising. More importantly, interaction of TRB1 with TERT, together with its affinity for telomeric DNA, indicates a possible role of TRB1 in telomerase recruitment to telomeres. This also explains the observed absence of any direct effect of TRB1 on telomerase activity in the TRAP assay, as this in vitro assay uses a non-telomeric template oligonucleotide that is not recognized by the Myb-like domain of TRB1 (Mozgova et al., 2008). However, POT proteins are not the only putative singlestranded DNA telomere-binding proteins in Arabidopsis, as several other proteins have been identified, e.g. STEP1 (Kwon and Chung, 2004), WHY1 (Yoo et al., 2007a) or CST complex components (Price et al., 2010). Similarly, in addition to the TRB family of proteins, there are also other candidate double-stranded DNA telomere-binding proteins in Arabidopsis, such as TRFL family proteins (Karamysheva et al., 2004). Association of these proteins with telomeres appears not to be mutually exclusive. Presumably, dynamic changes in the composition of telomeric nucleoprotein complexes may reflect the different functional states of telomeres. Two types of plant chromosome ends have been proposed: those with G–overhangs and blunt-ended ones that are recognized by the KU70/80 dimer (Riha et al., 2000; Gallego et al., 2003; Kazda et al., 2012; Nelson and Shippen, 2012a). Thus, the apparently redundant proteins may operate concurrently at telomeres with respect to cell cycle, developmental stage or type of chromosome ends. For example, localization of TRB1 is quite consistent but highly dynamic during interphase; moreover, the level of nuclearassociated TRB1 diminishes during mitotic entry, and it progressively re-associates with chromatin during anaphase/ telophase (Dvorackova et al., 2010). Interestingly, our BiFC assay also showed interaction of N–terminal fragments of TERT with TRP1, a member of the TRFL I family (Figure S10) (Hwang et al., 2001; Karamysheva et al., 2004). Importantly, TRP1 also interacts with KU70 (Kuchar and Fajkus, 2004), which is presumably involved in protection of blunt chromosome ends and may also therefore be an integral part of the plant telomere protection complex (Figure 6). Although it is tempting to draw possible analogies between mammalian shelterin components and the TRB and POT1b proteins involved in a similar plant complex, an alternative interpretation of the function of TRB is possible when considering our data in connection with a recent description of mammalian HOT1 protein (Kappei et al., 2013). This protein shows strikingly similar interactions and functions: it specifically binds double-stranded telomeric DNA repeats, localizes to a subset of telomeres (presumably those that are being elongated), and associates with active telomerase. Thus, HOT1 contributes to the association of telomerase with telomeres and to telomere length maintenance (Kappei et al., 2013). Our findings suggest that TRB proteins may perform similar functions in plant telomeres, i.e. as direct telomere-binding proteins that act as positive regulators of telomere length. EXPERIMENTAL PROCEDURES Primers The sequences of all primers and probes used in this study are provided in Table S2. Plant material and construct generation The 35Spro:GFP-TRB1 plants and construct have been described previously (Dvorackova et al., 2010). The TRBpro:TRB1-GFP construct was prepared as follows: genomic DNA from A. thaliana Col–0 was isolated using a DNeasy plant mini kit (Qiagen, http:// www.qiagen.com/), and used as a template for PCR to amplify the Figure 6. Schematic diagram of observed protein–protein interactions at telomeric ends. Half the telomeric ends in A. thaliana are blunt-ended (Riha et al., 2000; Kazda et al., 2012). Here we show a simplified chart of interactions associated with telomeres. Solid arrows indicate protein–protein interactions that were verified in this study or in previous studies (Kuchar and Fajkus, 2004; Kannan et al., 2008; Schrumpfova et al., 2008; Cifuentes-Rojas et al., 2012; Kazda et al., 2012) using at least two independent approaches (i.e. BiFC, pull-down or yeast-two hybrid assay). The grey arrow indicates a TERT– TRP1 interaction observed only by BiFC. The dashed black arrow shows a presumed interaction between POT1b and telomere single-stranded DNA that has not yet been directly demonstrated. The interaction between POT1b and telomerase is specific for the TER2 isoform of TER, while the other interactions with the telomerase complex are dependent on the catalytic TERT subunit. The diagram suggests the existence of distinct telomerase recruitment pathways for blunt-ended telomeres and telomeres with a G– overhang. © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 TRB interactions with telomeres and telomerase 777 TRB1 genomic sequence including the 5′ UTR. The 3′ UTR was amplified from BAC clone FJ10.16 (Arabidopsis Information Resource, http://www.arabidopsis.org/). We used 0.25 units of Hot Start Phusion polymerase (Finnzymes, http://www.thermoscienti ficbio.com/finnzymes/) with 0.2 mM dNTPs, 19 HF reaction buffer (Phusion Hot Start II high fidelity DNA polymerase; http:// www.thermoscientificbio.com/), 3% dimethylsulfoxide and 0.5 lM of each primer (5′ UTRFw + TRB1 Rev or 3′ UTR Fw + 3′ UTR Rev). The conditions used were in accordance with the manufacturer’s instructions (Finnzymes). PCR products were precipitated using poly(ethylene glycol), and cloned into a Gateway multi-site system (Invitrogen, http://www.lifetechnologies.com), together with the GFP tag (GFP in pDONR221, provided by Keke Yi, College of life Sciences, Zhejiang University, China). pKm43GW (Karimi et al., 2005) was used as the destination vector. A. thaliana Col–0 was subsequently transformed by floral dipping (Clough and Bent, 1998), and transformants selected on MS medium containing 30 lg/ml kanamycin were scored for GFP expression. PCR-based genotyping of plant lines T–DNA insertion mutant plants of trb1 (SALK_025147) in the Col–0 background were used. To distinguish between wild-type plants and those that were heterozygous or homozygous for the T–DNA insertion in the trb1 gene, we isolated genomic DNA from leaves using NucleoSpin Plant II (Machery Nagel, http://www.mn-net. com/). The genomic DNA was used for PCR analysis with MyTaq DNA polymerase (Bioline, http://www.bioline.com). The conditions used were in accordance with the manufacturer’s instructions. The primers used were specific for T–DNA (P3 + P2 primers) or the TRB1 gene (P1 + P2 primers). Cycling conditions were 98°C for 1 min (initial denaturation), followed by 30 cycles of 94°C for 30 sec, 58°C for 45 sec and 72°C for 2 min, with a final extension at 72°C for 10 min. Rt–pcr Total RNA was extracted from approximately 50 mg of frozen plant tissue using an RNeasy plant mini kit (Qiagen), and RNA samples were treated with TURBO DNA-free (Applied Biosystems/Ambion, http://www.lifetechnologies.com TURBO DNA-free). The quality and quantity of RNA were determined by electrophoresis on 1% w/v agarose gels and by measurement of absorbance using an Implen nanophotometer (http://www.implen.de/). Reverse transcription was performed using random hexamers (Sigma-Aldrich, http://www.sigmaaldrich.com) with 1 lg RNA and Mu-MLV reverse transcriptase (New England Biolabs, https://www.neb.com/). The cDNA obtained was screened by PCR analysis for the presence of trb1 transcripts using MyTaq DNA polymerase (Bioline) with primers P4 and P5. Thermal conditions were 95°C for 1 min (initial denaturation), followed by 30 cycles of 94°C for 45 sec, 55°C for 45 sec and 72°C for 2 min, with a final extension at 72°C for 10 min. Nicotiana benthamiana transformation, nuclei isolation and FISH Leaves of 5-week-old N. benthamiana plants were infiltrated with Agrobacterium tumefaciens containing 35Spro:GFP-TRB1 (vector pGWB6, strain LBA4404) (Dvorackova et al., 2010), and 35Spro:p19 (Silhavy et al., 2002) as described by Voinnet et al. (2003). The infiltration medium contained 10 mM MES (pH approximately 5.7) and 10 mM MgCl2. After 3–4 days, leaf discs were checked under a fluorescence microscope, and protoplasts were prepared as described by Yoo et al. (2007b); the digestion medium contained also 0.25% Pectolyase Y23 (Duchefa, http://www.duchefa-biochemie.nl/) in addition to celullase and macerozyme and a 119 lm filter was used for filtration. Protoplasts in W5 buffer were collected by centrifugation at 50 g, and resuspended in NIB (Nuclei Isolation Buffer; 10 mM MES, 0.2M Sucrose, 2.5 mM EDTA,10 mM NaCl, 10 mM KCl 2.5 mM DTT, 0.1 mM Spermine, 0.5 mM Spermidine) to extract nuclei as described by McKeown et al. (2008). Isolated nuclei were then fixed in 4% paraformaldehyde for 15 min, resuspended in wash buffer (50 mM Tris/Cl, pH 8.5, 5 mM MgCl2, 20% glycerol), 4°C, spun down at 300g and stored in storage buffer (50 mM Tris/ Cl, pH 8.5, 5 mM MgCl2, 50% glycerol) at -20°C until use. Then 20 ll of nuclei were spun on the Superfrost plus microscopic slide (http://www.menzel.de/) at 56 g, and re-fixed in 4% pformaldehyde in 19 PBS/0.05% Triton X-100 for 15 min. Slides were then treated with RNase (100 lg/ml) for 1 h at 37°C, and hybridized with telomeric Cy3-labelled peptide nucleic acid probe in 65% formamide/20% dextran sulfate/29 SSC at 37°C overnight. Post-hybridization washes were performed at 37°C using 29 SSC. Slides were counter-stained using 4,6–diamidino-2–phenylindole (1 lg/ml), and observed on a Zeiss (http://www.zeiss.cz/) Axioimager Z1 using an AHF filter set. Immunolocalization Arabidopsis seeds expressing TRB1pro:TRB1-GFP under the control of the native promoter and trb1 seeds were bleach-sterilized for 10 min, washed in water and sown onto half-strength MS medium/1% agar plates. Seedlings grown under the constant light, at 22°C for 2 weeks, then chopped into small pieces. Protoplasts were prepared as described by Yoo et al. (2007b), and the nuclei and immunolocalization protocols were adapted from those described by McKeown et al. (2008). Slides were first blocked in a mixture of 29 block solution (Roche, http://www.roche.cz)/19 PBS/ 5% goat serum at room temperature for 30 min, then incubated with primary antibodies [mouse anti-TRB 1.2 or 5.2 or, anti-GFP (Abcam ab290, http://www.abcam.com/), all diluted 1:300] for 2 h at 37°C, and visualized using secondary antibodies A11001 and A21207 (Invitrogen) at 1:500 dilution. Immunblot analysis To determine the level of TRB1 protein in plants, we isolated nuclei as described by Bowler et al. (2004). The nuclei were lysed using SDS loading buffer (250 mM Tris/Cl, pH 6.8, 4% w/v SDS, 0.2% w/v bromophenol blue, 20% v/v glycerol, 200 mM b–mercaptoethanol), heated at 80°C for 10 min, and protein extracts were analysed by SDS–PAGE. Proteins were electrophoretically transferred to a nitrocellulose membrane at 360 mA for 1 hour in 192 mM glycine, 25 mM Tris, 0.5% SDS and 10% (v/v) methanol in a Bio-Rad Mini Trans-Blot cell. Ponceau S staining was performed to check the quality of the extracts and to ensure equal gel loading for immunodetection. Membranes were blocked with 5% non-fat dry milk in Tris-buffered saline/Tween, and probed using the monoclonal anti-TRB1 1.2 antibody and the secondary polyclonal horseradish peroxidase-conjugated rabbit anti-mouse immunoglobulins (DAKO, http://www.dako.com), both diluted 1:5000. Immunoreactive bands were visualized using LumiGLO reagent and peroxide (Cell Signaling Technology, http://www.cellsignal.com) on a Fujifilm LAS-3000 CCD system (http://www.fujifilm.com/). Chromatin immunoprecipitation assay The ChIP assay was performed as described by Bowler et al. (2004) with modifications. Chromatin extracts were prepared from seedlings treated with 1% formaldehyde. The chromatin from isolated nuclei was sheared to a mean length of 250–500 bp by soni© 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 778 Petra Prochazkova Schrumpfova et al. cation using a Bioruptor (Diagenode, http://www.diagenode.com) and centrifuged (16 000 g/5 min/4°C). The matrix GFP-Trap A (Chromtec, http://www.chromotek.com) was blocked against nonspecific interaction using 200 mM ethanolamine, 1% BSA and the DNA sequences TR10–24–G and TR10–24–C (Table S2), which are not recognized by TRB proteins (Schrumpfova et al., 2004). The pre-treated matrix was incubated with chromatin diluted with ChIP dilution buffer (16.7 mM Tris/Cl, pH 8,0, 1.2 mM EDTA, 167 mM NaCl, 0.1% Triton X–100, phenylmethanesulfonyl fluoride and protease inhibitors) at 4°C for 4 h, and subsequently washed with low-salt, high-salt, LiCl and 10 mM Tris (pH = 8,0), 1 mM EDTA (TE) buffers. In contrast to Bowler et al. (2004), the levels of detergents (Triton X–100, Nonidet P-40 and sodium deoxycholate) were reduced to 0.1%. The cross-linking was reversed using 0.2 M NaCl overnight, and was followed by treatment with proteinase K (Serva, http://www.serva.de) treatment, phenol/chlorophorm extraction and treatment with RNase A (Serva) as described by Bowler et al. (2004). ChIP assays were repeated using three biological replicates (plants grown at different times). Dot-blot assay DNA isolated using ChIP was diluted into 200 ll of 400 mM NaOH and 10 mM EDTA, and samples were denatured at 95°C for 10 min and cooled on ice. They were then spotted onto Hybond XL membrane (GE Healthcare, http://www3.gehealthcare.com) and subjected to hybridization with sequence-specific probe TR–4C (Table S2). The probe was hybridized in 250 M sodium phosphate, pH 7.5, 7% SDS and 16 mM EDTA overnight at 55°C, and washed with 0.29 SSC + 0.1% SDS. The signal was evaluated using MultiGauge software (Fujifilm). All experiments were performed using three independent biological replicates. Re-hybridization with centromeric and 18S rDNA probes was performed as described previously (Mozgova et al., 2010). TRAP assay Protein extracts from 2-week-old seedlings were prepared as described by Fitzgerald et al. (1996). These extracts were subjected to the TRAP assay as described by Fajkus et al. (1998). TS21 was used as the substrate primer for extension by telomerase, and TEL-PR was used as the reverse primer in the subsequent PCR. TRF analysis TRF analysis was performed as described previously (Ruckova et al., 2008) using 500 ng genomic DNA isolated from 5–7-weekold rosette leaves using NucleoSpin Plant II (Machery Nagel). Southern hybridization was performed using the end-labelled telomere-specific probe TR–4C (Table S2). Telomeric signals were visualized using an FLA7000 imager (Fujifilm), and a grey-scale intensity profile was generated using MultiGauge software (Fujifilm). Evaluation of fragment lengths was performed using a Gene Ruler 1 kb DNA ladder (Fermentas, http://www.thermoscientific bio.com/fermentas/) as the standard. Mean telomere lengths were calculated as described by Grant et al. (2001). Yeast two-hybrid analysis Yeast two-hybrid experiments were performed using the Match- makerTM GAL4-based two-hybrid system (Clontech, http:// www.clontech.com/). cDNA sequences encoding TERT N – terminal fragments comprising amino acids 1-271 and 1-582 were subcloned from pDONR/Zeo entry clones (Zachova et al., 2013) into the Gateway-compatible destination vector pGBKT7-DEST (bait vector). The pGBKT7-DEST destination vector that was used in this study was created by Horak et al. (2008) who introduced the Gateway conversion cassette into the original Matchmaker system vector pGBKT7 (Clontech). The pGADT7 prey vectors (Clontech) carrying TRB1, TRB2 and TRB3 have been described previously (Schrumpfova et al., 2008). Each bait/prey combination was cotransformed into Saccharomyces cerevisiae PJ69–4a, and colonies were inoculated into YPD medium and cultivated overnight. Successful co-transformation was confirmed on SD medium lacking Leu and Trp, and positive interactions were selected on SD medium lacking Leu, Trp and His or SD medium lacking Leu, Trp and Ade. Co-transformation with an empty vector served as a negative control for auto-activation. Each test was performed three times using two replicates at a time. In addition, the protein expression levels were verified by immunoblotting. Bimolecular fluorescence complementation For PCR amplification of sequences encoding the tested proteins, and to generate restriction site overhangs, Phusion HF DNA polymerase (Finnzymes) was used. The conditions used were in accordance with the manufacturer’s instructions. The primers used were F–TRB1/2/3_BstBI, R–TRB1/2/3_SmaI, F–TERT_KpnI, R– RID1+BamHI, F–F2N_KpnI and R–F2N+BamHI, and plasmids encoding the tested proteins were used as templates. The amplified DNA fragments were gel-purified, digested with BstBI/SmaI or KpnI/BamHI (New England Biolabs), and ligated into vectors pSAT1-nEYFP and pSAT1-cEYFP. As a negative control, we used an AtGaut10-cEYFP construct. To quantify transformation efficiency and to label cell nuclei, we co-transfected a plasmid expressing mRFP fused to the nuclear localization signal of the VirD2 protein of A. tumefaciens (mRFP-VirD2NLS; Citovsky et al., 2006). The vectors and the mRFP-VirD2NLS and AtGaut10-cEYFP constructs were kindly provided by Stanton Gelvin (Department of Biological Sciences, Purdue University, IN, USA). Arabidopsis thaliana leaf protoplasts were prepared and transfected as described by Wu et al. (2009). DNA (10 lg of each construct) was introduced into 1 9 105 protoplasts. Transfected protoplasts were incubated in the light at room temperature overnight, and then observed for fluorescence using a Zeiss AxioImager Z1 epifluorescence microscope equipped with filters for YFP (Alexa Fluor 488), RFP (Texas Red) and CY5 (chloroplast autofluorescence). In vitro translation and co-immunoprecipitation Proteins were expressed from the same constructs as used in the yeast two-hybrid system with a haemagglutinin tag (pGADT7; TRB1, 2 and 3 proteins) or a Myc tag (pGBKT7; TERT fragments) using a TNT quick coupled transcription/translation system (Promega, https://www.promega.com) in 50 ll reaction volumes according to the manufacturer’s instructions. The TRB proteins were radioactively labelled using 35 S-Met. The co-immunoprecipitation procedure was performed as described by Schrumpfova et al. (2011). Input, unbound and bound fractions were separated by 12% SDS–PAGE, and analysed using an FLA7000 imager (Fuji- film). Accession numbers Sequence data have been deposited in the Arabidopsis Genome Initiative or GenBank/EMBL databases under the following accession numbers: At1g49950 (TRB1), At5g67580 (TRB2, formerly TBP3), At3g49850 (TRB3, formerly TBP2,), At5g16850.1 (TERT), At2g20810 (Gaut10) and At5g59430 (TRP1). © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 TRB interactions with telomeres and telomerase 779 ACKNOWLEDGEMENTS We would thank to Dorothy E. Shippen (Department of Biochemistry,Texas A&M University,TX, USA) for donation of TRFL contructs, Stanton B. Gelvin (Department of Biological Sciences, purdue University, IN, USA) for donation of mRFP-VirD2NLS and AtGaut10-cEYFP constructs and his help with initial analysis of BiFC experiments, and Iva Mozgova (Department of Plant Biology and Forest Genetics, Swedish University of Agricultural Sciences, SW) for discussion and support. The research was supported by the Czech Science Foundation (13-06943S), by project CEITEC (CZ.1.05/1.1.00/02.0068) of the European Regional Development Fund, and project CZ1.07/2.3.00/30.0009 co-financed from European Social Fund and the state budget of the Czech Republic. SUPPORTING INFORMATION Additional Supporting Information may be found in the online version of this article. Figure S1. Detection of TRB1–GFP protein by specific anti-TRB1 1.2 antibody in ChIP fractions. Figure S2. Telomeric sequence is highly enriched compared to centromeric DNA or 18S rDNA. Figure S3. Location of antibody recognition sites within the structure of TRB1 protein. Figure S4. Anti-TRB 1.2 and 5.2 antibodies were unable to detect proteins from TRFL family. Figure S5. Anti-TRB1 1.2 does not detect any signal on trb1À/À plants, but 5.2 recognizes some epitopes in trb1À/Àmutant plant lines. Figure S6. Telomere shortening in trb1À/Àplants is progressive. Figure S7. Whole images of protoplasts (whose segments are shown in Figure 5C) obtained by bimolecular fluorescence com- plementation. Figure S8. TRB1, 2 and 3 proteins are able to pull-down TERT frag- ments. Figure S9. Telomerase activity or processivity in vitro is not changed in response to TRB1 status. Figure S10. TRP1 protein interacts with plant telomerase (TERT). Table S1. Quantification of TRB1 foci co-localized/associated with telomeric foci. Table S2. Sequences of all primers and probes used in this study. REFERENCES Armstrong, S.J., Franklin, F.C. and Jones, G.H. (2001) Nucleolus-associated telomere clustering and pairing precede meiotic chromosome synapsis in Arabidopsis thaliana. J. Cell Sci. 114, 4207–4217. Baumann, P., Podell, E. and Cech, T.R. (2002) Human Pot1 (Protection of telomeres) protein: cytolocalization, gene structure, and alternative splicing. Mol. Cell. Biol. 22, 8079–8087. Bilaud, T., Koering, C.E., Binet-Brasselet, E., Ancelin, K., Pollice, A., Gasser, S.M. and Gilson, E. (1996) The telobox, a Myb-related telomeric DNA binding motif found in proteins from yeast, plants and human. Nucleic Acids Res. 24, 1294–1303. Blackburn, E.H. and Gall, J.G. (1978) Tandemly repeated sequence at termini of extrachromosomal ribosomal RNA genes in Tetrahymena. J. Mol. Biol. 120, 33–53. Bowler, C., Benvenuto, G., Laflamme, P., Molino, D., Probst, A.V., Tariq, M. and Paszkowski, J. (2004) Chromatin techniques for plant cells. Plant J. 39, 776–789. Brown, J.W. and Shaw, P.J. (2008) The role of the plant nucleolus in premRNA processing. Curr. Top. Microbiol. Immunol. 326, 291–311. Chen, L.Y., Redon, S. and Lingner, J. (2012) The human CST complex is a terminator of telomerase activity. Nature, 488, 540–544. Cifuentes-Rojas, C., Kannan, K., Tseng, L. and Shippen, D.E. (2011) Two RNA subunits and POT1a are components of Arabidopsis telomerase. Proc. Natl Acad. Sci. USA 108, 73–78. Cifuentes-Rojas, C., Nelson, A.D.L., Boltz, K.A., Kannan, K., She, X.T. and Shippen, D.E. (2012) An alternative telomerase RNA in Arabidopsis modulates enzyme activity in response to DNA damage. Genes Dev. 26, 2512–2523. Citovsky, V., Lee, L.Y., Vyas, S., Glick, E., Chen, M.H., Vainstein, A., Gafni, Y., Gelvin, S.B. and Tzfira, T. (2006) Subcellular localization of interacting proteins by bimolecular fluorescence complementation in planta. J. Mol. Biol. 362, 1120–1131. Clough, S.J. and Bent, A.F. (1998) Floral dip: a simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J. 16, 735–743. Dvorackova, M., Rossignol, P., Shaw, P.J., Koroleva, O.A., Doonan, J.H. and Fajkus, J. (2010) AtTRB1, a telomeric DNA-binding protein from Arabidopsis, is concentrated in the nucleolus and shows highly dynamic association with chromatin. Plant J. 61, 637–649. Fajkus, J., Fulneckova, J., Hulanova, M., Berkova, K., Riha, K. and Matyasek, R. (1998) Plant cells express telomerase activity upon transfer to callus culture, without extensively changing telomere lengths. Mol. Gen. Genet. 260, 470–474. Fitzgerald, M.S., McKnight, T.D. and Shippen, D.E. (1996) Characterization and developmental patterns of telomerase expression in plants. Proc. Natl Acad. Sci. USA 93, 14422–14427. Gallego, M.E., Bleuyard, J.Y., Daoudal-Cotterell, S., Jallut, N. and White, C.I. (2003) Ku80 plays a role in non-homologous recombination but is not required for T-DNA integration in Arabidopsis. Plant J. 35, 557–565. Gao, H., Cervantes, R.B., Mandell, E.K., Otero, J.H. and Lundblad, V. (2007) RPA-like proteins mediate yeast telomere function. Nat. Struct. Mol. Biol. 14, 208–214. Giraud-Panis, M.J., Teixeira, M.T., Geli, V. and Gilson, E. (2010) CST meets shelterin to keep telomeres in check. Mol. Cell 39, 665–676. Grant, J.D., Broccoli, D., Muquit, M., Manion, F.J., Tisdall, J. and Ochs, M.F. (2001) Telometric: a tool providing simplified, reproducible measurements of telomeric DNA from constant field agarose gels. Biotechniques, 31(1314–1316), 1318. Greider, C.W. and Blackburn, E.H. (1985) Identification of a specific telomere terminal transferase activity in Tetrahymena extracts. Cell, 43, 405–413. Greider, C.W. and Blackburn, E.H. (1989) A telomeric sequence in the RNA of Tetrahymena telomerase required for telomere repeat synthesis. Nature, 337, 331–337. Hofr, C., Sultesova, P., Zimmermann, M., Mozgova, I., Schrumpfova, P.P., Wimmerova, M. and Fajkus, J. (2009) Single-Myb-histone proteins from Arabidopsis thaliana: a quantitative study of telomere-binding specificity and kinetics. Biochem. J. 419, 221–228. Horak, J., Grefen, C., Berendzen, K.W., Hahn, A., Stierhof, Y.D., Stadelhofer, B., Stahl, M., Koncz, C. and Harter, K. (2008) The Arabidopsis thaliana response regulator ARR22 is a putative AHP phospho-histidine phosphatase expressed in the chalaza of developing seeds. BMC Plant Biol. 8, 77. Hwang, M.G., Chung, I.K., Kang, B.G. and Cho, M.H. (2001) Sequence-specific binding property of Arabidopsis thaliana telomeric DNA binding protein 1 (AtTBP1). FEBS Lett. 503, 35–40. Hwang, M.G., Kim, K., Lee, W.K. and Cho, M.H. (2005) AtTBP2 and AtTRP2 in Arabidopsis encode proteins that bind plant telomeric DNA and induce DNA bending in vitro. Mol. Genet. Genomics 273, 66–75. Jurczyluk, J., Nouwens, A.S., Holien, J.K., Adams, T.E., Lovrecz, G.O., Parker, M.W., Cohen, S.B. and Bryan, T.M. (2011) Direct involvement of the TEN domain at the active site of human telomerase. Nucleic Acids Res. 39, 1774–1788. Kannan, K., Nelson, A.D. and Shippen, D.E. (2008) Dyskerin is a component of the Arabidopsis telomerase RNP required for telomere maintenance. Mol. Cell. Biol. 28, 2332–2341. Kappei, D., Butter, F., Benda, C. et al. (2013) HOT1 is a mammalian direct telomere repeat-binding protein contributing to telomerase recruitment. EMBO J. 32, 1681–1701. Karamysheva, Z.N., Surovtseva, Y.V., Vespa, L., Shakirov, E.V. and Shippen, D.E. (2004) A C–terminal Myb extension domain defines a novel family of double-strand telomeric DNA-binding proteins in Arabidopsis. J. Biol. Chem. 279, 47799–47807. © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 780 Petra Prochazkova Schrumpfova et al. Karimi, M., De Meyer, B. and Hilson, P. (2005) Modular cloning in plant cells. Trends Plant Sci. 10, 103–105. Kazda, A., Zellinger, B., Rossler, M., Derboven, E., Kusenda, B. and Riha, K. (2012) Chromosome end protection by blunt-ended telomeres. Genes Dev. 26, 1703–1713. Kuchar, M. and Fajkus, J. (2004) Interactions of putative telomere-binding proteins in Arabidopsis thaliana: identification of functional TRF2 homolog in plants. FEBS Lett. 578, 311–315. Kwon, C. and Chung, I.K. (2004) Interaction of an Arabidopsis RNA-binding protein with plant single-stranded telomeric DNA modulates telomerase activity. J. Biol. Chem. 279, 12812–12818. Lai, C.K., Mitchell, J.R. and Collins, K. (2001) RNA binding domain of telomerase reverse transcriptase. Mol. Cell. Biol. 21, 990–1000. de Lange, T. (2005) Shelterin: the protein complex that shapes and safeguards human telomeres. Genes Dev. 19, 2100–2110. de Lange, T. (2009) How telomeres solve the end-protection problem. Science, 326, 948–952. Lo, S.J., Lee, C.C. and Lai, H.J. (2006) The nucleolus: reviewing oldies to have new understandings. Cell Res. 16, 530–538. Lue, N.F. (2005) A physical and functional constituent of telomerase anchor site. J. Biol. Chem. 280, 26586–26591. Marian, C.O., Bordoli, S.J., Goltz, M., Santarella, R.A., Jackson, L.P., Danilevskaya, O., Beckstette, M., Meeley, R. and Bass, H.W. (2003) The maize Single myb histone 1 gene, Smh1, belongs to a novel gene family and encodes a protein that binds telomere DNA repeats in vitro. Plant Physiol. 133, 1336–1350. McKeown, P., Pendle, A.F. and Shaw, P.J. (2008) Preparation of Arabidopsis nuclei and nucleoli. Methods Mol. Biol. 463, 67–75. Mozgova, I., Schrumpfova, P.P., Hofr, C. and Fajkus, J. (2008) Functional characterization of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. Phytochemistry, 69, 1814–1819. Mozgova, I., Mokros, P. and Fajkus, J. (2010) Dysfunction of chromatin assembly factor 1 induces shortening of telomeres and loss of 45S rDNA in Arabidopsis thaliana. Plant Cell, 22, 2768–2780. Nandakumar, J., Bell, C.F., Weidenfeld, I., Zaug, A.J., Leinwand, L.A. and Cech, T.R. (2012) The TEL patch of telomere protein TPP1 mediates telomerase recruitment and processivity. Nature, 492, 285–289. Nelson, A.D. and Shippen, D.E. (2012a) Blunt-ended telomeres: an alternative ending to the replication and end protection stories. Genes Dev. 26, 1648–1652. Nelson, A.D. and Shippen, D.E. (2012b) Surprises from the chromosome front: lessons from Arabidopsis on telomeres and telomerase. Cold Spring Harb. Symp. Quant. Biol. 77, 7–15. Peska, V., Sykorova, E. and Fajkus, J. (2008) Two faces of Solanaceae telomeres: a comparison between Nicotiana and Cestrum telomeres and telomere-binding proteins. Cytogenet. Genome Res. 122, 380–387. Peska, V., Schrumpfova, P.P. and Fajkus, J. (2011) Using the telobox to search for plant telomere binding proteins. Curr. Protein Peptide Sci. 12, 75–83. Pinto, A.R., Li, H., Nicholls, C. and Liu, J.P. (2011) Telomere protein complexes and interactions with telomerase in telomere maintenance. Front. Biosci. 16, 187–207. Price, C.M., Boltz, K.A., Chaiken, M.F., Stewart, J.A., Beilstein, M.A. and Shippen, D.E. (2010) Evolution of CST function in telomere maintenance. Cell Cycle 9, 3157–3165. Riha, K., Fajkus, J., Siroky, J. and Vyskot, B. (1998) Developmental control of telomere lengths and telomerase activity in plants. Plant Cell, 10, 1691–1698. Riha, K., McKnight, T.D., Fajkus, J., Vyskot, B. and Shippen, D.E. (2000) Analysis of the G–overhang structures on plant telomeres: evidence for two distinct telomere architectures. Plant J. 23, 633–641. Rotkova, G., Sykorova, E. and Fajkus, J. (2009) Protect and regulate: recent findings on plant POT1-like proteins. Biol. Plant. 53, 1–4. Ruckova, E., Friml, J., Prochazkova Schrumpfova, P. and Fajkus, J. (2008) Role of alternative telomere lengthening unmasked in telomerase knockout mutant plants. Plant Mol. Biol. 66, 637–646. Santos, J.H., Meyer, J.N., Skorvaga, M., Annab, L.A. and Van Houten, B. (2004) Mitochondrial hTERT exacerbates free-radical-mediated mtDNA damage. Aging Cell, 3, 399–411. Schrumpfova, P., Kuchar, M., Mikova, G., Skrisovska, L., Kubicarova, T. and Fajkus, J. (2004) Characterization of two Arabidopsis thaliana Myb-like proteins showing affinity to telomeric DNA sequence. Genome, 47, 316–324. Schrumpfova, P.P., Kuchar, M., Palecek, J. and Fajkus, J. (2008) Mapping of interaction domains of putative telomere-binding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. FEBS Lett. 582, 1400–1406. Schrumpfova, P.P., Fojtova, M., Mokros, P., Grasser, K.D. and Fajkus, J. (2011) Role of HMGB proteins in chromatin dynamics and telomere maintenance in Arabidopsis thaliana. Curr. Protein Peptide Sci. 12, 105–111. Sealey, D.C., Zheng, L., Taboski, M.A., Cruickshank, J., Ikura, M. and Harrington, L.A. (2010) The N–terminus of hTERT contains a DNA-binding domain and is required for telomerase activity and cellular immortalization. Nucleic Acids Res. 38, 2019–2035. Sealey, D.C., Kostic, A.D., LeBel, C., Pryde, F. and Harrington, L. (2011) The TPR-containing domain within Est1 homologs exhibits species-specific roles in telomerase interaction and telomere length homeostasis. BMC Mol. Biol. 12, 45. Sfeir, A., Kosiyatrakul, S.T., Hockemeyer, D., MacRae, S.L., Karlseder, J., Schildkraut, C.L. and de Lange, T. (2009) Mammalian telomeres resemble fragile sites and require TRF1 for efficient replication. Cell, 138, 90–103. Shakirov, E.V., Surovtseva, Y.V., Osbun, N. and Shippen, D.E. (2005) The Arabidopsis Pot1 and Pot2 proteins function in telomere length homeostasis and chromosome end protection. Mol. Cell. Biol. 25, 7725–7733. Silhavy, D., Molnar, A., Lucioli, A., Szittya, G., Hornyik, C., Tavazza, M. and Burgyan, J. (2002) A viral protein suppresses RNA silencing and binds silencing-generated, 21- to 25-nucleotide double-stranded RNAs. EMBO J. 21, 3070–3080. Surovtseva, Y.V., Churikov, D., Boltz, K.A., Song, X.Y., Lamb, J.C., Warrington, R., Leehy, K., Heacock, M., Price, C.M. and Shippen, D.E. (2009) Conserved telomere maintenance component 1 interacts with STN1 and maintains chromosome ends in higher eukaryotes. Mol. Cell 36, 207–218. Surovtseva, Y. V., Shakirov, E. V., Vespa, L., Osbun, N., Song, X. and Shippen, D. E. (2007) Arabidopsis POT1 associates with the telomerase RNP and is required for telomere maintenance. EMBO J. 26, 3653–3661. Sykorova, E. and Fajkus, J. (2009) Structure–function relationships in telomerase genes. Biol. Cell 101, 375–392. Tani, A. and Murata, M. (2005) Alternative splicing of Pot1 (Protection of telomere)-like genes in Arabidopsis thaliana. Genes Genet. Syst. 80, 41–48. Voinnet, O., Rivas, S., Mestre, P. and Baulcombe, D. (2003) An enhanced transient expression system in plants based on suppression of gene silencing by the p19 protein of tomato bushy stunt virus. Plant J. 33, 949–956. Wu, F.H., Shen, S.C., Lee, L.Y., Lee, S.H., Chan, M.T. and Lin, C.S. (2009) Tape-Arabidopsis sandwich – a simpler Arabidopsis protoplast isolation method. Plant Methods, 5, 16. Wu, P., Takai, H. and de Lange, T. (2012) Telomeric 3’ overhangs derive from resection by Exo1 and Apollo and fill-in by POT1b-associated CST. Cell, 150, 39–52. Wyatt, H.D., Lobb, D.A. and Beattie, T.L. (2007) Characterization of physical and functional anchor site interactions in human telomerase. Mol. Cell. Biol. 27, 3226–3240. Yoo, H.H., Kwon, C., Lee, M.M. and Chung, I.K. (2007a) Single-stranded DNA binding factor AtWHY1 modulates telomere length homeostasis in Arabidopsis. Plant J. 49, 442–451. Yoo, S.D., Cho, Y.H. and Sheen, J. (2007b) Arabidopsis mesophyll protoplasts: a versatile cell system for transient gene expression analysis. Nat. Protoc. 2, 1565–1572. Zachova, D., Fojtova, M., Dvorackova, M., Mozgova, I., Lermontova, I., Peska, V., Schubert, I., Fajkus, J. and Sykorova, E. (2013) Structure–function relationships during transgenic telomerase expression in Arabidopsis. Physiol. Plant. 149, 114–126. © 2014 The Authors. The Plant Journal published by Society for Experimental Biology and John Wiley & Sons Ltd., The Plant Journal, (2014), 77, 770–781 TRB interactions with telomeres and telomerase 781 Supporting Figures and Tables Figure S1 Detection of TRB1-GFP protein by specific anti-TRB1 1.2 antibody in ChIP fractions 20 µl of each ChIP fraction or 5 µl of GFP-TRAP_A Matrix were separated with SDS-PAGE, transferred to nitrocellulose membrane using standard wet western blot protocol and visualised with specific anti-TRB1 1.2 antibody (1:5 000) and secondary antibody anti-mouse HRP (DAKO) (1:10 000) 1. Input (chromatin after sonication); 2. Unbound (unbound chromatin that was previously diluted with ChIP buffer); 3. Matrix after chromatin binding before wash procedure; 4. Matrix after chromatin binding after whole wash procedure; 5. Matrix after elution; 6. Elution. Figure S2 Telomeric sequence is highly enriched compared to the centromeric DNA or 18S rDNA DNA cross-linked with TRB1 protein was isolated by ChIP analysis using GFP-TRAP_A matrix from Wt and TRB1pro:TRB1-GFP plants. Subsequent hybridisation of isolated TRB1associated DNA with radioactively labelled telomeric, centromeric 18S rDNA probes in three biologically and technically replicated experiments has shown marked enrichment of telomeric DNA contrasting with only negligible enrichment of the centromeric DNA or 18S rDNA. The relative difference of the probed DNA co-precipitated with TRB1-GFP compared to the Wt was measured in each replicate (1, 2, 3; enrichment was related to Wt=1) using Multi Gauge software. Figure S3 Location of antibody recognition sites within the structure of TRB1 protein with ELISA Construction of vectors and protein expression/purification protocols of TRB1 - 3 proteins or TRB1 domains were previously described in (Schrumpfova, Kuchar et al. 2004, Mozgova, Schrumpfova et al. 2008). Purified proteins as antigens were coated onto a 96-well microtiter plate (Nunc MaxiSorp) with 500 ng / well of antigen. As primary antibody, we used supernatants of tested monoclonal hybridomes anti-TRB 1.2 and 5.2 (1:200). Colouring intensity of secondary antibody anti-mouse HRP (DAKO) (1:10 000) visualised with TMB (Test line) was measured as absorbance at 450 nm after 30min. Values of measured absorbance are listed in the table. Figure S4 Anti-TRB 1.2 and 5.2 antibodies were unable to detect proteins from TRFL family Genes coding for TRFL2, TRFL9 and TRP1 proteins were cloned into pET28 vector (Karamysheva, Surovtseva et al. 2004). Radioactively labeled proteins were synthesized by in vitro transcription and translation using rabbit reticulocyte system (Promega), separated with SDS-PAGE and transferred to nitrocellulose membrane (as described in (Schrumpfova, Kuchar et al. 2008)). Proteins were visualised using autoradiographic analysis with Typhoon FLA7000 (GE Healthcare) (35 S-labeled) or probed with the monoclonal antibodies (anti-His (Sigma); anti-TRB1 1.2; anti-TRB1 5.2 (1:5 000)). As a second antibody, polyclonal anti-mouse rabbit HRP (DAKO) was used. Both antibodies were diluted 1:5 000. Immunoreactive bands were visualised with LumiGLO Reagent and Peroxide (Cell Signaling Technology) a Fujifilm LAS-3000 CCD system. Extracts from isolated nuclei used in ChIP experiment were used as positive control (see Figure 3D). Figure S5 Anti-TRB1 1.2 does not detect any signal on trb1-/- plants, while 5.2 recognises some epitopes also in trb1-/- mutant plant lines Antibodies 1.2 and 5.2. (green) were tested on isolated nuclei from trb1-/- and TRB1pro:TRB1-GFP plants (TRB1-GFP). Nuclei are counterstained with DAPI (magenta). (a) 1.2 is specific to TRB1 protein since it recognises epitopes in the sample from TRB1-GFP plants, but not in trb1-/-. (b) 5.2 antibody recognises also other members of the SMH protein family, thus producing a detectable signal in trb1-/- nuclei. Figure S6 Telomere shortening in trb1-/- plants is moderately progressive. Terminal restriction fragment analyses of trb1-/- plant lines A, B and C were performed in their 5th generation (G5) and in earlier generations (G2 or G4). Results are expressed relatively to the Wt control (Wt). Figure S7 Whole images of protoplasts (whose segments are shown in Figure 5C) obtained by bimolecular fluorescence complementation. Interactions of TRB proteins with TERT fragments are depicted with the arrows. Plasmids encoding nEYFP-tagged TRB clones, cEYFP-tagged TERT fragments or Gaut10 (as negative control), and mRFP-VirD2NLS (to mark cell nuclei and to determine transfection efficiency). [YFP fluorescence (yellow), mRFP fluorescence (red), and chloroplasts (green pseudocolor); chloroplast autofluorescence also visible in the YFP chanell]. Bar = 20 µm. Figure S8 TRB1, 2 and 3 proteins are able to pull-down TERT fragments The TNT expressed TRB1, 2 and 3 (35S-labelled*) were mixed with fragments of TERT (myc-tag) and incubated with anti-myc antibody. In the control experiment, the TRB proteins were incubated with myc-antibody and beads in the absence of partner protein. Input (I), unbound (U), and bound (B) fractions were collected and run in SDS–12% PAGE gels. Interactions of TRB2 and 3 proteins with TERT fragments appear to be relatively weaker than interaction between TRB1 and TERT fragments. Together, these results support yeast two-hybrid or bimolecular fluorescence complementation experiments. Figure S9 Telomerase activity or processivity in vitro is not changed in response to TRB1 status. Ladder of telomeric repeats produced in TRAP assay (telomere repeat amplification protocol) has shown no differences in the trb1 mutant or TRB1-GFP plant lines compared to wild-type lines. In negative control (-), no protein extract was used Figure S10 TRP1 protein interacts with plant telomerase (TERT) Bimolecular fluorescence complementation has shown interaction of TRP1 protein with TERT fragments. Arabidopsis leaf protoplasts were transfected with 10 μg of each of plasmids encoding nEYFP-tagged TRP1 clones, cEYFP-tagged TERT fragments or Gaut10 (as negative control), and mRFP-VirD2NLS (to mark cell nuclei and to determine transfection efficiency). The cells were imaged by epifluorescence microscopy after overnight incubation. Protoplast image [YFP fluorescence (yellow), mRFP fluorescence (red), and chloroplasts (green pseudocolor); chloroplast autofluorescence also visible in the YFP chanell] we can observe clear nuclear interactions of TRP1 proteins with TERT fragments depicted with the arrows. Bar = 20 µm. Table S1 Quantification of TRB1 foci co-localised/associated with telomeric foci Isolated nuclei from leaves transformed with GFP-TRB1 were used for telomere PNA FISH. Number of fluorescence foci of GFP-TRB1 signal co-localised with telomeric signal is numbered in this Table. These FISH results showed that TRB1 speckles and telomere signals are associated (31%) or co-localised (59%) cases Table S2 The sequences of all primers and probes used in this study. Name Sequence (5 ′→ 3 ′ ) P1 GAGAGGAGAAGATAAAGATGTCACC P2 CGTTCTCCCTTCCTAAACAGG P3 CAACACTCAACCCTATCTCGG P4 AGTATTCATATGGGTGCTCCTAAGCAGA P5 CGGGATCCTCAGGCACGGATCATCATT TR10-24-G AGTACCAGCCATGACCAGCCATGA TR10-24-C TCATGGCTGGTCATGGCTGGTACT TR-4C CTAAACCCTAAACCCTAAACCCTAAACC TS21 GACAATCCGTCGAGCAGAGTT TEL-PR CCGAATTCAACCCTAAACCCTAAACCCTAAACCC 5´UTR Fw GGGGACAACTTTGTATAGAAAAGTTGTCCACCCATTAGAGGGACGAGTATGG TRB1 Rev GGGGAC TGC TTT TTTGTACAA ACTTGCGGCACGGATCATCTGTCGAAT 3´UTR Fw GGGGACAGCTTTCTTGTACAAAGTGGCCggtaatggaaagcgagagaagaag 3´UTR Rev GGGGACAACTTTGTATAATAAAGTTGCTATTTTAGTATGTCA AATTTCGGATGA F-TRB1_BstBI CCTTCGAAATGGGTGCTCCTAAGCAGAAATG F-TRB2_BstBI CCTTCGAAATGGGTGCACCAAAGCAGAAG F-TRB3_BstBI CCTTCGAAATGGGAGCTCCAAAGCTGAAG R-TRB1_SmaI ATCCCGGGGGCACGGATCATCATTTTGCAG R-TRB2_SmaI ATCCCGGGCCAAGGATGATTACGGATCCTG R-TRB3_SmaI ATCCCGGGCCGAGTTTGGCTATGCATTCTATAC F-TERT_KpnI atggtaccATGCCGCGTAAACCTAGACATC R-RID1+BamHI GAGGATCCTTAGGGAGTTATACAAGGAGCATTAC F-F2N_KpnI taggtaccGGCGAGGATGTAGACCAACAT R-F2N+BamHI GAGGATCCCTACCAGCTCCTTTTCCGGTA Supplement I Schrumpfová, P.P., Vychodilová, I., Hapala, J., Schořová, Š., Dvořáček, V., Fajkus, J., 2016. Telomere binding protein TRB1 is associated with promoters of translation machinery genes in vivo. Plant Mol.Biol. 90, 189–206. P.P.S. participated in the design of experiments, significantly involved in the experimental part, data evaluation, and in the ms writing and editing Telomere binding protein TRB1 is associated with promoters of translation machinery genes in vivo Petra Procha´zkova´ Schrumpfova´1,2 • Ivona Vychodilova´1,2 • Jan Hapala1,2 • Sˇa´rka Schorˇova´1,2 • Vojteˇch Dvorˇa´cˇek1,3 • Jirˇı´ Fajkus1,2,3 Received: 29 July 2015 / Accepted: 16 November 2015 / Published online: 23 November 2015 Ó Springer Science+Business Media Dordrecht 2015 Abstract Recently we characterised TRB1, a protein from a single-myb-histone family, as a structural and functional component of telomeres in Arabidopsis thaliana. TRB proteins, besides their ability to bind specifically to telomeric DNA using their N-terminally positioned myb-like domain of the same type as in human shelterin proteins TRF1 or TRF2, also possess a histone-like domain which is involved in protein–protein interactions e.g., with POT1b. Here we set out to investigate the genome-wide localization pattern of TRB1 to reveal its preferential sites of binding to chromatin in vivo and its potential functional roles in the genome-wide context. Our results demonstrate that TRB1 is preferentially associated with promoter regions of genes involved in ribosome biogenesis, in addition to its roles at telomeres. This preference coincides with the frequent occurrence of telobox motifs in the upstream regions of genes in this category, but it is not restricted to the presence of a telobox. We conclude that TRB1 shows a specific genome-wide distribution pattern which suggests its role in regulation of genes involved in biogenesis of the translational machinery, in addition to its preferential telomeric localization. Keywords Telomere repeat binding (TRB) Á ChIP-seq Á Arabidopsis thaliana Á Ribosome Á snoRNA Á Translation machinery Introduction Telomere binding proteins and their complexes, exemplified by the shelterin complex in vertebrates (de Lange 2005), perform essential functions at chromosome ends. Primarily, they inhibit DNA-damage responses at telomeres to protect them from being mis-recognized as unrepaired chromosome breaks, thus solving the so-called endprotection problem (Sfeir and de Lange 2012) of linear chromosomes. Other functions of telomere proteins include, for example, telomerase recruitment and coordination of telomere elongation by telomerase with lagging strand synthesis by DNA polymerase during telomere replication (thereby solving the end-replication problem) (Latrick and Cech 2010; Sfeir et al. 2009; Soudet et al. 2014). However, the functions attributed to telomerebinding proteins are presumably not the original functions of these proteins in earlier stages of evolution, which preceded the onset of linear chromosomes connected with the necessity to solve the end-replication problem (Fajkus et al. 2005; Louis and Vershinin 2005; Nosek and Tomaska 2003; Valach et al. 2011). Examples supporting this notion can be seen, for example, in DNA repair proteins which paradoxically are also implicated in the control of telomere organization and length although their Electronic supplementary material The online version of this article (doi:10.1007/s11103-015-0409-8) contains supplementary material, which is available to authorized users. & Jirˇı´ Fajkus fajkus@sci.muni.cz 1 Mendel Centre for Plant Genomics and Proteomics, CEITEC – Central European Institute of Technology, Masaryk University, Kamenice 5, 625 00 Brno, Czech Republic 2 Laboratory of Functional Genomics and Proteomics, National Centre for Biomolecular Research, Faculty of Science, Masaryk University, Kamenice 5, 625 00 Brno, Czech Republic 3 Institute of Biophysics, Academy of Sciences of the Czech Republic, v.v.i., Kra´lovopolska´ 135, 61265 Brno, Czech Republic 123 Plant Mol Biol (2016) 90:189–206 DOI 10.1007/s11103-015-0409-8 presence at telomeres apparently contradicts the end-protective telomere functions (Kazda et al. 2012; Weaver 1998). One of the best known telomere-binding proteins in budding yeast, the repressor-activator protein 1 (RAP1), is also well known for its involvement in gene activation and repression and in DNA replication. Further studies have examined additional roles for RAP1 in heterochromatin boundary-element formation, creation of hotspots for meiotic recombination, and chromatin opening (reviewed in Morse 2000). The TTAGGG DNA repeat-binding proteins 1 and 2 (TRF1 and TRF2) bind to mammalian telomeres as part of the shelterin complex and are essential for maintaining chromosome end stability. While most of their binding sites identified in a chromatin immunoprecipitation sequencing (ChIP-seq) study corresponded to telomeric regions, these two proteins also localize to extratelomeric sites (Simonet et al. 2011) of which the vast majority contain interstitial telomeric sequences (ITSs). However, non-ITS sites were also identified which correspond to centromeric and pericentromeric satellite DNA, and these TRF-binding sites are often located in the proximity of genes or within introns. It was thus suggested that TRF1 and TRF2 may couple the functional state of telomeres to the long-range organization of chromosomes and gene regulation networks by binding to extratelomeric sequences (Simonet et al. 2011). Even the specific telomere-elongation tool, telomerase, is involved in a number of non-telomeric processes (reviewed in Majerska et al. 2011). These examples demonstrate that the function of telomere-localized proteins may not be exclusively telomere-associated and that telomere metabolism/protection may be mediated by proteins which play more general roles in the genome. Understanding of the composition and function of telomere-binding protein complexes in plants lags behind that in animals and yeasts. Nevertheless, we have recently characterized a key candidate shelterin-like component belonging to the plant-specific single-myb-histone group of proteins termed TRB (Telomere repeat binding) (Schrumpfova et al. 2014). In addition to our earlier studies which demonstrated specific binding of proteins from this group to telomeric DNA in vitro (Schrumpfova et al. 2004) and characterized their DNA-protein and protein–protein interactions in detail (Hofr et al. 2009; Mozgova et al. 2008; Prochazkova Schrumpfova et al. 2008), our recent study revealed preferential co-localization of a member of this group, TRB1, with telomeres in situ and in vivo, telomere shortening in trb1 knockout mutants, and moreover its physical interaction with the catalytic subunit of telomerase, TERT (Prochazkova Schrumpfova et al. 2014). These results, together with our previous findings of TRB1 interaction with POT1b protein, one of the paralogs of Protection Of Telomeres (POT1) protein (Kuchar and Fajkus 2004; Schrumpfova et al. 2008) and the data on interaction of POT1 proteins with telomerase (CifuentesRojas et al. 2011; Rossignol et al. 2007), make this protein currently the best-established component of a putative plant shelterin complex. In addition to the results demonstrating sequencespecific binding of TRB1 to telomeric DNA and corresponding telomere-specific functions, this protein is also capable of binding to chromatin through protein–protein interactions or sequence-non-specific interactions with DNA via its H1/H5-like domain (Mozgova et al. 2008). In vivo, the protein shows highly dynamic association with chromatin and preferential localization to the nucleus and the nucleolus during interphase (Dvorackova et al. 2010). TRB1 localization is cell cycle-regulated, as the level of nuclear-associated TRB1 diminishes during mitotic entry and it progressively re-associates with chromatin during anaphase/telophase. Using fluorescence recovery after photobleaching and fluorescence loss in photobleaching, we determined that TRB1 interaction with chromatin is regulated at two levels at least, one of which is coupled with cell-cycle progression with the other involving rapid exchange (Dvorackova et al. 2010). These results strongly suggest additional roles for TRB1 connected with chromatin function. In this study, we thus set out to investigate the genomewide localization pattern of TRB1 to examine its preferential sites of binding to DNA in vivo and its potential functional roles in the genome-wide context. Using ChIP followed by next generation sequencing (ChIP-seq), we show that TRB1 associates with promoter regions of certain genes in addition to binding long telomeric repeats. Classification of these genes using GO analysis revealed a strong link between TRB1 binding and promoters of translation machinery-related genes. Materials and methods Plant material and construct generation The TRBpro:GFP-TRB1 and trb1-/- plants and constructs were described previously (Schrumpfova et al. 2014). All the A. thaliana plants used in this study have a Col0 background. Chromatin immunoprecipitation assay The ChIP assay was performed as described (Bowler et al. 2004) with modifications described in (Schrumpfova et al. 2014). Chromatin extracts were prepared from seedlings treated with 1 % formaldehyde. The chromatin from 190 Plant Mol Biol (2016) 90:189–206 123 isolated nuclei was sheared to an average length of 250–500 bp by sonication (Bioruptor, Diagenode) and centrifuged. Anti-TRB1 5.2 antibody (Schrumpfova et al. 2014) was bound to a Protein G agarose matrix (Pierce) for 3 h at 4 °C which was subsequently washed with ChIP dilution buffer (16.7 mM Tris–HCl, pH 8.0, 1.2 mM EDTA, 167 mM NaCl, 0.1 % Triton X-100, 0.1 mM PMSF and protease inhibitors). The experimental procedures for isolating TRB-GFP protein on GFP-TRAP_A (Chromtec) and native TRB1 protein on an Protein G agarose matrix (Pierce) with bound anti-TRB1 antibody were as described in (Schrumpfova et al. 2014). Both matrices were blocked against non-specific interaction with 200 mM ethanolamine, 1 % BSA and TR10-24-G and TR10-24-C (Schrumpfova et al. 2014), incubated with chromatin diluted with ChIP dilution buffer, and subsequently washed with low salt; high salt; LiCl; TE buffers as in Bowler (Bowler et al. 2004). The cross-linking was reversed by 0.2 M NaCl overnight followed by proteinase K (Serva) treatment, phenol/chlorophorm extraction, and RNase A (Serva) treatment. ChIP-qPCR and RT-qPCR reactions Primers used for ChIP-qPCR are described in Fig. 4 and are listed in Supplementary Table S1a. Three ul of ChIP DNA was added to the 20 ul reaction mix of FastStart SYBR Green Master (Roche) and the final concentration of each forward and reverse primer was 0.4 uM. Reactions were done in triplicate; the PCR cycle consisted of 5 min of initial denaturation at 95 °C followed by 40 cycles of 5 s at 95 °C, 20 s at 60 °C and 15 s at 72 °C. At least two biological replicates in two technical replicates were analysed. The relative copy number of each selected gene in Wt (-) or TRB-GFP (?) samples compared to that in the genomic input fraction was calculated by the 2-DCt method (Pfaffl 2004). Total RNA was isolated from Arabidopsis seedlings using the RNeasy Plant Mini Kit (Qiagen) followed by the DNaseI treatment (TURBO DNA-free, Applied Biosystems/Ambion) according to the manufacturers’ instructions. The quality and quantity of RNA was checked by electrophoresis on 1 % (w/v) agarose gels and by absorbance measurement (NanoPhotometr IMPLEN). cDNA was prepared by reverse transcription of 1 lg of RNA using M-MuLV reverse transcriptase (NEB) and Random Nonamers (Sigma). Quantification of the transcripts of translation machinery related genes was related to the ubiquitin reference gene and was done using qPCR GreenMaster with UNG/lowROX (Jena Bioscience) by the Rotorgene6000 (Qiagen) machine. Four ll of 129 diluted cDNA was added to the 20 ll reaction mix; the final concentration of each forward and reverse primer was 0.3 lM (Supplementary Table S1b). Reactions were done in triplicates; the PCR program consisted of 5 min of initial denaturation at 95 °C followed by 40 cycles of 15 s at 95 °C, 20 s at 58 °C and 30 s at 72 °C. Analyses were performed for at least three biological replicates in three technical replicates. Transcript levels of chosen genes in the trb1-/- (Schrumpfova et al. 2014) were normalized to Ubq10 transcript and presented ‘‘relative transcript levels’’ were calculated as the fold increase/decrease relative to respective wild-type seedlings using (2-DDCt ) method (Pfaffl 2004). Library preparation and sequencing Library preparation and sequencing was done by the European Molecular Biology Laboratory (EMBL) Genomics Core Facility, Heidelberg, Germany. Fifty microliters of immunoprecipitated DNA [0.2–6 ng DNA measured by Qubit dsDNA HS Assay Kit (Invitrogen)] was used for library preparation using NEBNext ChIP Seq Library Prep Master Mix (NEB), with Agencourt XP beads (Beckman) in the ratios described in the protocol with only one exception, dilution of the Adapter 1:17 in water. DNA fragments of 270–300 bp were selected on 2 % e-gels (Invitrogen). Subsequently, a PCR reaction (18 cycles) with indexed Primers 1–11 from the NEBNext Multiplex Oligo set (NEB) was performed. The quality of the final libraries was checked on the Bioanalyzer (Agilent) and the quantity with the Qubit dsDNA HS Assay Kit (Invitrogen). The libraries were pooled equimolar and diluted to 10 pM (denaturation in NaOH) in Hyb Buffer 1 (HT1) from the TruSeq SR Cluster Kit v3-cBot-HS (Illumina). Libraries were sequenced (50 bp single-end reads) on an Illumina HiSeq 2500 using TruSeq SBS Kit v3-HS (Illumina) sequencing reagents. Bioinformatic methods Sequence logo construction The sequence logo was created using a k-mer analysis. We counted all substrings of length k in two datasets, the 50 UTR sequences and the 50 UTR peak-covered regions only. We decided to use 8-mers sorted by their number of occurrence, and for fragment reconstruction used only those whose number was at least twice as high in peak regions as in the whole 50 UTR dataset. The fragment was constructed by a modified version of the algorithm described in (Macas et al. 2010), and instead of extending the whole prefix and subsequently the whole suffix we extended both the prefix and the suffix alternately by one nucleotide. As fragments are composed of k-mers with diverse frequencies, every base of each fragment was Plant Mol Biol (2016) 90:189–206 191 123 assigned a weight representing the frequency of its k-mer of origin. For better matching, some fragments were used as their reverse complements with preference for C-rich variants of DNA. Fragments were aligned with MUSCLE (Edgar 2004). The sequence logo (Schneider and Stephens 1990) was then generated from fragments with mean weight higher than 20. The y-axis displays position weights totalled from the aligned fragments. Read mapping and filtration We mapped the Illumina fastq reads onto the reference genome of Arabidopsis thaliana (A_thaliana_Jun_2009) with Bowtie2 (using the very-sensitive option) (Langmead and Salzberg 2012). Unmapped reads or reads with a low quality mapping score (\25) were removed. Biological replicates were merged into single files. The coverage of the mitochondrial DNA was negligible compared to the chromosomal DNA, while the chloroplast DNA was represented in the data more frequently. Therefore, we separated the reads from each dataset into three files, genomic, mitochondrial and chloroplast sequences, and peaks were called for all groups independently. However, we did not detect any peaks in mitochondria and chloroplasts. We used average profiles and heatmaps to illustrate the distribution of ChIP-seq reads along genes (±2000 bp). The pictures were produced by ngs.plot (Shen et al. 2014) with parameters –R genebody, –F protein_coding. Peak calling We called peaks in the data files using two programs, MACS v. 1.4.3 20131216 and PePr v. 1.0.2 (Zhang et al. 2008b, 2014). MACS is one of the most widely used programs for peak calling in sequencing data while PePr is a new peak-calling software which detected peaks visually more accurately in our data. With MACS (parameters: effective genome size—g: 90000000, band width—bw 250, keeping all duplicate reads—keep-dup all, ChIP/control scaling ratio—ratio 1.215273 and 1.742786 for the first and second replicate, respectively) we obtained a set of peaks for each of the two replicates. The ChIP/control scaling ratios were calculated with NCIS (Liang and Keles 2012). We selected peaks with a false discovery rate (FDR) below 5 % from each replicate and calculated an intersection of these two sets. We produced another set of peaks with PePr (parameters: –peaktype = sharp) which accepts two replicates (both input and control files) and returns only peaks present in both of them. PePr peaks with too short (\30 bp) estimated fragment lengths in any of the two samples were removed (1879 peaks). Finally, we made an intersection (4995) of the PePr (24,219) and MACS (3690) peaks. Peak analysis We studied the relative amount of peaks in genomic regions with different functions. We downloaded annotated datasets provided by TAIR (ftp://ftp.arabidopsis.org/home/ tair/Sequences/blast_datasets/TAIR10_blastsets/) that contain sequences separated with respect to their position within or outside of genes. We converted these FASTA files into genomic coordinate files (BED) and calculated intersections with the 4995 TRB1 genomic peaks. Reference datasets We examined intersections with 50 UTR, 30 UTR (variable length), upstream and downstream sequences (500 bp) and intergenic regions (variable length, trimmed by 1000 bp at both ends). In addition, we created a dataset of coding sequences derived from a coordinate file TAIR10_GFF3_genes.gff also provided by TAIR. Coverage quantification We calculated dataset coverage as the length of each reference dataset relative to the genome size. Overlapping base pairs of sequences of the same type (i.e. those on opposite strands) were counted once. In addition, for each dataset peak coverage was obtained as the total length of the dataset-peak intersection relative to the total size (bp) of the dataset. Finally, peak relative occurrence represents the number of peaks intersecting with any sequence of the particular dataset relative to the total size of the dataset (presented in Mb). In all cases, two elements intersect each other if they have at least 1 bp-overlap. Quantification of telomeric, centromeric and 18SrDNA sequences We analysed the unprocessed Illumina files for the occurrence of telomeric, centromeric and 18SrDNA sequences by counting the number of reads containing their respective sequence motifs, in particular AAACCCTAAACCC TAAACCCT for telomeres, TATGAGTCTTTGGCTTT GTGTCTT for centromeres, and CTAGAGCTAATACGT GCAACAAAC for 18SrDNA. Telo-box occupation We examined sequences from all reference datasets for the presence of a telobox (any of AAACCCTA, AACCCTAA, ACCCTAAA, TAAACCCT, TAGGGTTT, TTAGGGTT, TTTAGGGT and AGGGTTTA). Sequences containing one or more telo-boxes were included. Next, we counted 192 Plant Mol Biol (2016) 90:189–206 123 telo-boxes intersected by a TRB1 peak and calculated telobox peak coverage as the ratio of these two numbers. Gene ontology (GO) enrichment We analysed the GO database using the program GoMiner (v. 2011-01) (Zeeberg et al. 2003, 2005). Only GO categories with a p value higher than 0.05 and categories with size 5–500 genes were arranged graphically in CIMminer software (http://discover.nci.nih.gov/cimminer/). Data availability The data have been deposited in NCBI’s Gene Expression Omnibus (Edgar et al. 2002) and are accessible through GEO Series accession number GSE69431 (http://www. ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE69431). Results TRB1 is associated with telomeric sequences in vivo Previously, we described the co-localization of TRB1 protein with telomeres in situ and showed that TRB1 can bind DNA containing telomeric sequences in vivo (Schrumpfova et al. 2014). Since TRB1 is capable of binding telomeric (via its myb-like domain) as well as nontelomeric DNA (via its histone-like domain) (Hofr et al. 2009; Mozgova et al. 2008; Schrumpfova et al. 2014), we pondered whether TRB1 is localized exclusively to telomeres or has also other binding sites in the genome in vivo. To address this question, we performed nextgeneration sequencing of DNA isolated by ChIP of native or GFP-tagged TRB1. As source material, we used formaldehyde cross-linked Arabidopsis seedlings stably transformed with a construct driven by the native promoter TRB1pro:TRB1-GFP [TRB1-GFP (?)] (Dvorackova et al. 2010). Wild-type plants were used as negative controls [Wt (-)]. The experimental setup is depicted schematically in Fig. 1a for anti-GFP-TRB1. TRB1-GFP protein was immunoprecipitated from purified nuclei using a GFP-Trap A matrix which contains a single variable antibody domain that recognizes GFP, as described in (Schrumpfova et al. 2014). Furthermore, we performed another ChIP analysis using anti-TRB1 (5.2) antibody that specifically recognizes the conserved region of the myb domain of TRB proteins (Schrumpfova et al. 2014) (Fig. 1b). Native TRB1 protein was isolated from wild-type plants [TRB1 (?)] using Protein G magnetic beads. In the negative control, no antibody was used in a parallel ChIP [Wt (-)]. DNA recovered from ChIP underwent ultra-high-throughput Illumina sequencing. For evaluation purposes, arrays of at least three perfect telomeric repeats [(AAACCCT)3] were considered as ‘‘long telomeric repeats’’ and reads containing them were counted in each sequenced sample. The relative differences in the numbers of long telomeric repeats in TRB1-GFP (?) or TRB1 (?) samples were expressed relative to the relevant Wt (-) = 1 (Fig. 1c). Absolute values confirmed our previous results (Schrumpfova et al. 2014) and demonstrated that the observed enrichment of long telomeric sequences in these samples is not due to the high copy number of telomeric DNA in the Arabidopsis genome but due to sequence-specific association of TRB1 with telomeric sequences. No enrichment of other repetitive sequences examined—centromeric DNA or 18S rDNA—was found in the ChIP-seq data. Association of TRB1 with euchromatin in vivo To define the genomic regions that are associated with TRB1 protein, we mapped the ChIP-seq data onto the TAIR A_thaliana_Jun_2009 assembly (TAIR10 annotation) and visualised the data using Integrative Genomics Viewer (IGV). To limit falsely identified peaks in the data, we used two independent peak calling programs, MACS and PePr (for details see Methods). We examined peaks in all five chromosomes of A. thaliana. Only peaks detected by both programs in both ChIP approaches (anti-GFPTRB1 and anti-TRB1) were used in subsequent analysis. This set contains 4995 peaks and here we term it ‘‘TRB1 genomic peaks’’. No peaks were detected in mitochondrial or chloroplast DNA. However, mitochondrial DNA coverage was negligible compared to the nuclear DNA. TRB1 genomic peaks are absent in centromeric regions or heterochromatic knobs localized on chromosomes 4 and 5, but appear associated with euchromatic genes in both variants of ChIP (Fig. 2a, b, and Fig. S1). As the TAIR9 assembly lacks clusters of repetitive sequences, long telomeric tracts or 45S rDNA repetitive sequences are not visible in the IGV viewer. Figure 2c shows the detailed enrichment of DNA regions in TRB1-GFP (?) and TRB1 (?) samples that were immunoprecipitated by ChIP and sequenced. It can be seen that only DNA regions in TRB1GFP (?) or TRB1 (?) samples which were highly enriched with respect to the Wt (-) were considered as TRB1 genomic peaks (marked with an asterisk). The data have been deposited in NCBI’s Gene Expression Omnibus (Edgar et al. 2002) and are accessible through GEO Series accession number GSE69431 (http://www.ncbi.nlm.nih. gov/geo/query/acc.cgi?acc=GSE69431). Plant Mol Biol (2016) 90:189–206 193 123 Analysis of TRB1 targets To examine preferential targeting of TRB1 protein to specific Arabidopsis genome loci we visualized distribution of ChIP-seq reads along protein-coding genes genes. Clear enrichment of TRB1 was observed in the vicinity of transcription start sites (TSS) and transcription end sites (TES) using both purification approaches—anti-GFP-TRB1 and anti-TRB1 (Fig. 3). For a more detailed resolution of TRB1 association with specific parts of genes, gene families and also non-protein coding genes we established various datasets covering specific parts of the A. thaliana genome according to TAIR10 blastsets (ftp://ftp.arabidopsis.org/home/tair/ Sequences/blast_datasets/TAIR10_blastsets/). The major division of these datasets is into Protein coding genes and Non-protein coding genes (Fig. 4a, yellow boxes). The group of Non-protein coding genes includes rRNA, snoRNA, snRNA, tRNA, ncRNA, miRNA, other RNAs, pseudogenes and transposable element genes. In the vicinity of both Protein coding and Non-protein coding gene categories, datasets designated as Upstream 500 bp (from TSSs) or Downstream 500 bp (from TES) were classified. Moreover, four extra datasets close to the Protein coding genes sets were analysed: 50 UTR, 30 UTR, Upstream translation start 500 bp, and Downstream translation stop 500 bp. To clearly distinguish associations of all datasets that closely surround coding genes with which the TRB1 protein is associated, we designed a further dataset named Intergenic covering regions beginning more than 500 bp in front of the Upstream 500 bp or more than 500 bp behind the Downstream 500 bp datasets, i.e. 1000 bp distant from the transcription start or stop site of Protein or of Non-protein coding genes. The relative coverage of the whole A. thaliana genome by each dataset (e.g., the dataset Upstream 500 bp of Fig. 1 Long telomeric tracts are tandem repeats preferentially targeted by TRB1 genome-wide. Schematic illustration of the two different approaches used in ChIP analysis: a first, TRB1 tagged with GFP was isolated using a GFP-Trap A matrix from crosslinked wildtype [Wt (-)] and TRB1pro:TRB1-GFP [TRB-GFP (?)] seedlings. b Second, native-TRB1 [TRB1 (?)] protein was isolated from crosslinked seedlings using anti-TRB1 antibody linked to a Protein G matrix; wild-type seedlings [Wt (-)] were used as a negative control. The associated DNA was subjected to next-generation sequencing (NGS) using the Illumina platform. Two biological replicates were analysed by both approaches. c Marked enrichment of long telomeric DNA repeats (AAACCCT)3 compared to other repetitive sequences (centromeric or 18S rDNA) in ChIP-seq samples showing preferred association of TRB1 with telomeric sequence. The relative difference in the amount of the selected repetitive sequence was measured in each sequenced sample (biological replica) separately and enrichment was expressed relative to the relevant [Wt (-) = 1] 194 Plant Mol Biol (2016) 90:189–206 123 Protein coding genes covers 10.8 % of the whole genome) and the coverage of individual datasets by TRB1 genomic peaks obtained by combining both purification approaches (see above), is listed in Fig. 4b. It can be seen that in the vicinity of Protein coding genes, approximately 10 % of Upstream datasets (green boxes, 9.7–9.9 %) are covered by the TRB1 protein. This is approximately twice the value for TRB1 association with the entire Protein coding genes (yellow box, 3.9 %) or Downstream datasets (blue boxes, 4.8–6 %) although generally Upstream and Downstream datasets cover roughly the same proportion of the A. thaliana genome. The enrichment of TRB1 protein association with the Upstream in comparison to the Downstream genomic loci is obvious even in the Non-protein coding genes (3.7 % for Upstream compared to 0.8 % for Non-protein coding genes and 2.9 % for Downstream genomic loci). In contrast, the TRB1 protein is associated only with 1.5 % of the Intergenic sequences although these sequences cover almost 14 % of the genome. The preferential association of the TRB1 protein with sequences in Upstream coding genes is shown in the graph in Fig. 4c featuring the relative occurrence of TRB1 genomic peaks per sequence (dataset) length (displayed per Mb). The total bp size and the number of TRB1 genomic peaks in each dataset are listed in the Supplemental Table S2. The overall summary of sequences belonging to each dataset is given in the Supplemental Table S3. Taken together, our results demonstrate that the TRB1 protein is preferentially associated in vivo with promoter regions of coding genes compared to gene bodies or Downstream genomic loci. Association of TRB1 with Intergenic sequences is very low. This detailed analysis Fig. 2 TRB1 is preferentially associated with euchromatic regions. a IGV view of All coding genes, i.e. Protein coding genes and Nonprotein coding genes (transposable element genes, rRNA, snoRNA, snRNA, tRNA, ncRNA, miRNA and pseudogenes) on chromosome 4. Regions where TRB1 binding was enriched are depicted as TRB1 genomic peaks. The absence of TRB1 association in the centromeric region and the heterochromatic knob (grey box) is visible. b IGV view of a 1.4 Mb region of the short arm of chromosome 4 (position 0.9–2.3 Mb) where the heterochromatin knob is located. Only Protein coding genes or ‘‘Transposable element genes’’ are depicted in separate lines below (modified from GBrowse). Significant enrichment of the TRB1-associated regions within euchromatin is visible not only as increasing peaks in IGV viewer [boxes: TRB1-GFP (?) and TRB1 (?)] but also in the line TRB1 genomic peaks. c The IGV view of a 620 kb region displays regions highly associated with TRB1 proteins used for further analyses (asterisk) Plant Mol Biol (2016) 90:189–206 195 123 localizes extratelomeric TRB1-DNA association in euchromatin, especially in promoter regions. Validation of TRB1 binding sites by ChIP-qPCR To verify the TRB1 binding sites identified by ChIP-seq (Fig. 5a), we performed an independent ChIP experiment using the same seedlings as those used for the ChIP-seq analysis and purified DNA, and quantified the abundance of selected loci by qPCR. Preferential association of TRB1 protein with DNA regions that were identified as TRB1 genomic peaks was verified for several selected genomic loci (examples are shown in Fig. 5b). We confirmed by qPCR that all selected TRB1 binding sites are enriched in TRB1-GFP (?) samples with respect to the Wt (-) controls (examples of ribosomal protein L34, S5 and snoRNA Fig. 3 TRB1 specifically associates with TSS and TES of genes. Distribution of ChIP-seq reads obtained using a immunoprecipitation of TRB1-GFP protein relative to negative control or b immunoprecipitation of native TRB1 protein relative to negative control, illustrated by average gene profiles (upper panels) and heatmaps (lower panels) 196 Plant Mol Biol (2016) 90:189–206 123 Fig. 4 Detailed analysis of TRB1-bound regions. a Schematic representation of the selected ‘‘datasets’’ (coloured boxes) used in detailed analysis of the DNA regions associated with TRB1 protein. These datasets were designed according to the TAIR10 annotation. The group of Non-protein coding genes contains genes depicted in TAIR10 as rRNA, snoRNA, snRNA, tRNA, ncRNA, miRNA, pseudogenes, transposable element genes and other RNAs. The Intergenic dataset is a collection of regions located at least 1000 bp from the transcription start or stop site (Upstream or Downstream 500 bp) of the Protein coding genes or of the Non-protein coding genes to avoid interference with other datasets. b An overview shows the percentage representation of each analysed dataset within the A. thaliana genome: ‘‘Genome coverage by whole dataset’’. ‘‘Dataset coverage by TRB1 genomic peaks’’ represents a portion of the dataset overlapping with the TRB1-enriched regions. c The number of the TRB1 genomic peaks counted per size unit (1 Mb) of each dataset Fig. 5 Confirmation of the TRB1 protein binding sites by qPCR. Two ribosomal protein coding genes and one snoRNA coding region identified among TRB1 genomic peaks, with ACTIN 7 as negative control, were chosen for quantitativePCR (qPCR) verification of regions enriched in ChIP-seq and defined as TRB1 genomic peaks. a IGV view of the TRB1 genomic peaks in Wt (-) or TRB-GFP (?) ChIP-seq samples where PCR amplicons are depicted by triangles below the simplified gene models. b Yaxis values represent the abundance of the DNA recovered from each locus related to the input DNA. Mean ± SD of three technical replicates are shown Plant Mol Biol (2016) 90:189–206 197 123 are shown). In contrast, a site that was not included in the list of TRB1 genomic peaks (ACTIN 7) did not show any significant difference between TRB1-GFP (?) and Wt (-) control. The telomeric repeat is the most conserved motif associated with TRB1 To determine the most frequent motif associated with TRB1 protein, we selected the dataset with the highest relative occurrence of TRB1 genomic peaks per 1 Mb, i.e. the 50 UTR dataset, and extracted regions associated with the TRB1 protein for further analysis. As promoter regions could be biased for some motifs, the TRB1-binding motif was reconstructed as a sequence logo from 8-mers twice as abundant in the TRB1-associated regions as in the whole 50 UTR dataset (see Table S4a, b). The sequence logo shows that the telomeric repeat is highly represented in the sequences recognised by the TRB1 protein (Fig. 6a). If the plant telomeric repeat unit (CCCTAAA) and its circular permutations are removed (Fig. 6b), a related motif (i.e. sequences enriched for CCTA) is still clearly overrepresented (Fig. 6c). TRB1 is bound to short telomeric sequences in promoters Binding of TRB proteins to the telomeric DNA sequences was initially shown in vitro (Schrumpfova et al. 2004) and binding of TRB proteins expressed in E. coli to the synthetic telomeric oligo was described in detail (Hofr et al. 2009; Mozgova et al. 2008; Schrumpfova et al. 2004). The present study confirmed recent observation by hybridisation with a radioactively labelled telomeric probe which showed that the DNA immunoprecipitated with TRB1 protein is enriched for long telomeric sequences (Schrumpfova et al. 2014). The Arabidopsis genome contains short interspersed segments of the telomeric sequence both in terminal and interstitial positions, as well as the long telomeric repeats. These short interstitial telomere motifs, termed telo-boxes, exhibit a non-random distribution. They were described in the promoters of genes coding for translation elongation factor EF1a (Liboz et al. 1990), promoters of many ribosomal protein coding genes (Tremousaygue et al. 1999), and promoters of genes involved in the biogenesis of the translation machinery (Gaspin et al. 2010). The occurrence of telo-box motifs in the TRB1-associated regions from the above mentioned datasets was examined. The motifs analysed, AAACCCTA, AACCCTAA or ACCCTAAA (and their reverse complements), represent the most frequent permutations of the shortest telo-box motifs, as described in (Gaspin et al. 2010). Only telo-boxes directly covered by a TRB1 genomic peak were counted. Almost 28 % of telobox-sequences located in the 50 -UTR region of the Protein coding genes are covered by TRB1 peaks (Fig. 7 and Table S5). The total number of telo-boxes from each dataset and the number of TRB1 genomic peaks that Fig. 6 The most conserved DNA motif associated with the TRB1 protein. Sequence logos were constructed from the most frequent 8-mers. a The total logo was created from the most frequent 8-mers present in the 50 UTR regions and covered by the TRB1 genomic peaks. b Only 8-mers containing any permutation of at least one plant telomeric repeat were used for sequence logo construction with telomeric repeat. c The rest of the 8-mers, e.g. without any permutation of the plant telomeric repeat, were used for sequence logo construction without telomeric repeat 198 Plant Mol Biol (2016) 90:189–206 123 overlap these telo-boxes are listed in the supplementary Table S6. In general, it is evident that the TRB1 protein is associated at least twice as frequently with telo-box sequences in the Upstream datasets (22–28 %; green boxes) in comparison with the Downstream datasets (10–15 %; blue boxes) or with the low number (4 %) of Interstitial sequences. The number of telo-box genes with which the TRB1 protein is associated is presumably underestimated here, because the two independent immunoprecipitation approaches and the very stringent detection and refinement of the TRB1 genomic peaks by two independent programs limits false positives to a minimum, while also discarding many true peaks. Therefore, although TRB1 is visibly associated with e.g., all three promoters of eEF1 alpha-1, - 2 and -3 telo-box-containing genes (At1g07940, At1g07930 and At1g07920) (Figure S2), a TRB1 genomic peak was reported only in the eEF1-alpha 1 gene (At1G07940). Our strict policy of selection of TRB1 genomic peaks is also visible in the Supplementary Figure S3, where the genes coding for ribosomal proteins RPS15B, RPS15C, RPS15D (At5g09490, At5g09500, At5g09510) are shown. TRB1 protein is associated with genes related to ribosome biogenesis As most of the TRB1-associated loci are located in gene promoter regions, we wanted to know whether these genes are related in their function, biological processes, or subcellular localization of their products. We selected genes with TRB1-enrichment in their upstream region (Upstream 500 bp dataset) and thus both the categories Protein coding and also Non-protein coding genes are included in this summary. The Arabidopsis Genome Initiative numbers (AGI numbers) of these genes were compared to the set of AGI numbers of all genes in the Upstream 500 bp datasets using Gene Ontology miner (GOMiner) (Zeeberg et al. 2003, 2005). The table of GO subcategories GO:0005575 Cellular component and GO:0008150 Biological process arranged graphically by Clustered Image Maps miner (CIMminer) demonstrates a statistically significant enrichment of these GO subcategories (p B 0.05) (Fig. 8). The total output of GOMiner and also of CIMminer with all genes and detailed description (e.g. name, gene type, function) is shown in Supplementary Table S7. As is clearly visible in Fig. 7, protein coding genes with whose promoters the TRB1 protein is associated are enriched in GO:0005575 Cellular component in categories like GO:0005730 nucleolus/GO:0031981 nuclear lumen/ GO:0005654 nucleoplasm/GO:0016604 nuclear body, etc. (GO categories highlighted in light orange). Also, terms associated with small and large ribosomal subunit (GO:0005732 small nucleolar ribonucleoprotein complex/ GO:0015934 large ribosomal subunit/GO:0022626 cytosolic ribosome) (dark orange) are highly represented. In the category GO:0008150 Biological process, many genes coding for ribosomal proteins were statistically enriched in subcategories GO:0071843 cellular component biogenesis at cellular level/GO:0042254 ribosome biogenesis/GO:0022613 ribonucleoprotein complex biogenesis (highlighted in dark orange). Names and functions of individual genes are visible in Supplementary Table S7, list ‘‘CIM with genes’’. Also, a significant number of intergenic snoRNA genes are bound by TRB1 protein in their Fig. 7 The TRB1 protein preferentially occupies telo-box sequences. Sequences with at least one permutation of the telomeric repeat were extracted from each dataset. The chart shows the proportion of the sequences containing a telo-box motif that are covered by a TRB1 genomic peak. The clear preference of the TRB1 protein for telo-box sequences is especially apparent in the 50 -UTR dataset in which almost 28 % of the telo-box sequences located in the 50 -UTR are recognized by the TRB1 protein Plant Mol Biol (2016) 90:189–206 199 123 Upstream 500 bp region and are included in the CIMminer Table. These snoRNA genes are highly enriched in categories such as e.g., GO:0034660 ncRNA metabolic process/GO:0000154 rRNA modification/GO:0006364 rRNA processing, etc. Discussion After recent findings, the previously-established concept that the shelterin proteins are located exclusively on the chromosome ends, the telomeres, has been superseded. Fig. 8 Schematic view of Gene Ontology (GO) classification of categories significantly enriched in TRB1 protein. All coding genes (Protein coding genes together with Non-protein coding genes) that contain a TRB1 genomic peak in their 500 bp Upstream proximity (datasets: Upstream 500 bp) were analysed by the GOMiner software. The set of genes in which TRB1 protein is enriched in their 500 bpupstream region was compared to the set of all coding genes from A. thaliana. General tables of the GO subcategories ‘‘GO:0005575 Cellular components’’ and ‘‘GO:0008150 Biological processes’’ were arranged graphically by CIMminer. The horizontal axes represent significantly enriched GO subcategories (p B 0.05) and the vertical axes represent individual genes, in clusters. Genes that contain a TRB1 genomic peak within 500 bp from the transcription start site and are significantly enriched in the GO subcategories are highlighted as red boxes. For a detailed description of individual genes, see Supplementary Table S7 200 Plant Mol Biol (2016) 90:189–206 123 Emerging evidence indicates that the shelterin components also possess non-telomeric functions such as transcriptional regulation (Martinez et al. 2010; Zhang et al. 2008a), DNA repair (Bradshaw et al. 2005), NF-jB activation (Teo et al. 2010), Epstein-Barr virus replication (Deng et al. 2002), or regulation of mitochondrial oxidative phosphorylation (Chen et al. 2012). Some of these non-telomeric functions could be explained, at least partially, by their binding to ITSs (Bosco and de Lange 2012; Krutilina et al. 2003; Mignon-Ravix et al. 2002; Simonet et al. 2011) or even to unrelated DNA sequences which remain to be identified in future studies. Association of TRB1 with telomeric sequences Besides the terminally located long telomeric repeats, the A. thaliana genome contains two long interstitial telomeric tracts that consist of degenerate telomeric repeats with islands of perfect plant telomeric sequences (Uchida et al. 2002). In a previous study we speculated on the telomeric association of GFP–TRB1 speckles in A. thaliana (Dvorackova et al. 2010), and using a plant system with longer telomeres, N. benthamiana, we showed clear in situ colocalization of TRB1-GFP with telomeres in leaves (Schrumpfova et al. 2014). However, localization of TRB1 in many speckles in A. thaliana nuclei, together with the fact that the TAIR9 assembly, used as the reference genome, lacks not only telomeres but also interstitial telomeric regions, did not allow us to determine whether the TRB1associated telomeric tracts are terminally or interstitially located. In the current study we describe a significant enrichment of perfect long plant telomeric repeats associated with the TRB1 protein using ChIP-seq. TRB1 specific binding to telomeric sequence Many different telomere-binding proteins possess a single myb-related DNA-binding domain with Helix-Turn-Helix (HTH) organization (reviewed in Peska et al. 2011). In the case of human TRF1 and TRF2, the X-ray crystal structures of complexes with telomeric DNA show that they recognize the same AACCCTA binding site by means of homeodomains, as does the yeast telomeric protein Rap1p; TRF dimers specifically recognize two of the three G-C base pairs in the major groove that characterize telomeric repeats (Court et al. 2005; Hanaoka et al. 2005; Nishikawa et al. 2001). Our search for a sequence-specific motif recognized by TRB1 revealed that the telo-box sequence is highly abundant in promoter regions, and moreover the core of this sequence (CCTA) containing two G-C base pairs is still present if we focus only on sequences without the complete telo-box motif. Thus we conclude that potential TRB1-binding motifs are associated with short interstitial telomere sequences as well as with long telomeric tracts, and furthermore the TRB1 protein binds to the core segment of telomeric repeats (CCTA) through its myb-domain in the same manner as human TRF proteins. Preferential association of TRB1 with promoter regions The landscape of the A. thaliana epigenome can be classified into four main chromatin states (Roudier et al. 2011) which have been further subdivided in a recent more detailed study, providing a total of nine chromatin states (Sequeira-Mendes et al. 2014). Only a small proportion of the TRB1 protein is associated with two heterochromatin states, mostly located in the centromeric regions or knobs on the short arm of chromosome 4 or the long arm of chromosome 5. In contrast, most of the TRB1 protein shows a preferential euchromatic localization, especially with sequences in the Upstream regions, mostly 50 UTR regions. Association of TRB1 protein with entire gene bodies or intergenic regions is markedly lower. A similar distribution of the TRB1 protein preferring DNA Upstream from TSSs was observed in both Protein coding and Nonprotein coding genes. The partial increase in TRB1 association with the genomic loci behind the transcription stop, Downstream, may be due to the fact that neighbouring downstream genes transcribed in the same direction are frequent at a distance of 100–700 bp (Alexandrov et al. 2006), and thus an overlap of the genomic regions located Upstream from TSSs or Downstream from the transcriptional stop may result in unspecific enrichment in one of these datasets. However, association of TRB1 with the upstream regions is markedly higher. In accordance with the localization of TRB1 in euchromatin, we found that the overlap between the list of the chromosome 4 genes associated with marks of silent euchromatin (H3K9me3) and Polycomb-regulated chromatin (H3K27me3) (Turck et al. 2007) and the genes with a TRB1 genomic peak in their 500 bp upstream vicinity (Upstream 500 bp of Protein and Non-protein coding genes datasets) is considerably higher (20 and 9 %, respectively) than the overlap between the genes associated with the heterochromatin mark H3K9me2 and the genes with a TRB1 genomic peak in proximity to their 500 bp upstream region (0.3 %) (Turck et al. 2007). TRB1 target genes are connected with ribosome biogenesis The distribution of short interspersed telomeric repeats or telo-boxes within the Arabidopsis genome is not uniform and their frequency is higher within 50 flanking regions (Gaspin et al. 2010). In this study we show that telo-boxes Plant Mol Biol (2016) 90:189–206 201 123 located Upstream from TSSs are more likely to be associated with the TRB1 protein than those located Downstream from the transcription stop region or even in entire Protein coding or Non-protein coding gene bodies or within the Intergenic range. The occurrence of telo-boxes is enriched in promoter regions of genes participating in translation, e.g. genes coding for ribosomal proteins, translation elongation factors (EF1 a), eukaryotic initiation factors (eIFs), and small nucleolar RNAs (snoRNAs), or in the promoters of genes involved in rRNA processing (Gaspin et al. 2010; Liboz et al. 1990; Tremousaygue et al. 1999). The majority of telo-boxes of plant translation-related genes are located within a narrow window located between -50 and ?50 relative to the TSS (Gaspin et al. 2010). Telo-boxes are not able, by themselves, to activate gene expression in transgenic plants but act in synergy with other cis-acting elements like site II motifs or TEF boxes that show conservative topological association with teloboxes in Arabidopsis and O. sativa (Gaspin et al. 2010; Manevski et al. 2000; Tremousaygue et al. 2003). In agreement with this observation, we found that TRB1 shows a statistically significant enrichment in 500 bp Upstream regions of the genes connected in GO with terms describing small and large subunit of the ribosome, nucleolus, or rRNA modification. Since many snoRNA genes that are involved in the processing of rRNA are transcribed in clusters or are located in introns (Brown et al. 2008), the total number of snoRNA genes with TRB1-binding sites in their promoter is underestimated because the distal snoRNA gene in a transcribed cluster may be farther than 500 bp away from the transcription start site and, consequently, is not included in our analysis. Another similar case is intronic snoRNA genes, which are indistinguishable from the protein coding genes. For details, see supplementary Figure S4 (snoRNA clusters) which displays the IGV view of snoRNA clusters significantly associated with TRB1 protein in their shared promoter. We proved that only the snoRNAs category is associated with the TRB1 protein in its 500 bp Upstream proximity (Fig. 9a). Other categories from the Non-protein coding dataset e.g., rRNA, snRNA, tRNA, ncRNA, miRNA, other RNA, pseudogenes and transposable element do not exhibit increased association with TRB1. We found that more than 80 % of intergenic orphan snoRNA genes (not transcribed in clusters) or intergenic snoRNAs with a telo-box in proximity to their their 500 bp Upstream are covered by TRB1 genomic peaks (Fig. 9b, c). Our analysis of 192 ribosomal protein coding genes, used for a transcription level study (Savada and BonhamSmith 2014), revealed that almost 85 % of them contain a telo-box in proximity to their 500 bp Upstream translation start and 65 % of these telo-box sequences are covered by TRB1 genomic peaks. At least nine of ten promoters of the most frequently transcribed ribosomal genes in A. thaliana seedlings (Savada and Bonham-Smith 2014) are recognised by the TRB1 protein. By contrast, promoters of the ribosomal protein-coding genes (which are transcribed at the lowest level in A. thaliana seedlings (Savada and BonhamSmith 2014)) showed either lower or a zero level of TRB1 binding. Analogous estimation of eIFs with telo-boxes which are associated with the TRB1 protein was derived from a list of eIFs (https://www.arabidopsis.org/browse/genefamily/eIF. jsp). Almost 60 % of eIFs contain a telo-box sequence in proximity to their 500 bp Upstream translation start and 50 % of these telo-box sequences are covered by TRB1 genomic peaks. The pattern of TRB1 association with chromatin suggests its role as a transcription factor In humans, some core shelterin subunits are documented to modulate gene expression outside of telomeres; for example, TRF2 interacts with the repressor element 1-silencing transcription factor (REST), a repressor of genes devoted to neuronal functions (Zhang et al. 2008a). Another example is mammalian RAP1 which is known to play a role in repression of subtelomeric genes, and more than 70 % of its binding sites are found at intragenic positions or in the vicinity of gene-coding chromatin (Martinez et al. 2010). Strong nucleolar localization of the TRB1 protein, besides weaker nuclear and cytoplasmic localization, has been already shown (Dvorackova et al. 2010). Genes coding for 45S ribosomal RNA (rDNA) which organize nucleoli are close neighbours of telomeres in chromosomes of many eukaryotes including Arabidopsis, and also show a number of functional associations (Dvorackova et al. 2015; Fransz et al. 2002). A proteomic analysis has revealed that a significant portion of the nucleolar protein pool consists of ribosomal proteins (RPSs or RPLs), RNA modifying factors (snRNA or snoRNA binding), and proteins participating in translation (EF1s, eIFs) (Brown et al. 2005; Pendle et al. 2005). The majority of non-ribosomal nucleolar proteins occur in the nucleolus only transiently, since many of these factors fluctuate between the nucleus and nucleolus (Dundr et al. 1997; Snaar et al. 2000; Sprague and McNally 2005), and similar behaviour was observed for the TRB1 protein (Dvorackova et al. 2010) that is largely dispersed at prophase, coinciding with nucleolar disassembly, and re-localized in early anaphase after cytokinesis. TRB1 association was detected in the promoters of the H2AX A and H2AX B genes, which also contain telo-box sequences in their promoter region. Phosphorylated products of these genes are involved in response to DNA double 202 Plant Mol Biol (2016) 90:189–206 123 strand breaks. All these connections indicate the importance to examine possible involvement of the TRB1 protein in regulation of transcription. Whole-genome duplication events in Arabidopsis phylogeny often resulted in increased number of genes of the same family (Mandakova and Lysak 2008; Nelson et al. 2014). TRB1 protein belongs to the single-myb-histone family with five members (Marian et al. 2003; Schrumpfova et al. 2004, 2014). Although these proteins slightly differ in binding properties (Hofr et al. 2009; Schrumpfova et al. 2004, 2014) their partial functional redundancy in vivo cannot be excluded. Therefore, no observation of significant Fig. 9 Analysis of individual categories of the Upstream 500 bp of Non-protein coding genes dataset. The group of Non-protein coding genes includes rRNA, snoRNA, snRNA, tRNA, ncRNA, miRNA, other RNA, pseudogenes and transposable element genes categories. a The pie chart shows the proportional representation of individual categories. The most represented categories transposable elements, pseudogenes, pre-tRNA, other RNA cover nearly 64, 15, 10, and 6.5 %, respectively of the Non-coding genes dataset. By contrast, categories miRNA, snoRNA, snRNA, rRNA occupy only 2.9, 1.2, 0.2, and 0.1 %, respectively. b Dark grey represents the proportion of the genes in each category where any TRB1 genomic peak interferes with the genomic region located in region 500 bp upstream from the transcription start site of the relevant non-protein coding gene. c The dark grey part represents the proportion of the genes in each category that harbour at least one telo-box in their 500 bp upstream region. The hatched part represents the subset of the telobox-containing genes that are covered by TRB1 genomic peaks Plant Mol Biol (2016) 90:189–206 203 123 changes in transcript levels of chosen genes in trb1-/mutant plants (Fig. S5) is not surprising and points to the need for using multiple trb mutants in further analyses. In conclusion, this report demonstrates that the TRB1 protein from A. thaliana, whose N-terminal myb-domain shows high sequence homology with the human TRF1 or TRF2 C-terminal myb-domain, is associated in vivo with a subset of interstitial sites in the Arabidopsis genome besides its major location at telomeres (Schrumpfova et al. 2014). Our finding of an association of TRB1 with the 50 flanking region of protein or non-protein coding regions, especially with sequences located upstream of the transcription start site of the ribosomal protein coding genes, snoRNA genes, and genes coding for elongation factors (eEF-1) and eukaryotic initiation factors (eIFs), correlated with a higher frequency of short telomere-like sequences (telo-boxes) in their promoters, together with its cell-cycle regulated localization (Dvorackova et al. 2010), suggest that the TRB1 protein may be functional also as a transcription factor (see concluding overview in Fig. 10). In this feature, TRB1 resembles other analogous shelterin proteins from diverse organisms which, besides their telomeric localization and functions, can bind to nontelomeric regions and have extra-telomeric functions in gene regulation networks. Acknowledgments We thank Vladimir Benes from the European Molecular Biology Laboratory (EMBL), Genomics Core Facility, Heidelberg, Germany, for valuable comments and recommendations on our ChIP procedure. We also thank Ronald Hancock, Hoˆtel-Dieu de Que´bec for critical reading of the manuscript. Access to computing and storage facilities owned by parties and projects contributing to the National Grid Infrastructure MetaCentrum, provided under the programme ‘‘Projects of Large Infrastructure for Research, Development, and Innovations’’ (LM2010005), is greatly appreciated. This research was supported by the Czech Science Foundation (13-06943S) and by project CEITEC (CZ.1.05/1.1.00/02.0068) of the European Regional Development Fund. Fig. 10 Concluding overview of telomeric and non-telomeric locations and roles of TRB1 protein. TRB1 protein from A. thaliana, whose myb-domain shows high sequence similarity to the human TRF1 or TRF2 myb-domains, acts as a component of plant telomereprotection complex. TRB1 is co-localised with telomeric tracts in vivo, interacts with POT1b protein, telomerase reverse transcriptase (TERT) subunit and the loss of TRB1 protein leads to evident telomere shortening (Kuchar and Fajkus 2004; Schrumpfova et al. 2008, 2014). Furthermore, our present finding demonstrated association of TRB1 also to the 50 flanking region of protein- or non-protein coding regions related to the translation machinery genes e.g.: ribosomal protein coding genes, snoRNA genes, and genes coding for eEF-1 and eIFs, correlated with a higher frequency of short telomere-like sequences (telo-boxes) in their promoters. Physical association of telomeres to nucleoli (depicted in the upper part) may hypothetically correspond to functional relevance of TRB1 for coordinated regulation of translation machinery genes 204 Plant Mol Biol (2016) 90:189–206 123 References Alexandrov NN, Troukhan ME, Brover VV, Tatarinova T, Flavell RB, Feldmann KA (2006) Features of Arabidopsis genes and genome discovered using full-length cDNAs. Plant Mol Biol 60:69–85. doi:10.1007/s11103-005-2564-9 Bosco N, de Lange T (2012) A TRF1-controlled common fragile site containing interstitial telomeric sequences. Chromosoma 121:465–474. doi:10.1007/s00412-012-0377-6 Bowler C, Benvenuto G, Laflamme P, Molino D, Probst AV, Tariq M, Paszkowski J (2004) Chromatin techniques for plant cells. Plant J 39:776–789. doi:10.1111/j.1365-313X.2004.02169.x Bradshaw PS, Stavropoulos DJ, Meyn MS (2005) Human telomeric protein TRF2 associates with genomic double-strand breaks as an early response to DNA damage. Nat Genet 37:193–197. doi:10.1038/ng1506 Brown JW, Shaw PJ, Shaw P, Marshall DF (2005) Arabidopsis nucleolar protein database (AtNoPDB). Nucleic Acids Res 33:D633–D636 Brown JW, Marshall DF, Echeverria M (2008) Intronic noncoding RNAs and splicing. Trends Plant Sci 13:335–342. doi:10.1016/j. tplants.2008.04.010 Chen LY et al (2012) Mitochondrial localization of telomeric protein TIN2 links telomere regulation to metabolic control. Mol Cell 47:839–850. doi:10.1016/j.molcel.2012.07.002 Cifuentes-Rojas C, Kannan K, Tseng L, Shippen DE (2011) Two RNA subunits and POT1a are components of Arabidopsis telomerase. Proc Natl Acad Sci USA 108:73–78. doi:10.1073/ pnas.1013021107 Court R, Chapman L, Fairall L, Rhodes D (2005) How the human telomeric proteins TRF1 and TRF2 recognize telomeric DNA: a view from high-resolution crystal structures. EMBO Rep 6:39–45. doi:10.1038/sj.embor.7400314 de Lange T (2005) Shelterin: the protein complex that shapes and safeguards human telomeres. Genes Dev 19:2100–2110. doi:10. 1101/Gad.1346005 Deng Z, Lezina L, Chen CJ, Shtivelband S, So W, Lieberman PM (2002) Telomeric proteins regulate episomal maintenance of Epstein-Barr virus origin of plasmid replication. Mol Cell 9:493–503 Dundr M, Meier UT, Lewis N, Rekosh D, Hammarskjold ML, Olson MO (1997) A class of nonribosomal nucleolar components is located in chromosome periphery and in nucleolus-derived foci during anaphase and telophase. Chromosoma 105:407–417 Dvorackova M, Rossignol P, Shaw PJ, Koroleva OA, Doonan JH, Fajkus J (2010) AtTRB1, a telomeric DNA-binding protein from Arabidopsis, is concentrated in the nucleolus and shows highly dynamic association with chromatin. Plant J 61:637–649. doi:10. 1111/j.1365-313X.2009.04094.x Dvorackova M, Fojtova M, Fajkus J (2015) Chromatin dynamics of plant telomeres and ribosomal genes. Plant J. doi:10.1111/tpj. 12822 Edgar RC (2004) MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32:1792–1797. doi:10.1093/nar/gkh340 Edgar R, Domrachev M, Lash AE (2002) Gene expression omnibus: NCBI gene expression and hybridization array data repository. Nucleic Acids Res 30:207–210. doi:10.1093/Nar/ 30.1.207 Fajkus J, Sykorova E, Leitch AR (2005) Telomeres in evolution and evolution of telomeres. Chromosome Res 13:469–479. doi:10. 1007/s10577-005-0997-2 Fransz P, De Jong JH, Lysak M, Castiglione MR, Schubert I (2002) Interphase chromosomes in Arabidopsis are organized as well defined chromocenters from which euchromatin loops emanate. Proc Natl Acad Sci USA 99:14584–14589. doi:10.1073/pnas. 212325299 Gaspin C, Rami JF, Lescure B (2010) Distribution of short interstitial telomere motifs in two plant genomes: putative origin and function. BMC Plant Biol 10:283. doi:10.1186/1471-2229-10- 283 Hanaoka S, Nagadoi A, Nishimura Y (2005) Comparison between TRF2 and TRF1 of their telomeric DNA-bound structures and DNA-binding activities. Protein Sci 14:119–130. doi:10.1110/ps. 04983705 Hofr C, Sultesova P, Zimmermann M, Mozgova I, Schrumpfova PP, Wimmerova M, Fajkus J (2009) Single-Myb-histone proteins from Arabidopsis thaliana: a quantitative study of telomerebinding specificity and kinetics. Biochem J 419:221–228. doi:10. 1042/BJ20082195 Kazda A, Zellinger B, Rossler M, Derboven E, Kusenda B, Riha K (2012) Chromosome end protection by blunt-ended telomeres. Genes Dev 26:1703–1713. doi:10.1101/gad.194944.112 Krutilina RI et al (2003) Protection of internal (TTAGGG)n repeats in Chinese hamster cells by telomeric protein TRF1. Oncogene 22:6690–6698. doi:10.1038/sj.onc.1206745 Kuchar M, Fajkus J (2004) Interactions of putative telomere-binding proteins in Arabidopsis thaliana: identification of functional TRF2 homolog in plants. FEBS Lett 578:311–315. doi:10.1016/ j.febslet.2004.11.021 Langmead B, Salzberg SL (2012) Fast gapped-read alignment with Bowtie 2. Nat Methods 9:357–359. doi:10.1038/nmeth.1923 Latrick CM, Cech TR (2010) POT1-TPP1 enhances telomerase processivity by slowing primer dissociation and aiding translocation. EMBO J 29:924–933. doi:10.1038/emboj.2009.409 Liang K, Keles S (2012) Normalization of ChIP-seq data with control. BMC Bioinformatics 13:199. doi:10.1186/1471-2105-13-199 Liboz T, Bardet C, Le Van Thai A, Axelos M, Lescure B (1990) The four members of the gene family encoding the Arabidopsis thaliana translation elongation factor EF-1 alpha are actively transcribed. Plant Mol Biol 14:107–110 Louis EJ, Vershinin AV (2005) Chromosome ends: different sequences may provide conserved functions. BioEssays 27:685–697. doi:10.1002/Bies.20259 Macas J, Neumann P, Novak P, Jiang J (2010) Global sequence characterization of rice centromeric satellite based on oligomer frequency analysis in large-scale sequencing data. Bioinformatics 26:2101–2108. doi:10.1093/bioinformatics/btq343 Majerska J, Sykorova E, Fajkus J (2011) Non-telomeric activities of telomerase. Mol BioSyst 7:1013–1023. doi:10.1039/ c0mb00268b Mandakova T, Lysak MA (2008) Chromosomal phylogeny and karyotype evolution in x = 7 crucifer species (Brassicaceae). Plant Cell 20:2559–2570. doi:10.1105/tpc.108.062166 Manevski A, Bertoni G, Bardet C, Tremousaygue D, Lescure B (2000) In synergy with various cis-acting elements, plant insterstitial telomere motifs regulate gene expression in Arabidopsis root meristems. FEBS Lett 483:43–46 Marian CO et al (2003) The maize Single myb histone 1 gene, Smh1, belongs to a novel gene family and encodes a protein that binds telomere DNA repeats in vitro. Plant Physiol 133:1336–1350. doi:10.1104/pp.103.026856 Martinez P et al (2010) Mammalian Rap1 controls telomere function and gene expression through binding to telomeric and extratelomeric sites. Nat Cell Biol 12:768–780. doi:10.1038/ ncb2081 Mignon-Ravix C, Depetris D, Delobel B, Croquette MF, Mattei MG (2002) A human interstitial telomere associates in vivo with specific TRF2 and TIN2 proteins. Eur J Hum Genet 10:107–112. doi:10.1038/sj.ejhg.5200775 Plant Mol Biol (2016) 90:189–206 205 123 Morse RH (2000) RAP, RAP, open up! New wrinkles for RAP1 in yeast. Trends Genet 16:51–53. doi:10.1016/S0168- 9525(99)01936-8 Mozgova I, Schrumpfova PP, Hofr C, Fajkus J (2008) Functional characterization of domains in AtTRB1, a putative telomerebinding protein in Arabidopsis thaliana. Phytochemistry 69:1814–1819. doi:10.1016/j.phytochem.2008.04.001 Nelson AD, Forsythe ES, Gan X, Tsiantis M, Beilstein MA (2014) Extending the model of Arabidopsis telomere length and composition across Brassicaceae. Chromosome Res 22:153–166. doi:10.1007/s10577-014-9423-y Nishikawa T, Okamura H, Nagadoi A, Konig P, Rhodes D, Nishimura Y (2001) Solution structure of a telomeric DNA complex of human TRF1. Structure 9:1237–1251 Nosek J, Tomaska L (2003) Mitochondrial genome diversity: evolution of the molecular architecture and replication strategy. Curr Genet 44:73–84. doi:10.1007/s00294-003-0426-z Pendle AF et al (2005) Proteomic analysis of the Arabidopsis nucleolus suggests novel nucleolar functions. Mol Biol Cell 16:260–269 Peska V, Schrumpfova PP, Fajkus J (2011) Using the telobox to search for plant telomere binding proteins. Curr Protein Pept Sci 12:75–83 Pfaffl MW (2004) Quantification strategies in real-time PCR. International University Line (IUL), La Jolla Rossignol P, Collier S, Bush M, Shaw P, Doonan JH (2007) Arabidopsis POT1A interacts with TERT-V(I8), an N-terminal splicing variant of telomerase. J Cell Sci 120:3678–3687 Roudier F et al (2011) Integrative epigenomic mapping defines four main chromatin states in Arabidopsis. EMBO J 30:1928–1938. doi:10.1038/emboj.2011.103 Savada RP, Bonham-Smith PC (2014) Differential transcript accumulation and subcellular localization of Arabidopsis ribosomal proteins. Plant Sci 223:134–145. doi:10.1016/j.plantsci.2014.03. 011 Schneider TD, Stephens RM (1990) Sequence logos: a new way to display consensus sequences. Nucleic Acids Res 18:6097–6100 Schrumpfova P, Kuchar M, Mikova G, Skrisovska L, Kubicarova T, Fajkus J (2004) Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. Genome 47:316–324. doi:10.1139/g03-136 Schrumpfova PP, Kuchar M, Palecek J, Fajkus J (2008) Mapping of interaction domains of putative telomere-binding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. FEBS Lett 582:1400–1406. doi:10.1016/j.febslet.2008.03.034 Schrumpfova PP, Vychodilova I, Dvorackova M, Majerska J, Dokladal L, Schorova S, Fajkus J (2014) Telomere repeat binding proteins are functional components of Arabidopsis telomeres and interact with telomerase. Plant J 77:770–781. doi:10.1111/Tpj.12428 Sequeira-Mendes J et al (2014) The functional topography of the Arabidopsis genome is organized in a reduced number of linear motifs of chromatin states. Plant Cell 26:2351–2366. doi:10. 1105/tpc.114.124578 Sfeir A, de Lange T (2012) Removal of shelterin reveals the telomere end-protection problem. Science 336:593–597. doi:10.1126/ science.1218498 Sfeir A, Kosiyatrakul ST, Hockemeyer D, MacRae SL, Karlseder J, Schildkraut CL, de Lange T (2009) Mammalian telomeres resemble fragile sites and require TRF1 for efficient replication. Cell 138:90–103. doi:10.1016/j.cell.2009.06.021 Shen L, Shao N, Liu X, Nestler E (2014) ngs.plot: quick mining and visualization of next-generation sequencing data by integrating genomic databases. BMC Genom 15:284. doi:10.1186/1471- 2164-15-284 Simonet T et al (2011) The human TTAGGG repeat factors 1 and 2 bind to a subset of interstitial telomeric sequences and satellite repeats. Cell Res 21:1028–1038. doi:10.1038/cr.2011.40 Snaar S, Wiesmeijer K, Jochemsen AG, Tanke HJ, Dirks RW (2000) Mutational analysis of fibrillarin and its mobility in living human cells. J Cell Biol 151:653–662 Soudet J, Jolivet P, Teixeira MT (2014) Elucidation of the DNA endreplication problem in Saccharomyces cerevisiae. Mol Cell 53:954–964. doi:10.1016/j.molcel.2014.02.030 Sprague BL, McNally JG (2005) FRAP analysis of binding: proper and fitting. Trends Cell Biol 15:84–91 Teo H et al (2010) Telomere-independent Rap1 is an IKK adaptor and regulates NF-kappaB-dependent gene expression. Nat Cell Biol 12:758–767. doi:10.1038/ncb2080 Tremousaygue D, Manevski A, Bardet C, Lescure N, Lescure B (1999) Plant interstitial telomere motifs participate in the control of gene expression in root meristems. Plant J 20:553–561 Tremousaygue D, Garnier L, Bardet C, Dabos P, Herve C, Lescure B (2003) Internal telomeric repeats and ‘TCP domain’ proteinbinding sites co-operate to regulate gene expression in Arabidopsis thaliana cycling cells. Plant J 33:957–966 Turck F et al (2007) Arabidopsis TFL2/LHP1 specifically associates with genes marked by trimethylation of histone H3 lysine 27. PLoS Genet 3:e86. doi:10.1371/journal.pgen.0030086 Uchida W, Matsunaga S, Sugiyama R, Kawano S (2002) Interstitial telomere-like repeats in the Arabidopsis thaliana genome. Genes Genet Syst 77:63–67 Valach M et al (2011) Evolution of linear chromosomes and multipartite genomes in yeast mitochondria. Nucleic Acids Res 39:4202–4219. doi:10.1093/nar/gkq1345 Weaver DT (1998) Telomeres: moonlighting by DNA repair proteins. Curr Biol 8:R492–R494 Zeeberg BR et al (2003) GoMiner: a resource for biological interpretation of genomic and proteomic data. Genome Biol 4:R28 Zeeberg BR et al (2005) High-throughput GoMiner, an ‘industrialstrength’ integrative gene ontology tool for interpretation of multiple-microarray experiments, with application to studies of Common Variable Immune Deficiency (CVID). BMC Bioinform 6:168. doi:10.1186/1471-2105-6-168 Zhang P, Pazin MJ, Schwartz CM, Becker KG, Wersto RP, Dilley CM, Mattson MP (2008a) Nontelomeric TRF2-REST interaction modulates neuronal gene silencing and fate of tumor and stem cells. Curr Biol 18:1489–1494. doi:10.1016/j.cub.2008.08.048 Zhang Y et al (2008b) Model-based analysis of ChIP-seq (MACS). Genome Biol 9:R137. doi:10.1186/gb-2008-9-9-r137 Zhang Y, Lin YH, Johnson TD, Rozek LS, Sartor MA (2014) PePr: a peak-calling prioritization pipeline to identify consistent or differential peaks from replicated ChIP-seq data. Bioinformatics 30:2568–2575. doi:10.1093/bioinformatics/btu372 206 Plant Mol Biol (2016) 90:189–206 123 Plant Molecular Biology Telomere Binding Protein TRB1 is Associated with Promoters of Translation Machinery Genes in vivo Petra Procházková Schrumpfová1,2 , Ivona Vychodilová1,2 , Jan Hapala1,2 , Šárka Schořová1,2 , Vojtěch Dvořáček and Jiří Fajkus1,2,3§ 1 Mendel Centre for Plant Genomics and Proteomics, CEITEC – Central European Institute of Technology, Masaryk University, Kamenice 5, CZ-625 00 Brno, Czech Republic; 2 Laboratory of Functional Genomics and Proteomics, National Centre for Biomolecular Research, Faculty of Science, Masaryk University, Kamenice 5, CZ-625 00 Brno, Czech Republic; 3 Institute of Biophysics, Academy of Sciences of the Czech Republic, v.v.i., Královopolská 135, CZ-61265 Brno, Czech Republic Coresponding author: e-mail: fajkus@sci.muni.cz tel: +420 549 49 4003 fax: +420 549 492 654 Supplementary Figures and Tables Figure S1- Overview of all TRB1 genomic peaks in each chromosome The IGV view of all five chromosomes from A. thaliana visualizes the preference of the TRB1 protein binding outside of heterochromatic regions (centromeres and heterochromatin knobs). The All coding genes category represents both Protein coding genes and Non-protein coding genes. TRB1 genomic peaks represents TRB1enriched regions. The heterochromatin knobs on chromosome 4 and 5 are indicated by grey boxes. Figure S2 - TRB1 association with the EF1-alpha genes Three EF1-alpha 1, 2 and 3 genes that contain telo-box sequence in Upstream 500 bp distance (At1g07940, At1g07930 and At1g07920) are visualised with IGV viewer. Telo-box sequences are highlighted in separate rows. Although the pattern of TRB1 occupancy would visually suggest that TRB1 is occupying regions around TSSs of all three loci, only one (EF1-alpha 1 gene (At1G07940)) has been detected as a TRB1 genomic peak under our stringent threshold criteria. So the number of the TRB1 associated regions should be actually higher. Figure S3 – TRB1 association with genes coding for ribosomal proteins Three genes coding for ribosomal proteins RPS15B, RPS15C, RPS15D (At5g09490, At5g09500, At5g09510) are visualised with IGV viewer. Telo-box sequences are highlighted in separate rows. Although the pattern of TRB1 occupancy would visually suggest that TRB1 is occupying regions around TSSs of all three loci, only one (RPS15D gene (At5g09510)) has been detected as TRB1 genomic peak under our stringent threshold criteria. So the number of the TRB1 associated regions should be actually higher. Figure S4 – Schematic view of the snoRNA clusters An IGV view of three examples of association of the TRB1 protein with promoter sequences of the snoRNA genes that are clustered into one transcript (blue boxes). TRB1 genomic peaks are highlighted. Figure S5 – Effect of TRB1 on transcription of selected TRB1-associated genes Several genes with promoters visibly associated with TRB1 protein, coding for proteins of large or small ribosome subunits, and genes coding for eEFs and H2AX proteins were chosen for analysis of transcript levels in trb1-/- plants and compared to the respective wt control. For comparison, genes coding for Pheres (PHE) and RuvB-like AAA ATPase 1 (RuvBL1), lacking association with TRB1 protein (light grey), were also analysed. Y-axis values represent relative transcript levels (Wt=1). Table S1 - List of primers used in the qChIP-qPCR and RT-PCR analysis. List of primers used for ChIP-qPCR described in (a) Figure 4 and (b) Figure S5 A Actin 7 (AT5G09810) Forward GGAAACATCGTTCTCAGTGGT Reverse CTTGATCTTCATGCTGCTAGGT Ribosomal protein L34 (AT3G28900) Forward AGGTGATCTCTTGAGCCCTAG Reverse AAAGATAAAAAACACTACAATCAATCTGAG Ribosomal protein S5 (AT2G41840) Forward AGGCCTTGTTGGGTTTGT Reverse TTGTTCTGATTAACGTGTGACATTAG snoRNA (AT4G15258 ) Forward CGAAACCTTATAAATACACAGACACAG Reverse TTGGGCCGAGAACCTAAAATAG B RPL37aC (at3g60245) Forward AGGTTGGAATCGTCGGCAAA Reverse TCACTCCGTACTTGCCACAG RPL10C (at1g66580) Forward ACCGTGCTGAGTACACGAAG Reverse TCATTCTTATTCGCTAGTGGCTGA RPL23C (at3g04400) Forward GCCACTGTGAACTGTGCTGA Reverse CAACACACGCTGATGGCAAA RPL17B (at1g67430) Forward GTACTCGCAAGAACCCGACA Reverse TGGTAGCTTCCTGATTGCGT RPL18aB (at2g34480) Forward GGACAGATGCTCGCCATCAA Reverse CATCTGCTCAACAGCTCCGT eEF1-alpha 2 (at1g07930 ) Forward GTCTGTTGAGATGCACCACG Reverse CCCTCTCTTAAGATCCTTCACGG eEF1-alpha 3 (at1g07920) Forward GCTGCTAACTTCACCTCCCA Reverse TCTCCTTACCAGAACGCCTG eEF1a-4 (at5g60390) Forward ACAAGCGTGTCATCGAGAGG Reverse TCACGCTCGGCCTTAAGTTT γH2AX A (at1g08880) Forward ATGAGTACAGGCGCAGGAAG Reverse CTAGCGATTCTTCCGACGGG γH2AX B (at1g54690) Forward ACAACTAAAGGTGGCAGAGGA Reverse TCGGCGTATTTACCGGCTTTA eEF1-alpha 3 (at1g07920) Forward GCTGCTAACTTCACCTCCCA Reverse TCTCCTTACCAGAACGCCTG eEF1a-4 (at5g60390) Forward ACAAGCGTGTCATCGAGAGG Reverse TCACGCTCGGCCTTAAGTTT γH2AX A (at1g08880) Forward ATGAGTACAGGCGCAGGAAG Reverse CTAGCGATTCTTCCGACGGG γH2AX B (at1g54690) Forward ACAACTAAAGGTGGCAGAGGA Reverse TCGGCGTATTTACCGGCTTTA PHE (At1g65330) Forward GATCGCCAAAGAAACAGAACG Reverse ATCTCAACCCTACGAATAACACC RuvBL1 (At5g22330) Forward CGGATTGCTACTCACACCCA Reverse GCTGCCTCTCTAGCCTCAAG UBQ10 (At4g05320) Forward AACGGGAAAGACGATTAC Reverse ACAAGATGAAGGGTGGAC Table S2 – Computation of the TRB1 protein coverage dataset per size unit of 1 Mb The size of each dataset analysed in this study (bp) is listed in the column “whole dataset size”. Overlapping base pairs were counted once. The proportional representation of these numbers is shown in Figure 3b as “Genome coverage by whole dataset”. The section “Number of TRB1 genomic peaks in dataset” represents the absolute number of DNA sequences from each dataset that were covered at least partly by one (or more) TRB1 genomic peak. The “TRB1 genomic peaks per size unit of 1 Mb” column is illustrated graphically in the Figure 3c and represents the frequency of TRB1 genomic peaks per one megabase of each dataset. Table S3 - Overall summary of sequences belonging to each dataset (In separate Supplementary .xls file) A table of DNA sequences in each dataset that are significantly covered with the TRB1 protein (TRB1 genomic peaks). Table S4 - List of 8-mers and List of fragments used in the logo construction (In separate Supplementary .xls file) (A) A list of every unique substring of the length k (8) in the dataset 5’UTR covered by TRB1 genomic peaks. The proportion of purple 8-mers is twice as high in the dataset 5’UTR covered by TRB1 genomic peaks as in the whole 5’UTR dataset. Only these purple 8-mers were used for fragment construction of the total sequence logo (Fig. 5a). The green boxes highlight 8-mers with a permutation of at least one plant telomeric repeat and used for partial sequence logo construction with telomeric repeat (see Fig. 5b). The rest of the purple 8-mers do not contain any telomeric repeat and were used for partial sequence logo construction without telomeric repeat (Fig. 5c). (B) Here are listed fragments constructed from 8-mers separately for each sequence logo (total, with telomeric repeat, without telomeric repeat (Fig. 5)). Fragments in their forward or reverse complementary version (orange) were aligned by MUSCLE. The mean weight of each fragment or the relative weight for each base in the fragment is included. Table S5 - Table of Telo-box sequences in each dataset The number of all sequences in the datasets which contain at least one telo-box are listed in the column “All telobox containing sequences”. The column “Telo-box covered by TRB1 genomic peaks” lists the number of sequences where at least one telo-box is preferentially recognized by the TRB1 protein and thus termed a TRB1 genomic peak. The last column corresponds to the graphical illustration in Figure 6 where the percentages of all sequences from each dataset covered by TRB1 genomic peak are shown. Table S6 - List of telo-box sequences belonging to each dataset (In separate Supplementary .xls file) Genomic coordinates of all telo-boxes and telo-boxes with TRB1 genomic peaks. Table S7- Detailed reports from GOMiner and CIMminer (In separate Supplementary .xls file) Two main GO categories, GO:0008150 Biological process and GO:0005575 Cellular component, were individually analyzed in GOMiner and CIMminer. The list “CIM” contains GO subcategories of these GO categories which are statistically enriched (p ≤ 0.05). For this table, only GO subcategories (listed in horizontal rows) containing 5−500 members were selected. Separate genes are listed vertically (AGI gene codes), so this table corresponds to Figure 7 in the main text. The list “CIM with genes” contains highly enriched GO subcategories (p ≤ 0.05), ordered vertically, and selected genes where AGI gene codes are presented with detailed description of the gene names and function in rows. The list “Total vs. Total Report” contains a comparison of All coding genes from A. thaliana (Total file; Protein coding genes together with Non-protein coding genes) together with the set of genes where the TRB1 protein is enriched in the region 500 bp upstream from the transcription start site ("Changed" file). All coding genes involved in each GO subcategory are listed and marked whether they were enriched in the region 500 bp upstream from the transcription start site or not ("no change/changed"). GO subcategories are ordered according to the number of members. The list “Gene Category Report” contains a detailed analysis of individual genes involved in each GO subcategory. GO subcategories are ordered according to the p-value so the GO subcategories where the amount of genes associated with TRB1 protein in their Upstream 500 bp proximity is highest are listed in the upper part of the Table. The list “Category Summary Report” summarizes the results for all GO subcategories (Gene Category Report). GO subcategories are ordered by p-value. Supplement J Schrumpfová, P.P.*, Schořová, Š., Fajkus, J., 2016. Telomere- and Telomerase-Associated Proteins and Their Functions in the Plant Cell. Front. Plant Sci. 7:851 P.P.S. wrote the ms and was significantly involved the ms editing REVIEW published: 28 June 2016 doi: 10.3389/fpls.2016.00851 Edited by: Anne-Catherine Schmit, Institut de Biologie Moléculaire des Plantes, CNRS UPR2357, France Reviewed by: Takashi Murata, National Institute for Basic Biology, Japan Franziska Katharina Turck, Max Planck Society, Germany *Correspondence: Petra Procházková Schrumpfová schpetra@centrum.cz Specialty section: This article was submitted to Plant Cell Biology, a section of the journal Frontiers in Plant Science Received: 12 November 2015 Accepted: 31 May 2016 Published: 28 June 2016 Citation: Procházková Schrumpfová P, Schoˇrová Š and Fajkus J (2016) Telomereand Telomerase-Associated Proteins and Their Functions in the Plant Cell. Front. Plant Sci. 7:851. doi: 10.3389/fpls.2016.00851 Telomere- and Telomerase-Associated Proteins and Their Functions in the Plant Cell Petra Procházková Schrumpfová1,2 *, Šárka Schoˇrová2 and Jiˇrí Fajkus1,2,3 1 Mendel Centre for Plant Genomics and Proteomics, Central European Institute of Technology, Masaryk University, Brno, Czech Republic, 2 Laboratory of Functional Genomics and Proteomics, National Centre for Biomolecular Research, Faculty of Science, Masaryk University, Brno, Czech Republic, 3 Institute of Biophysics, Academy of Sciences of the Czech Republic, v.v.i., Brno, Czech Republic Telomeres, as physical ends of linear chromosomes, are targets of a number of specific proteins, including primarily telomerase reverse transcriptase. Access of proteins to the telomere may be affected by a number of diverse factors, e.g., protein interaction partners, local DNA or chromatin structures, subcellular localization/trafficking, or simply protein modification. Knowledge of composition of the functional nucleoprotein complex of plant telomeres is only fragmentary. Moreover, the plant telomeric repeat binding proteins that were characterized recently appear to also be involved in nontelomeric processes, e.g., ribosome biogenesis. This interesting finding was not totally unexpected since non-telomeric functions of yeast or animal telomeric proteins, as well as of telomerase subunits, have been reported for almost a decade. Here we summarize known facts about the architecture of plant telomeres and compare them with the well-described composition of telomeres in other organisms. Keywords: telomere, telomerase, telomeric proteins, shelterin, telomeric repeat binding (TRB), plant TELOMERES AS NUCLEOPROTEIN STRUCTURES Telomeres are nucleoprotein structures at the ends of eukaryotic chromosomes that protect linear chromosomes against damage by endogenous nucleases and erroneous recognition as unrepaired chromosomal breaks. It is now known that telomeric structures are formed by telomeric DNA, histone octamers, and a number of proteins that bind telomeric DNA, either directly or indirectly, and together, form the protein telomere cap (Fajkus and Trifonov, 2001; de Lange, 2005; Bianchi and Shore, 2008; Sfeir, 2012). The telomeric cap proteins of diverse organisms are less conserved than one might expect. Even within a single taxonomic class, such as mammals, telomeric proteins display less conservation than other chromosomal proteins (Linger and Price, 2009). On the other hand, in many plant families, whole-genome duplication events have occurred, resulting in a multitude of genomic changes, such as deletions of large fragments of chromosomes, silencing of duplicate genes, and recombining of homologous chromosomal segments, as was shown, e.g., in crucifer species (Mandakova and Lysak, 2008). Polyploidy can result in increased numbers of genes of the same family (Taylor and Raes, 2004; He and Zhang, 2005; Freeling, 2009), which may show sub-functionalization, neo-functionalization, and partial or full redundancy and complicates assignment of an actual and specific function for individual proteins in vivo. Gene duplications and losses in plant phylogeny can be traced also in telomere associated protein families Frontiers in Plant Science | www.frontiersin.org 1 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins (e.g., in Arabidopsis thaliana: single myb histone (SMH) family, TRF-like (TRFL) family, or Pot1-like family) (Nelson et al., 2014; Beilstein et al., 2015). In land plants, the telomere is mostly composed of Arabidopsis-type TTTAGGG repeats (Richards and Ausubel, 1988; Figure 1A). Known exceptions are species in the order Asparagales, starting from divergence of the Iridaceae family, which shares the human-type telomeric repeat (TTAGGG; probably caused by a mutation that altered the RNA template subunit of telomerase ∼80 Mya; Adams et al., 2001; Weiss and Scherthan, 2002; Sykorova et al., 2003). The human-type telomere is also shared by species of the Allioideae subfamily, except for the Allium genus (Sykorova et al., 2006), where novel telomeric sequence (CTCGGTTATGGG) was recently described (Fajkus et al., 2016). An unusual telomeric motif (TTTTTTAGGG) was found in the family Solanaceae, in Cestrum elegans and related species (Peska et al., 2015). Also some of the species from the carnivorous genus Genlisea display, instead normal Arabidopsistype of telomere, two intermingled sequence variants (TTCAGG and TTTCAGG; Tran et al., 2015). Moreover, across the Plantae kingdom, outside of land plants but including red algae, green algae, and Glaucophytes (Koonin, 2010), telomere types also vary (Figure 1B). For example, in algae, in addition to the Arabidopsis-type of telomeric repeat, the Chlamydomonas-type (TTTTAGGG), human-type (TTAGGG), and a novel TTTTAGG repeat have been described (Fulnecková et al., 2013; Fulneˇcková et al., 2015). The length of plant telomeric DNA at a single chromosomal arm can be as small as 500 base pairs (bp) in Physcomitrella patens (Shakirov et al., 2010; Fojtova et al., 2015), as long as 160 kb in Nicotiana tabacum (Fajkus et al., 1995), or 200 kb in Nicotiana sylvestris (Kovarik et al., 1996). Besides the remarkable variation in telomere lengths among diverse plant genera or orders, telomere lengths can also vary at the level of the species or ecotypes: e.g., Arabidopsis telomeres range from 1.5 to 9 kb, depending on the ecotype. Also in the long-living organism Betula pendula, telomeres in different genotypes varied from a minimum length of 5.9–9.6 kb to a maximum length of 15.3– 22.8 kb (Shakirov and Shippen, 2004; Maillet et al., 2006; Aronen and Ryynanen, 2014). Since telomeric DNA serves as a landing pad for a set of proteins, the total length or composition of telomeric tracts could markedly affect the number or selection of telomere-associated proteins and subsequently influence telomere packaging, structural transitions, or launch various biochemical pathways (see below). NUCLEAR LOCALIZATION AND DYNAMICS OF TELOMERES In some species during interphase, telomeres, and centromeres could be located at opposite sides of the nucleus, at the nuclear periphery, in limited regions or clusters; this is known as the Rabl organization (Rabl, 1885; for review, see Cowan et al., 2001). The Rabl organization (Wen et al., 2012) was observed in wheat, rye, barley, and oats. Other plant species, such as maize (Zea mays) and sorghum (Sorghum bicolor), despite having fairly large genomes, are not known to exhibit the Rabl configuration (Dong and Jiang, 1998). A recent study among Brachypodium species revealed a positive correlation between Rabl configuration and an increase in DNA content (resulting from replication) and a negative influence of increasing nuclear elongation (Idziak et al., 2015). A rosette-like organization of chromosomes in interphase nuclei was observed in Arabidopsis: telomeres show persistent clustering at the nucleolus while centromeres do not cluster (Armstrong et al., 2001; Tiang et al., 2012). Moreover, during early meiotic prophase, at the leptotene–zygotene transition, telomeres of most plant species cluster to form a bouquet (Bass et al., 1997; Martinez-Perez et al., 1999; Cowan et al., 2002; Corredor and Naranjo, 2007; Higgins et al., 2012; Phillips et al., 2012). Arabidopsis belongs to a small group of species that do not form telomeric bouquets (Armstrong et al., 2001). Chromatin attachment to the inner nuclear membrane in plants, as well as in other species, is mediated by a well conserved multi-protein complex gathered around SUN (Sad1-UNC-84 homology)-KASH (Klarsicht, ANC-1, and Syne homology) proteins [respectively AtSUN-AtSINE (SUN domain-interacting NE proteins) in A. thaliana; Starr et al., 2001; Zhou et al., 2014; Tamura et al., 2015]. In fission and budding yeasts, interactions during meiosis between telomeres and the nuclear envelope, via interactions between SUN domain proteins and telomere-binding proteins, was described: in Saccharomyces cerevisiae SUN-domain protein yMps3 (monopolar spindle protein 3) is needed for yKu80-mediated telomeric chromatin anchoring (Schober et al., 2009), while in Schizosaccharomyces pombe, interactions between telomeric protein pRap1 (repressor activator protein 1) and pSUN proteins are mediated by pBqt1 and pBqt2 (telomere bouquet protein 1 and 2; Chikashige et al., 2006). The tethering of human telomeres to the nuclear matrix was proposed to depend on an isoform of telomere repeat binding factor 1 (TRF1) interacting partner (hTIN2), named hTIN2L (Kaminker et al., 2009), or an A-type lamin (Ottaviani et al., 2008; for review, see Giraud-Panis et al., 2013). Various homologs of SUN domain proteins were identified in Arabidopsis or in maize. In Arabidopsis, they are also localized to the inner nuclear membrane in somatic cells (Graumann et al., 2010; Tamura et al., 2015), however, homologs of Bqt proteins or TIN2 proteins have not been found in plants and their sequences are poorly conserved. Telomeres are processed by a telomere-specific machinery that includes telomerase and its regulatory units, as well as nucleases, as exemplified by the exonuclease 1 (AtEXO1) ortholog in Arabidopsis (Kazda et al., 2012; Derboven et al., 2014). In plants, as well as in most of other kingdoms, replication of chromosomal ends results in single-stranded 3 DNA protrusions (G-overhangs) after degradation of the last RNA primer at the 5 terminus of a nascent strand. In Silene latifolia or A. thaliana, relatively short (20–30 nucleotides) G-overhangs were detected. Moreover, half of the Silene and Arabidopsis telomeres showed no overhangs or overhangs less than 12 nucleotides in length (Riha et al., 2000; Kazda et al., 2012). These G-overhangs are also thought to be required for chromosome end protection by forming secondary DNA structures such as Frontiers in Plant Science | www.frontiersin.org 2 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins FIGURE 1 | Summary of current knowledge on telomere DNA diversity in land plants (A) and green algae (B). The prevalent plant telomeric sequence motif TTTAGGG was first described in A. thaliana (Richards and Ausubel, 1988). Divergent telomeric sequences have been observed in Asparagales (Sykorova et al., 2003), in Cestrum spp. (Peska et al., 2015), in Genlisea (Tran et al., 2015), or in Allium (Fajkus et al., 2016). While Arabidopsis-type telomeric sequence is dominant in “green lineages” of algae Chlorophyta and Streptophyta, this ancestral motif was replaced several times with a novel motifs (reviewed in Fulne ˇcková et al., 2015). (The relationships between families and genera are adapted from the schematic phylogenetic tree presented in Fulne ˇcková et al., 2015 and Fajkus et al., 2016.) Frontiers in Plant Science | www.frontiersin.org 3 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins t-loops (reviewed in Tomaska et al., 2009). Although formation of t-loop structures was demonstrated among plants only in the garden pea (Pisum sativum; Cesare et al., 2003), it is believed that excision from a t-loop in Arabidopsis may result in t-circle formation and in telomere rapid deletion (Watson and Shippen, 2007). In tobacco cell culture, knockdown of one of three human hnRNP homologs, named NgGTBP1 (G-strand specific single-stranded telomere-binding protein 1), led to frequent formation of extrachromosomal t-circles, inhibition of single-stranded invasion into double-stranded telomeric DNA and the loss of protection of telomeres against inter-telomeric recombination (Lee and Kim, 2010, 2013). As well as in humans, mouse, or Caenorhabditis (Uringa et al., 2011; Vannier et al., 2012), the regulator of telomere elongation helicase 1 (AtRTEL1) plays a putative role in Arabidopsis in the destabilization of DNA loop structures such as t-loops or d-loops (Recker et al., 2014). However, a substantial portion of telomeres in Arabidopsis does not apparently undergo nucleolytic resection, and 3 ends produced by leading-strand replication remain blunt-ended (Riha et al., 2000). It is believed that bluntends in Arabidopsis are specifically recognized and protected by the AtKu70/80 heterodimer although in situ localization of Ku to telomeres remains elusive (Kazda et al., 2012). PROTEINS ASSOCIATED WITH TELOMERIC DNA Telomere-associated proteins can regulate lengths of telomere tracts by modulating access of telomerase or affecting conventional DNA replication machinery. In mammals, telomeric DNA associates with a six-protein complex called shelterin. The specific telomeric dsDNA binding is mediated by TRF1 and TRF2 (Broccoli et al., 1997; Court et al., 2005), through their Myb-like domain with an LKDKWRT amino acid motif that is also conserved in other telobox binding proteins, not only in mammals but also in plants (Bilaud et al., 1996; Feldbrugge et al., 1997). A bridge between proteins directly associated with DNA—TRF1, TRF2, and ssDNA binding protein Pot1 (Protection of telomeres 1)—is mediated by TIN2 and the oligosaccharide/oligonucleotide binding (OB)-fold domain of TPP1 (TINT1, PTOP, PIP1) protein (for review see Schmidt and Cech, 2015; Lazzerini-Denchi and Sfeir, 2016). Moreover, protein Rap1, the last component of shelterin, interacts with TRF2 (Arat and Griffith, 2012) and modulates its recruitment to telomeric DNA (Janouskova et al., 2015). A schematic model of mammalian telomere-associated proteins (Figure 2A) and a proposed model of the telomeric complex in A. thaliana (Figure 2B) summarizes recent knowledge in mammalian and plant telomere biology and provides a clear comparison of conserved structures at chromosome termini. In addition, a general overview of telomere-associated proteins that have been described in plants is given in Table 1. Detailed description of telomeric and putative telomeric dsDNA and ssDNA binding proteins from A. thaliana is shown in Table 2. Telomeric dsDNA-Associated Proteins Myb-like Proteins In plants, telomeric dsDNA sequence binding proteins with a Myb-like domain of a telobox (short telomeric motif) type can be classified into three main groups: (i) with a Myb-like domain at the N-terminus (SMH family), (ii) with a Myb-like domain at the C-terminus (TRFL family), and (iii) with a Myb-like domain at the C-terminus (AID family; reviewed in Peska et al., 2011; Du et al., 2013). The first group of proteins, with a Myb-like domain at the N-terminus, also contains a central histone-like domain with homology to the H1 globular domain found in the linker histones H1/H5, and is therefore called the SMH family (Marian et al., 2003; Schrumpfova et al., 2004). Proteins with an SMH motif are plant-specific but are well conserved throughout the plant kingdom (e.g., eudicots, monocots, moss, or red algae; Du et al., 2013). In A. thaliana, there are five members of the SMH family, named telomere repeat binding (AtTRB) proteins (Marian et al., 2003; Schrumpfova et al., 2004). AtTRB1 protein specifically binds plant telomeric repeats through a Myb-like domain in vitro (Mozgova et al., 2008), co-localizes with telomeres in situ, and physically interacts with AtTERT (Figure 2B). Moreover shortening of telomeres was observed in attrb1 knockout mutants (Schrumpfova et al., 2014). Also other members of this family, AtTRB2 and AtTRB3 (previously named AtTBP3 and AtTBP2, respectively; Schrumpfova et al., 2004), bind telomeric dsDNA as well as telomeric ssDNA in vitro as homo- or heteromultimers (Schrumpfova et al., 2004; Mozgova et al., 2008; Hofr et al., 2009; Lee W.K. et al., 2012; Yun et al., 2014). In Arabidopsis, AtTRB1 protein physically interacts via its histone-like domain with AtPot1b (Schrumpfova et al., 2008), an A. thaliana homolog of the G-overhang binding protein Pot1, and a component of an alternative telomerase holoenzyme complex (Tani and Murata, 2005; He et al., 2006; Surovtseva et al., 2007). Also other members of SMH family proteins in land plants show telomeric dsDNA binding capability: e.g., Oryza sativa OsTRBFs (Byun et al., 2008) or Z. mays ZmSMHs (Marian et al., 2003). In addition, proteins with Myblike domain of a telobox type in plants, adopt distinct nontelomeric functions, e.g., PcMYB1 from Petroselinum crispum acts only as a transcription factor (Feldbrugge et al., 1997). Recently it was shown that AtTRB1 from A. thaliana was not only telomere- and telomerase-binding but was also associated, in vivo, with promoters, mostly with a telo box motif of translation machinery genes (Figure 3; Schrumpfova et al., 2016). The AtTRB1 association with telo box motif was then proven by Zhou et al. (2016). Moreover AtTRB proteins seem to have a new role as chromatin modulators: AtTRB1 competes with LIKE HETEROCHROMATIN PROTEIN 1 (AtLHP1) to maintain downregulation of polycomb group (PcG) target genes (Zhou et al., 2016) and protein AtTRB2 directly interacts with histone deacetylases, HDT4 and HDA6, in vitro and in vivo (Lee and Cho, 2016). Deacetylase activity of HDT4 (Lee and Cho, 2016) and HDA6 (To et al., 2011) against H3K27ac, could be important for subsequent methylations of H3K27me3, that is among others target also for AtLHP1. Frontiers in Plant Science | www.frontiersin.org 4 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins FIGURE 2 | Nucleoprotein complexes associated with mammalian and A. thaliana telomeres. (A) Mammalian shelterin proteins (TRF1/2, Rap1, TIN2, TPP, and Pot1) modulate access to the telomerase complex and the ATR/ATM dependent DNA damage response pathway. The CST complex (CTC1–STN1–TEN1) affects telomerase and DNA polymerase α recruitment to the chromosomal termini, and thus coordinates G-overhang extension by telomerase with fill-in synthesis of the complementary C-strand (blue dashed line; figure adopted from Chen and Lingner, 2013). (B) Arabidopsis TRB1, 2 and 3 interact with the telomeric sequence due to the same Myb-like binding domain as mammalian TRF1/2 (Marian et al., 2003; Kuchar and Fajkus, 2004; Schrumpfova et al., 2004). TRB proteins interact with TERT and Pot1b and when localized at chromosomal ends they are eligible to function as components of the plant shelterin complex, mainly at telomeres with a G-overhang (Schrumpfova et al., 2008, 2014). An evolutionary conserved CST complex is suggested to coordinate the unique requirements for efficient replication of telomeric DNA in plants as well as in other organisms (Derboven et al., 2014). Blunt-ended telomeres are specifically recognized and protected by the KU70/80 heterodimer that directly interacts with TRP1, and by extension, with TERT (Kuchar and Fajkus, 2004; Kazda et al., 2012; Schrumpfova et al., 2014). Taken together, two lines of evidence classify the AtTRB proteins as novel epigenetic regulators that potentially impact transcription status of thousands of genes: (i) association of AtTRB1 with telo box DNA motif (Schrumpfova et al., 2016; Zhou et al., 2016) that is linked with PcG protein pathway (Deng et al., 2013; Wang et al., 2016; Zhou et al., 2016), (ii) involvement of AtTRB proteins in control of H3K27 epigenetic modifications (Lee and Cho, 2016; Zhou et al., 2016), that are also connected with PcG chromatin remodelers. The second group of proteins, with a Myb-like domain at the C-terminus, is also named TRFL. However a TRFL Myblike domain alone is not sufficient for telomere binding and requires a more extended domain—Myb-extension (Myb-ext)— for telomeric dsDNA interactions in vitro (Karamysheva et al., 2004; Ko et al., 2008). Consequently, two families of TRFL can be distinguished: TRFL family 1 with a Myb-ext, whose protein members bind telomeric dsDNA in vitro, and TRFL family 2 without a Myb-ext, whose protein members do not bind telomeric dsDNA specifically in vitro and they are usually not considered as telomeric proteins (Karamysheva et al., 2004). The first identification of a TRFL family protein from O. sativa— telomere-binding protein 1 (OsRTBP1; Yu et al., 2000) was soon followed by numerous other TRFL members: e.g., Nicotiana glutinosa (NgTRF1; Yang et al., 2003), Solanum lycopersicum (LeTBP1; Moriguchi et al., 2006), A. thaliana (AtTBP1, AtTRP1, AtTRFL2-10; Chen et al., 2001; Hwang et al., 2001; Karamysheva et al., 2004), Cestrum parqui (CpTBP; Peska et al., 2011). Even though O. sativa or N. glutinosa mutants for TRFL members exhibited markedly longer telomeres (Yang et al., 2004; Hong et al., 2007), in A. thaliana, a knockout of AtTRP1, member of TRFL family 1 with a Myb-ext, did not change telomere length significantly (Chen et al., 2005). Even multiple knock-out plant, deficient for all six proteins from TRFL family 1 in A. thaliana (AtTBP1, AtTRP1, AtTRFL1, AtTRFL2, AtTRFL4, and AtTRF9) did not exhibit changes in telomere length, or phenotypes Frontiers in Plant Science | www.frontiersin.org 5 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins TABLE1|Ageneraloverviewoftelomere/telomeraseassociatedproteinsdescribedinplants. A.thalianaO.sativaZ.maysN.glutinosa/ N.tabacum/ N.sylvestris H.vulgareS.lycopersicum/ S.tuberosum C.parquiP.crispum TelomericdsDNA-associatedproteins Myb-like proteins Myb-likedomain attheN-terminus (SMHfamily) AtTRB1-3(Schrumpfovaetal.,2004)OsTRBF1-3 (Byunetal.,2008; Heetal.,2013) ZmSMHs (Marianetal.,2003) PcMYB1 (Feldbruggeetal., 1997) Myb-likedomain attheC-terminus (TRFL-family) AtTBP1(Hwangetal.,2001) AtTRP1(Chenetal.,2001) AtTRFL2-10(Karamyshevaetal.,2004) OsRTBP1 (Yuetal.,2000) ZmIBP1 (LugertandWerr,1994) ZmIBP2 (Moore,2009) NgTRF1 (Yangetal.,2003) LeTBP1 (Moriguchietal., 2006) CpTBP1 (Peskaetal., 2011) PcBPF-1 (daCostaeSilva etal.,1993) Myb-likedomain attheC-terminus (AIDfamily) OsAID1 (Zhuetal.,2004; Heetal.,2013) ZmTacs1 (MarianandBass, 2005) CSTcomplexAtStn1(Songetal.,2008) AtTEN1(Leehyetal.,2013) AtCTC1(Surovtsevaetal.,2009) TelomericssDNA-associatedproteins OB-foldPot-likeAtPot1a-c(KucharandFajkus,2004;Tani andMurata,2005;Rossignoletal.,2007) ZmPot1a(Shakirov etal.,2009b) ZmPot1b (Shakirovetal.,2009b) HvPot1(Shakirov etal.,2009b) StPot1 (Shakirovetal., 2009b) CpPot1 (Peskaetal., 2008) Non-OBfoldWHYAtWhy1(Yooetal.,2007)HvWhy1 (Grabowskietal., 2008) RRM-motifAtSTEP1(KwonandChung,2004)Os08g0492100 (Heetal.,2013) Os08g0320100 (Heetal.,2013) NtGTBP1-3 (Hirataetal., 2004;Leeand Kim,2010) Telomeraseassociated TERTsubunitAtTERT(Fajkusetal.,1996)OsTERT (Oguchietal., 2004) ZmTERT(Sykorova etal.,2009) NtTERT,NsTERT (Sykorovaetal., 2012) TERT/TR associated proteins Myb-likedomain attheN-terminus (SMHfamily) AtTRB1-3(Schrumpfovaetal.,2014) Myb-likedomain attheC-terminus (TRFLfamily) AtTRP1(Schrumpfovaetal.,2014) Dyskerin-likeAtCBF5(Lermontovaetal.,2007) Pot-likeAtPot1a(Rossignoletal.,2007) RRM-motifAtRRM(LeeL.Y.etal.,2012) ARM-motifAtARM(LeeL.Y.etal.,2012) RNA-bindingAtG2p(Dokládaletal.,2015) Metallothionein- like AtMT2A(Dokládaletal.,2015) (Continued) Frontiers in Plant Science | www.frontiersin.org 6 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins TABLE1|Continued A.thalianaO.sativaZ.maysN.glutinosa/ N.tabacum/ N.sylvestris H.vulgareS.lycopersicum/ S.tuberosum C.parquiP.crispum DNAprocessingandrepair-associatedproteinsattelomerases HelicaseAtRTEL1(Reckeretal.,2014) ExonucleaseEXO1(Kazdaetal.,2012) KuAtKu70/80(Bundocketal.,2002; Rihaetal.,2002) OsKu70 (Hongetal.,2010) PI3kinaseAtATM(Amiardetal.,2011) AtATR(Amiardetal.,2011) MRNComplexAtRad50(GallegoandWhite,2001) AtMre11(BundockandHooykaas,2002) AtNbs1(NajdekrovaandSiroky,2012) Ku-independentEJpathwayAtRad1(Vannieretal.,2009) AtERCC1(Vannieretal.,2009) AtXRCC1(Amiardetal.,2014) Backup-NHEJKU-independentpathwayAtPARP1,2(Boltzetal.,2014) ProteinsthatwereprovenastelomericDNA-ortelomerase-associatedproteinsarelistedabove.Theirhomologuesinvolvedinnon-telomericprocesses,whiletheirassociationswithtelomeresortelomerasehavenot beenobserved,areshowningreen. SingleMybHistonefamily(SMH);TRF-like(TRFL-family);AntherIndehiscencefamily(AID-family);Cdc13/CTC1-Stn1-Ten1(CST);Oligonucleotide/OligosaccharideBindingfold(OB-fold);proteinswithout Oligonucleotide/OligosaccharideBindingfold(non-OB-fold);Protectionoftelomeres–like(Pot-like);Whirly(WHY);RNARecognitionMotifs(RRM-motif);TelomeraseReverseTranscriptase(TERT);TelomeraseRNA template(TR);Armadillo/β-catenin-likeRepeat-containingprotein(ARM);Phosphoinositide3-kinase(PI3kinase);Mre11/Rad50/Nbs1(MRN);End-joining(EJ);Non-homologousend-joining(NHEJ). Arabidopsisthaliana(At):TelomericRepeatBinding1-3(AtTRB1-3);TelomereBindingProtein1(AtTBP1);TelomereRepeatbindingFactor1(AtTRF1);TRF-like2-10(AtTRFL2-10);Suppressorofcdcthirteenhomolog (AtStn1);TelomericpathwaysinassociationwithStn1(AtTen);Conservedtelomeremaintenancecomponent1(AtCTC1);Protectionoftelomeres1a,b,c(AtPot1a,b,c);Whirly1(AtWhy1);Single-strandedTelomere- bindingProtein1(AtSTEP1);TelomeraseReverseTranscriptase(AtTERT);CajalBodiesFactor5(AtCBF5);RNARecognitionMotifs(AtRRM);Armadillo/β-catenin-likeRepeat-containingprotein(AtARM);Gene2(AtG2p); Metallothionein-like2A(AtMT2A);RegulatorofTelomereElongationHelicase1(AtRTEL1);Exonuclease1(AtEXO1);Ataxiatelangiectasiamutatedkinase(AtARM);ATM-andRAD3-relatedkinases(AtATR);DNArepair protein50(AtRad50);Meioticrecombination11(AtMre11);DNArepairprotein1(AtRad1);Excisionrepaircross-complementation1(AtERCC1);X-rayrepaircross-complementing1(XRCC1);Nijmegenbreakage syndrome1(AtNbs1);Poly(ADP-Ribose)polymerase1,2(AtPARP1,2). Orizasativa(Os):TelomereRepeatBindingFactor1(OsTRBF1);RiceTelomereBindingProtein1(RTBP1);AntherIndehiscence1(OsAID1);TelomeraseReverseTranscriptase(OsTERT). Zeamays(Zm):SingleMybHistone(ZmSMHs);Initiator-bindingprotein1(ZmIBP1);Initiator-bindingprotein2(ZmIBP2);TerminalacidicSANT1(ZmTacs1);Protectionoftelomeres1a,b(ZmPot1a,b);Telomerase ReverseTranscriptase(ZmTERT). Nicotianaglutinosa/tabacum/sylvestris(Ng/Nt/Ns):TelomereRepeatbindingFactor1(NgTRF1);TelomericssDNAbindingprotein1-3(NtGTBP1-3);TelomeraseReverseTranscriptase(Ng/NtTERT). Hordeumvulgare(Hv):Protectionoftelomeres1(HvPot1);Whirly1(HvWhy1). Solanumlycopersium/tuberosum(Le/St):TelomereBindingProtein1(LeTBP1);Protectionoftelomeres1(StPot1). Cestrumparqui(Cp):TelomereBindingProtein1(CpTBP1);Protectionoftelomeres1(CpPot1). Petroselinumcrispum(Pc):Myb-likeprotein1(PcMYB1);BoxP-bindingFactor1(PcBPF-1). Frontiers in Plant Science | www.frontiersin.org 7 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins TABLE2|TelomericandputativetelomericdsDNA-andssDNA-bindingproteinsfromArabidopsisthaliana. Protein abbreviation Accession number SchematicDepictionof ConservedDomains Telomeric- Sequence- Binding invitro Localization on Telomeres insitu Telomere length regulation Reference TelomericdsDNA-associatedproteins Myb-like proteins Myb-like domainatthe N-terminus (SMHfamily) AtTRB1At1g49950(+)(+)(+)Schrumpfovaetal.,2004:Schrumpfovaetal.,2014 AtTRB2At5g67580(+)(+/−)Schrumpfovaetal.,2004;Dvoráckováetal.,2010 AtTRB3At3g49850(+)(+/−)(+/−)Schrumpfovaetal.,2004;Dvoráckováetal.,2010; Zhouetal.,2016 Myb-like domainatthe C-terminus (TRFL-familyI) AtTBP1At5g13820(+)(−)Hwangetal.,2001;FulcherandRiha,2016 AtTRP1At5g59430(+)(−)Chenetal.,2001;FulcherandRiha,2016 AtTRFL1At3g46590(+)(−)Karamyshevaetal.,2004;Hwangetal.,2005; FulcherandRiha,2016 AITRFL2At1g07540(+)(−)Karamyshevaetal.,2004;FulcherandRiha,2016 AtTRFL4At3g53790(+)(−)Karamyshevaetal.,2004;FulcherandRiha,2016 AtTRFL9At3g12560(+)(−)Karamyshevaetal.,2004;Hwangetal.,2005; FulcherandRiha,2016 Myb-like domainatthe terminus (TRFL-familyII) AtTRFL3At1g17460(−)Karamyshevaetal.,2004 AtTRFL5At1g15720(−)Karamyshevaetal.,2004 AtTRFL6At1g72650(−)Karamyshevaetal.,2004 AtTRFL7At1g06910(−)Karamyshevaetal.,2004 AtTRFL8At2g37025(−)Karamyshevaetal.,2004 AtTRFL10At5g03780(−)Karamyshevaetal.,2004 CST complex AtStn1At1g07130(+)(+)Songetal.,2008 AtTen1At1g56260(+/−)(+)Leehyetal.,2013 AtCTC1At4g09680(+)(+)Surovtsevaetal.,2009 TelomericssDNA-associatedproteins OB-foldPot-likeAtPot1aAt2g05210(+/−)(+)KucharandFajkus,2004;TaniandMurata,2005; Shakirovetal.,2005,2009b;Surovtsevaetal., 2007 AtPot1bAt5g06310(+/−)(+)KucharandFajkus,2004;TaniandMurata,2005; Shakirovetal.,2005,2009b AtPot1cAt2g04395(+)Rossignoletal.,2007;Nelson,2012 Non-OB fold WHYAtWhy1At1g14410(+)(+)Yooetal.,2007 RRM-motifAtSTEP1At4g24770(+)KwonandChung,2004 Symbol(+)indicatespublishedevidenceofproteinbindingtotelomericsequence“Telomeric-sequenceBindinginvitro,”protein“LocalizationonTelomeresinsitu”,orthatproteinregulatestelomerelengthinplants “TelomereLengthRegulation.”Symbol(+/−)meansthatonlyindirectevidenceorambiguousresultsexist.Symbol(−)indicatesthatproteindoesnotspecificallyinteractwithtelomericDNA“Telomeric-sequenceBinding invitro,”proteinisnotlocalizedontelomeres“LocalizationonTelomeresinsitu”,orthatproteindoesnotregulatetelomerelengthinplants“TelomereLengthRegulation.”Myb-likedomain(Myb);Myb-extension(-ext); Histone-likedomain(Hl/5);CoiledCoilDomain(CCD);Oiigonucleotide/Oligosaccharide-BindingFolddomain(OB);Whirlydomain(Whirly);RNA-bindingdomain(RB);A.thaliana(At);TelomericRepeatBinding(AtTRB); TRF-likefamily(TRFLfamily);Suppressorofcdcthirteenhomolog(AtStn1);Conservedtelomeremaintenancecomponent1(AtCTC1);(CTC1-Stn1-Ten1)complex(CST);RNArecognitionmotifs(RRM);Protectionof telomeres1a,b,c(AtPot1a,b,c);Whirly1(Why1);Single-strandedtelomere-bindingprotein1(STEP1) Frontiers in Plant Science | www.frontiersin.org 8 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins FIGURE 3 | Association of shelterin proteins with extra-telomeric sequences. (A) Mammalian telomere-binding proteins TRF1/2, TIN2, or Rap1 associate not only with terminally localized telomeric repeats but also with interstitial telomeric sequences (ITS) and satellite repeats (Krutilina et al., 2001; Mignon-Ravix et al., 2002; Simonet et al., 2011; Yang et al., 2011). Mouse Rap1, together with TRF2, acts as a gene transcription regulator, including subtelomeric-gene (S-gene) silencing, and also binds to non-coding genomic regions enriched with TTAGGG repeats (Martinez et al., 2010). Moreover, TRF2 regulates neuronal genes by interaction of TRF2 with repressor element 1-silencing transcription factor (REST; Zhang et al., 2008). (B) Arabidopsis TRB1 protein was found not only as a component of the telomeric interactome (Schrumpfova et al., 2014), but also as a factor associated with 5 flanking regions (mostly comprising the telo box) of translation machinery genes (Schrumpfova et al., 2016; Zhou et al., 2016). associated with telomere dysfunction (Fulcher and Riha, 2016). Thus, although the AtTRFL proteins from A. thaliana specifically bind telomeric DNA in vitro and an interaction between AtTRP1 and AtKu70 was observed, suggesting a putative telomere function (Figure 2B; Kuchar and Fajkus, 2004), no functional evidence exists for their role at telomeres. Another member of this family—ZmIBP2 (initiator-binding protein) protein—binds not only telomeric repeats (Moore, 2009), but was originally identified as a promoter binding ligand (Lugert and Werr, 1994). Moreover, some members of this Myb-like family were identified exclusively based on their ability to bind promoter regions of certain genes: ZmIBP1 (Lugert and Werr, 1994), PcBPF-1 (box P-binding factor) from P. crispum (da Costa e Silva et al., 1993) or CrBPF from Catharanthus roseus (van der Fits et al., 2000). The third group with a Myb-like domain at the C-terminus (AID family) contains only a few members. The AID family is named according to anther indehiscence 1 (AID) protein from O. sativa—OsAID1 (Zhu et al., 2004). OsAID1 was initially identified as being involved in anther development. Another member of this family—ZmTacs1 (terminal acidic SANT) from Z. mays—may function in chromatin remodeling within the meristem (Marian and Bass, 2005). In an affinity pull-down technique, 80 proteins from O. sativa were identified for their ability to bind to a telomeric repeat (He et al., 2013). Among them, two of three previously reported proteins from the SMH family–OsTRBF1 and OsTRBF2 (Byun et al., 2008), and one protein with a Myb-domain at the C-terminus (AID family)—OsAID1 (Zhu et al., 2004; Du et al., 2013) were demonstrated, while no member with a Myb-domain at the C-terminus of the TRFL family could be found. From other ribonucleoproteins or RNA-binding proteins with putative telomere association, two homologs of N. tabacum telomeric ssDNA binding protein NtGTBP1 (Os08g0492100 and Os08g0320100), with RNA recognition motifs (RRM; see below; Lee and Kim, 2010), were also identified. Telomere-binding proteins in budding yeast (yRap1) or in mammals (TRF1, TRF2, Rap1, and TIN2) are associated with extra-telomeric sequences and thus participate in additional roles, e.g., gene activation and repression, DNA replication, heterochromatin boundary-element formation, creation of hotspots for meiotic recombination and chromatin opening (Figure 3A; Morse, 2000; Smogorzewska et al., 2000; Krutilina et al., 2001; Zhang et al., 2008; Martinez et al., 2010; Simonet et al., 2011; Yang et al., 2011; Maï et al., 2014; Ye et al., 2014). Frontiers in Plant Science | www.frontiersin.org 9 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins CST Complex An evolutionary conserved trimeric protein complex named CST (Cdc13/CTC1–Stn1–Ten1) is, similarly to Myb-like proteins, involved in several stages of telomere end formation. In yeast, these OB-fold proteins are required for recruitment of telomerase and DNA polymerase α to the chromosomal termini, and thus coordinate G-overhang extension by telomerase with the fillin synthesis of the complementary C-strand (Qi and Zakian, 2000; Grossi et al., 2004; Giraud-Panis et al., 2010; Wellinger and Zakian, 2012). In mammals, CST is primarily involved in the rescue of stalled replication forks at the telomere and elsewhere in the genome, and limits telomerase action at individual telomeres to approximately one binding and extension event per cell cycle (Figure 2A; Chen et al., 2012; Stewart et al., 2012; Chen and Lingner, 2013; Kasbek et al., 2013). In A. thaliana, a mutation in any CST subunits leads to severe morphological defects and is accompanied by a decrease in telomere length, single-strand G-overhang elongation, mostly subtelomere–subtelomere chromosomal fusions and the appearance of extra-chromosomal telomeric circles. Plants lacking Suppressor of cdc thirteen homolog (AtStn1) or Conserved telomere maintenance component 1 (AtCTC1) exhibit no change in telomerase activity whereas telomerase activity was elevated in atten1 mutants (Song et al., 2008; Surovtseva et al., 2009; Leehy et al., 2013). Although circumstantial evidence indicates that CST in plants is needed for telomere integrity, clear evidence is absent that would show any direct physical interaction of any component of the CST complex with plant telomeric DNA. As Arabidopsis AtCTC1 interacts with the catalytic subunit of DNA polymerase α (ICU2) in vitro (Price et al., 2010) and atstn1 mutant phenotypes can be partially phenocopied by impairment of DNA polymerase α, it was recently suggested that seemingly specific function(s) of CST in telomere protection may rather represent unique requirements for efficient replication of telomeric DNA (Figure 2B; Derboven et al., 2014). It seems that the CST complex controls access of telomerase, end-joining recombination and the ATR-dependent (ATM and Rad3-related) DNA damage response pathway at the chromosomal ends in wild-type plants (Boltz et al., 2012; Leehy et al., 2013; Amiard et al., 2014). Telomeric ssDNA-Associated Proteins Proteins with OB-fold The telomeric G-rich overhang is evolutionarily conserved and is a substrate for ssDNA binding proteins. The majority of ssDNA binding proteins bind through OB motifs (OB-fold) and are required for both chromosomal end protection and regulation of telomere length, e.g., telomere-binding protein subunit alpha/beta (TEBPαβ) from Oxytricha nova (telomere end binding protein; Price and Cech, 1987), Cell division cycle 13 (Cdc13p) from S. cerevisiae (Garvik et al., 1995) and Pot1, are present in diverse organisms including human, mouse, chicken, or S. pombe (Figure 2A; Baumann and Cech, 2001; Lei et al., 2002; Wei and Price, 2004; Wu et al., 2006). In A. thaliana, three Pot-like proteins have been named AtPot1a, AtPot1b, AtPot1c (Kuchar and Fajkus, 2004; Rossignol et al., 2007; previously also named as AtPOT1-1, AtPOT1-2 (Tani and Murata, 2005) or AtPot1, AtPot2 (Shakirov et al., 2005; see Rotkova et al., 2009 for an overview). However, descriptions of their functions and binding properties are not unanimously agreed. While a very weak, but specific affinity of AtPot1a and AtPot1b for plant telomeric ssDNA was originally described (Shakirov et al., 2005), later these authors could not demonstrate AtPot1a and AtPot1b binding to telomeric ssDNA in vitro (Shakirov et al., 2009a,b). Nevertheless, stable telomeric ssDNA binding was observed for two full-length plant Pot1 proteins: OlPot1 from the green alga Ostreococcus lucimarinus as well as for ZmPot1b from Z. mays (Shakirov et al., 2009b). Although Pot1 proteins from plant species as diverse as Hordeum vulgare (HvPot1; barley), Populus trichocarpa (poplar), Helianthus argophyllus (sunflower), Selaginella moellendorffii (spikemoss), Gossypium hirsutum (cotton), Pinus taeda (pine), Solanum tuberosum (StPot1; potato), Asparagus officinalis and Z. mays (ZmPot1a) failed to bind telomeric DNA when expressed in a rabbit reticulocyte lysate expression system in vitro and subjected to an electrophoretic mobility shift assay (Shakirov et al., 2009b), binding of plant Pot1 proteins to telomeric DNA under native conditions cannot be excluded. Plants expressing AtPot1a truncated by an N-terminal OB-fold, showed progressive loss of telomeric DNA. In contrast, telomere length was unperturbed in plants expressing analogously trimmed AtPot1b, although overexpression of C-terminally truncated AtPot1b resulted in telomere shortening (Shakirov et al., 2005). AtPot1a binds AtStn1 and AtCTC1 proteins (Figure 2B; Renfrew et al., 2014), associates with an N-terminally spliced variant of AtTERT (AtTERT-V(I8)) (Rossignol et al., 2007), TER1, one of the RNA subunits of Arabidopsis telomerase, and is required for maintenance of telomere length in vivo (Surovtseva et al., 2007). AtPot1b directly interacts with Myb-like proteins AtTRB1-3 from the SMH family (Schrumpfova et al., 2008), and associates with TER2 and TER2s, putative alternative RNA subunits of telomerase that negatively regulate the function of active telomerase particles (TER1-AtTERT; Cifuentes-Rojas et al., 2012). Nevertheless, AtPot1b does not seem to substantially contribute to telomere maintenance (Cifuentes-Rojas et al., 2012). Pot1-like proteins were also identified in plants with unusual telomeres (e.g., CpPot1 protein in C. parqui; Peska et al., 2008). Non-OB-fold Telomeric ssDNA Binding Proteins The transcriptional activator protein Whirly 1 (Why1), from a small protein family found mainly in land plants (Desveaux et al., 2000, 2002; Krause et al., 2005), was also identified in a fraction of telomere-binding proteins in A. thaliana, and an atwhy1 knockout mutant appeared to have shorter telomeres (Yoo et al., 2007). While proteins from A. thaliana (AtWhy1; Yoo et al., 2007) and from H. vulgare (HvWhy1; Grabowski et al., 2008) were found to bind plant telomeric repeat sequences in vitro, diverse organelle localization of other Why family members from O. sativa, A. thaliana, S. tuberosum (Krause et al., 2005; Schwacke et al., 2007) and proposed binding to ssDNA of melted promoter regions (Desveaux et al., 2002), rather indicate a role Frontiers in Plant Science | www.frontiersin.org 10 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins in communication between plastid and nuclear genes encoding photosynthetic proteins (Foyer et al., 2014; Comadira et al., 2015). A truncated derivative of chloroplast RNA-binding protein (AtCP31) with RRMs from A. thaliana, named AtSTEP1 (singlestranded telomere-binding protein 1), localizes exclusively to the nucleus, specifically binding single-stranded G-rich plant telomeric DNA sequences and inhibiting telomerase-mediated telomere extension (Kwon and Chung, 2004). A protein identified by gel mobility shift assay that specifically binds the G-strand of telomeric ssDNA from N. tabacum (NtGTBP1) also contains a tandem pair of RRMs (Hirata et al., 2004). NtGTBP1 is not only associated with telomeric sequences, as well as two additional GTBP paralogs (NtGTBP2 and NtGTBP3), but also inhibits telomeric strand invasion in vitro and leaves of knockdown tobacco plants contained longer telomeres with frequent formation of extrachromosomal t-circles (Lee and Kim, 2010). These observations correspond to a previously detected protein from tobacco nuclei that binds G-rich telomeric strands and reduces accessibility to telomerase or terminal transferase (Fulneckova and Fajkus, 2000). In addition to the above described proteins, various telomeric ssDNA binding proteins have also been reported in nuclear extracts from Glycine max, A. thaliana, O. sativa, or Vigna radiata (Zentgraf, 1995; Kim et al., 1998; Lee et al., 2000; Kwon et al., 2004). However, precise characterization of these proteins, identified by gel mobility shift assay, is mostly missing. DNA Repair Proteins and Telomeres Ku in plants, as well as in other eukaryotes, is a highly conserved complex, consisting of two polypeptides (Ku70 and Ku80; Mimori et al., 1981). Due to its high affinity for DNA ends, Ku has a generally conserved role across species in protecting DNA from nucleolytic degradation. Ku is important for several cellular mechanisms: the DNA double-stranded break (DSB) repair pathway by the Ku-dependent non-homologous endjoining (NHEJ) pathway, the DNA damage response machinery, or protection of telomere ends from being recognized as DSBs, thereby preventing their recombination and degradation (reviewed in Fell and Schild-Poulter, 2015). Human Ku directly interacts not only with the shelterin proteins hTRF1, hTRF2, and hRap1, but also with telomerase subunits hTERT and hTR (RNA template; reviewed in Fell and Schild-Poulter, 2015). In contrast to a massive loss of telomeric DNA that was observed in human cells (Wang et al., 2009), mutations in Ku70 and Ku80 in the dicotyledonous A. thaliana, as well as in the monocotyledonous O. sativa, resulted in longer telomeres, suggesting their conserved role in the negative regulation of plant telomerase (Bundock et al., 2002; Riha et al., 2002; Gallego et al., 2003; Hong et al., 2010). On the other hand, severe developmental defects were observed in O. sativa osku70 knockout mutants, but a similar mutation in A. thaliana atku70 showed no effect on plant development (Bundock et al., 2002; Hong et al., 2010). In S. latifolia and A. thaliana, Ku contributes to the integrity of blunt-ended telomeres by protecting them from nucleolytic resection (Kazda et al., 2012). AtKu specifically interacts with AtTRP1 protein (see above; Figure 2B; Kuchar and Fajkus, 2004) and also assembles with TER2 and TER2S into alternative telomerase complexes that cannot sustain telomere repeats on chromosomal ends (Cifuentes-Rojas et al., 2012). The mammalian shelterin complex is involved in the repression of the primary signal transducers of DNA breakage, two phosphatidylinositol-3-kinase-like (PI3K) protein kinases: ataxia telangiectasia mutated (ATM) and ATM- and RAD3related (ATR) kinases. Mice TRF2 acts mainly to protect telomeres against ATM activation (Celli and de Lange, 2005) and POT1 is principally involved in repression of the ATR pathway (Denchi and de Lange, 2007; Guo et al., 2007). Short telomeres in telomerase-deficient plants activate both the AtATM and AtATR, whereas absence of members of the CST complex initiates only AtATR-dependent, but not AtATM-dependent DNA damage response (Amiard et al., 2011; Boltz et al., 2012). In mammals as well as in other organisms, DSBs activate ATM kinase in a manner dependent on the meiotic recombination 11 (Mre11), DNA repair protein 50 (Rad50), and Nijmegen breakage syndrome 1 (Nbs1) named MRN complex. The MRN complex has been found to associate with telomeres and contributes to their maintenance (reviewed in Lamarche et al., 2010). A. thaliana AtRad50 mutant plant cells show a progressive shortening of telomeric DNA (Gallego and White, 2001), while in AtMre11 mutant plants, telomere lengthening was observed (Bundock and Hooykaas, 2002). Contrary to these observations, the absence of the third MRN subunit, AtNbs1, does not affect the length of telomeres (Najdekrova and Siroky, 2012). A. thaliana plants mutated in XPF (xeroderma pigmentosum group F-complementing) and ERCC1 (excision repair crosscomplementation group 1) orthologs that form a structurespecific endonuclease essential for nucleotide excision repair (known as AtRad1 and AtERCC1), develop normally and show wild-type telomere length. However, in the absence of telomerase, mutations in either of these genes induce a significantly earlier onset of chromosomal instability, thus indicating a protective role of AtERCC1/AtRad1 against a 3 G-strand overhang invasion of interstitial telomeric repeats (Vannier et al., 2009). In addition to the Ku proteins that are involved in Ku-dependent NHEJ, an alternative Ku-independent NHEJ pathway was described (reviewed in Decottignies, 2013). Members of the poly(ADPribose) polymerase family play a role not only in the base excision repair pathway and the backup-NHEJ KU-independent pathway (Decottignies, 2013) but were also studied in the context of telomere maintenance, association with shelterin proteins or modulation of telomerase activity (Smith et al., 1998; Cook et al., 2002; Beneke et al., 2008). However, analysis of Arabidopsis orthologs AtPARP1/AtPARP2 (poly(ADP-Ribose) polymerase) has revealed that, unlike in humans, AtPARPs play a minor role in telomere biology (Boltz et al., 2014). It was proposed that DSB repair pathways in A. thaliana are hierarchically organized and the Ku-dependent NHEJ restricts access and action of other DSB repair processes (Charbonnel et al., 2010, 2011). Furthermore the end-joining recombination proteins (AtKU80, AtXRCC1, AtRad1) restrict telomerase activity at deprotected telomeres (Amiard et al., 2014). It was found recently that structure-specific endonucleases AtMUS81 (MMS and UV-sensitive protein 81) and AtSEND1 (single-strand DNA endonuclease 1), which presumably act to repair potentially toxic Frontiers in Plant Science | www.frontiersin.org 11 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins structures produced by DNA replication and recombination, are essential for telomere stability in Arabidopsis. Combined absence of these endonucleases results in increased occurrence of histone γ-H2AX foci in S-phase and in loss of telomeric DNA (Olivier et al., 2016). PLANT TELOMERASE Telomere length in plants and various other organisms is maintained by telomerase, a specialized reverse transcriptase which, in addition to its catalytic subunit (TERT), carries its own RNA template (TR) and elongates telomeric tracts at the chromosomal terminus (Blackburn and Gall, 1978; Fajkus et al., 1996). TERT subunits consist of an N-terminal portion with telomerase-specific motifs important for binding the telomerase RNA subunit, catalytic domains with the telomerase reverse transcriptase (RT) motifs essential for enzyme activity, and the C-terminal extension, which is highly conserved among plants as well as vertebrates (Sykorova and Fajkus, 2009). Although most eukaryotes harbor only a single TERT gene, in the allotetraploid N. tabacum there are three NtTERT gene variants inherited from its diploid progenitor species N. sylvestris and Nicotiana tomentosiformis. All three NtTERT gene variants are transcribed. Alternative splicing provides a major source of protein diversity within a given organism. Alternatively spliced variants of TERT transcripts with out-of-frame and/or in-frame mutations were identified not only in humans, mouse, chicken, or Xenopus (reviewed in Hrdlickova et al., 2006), but also in many plant species, e.g., A. thaliana, Z. mays (ZmTERT), O. sativa (OsTERT), Iris tectorum, and tobacco [with human-type (TTAGGG) telomere motif; reviewed in Sykorova and Fajkus, 2009; Sykorova et al., 2012]. Isoforms generated by alternative splicing may show changes or loss of specific function(s) or subcellular localization of the respective product, or could be functionally important, as was suggested for the A. thaliana variant AtTERT V(I8) that exclusively interacts with AtPot1a (Rossignol et al., 2007). It has been proposed that human telomerase is subjected to posttranslational regulation such as phosphorylation (Kang et al., 1999). Putative phosphorylation sites were detected in the OsTERT sequences from O. sativa (Oguchi et al., 2004) or N. tabacum BY-2 cells (Yang et al., 2002) but not in AtTERT from A. thaliana (Oguchi et al., 2004). Telomerase-Associated Proteins Rich protein interactomes of yeast, mammalian or Ciliate TERT have been described, including the Ku heterodimer (Chai et al., 2002), HSP90 (heat-shock protein of 90 kDa; Holt et al., 1999; Grandin and Charbonneau, 2001), ATPases pontin and reptin (Venteicher et al., 2008), TEP1 (telomere protein 1; Harrington et al., 1997), and many others, in a broad study (Fu and Collins, 2007) and reviewed in a constantly updated telomerase database (Podlevsky et al., 2008). In AtTERT, a mitochondrial targeting signal, multiple nuclear localization signals or a nuclear export signal have been reported (Zachova et al., 2013). As AtTERT protein and its domains localize mainly within the nucleus and the nucleolus (Zachova et al., 2013), it can be assumed that most interacting protein partners relevant to telomeric functions will be found among nuclear or nucleolar proteins. In plants, a limited number of proteins that directly interact with TERT have been described. It was demonstrated by various direct methods that AtTRB proteins, a group of plant homologs of human TRF proteins with a Mybdomain at the N-terminus (see above), physically interact with N-terminal domains of AtTERT (Figure 2B; Schrumpfova et al., 2014). A mediated interaction between AtTRP1 protein that belongs to the TRFL family, and AtTERT, was also observed (Schrumpfova et al., 2014). Moreover, the N-terminal part of AtTERT exclusively interacts with AtPot1a but not AtPot1b (Rossignol et al., 2007). Also various proteins with an RRMmotif (AtRRM), an ARM-motif (armadillo/β-catenin-like repeatcontaining protein; AtARM), metallothionein-like (AtMT2A), or RNA-binding (AtG2p) motifs were found as AtTERT interacting partners in A. thaliana (Lee L.Y. et al., 2012; Dokládal et al., 2015). Indirect regulation of TERT by various proteins or hormones was further described in plants. In tobacco cell culture, phytohormones such as auxin or abscisic acid regulate phosphorylation of telomerase protein, which is required for the generation of a functional telomerase complex (Tamura et al., 1999; Yang et al., 2002). In A. thaliana, reduced endogenous concentrations of auxin in telomerase activator 1 (AtTAC1) mutant plants blocks the ability of this zinc-finger protein to induce AtTERT. However, AtTAC1 does not directly bind the AtTERT promoter (Ren et al., 2004, 2007). A minimal promoter region for AtTERT was proposed using a set of T-insertion mutant lines in the protein-coding region of the AtTERT gene or in lines with insertions at the 5 end of AtTERT (Fojtova et al., 2011). Moreover T-DNA insertions in the region upstream of the ATG start of AtTERT also led to the activation of putative regulatory elements (Fojtova et al., 2011). In vertebrates, only one TR per organism was described. The folding of the TR molecule offers interaction sites for various associating cofactors such as dyskerin, Ku, nucleolar protein 10 (NOP10), H/ACA ribonucleoprotein complex subunit 1 (GAR1), or subunit 2 (NHP2; Ting et al., 2005; for review, see Kiss et al., 2010). A single TR was also described among Brassicaceae family plants. However, in A. thaliana, two TRs were detected— TER1 and TER2, and the latter may be alternately spliced to a TER2s form (Beilstein et al., 2012). The Arabidopsis homolog of human dyskerin, named AtCBF5 (alias AtNAP57), is located within nucleoli and Cajal bodies (Lermontova et al., 2007), associates with active telomerase, and weakly with AtPOT1a, but not AtTERT or AtKu70 (Kannan et al., 2008). Telomerase-Independent Processes in Plant Telomere Dynamics Compared to the human model, knowledge of individual protein contributions to the maintenance of telomere length/ Frontiers in Plant Science | www.frontiersin.org 12 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins accessibility/folding in plants or telomerase biogenesis/regulation is still very limited. The process of telomere maintenance is complicated by the fact that besides the widespread system of telomere maintenance by telomerase (Fajkus et al., 1996; Heller et al., 1996) in plants as well as in other organisms, in the absence of telomerase, telomeres can be elongated by recombinationdependent and telomerase-independent alternative telomere lengthening (ALT) mechanisms (Fajkus et al., 2005). Moreover, in plants, the ALT events appear to participate in early plant development (Ruckova et al., 2008). It was shown that AtKu70 deficiency facilitates engagement of ALT lengthening in A. thaliana (Zellinger et al., 2007) and that ALT was suppressed in the absence of ATM protein (Vespa et al., 2007). Telomeric DNA of higher eukaryotes, including plants, is associated not only with specific proteins, but also with histone complexes that form nucleosomes (Figure 2; reviewed in Dvoˇráˇcková et al., 2015). In various organisms, as well as in plants, telomeric nucleosomes display an unusually short periodicity (157 bp in length), usually 20–40 bp shorter than bulk nucleosomes of the corresponding organism (Fajkus et al., 1995; Fajkus and Trifonov, 2001; reviewed in Pisano et al., 2008). Moreover, the plant telomeric repeat (CCCTAAA) is a natural target for plant-specific asymmetric methylation (Cokus et al., 2008) that was shown to be mediated by an siRNA pathway (Vrbsky et al., 2010). Analysis of telomeres in A. thaliana (Vrbsky et al., 2010) and N. tabacum (Majerova et al., 2011) has demonstrated that telomeric histones were associated with both heterochromatin- and euchromatin-specific marks. Recent data strongly support the involvement of various epigenetic mechanisms (DNA methylation, posttranslational modifications of histones, nucleosome assembly or levels of telomere-repeat containing RNA) in maintenance of telomere stability (reviewed in Dvoˇráˇcková et al., 2015) thus demonstrating complexity of telomere regulation. CONCLUSION The need for protection of chromosomal termini remains conserved across most species. Nevertheless, an extraordinary plasticity of mechanisms protecting telomeres has been described among various organisms (reviewed in Giraud-Panis et al., 2013). While individual capping proteins can differ greatly, common features such as homologous binding domains, structures, or interacting partners exist between seemingly different capping systems. Plant systems show certain distinct features of telomere maintenance, including the reversible regulation of telomerase in somatic cells and the absence of developmental telomere shortening (Fajkus et al., 1998; Riha et al., 1998). These distinctions promote further efforts to elucidate plant telomere interactomes. Only recently the first complexes of telomerebinding proteins were demonstrated and meanwhile it seems that the plant telomere-maintenance system shares similarities with that described in mammals. For example, in A. thaliana, one of the most studied plant model systems: (i) the core plant telomeric dsDNA binding proteins (AtTRBs, AtTRP, etc.) contain similar Myb-domains which are also present in human TRF1 or TRF2 proteins; (ii) homologs of human telomeric ssDNA binding hPot1 (AtPOT1a-c) were described; (iii) crossspecies conserved CST complexes (AtCTC1/AtTen1/AtStn1) retain its function in plants. The similarities between plant and mammalian telomeric DNA-associated proteins apply also to their roles in regulation of gene expression, which are independent of their roles in telomere capping (Lee and Cho, 2016; Schrumpfova et al., 2016; Zhou et al., 2016), as was previously described in their mammalian counterparts (reviewed in Maï et al., 2014; Ye et al., 2014). Elucidation of the composition of the plant version of shelterin and molecular dissection of its components and their roles will be important in the near future to assess the conservation and mechanisms of endprotection and end-replication processes in yeasts, plants and animals. AUTHOR CONTRIBUTIONS PPS contributed substantially to the writing of the manuscript, tables and drawing the figures; ŠS participated in preparation of tables; JF edited the manuscript. All authors read and approved the manuscript for publication. FUNDING This research was supported by the Czech Science Foundation (13-06943S and 16-01137S) and by the Ministry of Education, Youth and Sports of the Czech Republic under the project CEITEC 2020 (LQ1601). REFERENCES Adams, S. P., Hartman, T. P., Lim, K. Y., Chase, M. W., Bennett, M. D., Leitch, I. J., et al. (2001). Loss and recovery of Arabidopsis-type telomere repeat sequences 5 -(TTTAGGG)(n)-3 in the evolution of a major radiation of flowering plants. Proc. Biol. Sci. 268, 1541–1546. doi: 10.1098/rspb. 2001.1726 Amiard, S., Depeiges, A., Allain, E., White, C. I., and Gallego, M. E. (2011). Arabidopsis ATM and ATR kinases prevent propagation of genome damage caused by telomere dysfunction. Plant Cell 23, 4254–4265. doi: 10.1105/tpc.111.092387 Amiard, S., Olivier, M., Allain, E., Choi, K., Smith-Unna, R., Henderson, I. R., et al. (2014). Telomere stability and development of ctc1 mutants are rescued by inhibition of EJ recombination pathways in a telomerasedependent manner. Nucleic Acids Res. 42, 11979–11991. doi: 10.1093/nar/ gku897 Arat, N. O., and Griffith, J. D. (2012). Human Rap1 interacts directly with telomeric DNA and regulates TRF2 localization at the telomere. J. Biol. Chem. 287, 41583–41594. doi: 10.1074/jbc.M112.415984 Armstrong, S. J., Franklin, F. C., and Jones, G. H. (2001). Nucleolus-associated telomere clustering and pairing precede meiotic chromosome synapsis in Arabidopsis thaliana. J. Cell Sci. 114, 4207–4217. Frontiers in Plant Science | www.frontiersin.org 13 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins Aronen, T., and Ryynanen, L. (2014). Silver birch telomeres shorten in tissue culture. Tree Genet. Genomes 10, 67–74. doi: 10.1007/s11295-013- 0662-4 Bass, H. W., Marshall, W. F., Sedat, J. W., Agard, D. A., and Cande, W. Z. (1997). Telomeres cluster de novo before the initiation of synapsis: a three-dimensional spatial analysis of telomere positions before and during meiotic prophase. J. Cell Biol. 137, 5–18. doi: 10.1083/jcb.137.1.5 Baumann, P., and Cech, T. R. (2001). Pot1, the putative telomere endbinding protein in fission yeast and humans. Science 292, 1171–1175. doi: 10.1126/science.1060036 Beilstein, M. A., Brinegar, A. E., and Shippen, D. E. (2012). Evolution of the Arabidopsis telomerase RNA. Front. Genet. 3:188. doi: 10.3389/fgene.2012.00188 Beilstein, M. A., Renfrew, K. B., Song, X., Shakirov, E. V., Zanis, M. J., and Shippen, D. E. (2015). Evolution of the telomere-associated protein POT1a in Arabidopsis thaliana is characterized by positive selection to reinforce proteinprotein interaction. Mol. Biol. Evol. 32, 1329–1341. doi: 10.1093/molbev/ msv025 Beneke, S., Cohausz, O., Malanga, M., Boukamp, P., Althaus, F., and Burkle, A. (2008). Rapid regulation of telomere length is mediated by poly(ADPribose) polymerase-1. Nucleic Acids Res. 36, 6309–6317. doi: 10.1093/nar/ gkn615 Bianchi, A., and Shore, D. (2008). How telomerase reaches its end: mechanism of telomerase regulation by the telomeric complex. Mol. Cell. 31, 153–165. doi: 10.1016/j.molcel.2008.06.013 Bilaud, T., Koering, C. E., Binet-Brasselet, E., Ancelin, K., Pollice, A., Gasser, S. M., et al. (1996). The telobox, a Myb-related telomeric DNA binding motif found in proteins from yeast, plants and human. Nucleic Acids Res. 24, 1294–1303. doi: 10.1093/nar/24.7.1294 Blackburn, E. H., and Gall, J. G. (1978). A tandemly repeated sequence at the termini of the extrachromosomal ribosomal RNA genes in Tetrahymena. J. Mol. Biol. 120, 33–53. Boltz, K. A., Jasti, M., Townley, J. M., and Shippen, D. E. (2014). Analysis of poly(ADP-Ribose) polymerases in Arabidopsis telomere biology. PLoS ONE 9:e88872. doi: 10.1371/journal.pone.0088872 Boltz, K. A., Leehy, K., Song, X., Nelson, A. D., and Shippen, D. E. (2012). ATR cooperates with CTC1 and STN1 to maintain telomeres and genome integrity in Arabidopsis. Mol. Biol. Cell 23, 1558–1568. doi: 10.1091/mbc.E11-12-1002 Broccoli, D., Smogorzewska, A., Chong, L., and de Lange, T. (1997). Human telomeres contain two distinct Myb-related proteins, TRF1 and TRF2. Nat. Genet. 17, 231–235. doi: 10.1038/ng1097-231 Bundock, P., and Hooykaas, P. (2002). Severe developmental defects, hypersensitivity to DNA-damaging agents, and lengthened telomeres in Arabidopsis MRE11 mutants. Plant Cell 14, 2451–2462. doi: 10.1105/tpc. 005959 Bundock, P., van Attikum, H., and Hooykaas, P. (2002). Increased telomere length and hypersensitivity to DNA damaging agents in an Arabidopsis KU70 mutant. Nucleic Acids Res. 30, 3395–3400. doi: 10.1093/nar/gkf445 Byun, M. Y., Hong, J. P., and Kim, W. T. (2008). Identification and characterization of three telomere repeat-binding factors in rice. Biochem. Biophys. Res. Commun. 372, 85–90. doi: 10.1016/j.bbrc.2008.04.181 Celli, G. B., and de Lange, T. (2005). DNA processing is not required for ATMmediated telomere damage response after TRF2 deletion. Nat. Cell Biol. 7, 712–718. doi: 10.1038/ncb1275 Cesare, A. J., Quinney, N., Willcox, S., Subramanian, D., and Griffith, J. D. (2003). Telomere looping in P. sativum (common garden pea). Plant J. 36, 271–279. doi: 10.1046/j.1365-313X.2003.01882.x Chai, W., Ford, L. P., Lenertz, L., Wright, W. E., and Shay, J. W. (2002). Human Ku70/80 associates physically with telomerase through interaction with hTERT. J. Biol. Chem. 277, 47242–47247. doi: 10.1074/jbc.M208 542200 Charbonnel, C., Allain, E., Gallego, M. E., and White, C. I. (2011). Kinetic analysis of DNA double-strand break repair pathways in Arabidopsis. DNA Repair (Amst.) 10, 611–619. doi: 10.1016/j.dnarep.2011.04.002 Charbonnel, C., Gallego, M. E., and White, C. I. (2010). Xrcc1-dependent and Kudependent DNA double-strand break repair kinetics in Arabidopsis plants. Plant J. 64, 280–290. doi: 10.1111/j.1365-313X.2010.04331.x Chen, C. M., Wang, C. T., and Ho, C. H. (2001). A plant gene encoding a Myblike protein that binds telomeric GGTTTAG repeats in vitro. J. Biol. Chem. 276, 16511–16519. doi: 10.1074/jbc.M009659200 Chen, C. M., Wang, C. T., Kao, Y. H., Chang, G. D., Ho, C. H., Lee, F. M., et al. (2005). Functional redundancy of the duplex telomeric DNA-binding proteins in Arabidopsis. Bot. Bull. Acad. Sin. 46, 315–324. Chen, L. Y., and Lingner, J. (2013). CST for the grand finale of telomere replication. Nucleus 4, 277–282. doi: 10.4161/nucl.25701 Chen, L. Y., Redon, S., and Lingner, J. (2012). The human CST complex is a terminator of telomerase activity. Nature 488, 540–544. doi: 10.1038/nature11269 Chikashige, Y., Tsutsumi, C., Yamane, M., Okamasa, K., Haraguchi, T., and Hiraoka, Y. (2006). Meiotic proteins bqt1 and bqt2 tether telomeres to form the bouquet arrangement of chromosomes. Cell 125, 59–69. doi: 10.1016/j.cell.2006.01.048 Cifuentes-Rojas, C., Nelson, A. D., Boltz, K. A., Kannan, K., She, X., and Shippen, D. E. (2012). An alternative telomerase RNA in Arabidopsis modulates enzyme activity in response to DNA damage. Genes Dev. 26, 2512–2523. doi: 10.1101/gad.202960.112 Cokus, S. J., Feng, S., Zhang, X., Chen, Z., Merriman, B., Haudenschild, C. D., et al. (2008). Shotgun bisulphite sequencing of the Arabidopsis genome reveals DNA methylation patterning. Nature 452, 215–219. doi: 10.1038/nature06745 Comadira, G., Rasool, B., Kaprinska, B., Garcia, B. M., Morris, J., Verrall, S. R., et al. (2015). WHIRLY1 functions in the control of responses to nitrogen deficiency but not aphid infestation in barley. Plant Physiol. 168, 1140–1151. doi: 10.1104/pp.15.00580 Cook, B. D., Dynek, J. N., Chang, W., Shostak, G., and Smith, S. (2002). Role for the related poly(ADP-Ribose) polymerases tankyrase 1 and 2 at human telomeres. Mol. Cell. Biol. 22, 332–342. doi: 10.1128/MCB.22.1.332-342.2002 Corredor, E., and Naranjo, T. (2007). Effect of colchicine and telocentric chromosome conformation on centromere and telomere dynamics at meiotic prophase I in wheat-rye additions. Chromosome Res. 15, 231–245. doi: 10.1007/s10577-006-1117-7 Court, R., Chapman, L., Fairall, L., and Rhodes, D. (2005). How the human telomeric proteins TRF1 and TRF2 recognize telomeric DNA: a view from high-resolution crystal structures. EMBO Rep. 6, 39–45. doi: 10.1038/sj.embor.7400314 Cowan, C. R., Carlton, P. M., and Cande, W. Z. (2001). The polar arrangement of telomeres in interphase and meiosis. Rabl organization and the bouquet. Plant Physiol. 125, 532–538. Cowan, C. R., Carlton, P. M., and Cande, W. Z. (2002). Reorganization and polarization of the meiotic bouquet-stage cell can be uncoupled from telomere clustering. J. Cell Sci. 115, 3757–3766. doi: 10.1242/jcs.00054 da Costa e Silva, O., Klein, L., Schmelzer, E., Trezzini, G. F., and Hahlbrock, K. (1993). BPF-1, a pathogen-induced DNA-binding protein involved in the plant defense response. Plant J. 4, 125–135. de Lange, T. (2005). Shelterin: the protein complex that shapes and safeguards human telomeres. Genes Dev. 19, 2100–2110. doi: 10.1101/gad.1346005 Decottignies, A. (2013). Alternative end-joining mechanisms: a historical perspective. Front. Genet. 4:48. doi: 10.3389/fgene.2013.00048 Denchi, E. L., and de Lange, T. (2007). Protection of telomeres through independent control of ATM and ATR by TRF2 and POT1. Nature 448, 1068–1071. doi: 10.1038/nature06065 Deng, W., Buzas, D. M., Ying, H., Robertson, M., Taylor, J., Peacock, W. J., et al. (2013). Arabidopsis Polycomb Repressive Complex 2 binding sites contain putative GAGA factor binding motifs within coding regions of genes. BMC Genomics 14:593. doi: 10.1186/1471-2164-14-593 Derboven, E., Ekker, H., Kusenda, B., Bulankova, P., and Riha, K. (2014). Role of STN1 and DNA polymerase alpha in telomere stability and genome-wide replication in Arabidopsis. PLoS Genet. 10:e1004682. doi: 10.1371/journal.pgen.1004682 Desveaux, D., Allard, J., Brisson, N., and Sygusch, J. (2002). A new family of plant transcription factors displays a novel ssDNA-binding surface. Nat. Struct. Biol. 9, 512–517. doi: 10.1038/nsb814 Desveaux, D., Despres, C., Joyeux, A., Subramaniam, R., and Brisson, N. (2000). PBF-2 is a novel single-stranded DNA binding factor implicated in PR-10a gene activation in potato. Plant Cell 12, 1477–1489. doi: 10.1105/tpc.12.8.1477 Frontiers in Plant Science | www.frontiersin.org 14 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins Dokládal, L., Honys, D., Rana, R., Lee, L.-Y., Gelvin, S., and Sýkorová, E. (2015). cDNA library screening identifies protein interactors potentially involved in non-telomeric roles of Arabidopsis telomerase. Front. Plant Sci. 6:985. doi: 10.3389/fpls.2015.00985 Dong, F., and Jiang, J. (1998). Non-Rabl patterns of centromere and telomere distribution in the interphase nuclei of plant cells. Chromosome Res. 6, 551–558. doi: 10.1023/A:1009280425125 Du, H., Wang, Y. B., Xie, Y., Liang, Z., Jiang, S. J., Zhang, S. S., et al. (2013). Genome-wide identification and evolutionary and expression analyses of MYBrelated genes in land plants. DNA Res. 20, 437–448. doi: 10.1093/dnares/dst021 Dvoˇráˇcková, M., Fojtova, M., and Fajkus, J. (2015). Chromatin dynamics of plant telomeres and ribosomal genes. Plant J. 83, 18–37. doi: 10.1111/tpj. 12822 Dvorácková, M., Rossignol, P., Shaw, P. J., Koroleva, O. A., Doonan, J. H., and Fajkus, J. (2010). AtTRB1, a telomeric DNA-binding protein from Arabidopsis, is concentrated in the nucleolus and shows highly dynamic association with chromatin. Plant J. 61, 637–649. doi: 10.1111/j.1365-313X.2009. 04094.x Fajkus, J., Fulneckova, J., Hulanova, M., Berkova, K., Riha, K., and Matyasek, R. (1998). Plant cells express telomerase activity upon transfer to callus culture, without extensively changing telomere lengths. Mol. Gen. Genet. 260, 470–474. doi: 10.1007/s004380050918 Fajkus, J., Kovarik, A., and Kralovics, R. (1996). Telomerase activity in plant cells. FEBS Lett. 391, 307–309. doi: 10.1016/0014-5793(96)00757-0 Fajkus, J., Kovarik, A., Kralovics, R., and Bezdek, M. (1995). Organization of telomeric and subtelomeric chromatin in the higher plant Nicotiana tabacum. Mol. Gen. Genet. 247, 633–638. doi: 10.1007/BF00290355 Fajkus, J., Sykorova, E., and Leitch, A. R. (2005). Telomeres in evolution and evolution of telomeres. Chromosome Res. 13, 469–479. doi: 10.1007/s10577- 005-0997-2 Fajkus, J., and Trifonov, E. N. (2001). Columnar packing of telomeric nucleosomes. Biochem. Biophys. Res. Commun. 280, 961–963. doi: 10.1006/bbrc. 2000.4208 Fajkus, P., Peska, V., Sitova, Z., Fulnecková, J., Dvorackova, M., Gogela, R., et al. (2016). Allium telomeres unmasked: the unusual telomeric sequence (CTCGGTTATGGG)n is synthesized by telomerase. Plant J. 85, 337–347. doi: 10.1111/tpj.13115 Feldbrugge, M., Sprenger, M., Hahlbrock, K., and Weisshaar, B. (1997). PcMYB1, a novel plant protein containing a DNA-binding domain with one MYB repeat, interacts in vivo with a light-regulatory promoter unit. Plant J. 11, 1079–1093. doi: 10.1046/j.1365-313X.1997.11051079.x Fell, V. L., and Schild-Poulter, C. (2015). The Ku heterodimer: function in DNA repair and beyond. Mutat. Res. Rev. Mutat. Res. 763, 15–29. doi: 10.1016/j.mrrev.2014.06.002 Fojtova, M., Peska, V., Dobsakova, Z., Mozgova, I., Fajkus, J., and Sykorova, E. (2011). Molecular analysis of T-DNA insertion mutants identified putative regulatory elements in the AtTERT gene. J. Exp. Bot. 62, 5531–5545. doi: 10.1093/jxb/err235 Fojtova, M., Sykorova, E., Najdekrova, L., Polanska, P., Zachova, D., Vagnerova, R., et al. (2015). Telomere dynamics in the lower plant Physcomitrella patens. Plant Mol. Biol. 87, 591–601. doi: 10.1007/s11103-015- 0299-9 Foyer, C. H., Karpinska, B., and Krupinska, K. (2014). The functions of WHIRLY1 and REDOX-RESPONSIVE TRANSCRIPTION FACTOR 1 in cross tolerance responses in plants: a hypothesis. Philos. Trans. R. Soc. Lond. B Biol. Sci. 369:20130226. doi: 10.1098/rstb.2013.0226 Freeling, M. (2009). Bias in plant gene content following different sorts of duplication: tandem, whole-genome, segmental, or by transposition. Annu. Rev. Plant Biol. 60, 433–453. doi: 10.1146/annurev.arplant.043008.092122 Fu, D., and Collins, K. (2007). Purification of human telomerase complexes identifies factors involved in telomerase biogenesis and telomere length regulation. Mol. Cell 28, 773–785. doi: 10.1016/j.molcel.2007.09.023 Fulcher, N., and Riha, K. (2016). Using centromere mediated genome elimination to elucidate the funcional redudancy of candidate telomere binding proteins in Arabidopsis thaliana. Front. Genet. 6:349. doi: 10.3389/fgene.2015.00349 Fulneckova, J., and Fajkus, J. (2000). Inhibition of plant telomerase by telomerebinding proteins from nuclei of telomerase-negative tissues. FEBS Lett. 467, 305–310. doi: 10.1016/S0014-5793(00)01178-9 Fulnecková, J., Sevcíková, T., Fajkus, J., Lukesová, A., Lukes, M., Vlcek, C., et al. (2013). A broad phylogenetic survey unveils the diversity and evolution of telomeres in eukaryotes. Genome Biol. Evol. 5, 468–483. doi: 10.1093/gbe/evt019 Fulneˇcková, J., Ševˇcikova, T., Lukešová, A., and Sýkorová, E. (2015). Transtitions between the Arabidopsis-type and the human-type telomere sequence in green algae (clade Caudivolvoxa, Chlamydomonadales). Chromosoma. doi: 10.1007/s00412-015-0557-2 [Epub ahead of print]. Gallego, M. E., Jalut, N., and White, C. I. (2003). Telomerase dependence of telomere lengthening in Ku80 mutant Arabidopsis. Plant Cell 15, 782–789. doi: 10.1105/tpc.008623 Gallego, M. E., and White, C. I. (2001). RAD50 function is essential for telomere maintenance in Arabidopsis. Proc. Natl. Acad. Sci. U.S.A. 98, 1711–1716. doi: 10.1073/pnas.98.4.1711 Garvik, B., Carson, M., and Hartwell, L. (1995). Single-stranded DNA arising at telomeres in cdc13 mutants may constitute a specific signal for the RAD9 checkpoint. Mol. Cell. Biol. 15, 6128–6613. doi: 10.1128/MCB.15.11.6128 Giraud-Panis, M. J., Pisano, S., Benarroch-Popivker, D., Pei, B., Le, Du, M. H., et al. (2013). One identity or more for telomeres? Front. Oncol. 3:48. doi: 10.3389/fonc.2013.00048 Giraud-Panis, M. J., Teixeira, M. T., Geli, V., and Gilson, E. (2010). CST meets shelterin to keep telomeres in check. Mol. Cell. 39, 665–676. doi: 10.1016/j.molcel.2010.08.024 Grabowski, E., Miao, Y., Mulisch, M., and Krupinska, K. (2008). Singlestranded DNA-binding protein Whirly1 in barley leaves is located in plastids and the nucleus of the same cell. Plant Physiol. 147, 1800–1804. doi: 10.1104/pp.108.122796 Grandin, N., and Charbonneau, M. (2001). Hsp90 levels affect telomere length in yeast. Mol. Genet. Genom. 265, 126–134. doi: 10.1007/s004380000398 Graumann, K., Runions, J., and Evans, D. E. (2010). Characterization of SUNdomain proteins at the higher plant nuclear envelope. Plant J. 61, 134–144. doi: 10.1111/j.1365-313X.2009.04038.x Grossi, S., Puglisi, A., Dmitriev, P. V., Lopes, M., and Shore, D. (2004). Pol12, the B subunit of DNA polymerase alpha, functions in both telomere capping and length regulation. Genes Dev. 18, 992–1006. doi: 10.1101/gad. 300004 Guo, X., Deng, Y., Lin, Y., Cosme-Blanco, W., Chan, S., He, H., et al. (2007). Dysfunctional telomeres activate an ATM-ATR-dependent DNA damage response to suppress tumorigenesis. EMBO J. 26, 4709–4719. Harrington, L., McPhail, T., Mar, V., Zhou, W., Oulton, R., Bass, M. B., et al. (1997). A mammalian telomerase-associated protein. Science 275, 973–977. doi: 10.1126/science.275.5302.973 He, H., Multani, A. S., Cosme-Blanco, W., Tahara, H., Ma, J., Pathak, S., et al. (2006). POT1b protects telomeres from end-to-end chromosomal fusions and aberrant homologous recombination. EMBO J. 25, 5180–5190. doi: 10.1038/sj.emboj.7601294 He, Q., Chen, L., Xu, Y., and Yu, W. (2013). Identification of centromeric and telomeric DNA-binding proteins in rice. Proteomics 13, 826–832. doi: 10.1002/pmic.201100416 He, X. L., and Zhang, J. Z. (2005). Gene complexity and gene duplicability. Curr. Biol. 15, 1016–1021. doi: 10.1016/j.cub.2005.04.035 Heller, K., Kilian, A., Piatyszek, M. A., and Kleinhofs, A. (1996). Telomerase activity in plant extracts. Mol. Gen. Genet. 252, 342–345. doi: 10.1007/BF02 173780 Higgins, J. D., Perry, R. M., Barakate, A., Ramsay, L., Waugh, R., Halpin, C., et al. (2012). Spatiotemporal asymmetry of the meiotic program underlies the predominantly distal distribution of meiotic crossovers in barley. Plant Cell 24, 4096–4109. doi: 10.1105/tpc.112.102483 Hirata, Y., Suzuki, C., and Sakai, S. (2004). Characterization and gene cloning of telomere-binding protein from tobacco BY-2 cells. Plant Physiol. Biochem. 42, 7–14. doi: 10.1016/j.plaphy.2003.10.002 Hofr, C., Sultesova, P., Zimmermann, M., Mozgova, I., Prochazkova Schrumpfova, P., Wimmerova, M., et al. (2009). Single-Myb-histone proteins from Arabidopsis thaliana: a quantitative study of telomere-binding specificity and kinetics. Biochem. J. 419, 221–228. doi: 10.1042/BJ20082195 , 222 p following 228. Holt, S. E., Aisner, D. L., Baur, J., Tesmer, V. M., Dy, M., Ouellette, M., et al. (1999). Functional requirement of p23 and Hsp90 in telomerase complexes. Genes Dev. 13, 817–826. doi: 10.1101/gad.13.7.817 Frontiers in Plant Science | www.frontiersin.org 15 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins Hong, J. P., Byun, M. Y., An, K., Yang, S. J., An, G., and Kim, W. T. (2010). OsKu70 is associated with developmental growth and genome stability in rice. Plant Physiol. 152, 374–387. doi: 10.1104/pp.109.150391 Hong, J. P., Byun, M. Y., Koo, D. H., An, K., Bang, J. W., Chung, I. K., et al. (2007). Suppression of RICE TELOMERE BINDING PROTEIN 1 results in severe and gradual developmental defects accompanied by genome instability in rice. Plant Cell 19, 1770–1781. doi: 10.1105/tpc.107.051953 Hrdlickova, R., Nehyba, J., Liss, A. S., and Bose, H. R. Jr. (2006). Mechanism of telomerase activation by v-Rel and its contribution to transformation. J. Virol. 80, 281–295. doi: 10.1128/JVI.80.1.281-295.2006 Hwang, M. G., Chung, I. K., Kang, B. G., and Cho, M. H. (2001). Sequence-specific binding property of Arabidopsis thaliana telomeric DNA binding protein 1 (AtTBP1). FEBS Lett. 503, 35–40. doi: 10.1016/S0014-5793(01)02685-0 Hwang, M. G., Kim, K., Lee, W. K., and Cho, M. H. (2005). AtTBP2 and AtTRP2 in Arabidopsis encode proteins that bind plant telomeric DNA and induce DNA bending in vitro. Mol. Gen. Genom. 273, 66–75. doi: 10.1007/s00438-004-1096-3 Idziak, D., Robaszkiewicz, E., and Hasterok, R. (2015). Spatial distribution of centromeres and telomeres at interphase varies among Brachypodium species. J. Exp. Bot. 66, 6623–6634. doi: 10.1093/jxb/erv369 Janouskova, E., Necasova, I., Pavlouskova, J., Zimmermann, M., Hluchy, M., Marini, V., et al. (2015). Human Rap1 modulates TRF2 attraction to telomeric DNA. Nucleic Acids Res. 43, 2691–2700. doi: 10.1093/nar/gkv097 Kaminker, P. G., Kim, S. H., Desprez, P. Y., and Campisi, J. (2009). A novel form of the telomere-associated protein TIN2 localizes to the nuclear matrix. Cell Cycle 8, 931–939. doi: 10.4161/cc.8.6.7941 Kang, S. S., Kwon, T., Kwon, D. Y., and Do, S. I. (1999). Akt protein kinase enhances human telomerase activity through phosphorylation of telomerase reverse transcriptase subunit. J. Biol. Chem. 274, 13085–13090. doi: 10.1074/jbc.274.19.13085 Kannan, K., Nelson, A. D., and Shippen, D. E. (2008). Dyskerin is a component of the Arabidopsis telomerase RNP required for telomere maintenance. Mol. Cell. Biol. 28, 2332–2341. doi: 10.1128/MCB.01490-07 Karamysheva, Z. N., Surovtseva, Y. V., Vespa, L., Shakirov, E. V., and Shippen, D. E. (2004). A C-terminal Myb extension domain defines a novel family of double-strand telomeric DNA-binding proteins in Arabidopsis. J. Biol. Chem. 279, 47799–47807. doi: 10.1074/jbc.M407938200 Kasbek, C., Wang, F., and Price, C. M. (2013). Human TEN1 maintains telomere integrity and functions in genome-wide replication restart. J. Biol. Chem. 288, 30139–30150. doi: 10.1074/jbc.M113.493478 Kazda, A., Zellinger, B., Rossler, M., Derboven, E., Kusenda, B., and Riha, K. (2012). Chromosome end protection by blunt-ended telomeres. Genes Dev. 26, 1703–1713. doi: 10.1101/gad.194944.112 Kim, J. H., Kim, W. T., and Chung, I. K. (1998). Rice proteins that bind single-stranded G-rich telomere DNA. Plant Mol. Biol. 36, 661–672. doi: 10.1023/A:1005994719175 Kiss, T., Fayet-Lebaron, E., and Jady, B. E. (2010). Box H/ACA small ribonucleoproteins. Mol. Cell. 37, 597–606. doi: 10.1016/j.molcel.2010.01.032 Ko, S., Jun, S. H., Bae, H., Byun, J. S., Han, W., Park, H., et al. (2008). Structure of the DNA-binding domain of NgTRF1 reveals unique features of plant telomerebinding proteins. Nucleic Acids Res. 36, 2739–2755. doi: 10.1093/nar/gkn030 Koonin, E. V. (2010). Preview. The incredible expanding ancestor of eukaryotes. Cell 140, 606–608. doi: 10.1016/j.cell.2010.02.022 Kovarik, A., Fajkus, J., Koukalova, B., and Bezdek, M. (1996). Species-specific evolution of telomeric and rDNA repeats in the tobacco composite genome. Theor. Appl. Genet. 92, 1108–1111. doi: 10.1007/BF00224057 Krause, K., Kilbienski, I., Mulisch, M., Rodiger, A., Schafer, A., and Krupinska, K. (2005). DNA-binding proteins of the Whirly family in Arabidopsis thaliana are targeted to the organelles. FEBS Lett. 579, 3707–3712. doi: 10.1016/j.febslet.2005.05.059 Krutilina, R. I., Oei, S., Buchlow, G., Yau, P. M., Zalensky, A. O., Zalenskaya, I. A., et al. (2001). A negative regulator of telomere-length protein trf1 is associated with interstitial (TTAGGG)n blocks in immortal Chinese hamster ovary cells. Biochem. Biophys. Res. Commun. 280, 471–475. doi: 10.1006/bbrc.2000. 4143 Kuchar, M., and Fajkus, J. (2004). Interactions of putative telomere-binding proteins in Arabidopsis thaliana: identification of functional TRF2 homolog in plants. FEBS Lett. 578, 311–315. doi: 10.1016/j.febslet.2004.11.021 Kwon, C., and Chung, I. K. (2004). Interaction of an Arabidopsis RNA-binding protein with plant single-stranded telomeric DNA modulates telomerase activity. J. Biol. Chem. 279, 12812–12818. doi: 10.1074/jbc.M312011200 Kwon, C., Kwon, K., Chung, I. K., Kim, S. Y., Cho, M. H., and Kang, B. G. (2004). Characterization of single stranded telomeric DNA-binding proteins in cultured soybean (Glycine max) cells. Mol. Cells 17, 503–508. Lamarche, B. J., Orazio, N. I., and Weitzman, M. D. (2010). The MRN complex in double-strand break repair and telomere maintenance. FEBS Lett. 584, 3682–3695. doi: 10.1016/j.febslet.2010.07.029 Lazzerini-Denchi, E., and Sfeir, A., (2016). Stop pulling my strings – what telomeres taught us about the DNA damage response. Nat. Rev. Mol. Cell. Biol. 17, 364–378. doi: 10.1038/nrm.2016.43 Lee, J. H., Kim, J. H., Kim, W. T., Kang, B. G., and Chung, I. K. (2000). Characterization and developmental expression of single-stranded telomeric DNA-binding proteins from mung bean (Vigna radiata). Plant Mol. Biol. 42, 547–557. doi: 10.1023/A:1006373917321 Lee, L. Y., Wu, F. H., Hsu, C. T., Shen, S. C., Yeh, H. Y., Liao, D. C., et al. (2012). Screening a cDNA library for protein-protein interactions directly in planta. Plant Cell 24, 1746–1759. doi: 10.1105/tpc.112.097998 Lee, W. K., and Cho, M. H. (2016). Telomere-binding protein regulates the chromosome ends through the interaction with histone deacetylases in Arabidopsis thaliana. Nucl. Acids Res. 44, 4610–4624. doi: 10.1093/nar/gkw067 Lee, W. K., Yun, J. H., Lee, W., and Cho, M. H. (2012). DNA-binding domain of AtTRB2 reveals unique features of a single Myb histone protein family that binds to both Arabidopsis- and human-type telomeric DNA sequences. Mol. Plant 5, 1406–1408. doi: 10.1093/mp/sss063 Lee, Y. W., and Kim, W. T. (2010). Tobacco GTBP1, a homolog of human heterogeneous nuclear ribonucleoprotein, protects telomeres from aberrant homologous recombination. Plant Cell 22, 2781–2795. doi: 10.1105/tpc.110.076778 Lee, Y. W., and Kim, W. T. (2013). Telomerase-dependent 3 G-strand overhang maintenance facilitates GTBP1-mediated telomere protection from misplaced homologous recombination. Plant Cell 25, 1329–1342. doi: 10.1105/tpc.112.107573 Leehy, K. A., Lee, J. R., Song, X., Renfrew, K. B., and Shippen, D. E. (2013). MERISTEM DISORGANIZATION1 encodes TEN1, an essential telomere protein that modulates telomerase processivity in Arabidopsis. Plant Cell 25, 1343–1354. doi: 10.1105/tpc.112.107425 Lei, M., Baumann, P., and Cech, T. R. (2002). Cooperative binding of single-stranded telomeric DNA by the Pot1 protein of Schizosaccharomyces pombe. Biochemistry 41, 14560–14568. doi: 10.1021/ bi026674z Lermontova, I., Schubert, V., Bornke, F., Macas, J., and Schubert, I. (2007). Arabidopsis CBF5 interacts with the H/ACA snoRNP assembly factor NAF1. Plant Mol. Biol. 65, 615–626. doi: 10.1007/s11103-007-9226-z Linger, B. R., and Price, C. M. (2009). Conservation of telomere protein complexes: shuffling through evolution. Crit. Rev. Biochem. Mol. Biol. 44, 434–446. doi: 10.3109/10409230903307329 Lugert, T., and Werr, W. (1994). A novel DNA-binding domain in the Shrunken initiator-binding protein (IBP1). Plant Mol. Biol. 25, 493–506. doi: 10.1007/BF00043877 Maï, M. E., Wagner, K. D., Michiels, J. F., Ambrosetti, D., Borderie, A., Destree, S., et al. (2014). The telomeric protein TRF2 regulates Angiogenesis by binding and activating the PDGFRβ promoter. Cell Rep. 9, 1047–1060. doi: 10.1016/j.celrep.2014.09.038 Maillet, G., White, C. I., and Gallego, M. E. (2006). Telomere-length regulation in inter-ecotype crosses of Arabidopsis. Plant Mol. Biol. 62, 859–866. doi: 10.1007/s11103-006-9061-7 Majerova, E., Fojtova, M., Mozgova, I., Bittova, M., and Fajkus, J. (2011). Hypomethylating drugs efficiently decrease cytosine methylation in telomeric DNA and activate telomerase without affecting telomere lengths in tobacco cells. Plant Mol. Biol. 77, 371–380. doi: 10.1007/s11103-011-9816-7 Mandakova, T., and Lysak, M. A. (2008). Chromosomal phylogeny and karyotype evolution in x = 7 crucifer species (Brassicaceae). Plant Cell 20, 2559–2570. doi: 10.1105/tpc.108.062166 Marian, C. O., and Bass, H. W. (2005). The terminal acidic SANT 1 (Tacs1) gene of maize is expressed in tissues containing meristems and encodes an acidic SANT Frontiers in Plant Science | www.frontiersin.org 16 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins domain similar to some chromatin-remodeling complex proteins. Biochim. Biophys. Acta 1727, 81–86. doi: 10.1016/j.bbaexp.2004.12.010 Marian, C. O., Bordoli, S. J., Goltz, M., Santarella, R. A., Jackson, L. P., Danilevskaya, O., et al. (2003). The maize single myb histone 1 gene, Smh1, belongs to a novel gene family and encodes a protein that binds telomere DNA repeats in vitro. Plant Physiol. 133, 1336–1350. Martinez, P., Thanasoula, M., Carlos, A. R., Gomez-Lopez, G., Tejera, A. M., Schoeftner, S., et al. (2010). Mammalian Rap1 controls telomere function and gene expression through binding to telomeric and extratelomeric sites. Nat. Cell Biol. 12, 768–780. doi: 10.1038/ncb2081 Martinez-Perez, E., Shaw, P., Reader, S., Aragon-Alcaide, L., Miller, T., and Moore, G. (1999). Homologous chromosome pairing in wheat. J. Cell Sci. 112, 1761–1769. Mignon-Ravix, C., Depetris, D., Delobel, B., Croquette, M. F., and Mattei, M. G. (2002). A human interstitial telomere associates in vivo with specific TRF2 and TIN2 proteins. Eur. J. Hum. Genet. 10, 107–112. doi: 10.1038/sj.ejhg. 5200775 Mimori, T., Akizuki, M., Yamagata, H., Inada, S., Yoshida, S., and Homma, M. (1981). Characterization of a high molecular weight acidic nuclear protein recognized by autoantibodies in sera from patients with polymyositisscleroderma overlap. J. Clin. Invest. 68, 611–620. doi: 10.1172/JCI110295 Moore, J. (2009). Investigating the DNA Binding Properties of the Initiator Binding Protein 2 (IBP2) in Maize (Zea Mays). Master thesis, Florida State University, Tallahassee. Moriguchi, R., Kanahama, K., and Kanayama, Y. (2006). Characterization and expression analysis of the tomato telomere-binding protein LeTBP1. Plant Sci. 171, 166–174. doi: 10.1016/j.plantsci.2006.03.010 Morse, R. H. (2000). RAP, RAP, open up! New wrinkles for RAP1 in yeast. Trends Genet. 16, 51–53. Mozgova, I., Schrumpfova, P. P., Hofr, C., and Fajkus, J. (2008). Functional characterization of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. Phytochemistry 69, 1814–1819. doi: 10.1016/j.phytochem.2008.04.001 Najdekrova, L., and Siroky, J. (2012). NBS1 plays a synergistic role with telomerase in the maintenance of telomeres in Arabidopsis thaliana. BMC Plant Biol. 12:167. doi: 10.1186/1471-2229-12-167 Nelson, A. D. J. (2012). Telomerase regulation in Arabidopsis thaliana. Ph.D. thesis, A&M University, Texas. Nelson, A. D. J., Forsythe, E. S., Gan, X., Tsiantis, M., and Beilstein, M. A. (2014). Extending the model of Arabidopsis telomere length and composition across Brassicaceae. Chromosome Res. 22, 153–166. doi: 10.1007/s10577-014-9423-y Oguchi, K., Tamura, K., and Takahashi, H. (2004). Characterization of Oryza sativa telomerase reverse transcriptase and possible role of its phosphorylation in the control of telomerase activity. Gene 342, 57–66. doi: 10.1016/j.gene.2004.07.011 Olivier, M., Da Ines, O., Amiard, S., Serra, H., Goubely, C., White, C., et al. (2016)). The structure-specific endonucleases MUS81 and SEND1 ere essential for telomere stability in Arabidopsis. Plant Cell 28, 74–86. doi: 10.1105/tpc.15.00898 Ottaviani, A., Gilson, E., and Magdinier, F. (2008). Telomeric position effect: from the yeast paradigm to human pathologies? Biochimie 90, 93–107. doi: 10.1016/j.biochi.2007.07.022 Peska, V., Fajkus, P., Fojtova, M., Dvorackova, M., Hapala, J., Dvoracek, V., et al. (2015). Characterisation of an unusual telomere motif (TTTTTTAGGG)n in the plant Cestrum elegans (Solanaceae), a species with a large genome. Plant J. 82, 644–654. doi: 10.1111/tpj.12839 Peska, V., Schrumpfova, P. P., and Fajkus, J. (2011). Using the telobox to search for plant telomere binding proteins. Curr. Protein Pept. Sci. 12, 75–83. doi: 10.2174/138920311795684968 Peska, V., Sykorova, E., and Fajkus, J. (2008). Two faces of Solanaceae telomeres: a comparison between Nicotiana and Cestrum telomeres and telomere-binding proteins. Cytogenet. Genome. Res. 122, 380–387. doi: 10.1159/000167826 Phillips, D., Nibau, C., Wnetrzak, J., and Jenkins, G. (2012). High resolution analysis of meiotic chromosome structure and behaviour in barley (Hordeum vulgare L.). PLoS ONE 7:e39539. doi: 10.1371/journal.pone.0039539 Pisano, S., Galati, A., and Cacchione, S. (2008). Telomeric nucleosomes: forgotten players at chromosome ends. Cell Mol. Life. Sci. 65, 3553–3563. doi: 10.1007/s00018-008-8307-8 Podlevsky, J. D., Bley, C. J., Omana, R. V., Qi, X., and Chen, J. J. (2008). The telomerase database. Nucleic Acids Res. 36, D339–D343. doi: 10.1093/nar/gkm700 Price, C. M., Boltz, K. A., Chaiken, M. F., Stewart, J. A., Beilstein, M. A., and Shippen, D. E. (2010). Evolution of CST function in telomere maintenance. Cell Cycle 9, 3157–3165. doi: 10.4161/cc.9.16.12547 Price, C. M., and Cech, T. R. (1987). Telomeric DNA-protein interactions of Oxytricha macronuclear DNA. Genes Dev. 1, 783–793. doi: 10.1101/gad. 1.8.783 Qi, H., and Zakian, V. A. (2000). The Saccharomyces telomere-binding protein Cdc13p interacts with both the catalytic subunit of DNA polymerase alpha and the telomerase-associated est1 protein. Genes Dev. 14, 1777–1788. Rabl, C. (1885). Uber Zelltheilung. Morphologisches Jahrbuch 10, 214–330. Recker, J., Knoll, A., and Puchta, H. (2014). The Arabidopsis thaliana homolog of the helicase RTEL1 plays multiple roles in preserving genome stability. Plant Cell 26, 4889–4902. doi: 10.1105/tpc.114.132472 Ren, S., Johnston, J. S., Shippen, D. E., and McKnight, T. D. (2004). TELOMERASE ACTIVATOR1 induces telomerase activity and potentiates responses to auxin in Arabidopsis. Plant Cell 16, 2910–2922. doi: 10.1105/tpc.104.025072 Ren, S., Mandadi, K. K., Boedeker, A. L., Rathore, K. S., and McKnight, T. D. (2007). Regulation of telomerase in Arabidopsis by BT2, an apparent target of TELOMERASE ACTIVATOR1. Plant Cell 19, 23–31. doi: 10.1105/tpc.106.044321 Renfrew, K. B., Song, X., Lee, J. R., Arora, A., and Shippen, D. E. (2014). POT1a and components of CST engage telomerase and regulate its activity in Arabidopsis. PLoS Genet. 10:e1004738. doi: 10.1371/journal.pgen.1004738 Richards, E. J., and Ausubel, F. M. (1988). Isolation of a higher eukaryotic telomere from Arabidopsis thaliana. Cell 53, 127–136. doi: 10.1016/0092- 8674(88)90494-1 Riha, K., Fajkus, J., Siroky, J., and Vyskot, B. (1998). Developmental control of telomere lengths and telomerase activity in plants. Plant Cell 10, 1691–1698. doi: 10.2307/3870766 Riha, K., McKnight, T. D., Fajkus, J., Vyskot, B., and Shippen, D. E. (2000). Analysis of the G-overhang structures on plant telomeres: evidence for two distinct telomere architectures. Plant J. 23, 633–641. doi: 10.1046/j.1365- 313x.2000.00831.x Riha, K., Watson, J. M., Parkey, J., and Shippen, D. E. (2002). Telomere length deregulation and enhanced sensitivity to genotoxic stress in Arabidopsis mutants deficient in Ku70. EMBO J. 21, 2819–2826. doi: 10.1093/emboj/21.11.2819 Rossignol, P., Collier, S., Bush, M., Shaw, P., and Doonan, J. H. (2007). Arabidopsis POT1A interacts with TERT-V(I8), an N-terminal splicing variant of telomerase. J. Cell Sci. 120, 3678–3687. doi: 10.1242/jcs.004119 Rotkova, G., Sykorova, E., and Fajkus, J. (2009). Protect and regulate: Recent findings on plant POT1-like proteins. Biol. Plant. 53, 1–4. doi: 10.1007/s10535- 009-0001-7 Ruckova, E., Friml, J., Prochazkova Schrumpfova, P., and Fajkus, J. (2008). Role of alternative telomere lengthening unmasked in telomerase knock-out mutant plants. Plant Mol. Biol. 66, 637–646. doi: 10.1007/s11103-008-9295-7 Schmidt, J. C., and Cech, T. R. (2015). Human telomerase: biogenesis, trafficking, recruitment, and activation. Genes Dev. 29, 1095–1105. doi: 10.1101/gad.263863.115 Schober, H., Ferreira, H., Kalck, V., Gehlen, L. R., and Gasser, S. M. (2009). Yeast telomerase and the SUN domain protein Mps3 anchor telomeres and repress subtelomeric recombination. Genes Dev. 23, 928–938. doi: 10.1101/gad.1787509 Schrumpfova, P., Kuchar, M., Mikova, G., Skrisovska, L., Kubicarova, T., and Fajkus, J. (2004). Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. Genome 47, 316–324. doi: 10.1139/g03-136 Schrumpfova, P. P., Kuchar, M., Palecek, J., and Fajkus, J. (2008). Mapping of interaction domains of putative telomere-binding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. FEBS Lett. 582, 1400–1406. doi: 10.1016/j.febslet.2008.03.034 Schrumpfova, P. P., Vychodilova, I., Dvorackova, M., Majerska, J., Dokladal, L., Schorova, S., et al. (2014). Telomere repeat binding proteins are functional components of Arabidopsis telomeres and interact with telomerase. Plant J. 77, 770–781. doi: 10.1111/tpj.12428 Schrumpfova, P. P., Vychodilova, I., Hapala, J., Schorova, S., Dvoracek, V., and Fajkus, J. (2016). Telomere binding protein TRB1 is associated with promoters of translation machinery genes in vivo. Plant. Mol. Biol. 90, 189–206. doi: 10.1007/s11103-015-0409-8 Frontiers in Plant Science | www.frontiersin.org 17 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins Schwacke, R., Fischer, K., Ketelsen, B., Krupinska, K., and Krause, K. (2007). Comparative survey of plastid and mitochondrial targeting properties of transcription factors in Arabidopsis and rice. Mol. Genet. Genom. 277, 631–646. doi: 10.1007/s00438-007-0214-4 Sfeir, A. (2012). Telomeres at a glance. J. Cell Sci. 125, 4173–4178. doi: 10.1242/jcs.106831 Shakirov, E. V., McKnight, T. D., and Shippen, D. E. (2009a). POT1-independent single-strand telomeric DNA binding activities in Brassicaceae. Plant J. 58, 1004–1015. doi: 10.1111/j.1365-313X.2009.03837.x Shakirov, E. V., Perroud, P. F., Nelson, A. D., Cannell, M. E., Quatrano, R. S., and Shippen, D. E. (2010). Protection of telomeres 1 is required for telomere integrity in the moss Physcomitrella patens. Plant Cell 22, 1838–1848. doi: 10.1105/tpc.110.075846 Shakirov, E. V., Song, X., Joseph, J. A., and Shippen, D. E. (2009b). POT1 proteins in green algae and land plants: DNA-binding properties and evidence of co-evolution with telomeric DNA. Nucleic Acids Res. 37, 7455–7467. doi: 10.1093/nar/gkp785 Shakirov, E. V., and Shippen, D. E. (2004). Length regulation and dynamics of individual telomere tracts in wild-type Arabidopsis. Plant Cell 16, 1959–1967. doi: 10.1105/tpc.104.023093 Shakirov, E. V., Surovtseva, Y. V., Osbun, N., and Shippen, D. E. (2005). The Arabidopsis Pot1 and Pot2 proteins function in telomere length homeostasis and chromosome end protection. Mol. Cell. Biol. 25, 7725–7733. doi: 10.1128/MCB.25.17.7725-7733.2005 Simonet, T., Zaragosi, L. E., Philippe, C., Lebrigand, K., Schouteden, C., Augereau, A., et al. (2011). The human TTAGGG repeat factors 1 and 2 bind to a subset of interstitial telomeric sequences and satellite repeats. Cell Res. 21, 1028–1038. doi: 10.1038/cr.2011.40 Smith, S., Giriat, I., Schmitt, A., and de Lange, T. (1998). Tankyrase, a poly(ADP-ribose) polymerase at human telomeres. Science 282, 1484–1487. doi: 10.1126/science.282.5393.1484 Smogorzewska, A., van Steensel, B., Bianchi, A., Oelmann, S., Schaefer, M. R., Schnapp, G., et al. (2000). Control of human telomere length by TRF1 and TRF2. Mol. Cell. Biol. 20, 1659–1668. doi: 10.1128/MCB.20.5.1659-1668. 2000 Song, X., Leehy, K., Warrington, R. T., Lamb, J. C., Surovtseva, Y. V., and Shippen, D. E. (2008). STN1 protects chromosome ends in Arabidopsis thaliana. Proc. Natl. Acad. Sci. U.S.A. 105, 19815–19820. doi: 10.1073/pnas.08078 67105 Starr, D. A., Hermann, G. J., Malone, C. J., Fixsen, W., Priess, J. R., Horvitz, H. R., et al. (2001). unc-83 encodes a novel component of the nuclear envelope and is essential for proper nuclear migration. Development 128, 5039–5050. Stewart, J. A., Wang, F., Chaiken, M. F., Kasbek, C., Chastain, P. D. II, Wright, W. E., et al. (2012). Human CST promotes telomere duplex replication and general replication restart after fork stalling. EMBO J. 31, 3537–3549. doi: 10.1038/emboj.2012.215 Surovtseva, Y. V., Churikov, D., Boltz, K. A., Song, X., Lamb, J. C., Warrington, R., et al. (2009). Conserved telomere maintenance component 1 interacts with STN1 and maintains chromosome ends in higher eukaryotes. Mol. Cell. 36, 207–218. doi: 10.1016/j.molcel.2009.09.017 Surovtseva, Y. V., Shakirov, E. V., Vespa, L., Osbun, N., Song, X., and Shippen, D. E. (2007). Arabidopsis POT1 associates with the telomerase RNP and is required for telomere maintenance. EMBO J. 26, 3653–3661. doi: 10.1038/sj.emboj.7601792 Sykorova, E., and Fajkus, J. (2009). Structure-function relationships in telomerase genes. Biol. Cell 101, 375–392. doi: 10.1042/BC20080205. Sykorova, E., Fulneckova, J., Mokros, P., Fajkus, J., Fojtova, M., and Peska, V. (2012). Three TERT genes in Nicotiana tabacum. Chromosome Res. 20, 381–394. doi: 10.1007/s10577-012-9282-3 Sykorova, E., Leitch, A. R., and Fajkus, J. (2006). Asparagales telomerases which synthesize the human type of telomeres. Plant Mol. Biol. 60, 633–646. doi: 10.1007/s11103-005-5091-9 Sykorova, E., Lim, K. Y., Kunicka, Z., Chase, M. W., Bennett, M. D., Fajkus, J., et al. (2003). Telomere variability in the monocotyledonous plant order Asparagales. Proc. Biol. Sci. 270, 1893–1904. doi: 10.1098/rspb.2003.2446 Tamura, K., Goto, C., and Hara-Nishimura, I. (2015). Recent advances in understanding plant nuclear envelope proteins involved in nuclear morphology. J. Exp. Bot. 66, 1641–1647. doi: 10.1093/jxb/erv036 Tamura, K., Liu, H., and Takahashi, H. (1999). Auxin induction of cell cycle regulated activity of tobacco telomerase. J. Biol. Chem. 274, 20997–21002. doi: 10.1074/jbc.274.30.20997 Tani, A., and Murata, M. (2005). Alternative splicing of Pot1 (Protection of telomere)-like genes in Arabidopsis thaliana. Genes Genet. Syst. 80, 41–48. doi: 10.1266/ggs.80.41 Taylor, J. S., and Raes, J. (2004). Duplication and divergence: the evolution of new genes and old ideas. Annu. Rev. Genet. 38, 615–643. doi: 10.1146/annurev.genet.38.072902.092831 Tiang, C. L., He, Y., and Pawlowski, W. P. (2012). Chromosome organization and dynamics during interphase, mitosis, and meiosis in plants. Plant Physiol. 158, 26–34. doi: 10.1104/pp.111.187161 Ting, N. S., Yu, Y., Pohorelic, B., Lees-Miller, S. P., and Beattie, T. L. (2005). Human Ku70/80 interacts directly with hTR, the RNA component of human telomerase. Nucleic Acids Res. 33, 2090–2098. doi: 10.1093/nar/gki342 To, T. K., Kim, J. M., Matsui, A., Kurihara, Y., Morosawa, T., Ishida, J., et al. (2011). Arabidopsis HDA6 regulates locus-directed heterochromatin silencing in cooperation with MET1. PLoS Genet. 7:e1002055. doi: 10.1371/journal.pgen.1002055 Tomaska, L., Nosek, J., Kramara, J., and Griffith, J. D. (2009). Telomeric circles: universal players in telomere maintenance? Nat. Struct. Mol. Biol. 16, 1010– 1015. doi: 10.1038/nsmb.1660 Tran, D. T., Cao, H. X., Jovtchev, G., Neumann, P., Novak, P., Fojtova, M., et al. (2015). Centromere and telomere sequence alterations reflect the rapid genome evolution within the carnivorous plant genus Genlisea. Plant J. 84, 1087–1099. doi: 10.1111/tpj.13058 Uringa, E. J., Youds, J. L., Lisaingo, K., Lansdorp, P. M., and Boulton, S. J. (2011). RTEL1: an essential helicase for telomere maintenance and the regulation of homologous recombination. Nucleic Acids Res. 39, 1647–1655. doi: 10.1093/nar/gkq1045 van der Fits, L., Zhang, H., Menke, F. L., Deneka, M., and Memelink, J. (2000). A Catharanthus roseus BPF-1 homologue interacts with an elicitor-responsive region of the secondary metabolite biosynthetic gene Str and is induced by elicitor via a JA-independent signal transduction pathway. Plant Mol. Biol. 44, 675–685. doi: 10.1023/A:1026526522555 Vannier, J. B., Depeiges, A., White, C., and Gallego, M. E. (2009). ERCC1/XPF protects short telomeres from homologous recombination in Arabidopsis thaliana. PLoS Genet. 5:e1000380. doi: 10.1371/journal.pgen.1000380 Vannier, J. B., Pavicic-Kaltenbrunner, V., Petalcorin, M. I., Ding, H., and Boulton, S. J. (2012). RTEL1 dismantles T loops and counteracts telomeric G4-DNA to maintain telomere integrity. Cell 149, 795–806. doi: 10.1016/j.cell.2012.03.030 Venteicher, A. S., Meng, Z., Mason, P. J., Veenstra, T. D., and Artandi, S. E. (2008). Identification of ATPases pontin and reptin as telomerase components essential for holoenzyme assembly. Cell 132, 945–957. doi: 10.1016/j.cell.2008.01.019 Vespa, L., Warrington, R. T., Mokros, P., Siroky, J., and Shippen, D. E. (2007). ATM regulates the length of individual telomere tracts in Arabidopsis. Proc. Natl. Acad. Sci. U.S.A. 104, 18145–18150. doi: 10.1073/pnas.070 4466104 Vrbsky, J., Akimcheva, S., Watson, J. M., Turner, T. L., Daxinger, L., Vyskot, B., et al. (2010). siRNA-mediated methylation of Arabidopsis telomeres. PLoS Genet. 6:e1000986. doi: 10.1371/journal.pgen.1000986 Wang, H., Liu, C., Cheng, J., Liu, J., Zhang, L., He, C., et al. (2016). Arabidopsis flower and embryo developmental genes are repressed in seedlings by different combinations of polycomb group proteins in association with distinct sets of cis-regulatory elements. PLoS Genet. 12:e1005771. doi: 10.1371/journal.pgen.1005771 Wang, Y., Ghosh, G., and Hendrickson, E. A. (2009). Ku86 represses lethal telomere deletion events in human somatic cells. Proc. Natl. Acad. Sci. U.S.A. 106, 12430–12435. doi: 10.1073/pnas.0903362106 Watson, J. M., and Shippen, D. E. (2007). Telomere rapid deletion regulates telomere length in Arabidopsis thaliana. Mol. Cell. Biol. 27, 1706–1715. doi: 10.1128/MCB.02059-06 Wei, C., and Price, C. M. (2004). Cell cycle localization, dimerization, and binding domain architecture of the telomere protein cPot1. Mol. Cell. Biol. 24, 2091– 2102. doi: 10.1128/MCB.24.5.2091-2102.2004 Weiss, H., and Scherthan, H. (2002). Aloe spp. - plants with vertebratelike telomeric sequences. Chromosome Res. 10, 155–164. doi: 10.1023/A:1014905319557 Frontiers in Plant Science | www.frontiersin.org 18 June 2016 | Volume 7 | Article 851 Procházková Schrumpfová et al. Telomere- and Telomerase-Associated Proteins Wellinger, R. J., and Zakian, V. A. (2012). Everything you ever wanted to know about Saccharomyces cerevisiae telomeres: beginning to end. Genetics 191, 1073–1105. doi: 10.1534/genetics.111.137851 Wen, R., Moore, G., and Shaw, P. J. (2012). Centromeres cluster de novo at the beginning of meiosis in Brachypodium distachyon. PLoS ONE 7:e44681. doi: 10.1371/journal.pone.0044681 Wu, L., Multani, A. S., He, H., Cosme-Blanco, W., Deng, Y., Deng, J. M., et al. (2006). Pot1 deficiency initiates DNA damage checkpoint activation and aberrant homologous recombination at telomeres. Cell 126, 49–62. doi: 10.1016/j.cell.2006.05.037 Yang, D., Xiong, Y., Kim, H., He, Q., Li, Y., Chen, R., et al. (2011). Human telomeric proteins occupy selective interstitial sites. Cell Res. 21, 1013–1027. doi: 10.1038/cr.2011.39 Yang, S. W., Jin, E., Chung, I. K., and Kim, W. T. (2002). Cell cycledependent regulation of telomerase activity by auxin, abscisic acid and protein phosphorylation in tobacco BY-2 suspension culture cells. Plant J. 29, 617–626. doi: 10.1046/j.0960-7412.2001.01244.x Yang, S. W., Kim, D. H., Lee, J. J., Chun, Y. J., Lee, J. H., Kim, Y. J., et al. (2003). Expression of the telomeric repeat binding factor gene NgTRF1 is closely coordinated with the cell division program in tobacco BY-2 suspension culture cells. J. Biol. Chem. 278, 21395–21407. doi: 10.1074/jbc.M209973200 Yang, S. W., Kim, S. K., and Kim, W. T. (2004). Perturbation of NgTRF1 expression induces apoptosis-like cell death in tobacco BY-2 cells and implicates NgTRF1 in the control of telomere length and stability. Plant Cell 16, 3370–3385. doi: 10.1105/tpc.104.026278 Ye, J., Renault, V. M., Jamet, K., and Gilson, E. (2014). Transcriptional outcome of telomere signalling. Nat. Rev. Genet. 15, 491–503. doi: 10.1038/nrg3743 Yoo, H. H., Kwon, C., Lee, M. M., and Chung, I. K. (2007). Single-stranded DNA binding factor AtWHY1 modulates telomere length homeostasis in Arabidopsis. Plant J. 49, 442–451. doi: 10.1111/j.1365-313X.2006.02974.x Yu, E. Y., Kim, S. E., Kim, J. H., Ko, J. H., Cho, M. H., and Chung, I. K. (2000). Sequence-specific DNA recognition by the Myb-like domain of plant telomeric protein RTBP1. J. Biol. Chem. 275, 24208–24214. doi: 10.1074/jbc.M003250200 Yun, J. H., Lee, W. K., Kim, H., Kim, E., Cheong, C., Cho, M. H., et al. (2014). Solution structure of telomere binding domain of AtTRB2 derived from Arabidopsis thaliana. Biochem. Biophys. Res. Commun. 452, 436–442. doi: 10.1016/j.bbrc.2014.08.095 Zachova, D., Fojtova, M., Dvorackova, M., Mozgova, I., Lermontova, I., Peska, V., et al. (2013). Structure-function relationships during transgenic telomerase expression in Arabidopsis. Physiol. Plant. 149, 114–126. doi: 10.1111/ppl.12021 Zellinger, B., Akimcheva, S., Puizina, J., Schirato, M., and Riha, K. (2007). Ku suppresses formation of telomeric circles and alternative telomere lengthening in Arabidopsis. Mol. Cell. 27, 163–169. doi: 10.1016/j.molcel.2007. 05.025 Zentgraf, U. (1995). Telomere-binding proteins of Arabidopsis thaliana. Plant Mol. Biol. 27, 467–475. doi: 10.1007/BF00019314 Zhang, P., Pazin, M. J., Schwartz, C. M., Becker, K. G., Wersto, R. P., Dilley, C. M., et al. (2008). Nontelomeric TRF2-REST interaction modulates neuronal gene silencing and fate of tumor and stem cells. Curr. Biol. 18, 1489–1494. doi: 10.1016/j.cub.2008.08.048 Zhou, X., Graumann, K., Wirthmueller, L., Jones, J. D., and Meier, I. (2014). Identification of unique SUN-interacting nuclear envelope proteins with diverse functions in plants. J. Cell Biol. 205, 677–692. doi: 10.1083/jcb.201401138 Zhou, Y., Hartwig, B., James, G. V., Schneeberger, K., and Turck, F. (2016). Complementary activities of TELOMERE REPEAT BINDING proteins and polycomb group complexes in transcriptional regulation of target genes. Plant Cell 28, 87–101. doi: 10.1105/tpc.15. 00787 Zhu, Q. H., Ramm, K., Shivakkumar, R., Dennis, E. S., and Upadhyaya, N. M. (2004). The ANTHER INDEHISCENCE1 gene encoding a single MYB domain protein is involved in anther development in rice. Plant Physiol. 135, 1514–1525. doi: 10.1104/pp.104.041459 Conflict of Interest Statement: The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. The handling Editor declared a current collaboration as co-Topic Editor in a Frontiers Research Topic with one of the authors, JF, and states that the process nevertheless met the standards of a fair and objective review. This was also confirmed by the Specialty Chief Editor of section Plant Cell Biology, Simon Gilroy. Copyright © 2016 Procházková Schrumpfová, Schoˇrová and Fajkus. This is an openaccess article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms. Frontiers in Plant Science | www.frontiersin.org 19 June 2016 | Volume 7 | Article 851 Supplement K Schrumpfová, P.P., Majerská, J., Dokládal, L., Schořová, Š., Stejskal, K., Obořil, M., Honys, D., Kozáková, L., Polanská, P.S., Sýkorová, E., 2017. Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo. Protoplasma 254, 1547–1562 P.P.S. participated in the design of experiments, was significantly involved in the experimental part, data evaluation and participated in the ms writing This journal did not provide open access, hence the article is not freely available. Supplement L Schrumpfová, P.P., Majerská, J., Dokládal, L., Schořová, Š., Stejskal, K., Obořil, M., Honys, D., Kozáková, L., Polanská, P.S., Sýkorová, E., 2018. Correction to: Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo. Protoplasma 255, 715 In the published online version, P.P. Schrumpfová was not recognized as the first co-author, and the affiliations were mixed up. Corrections were made to both the first co-authorship and the affiliation section. CORRECTION Correction to: Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo Jana Majerská1,2,3 & Petra Procházková Schrumpfová2 & Ladislav Dokládal1 & Šárka Schořová2 & Karel Stejskal2 & Michal Obořil2 & David Honys4 & Lucie Kozáková2 & Pavla Sováková Polanská2 & Eva Sýkorová1 Published online: 14 February 2018 # Springer-Verlag GmbH Austria, part of Springer Nature 2018 Correction to: Protoplasma (2017) 254:1547–1562 https://doi.org/10.1007/s00709-016-1042-3 In the published online version, the affiliations were mixed up. Corrected affiliation section is shown below. Also, the update has also been reflected in the author group section above. Jana Majerská and Petra Procházková Schrumpfová are both co-authors The online version of the original article can be found at https://doi.org/ 10.1007/s00709-016-1042-3 * Eva Sýkorová evin@ibp.cz 1 Institute of Biophysics, Academy of Sciences of the Czech Republic, v.v.i, Královopolská 135, CZ-61265 Brno, Czech Republic 2 Central European Institute of Technology and Faculty of Science, Masaryk University, Kotlářská 2, CZ-61137 Brno, Czech Republic 3 Present address: Swiss Institute for Experimental Cancer Research (ISREC), School of Life Sciences, École Polytechnique Fédérale de Lausanne (EPFL), 1015 Lausanne, Switzerland 4 Institute of Experimental Botany, Academy of Sciences of the Czech Republic v.v.i, Rozvojová 263, CZ-16502 Prague, Czech Republic Protoplasma (2018) 255:715 https://doi.org/10.1007/s00709-018-1224-2 Supplement M Schrumpfová, P.P., Fojtová, M., Fajkus, J., 2019. Telomeres in Plants and Humans: Not So Different, Not So Similar. Cells 8 P.P.S. was significantly involved in the ms writing and editing cells Review Telomeres in Plants and Humans: Not So Different, Not So Similar Petra Procházková Schrumpfová 1, Miloslava Fojtová 1,2,3 and Jiˇrí Fajkus 1,2,3,* 1 Laboratory of Functional Genomics and Proteomics, National Centre for Biomolecular Research, Faculty of Science, Masaryk University, CZ-61137 Brno, Czech Republic; petra.proch.schrumpfova@gmail.com (P.P.S.); miloslava.fojtova@gmail.com (M.F.) 2 Mendel Centre for Plant Genomics and Proteomics, CEITEC, Masaryk University, CZ-62500 Brno, Czech Republic 3 Institute of Biophysics of the Czech Academy of Sciences, CZ-61265 Brno, Czech Republic * Correspondence: tttaggg@gmail.com; Tel.: +420-5494-94003 Received: 17 December 2018; Accepted: 7 January 2019; Published: 16 January 2019 Abstract: Parallel research on multiple model organisms shows that while some principles of telomere biology are conserved among all eukaryotic kingdoms, we also find some deviations that reflect different evolutionary paths and life strategies, which may have diversified after the establishment of telomerase as a primary mechanism for telomere maintenance. Much more than animals, plants have to cope with environmental stressors, including genotoxic factors, due to their sessile lifestyle. This is, in principle, made possible by an increased capacity and efficiency of the molecular systems ensuring maintenance of genome stability, as well as a higher tolerance to genome instability. Furthermore, plant ontogenesis differs from that of animals in which tissue differentiation and telomerase silencing occur during early embryonic development, and the “telomere clock” in somatic cells may act as a preventive measure against carcinogenesis. This does not happen in plants, where growth and ontogenesis occur through the serial division of apical meristems consisting of a small group of stem cells that generate a linear series of cells, which differentiate into an array of cell types that make a shoot and root. Flowers, as generative plant organs, initiate from the shoot apical meristem in mature plants which is incompatible with the human-like developmental telomere shortening. In this review, we discuss differences between human and plant telomere biology and the implications for aging, genome stability, and cell and organism survival. In particular, we provide a comprehensive comparative overview of telomere proteins acting in humans and in Arabidopsis thaliana model plant, and discuss distinct epigenetic features of telomeric chromatin in these species. Keywords: telomere; telomerase; human; Arabidopsis; aging; chromatin; epigenetics; review 1. Introduction Telomere biology, whose foundations were laid out in maize and Drosophila at the end of the 1930s and which developed at the molecular level in the 1980s, has flourished enourmously in the last 30 years. This interest in telomere biology follows from the generally attractive links between telomere functions, cell aging mechanisms, and the genesis of severe diseases in humans. Research in recent decades has elucidated the principles of protection of the ends of linear eukaryotic chromosomes from progressive shortening due to the incomplete replication (end-replication problem) [1] and from their erroneous recognition as unrepaired chromosome breaks (end-protection problem) [2–4]. In addition to these basic functions, other potential roles of telomeres have been suggested, such as a trap for reactive oxygen species [5,6]. Telomeres are composed of non-coding repetitive tandem repeats of (TTAGGG)n in humans and the other vertebrates, and (TTTAGGG)n in most plants. During human Cells 2019, 8, 58; doi:10.3390/cells8010058 www.mdpi.com/journal/cells Cells 2019, 8, 58 2 of 31 aging, telomeres in most somatic cells are shortened at each cell division and it is generally assumed that when telomeres reach a critical length, cells enter a senescent state and cell division ceases [7,8]. However, most human individuals do not reach this critical telomere length brink during their life course [8,9], e.g., the mean leukocyte telomere length (LTL) in newborns is 9.5 kb [10] whereas a length of ~5 kb was defined as the ‘telomeric brink’, which denotes a high risk of imminent death, but only 0.78% of people younger than 90 years display an LTL ≤ 5 kb [9]. So it is obvious, that the link between shortened telomeres and human longevity is more complex than mere reaching the critical telomere length. For instance, age-dependent telomere shortening might alter gene expression in sub-telomeric regions (telomere position effect, TPE) or double strand DNA breaks in telomeres might be inefficiently repaired and initiate cell senescence [11,12]. Furthermore, it has been suggested that even a single critically short telomere in a cell can induce cellular senescence, which potentially contributes to organismal senescence [13,14]. In humans, five short telomeres were reported to predict the onset of cell senescence [15]. Although the principles of protection and replication of telomeres are conserved and point to common evolutionary roots of eukaryotes, their implications for cell and organism survival, senescence, and aging are not shared among kingdoms. In particular, plants show specific features of their growth and development, which lead to confusion of terms like lifespan or aging as commonly used and understood in animals. First, a plant’s body plan is not fully established during embryogenesis and all tissues and organs are formed from proliferating meristem cells throughout the adult life. Second, plant growth is modular. Individual modules of the body (branches, flowers, leaves) are dispensable for survival, and their functions can be replaced by tissues newly differentiated from indefinitely proliferating meristems. This results in the enormous developmental plasticity of plants. Moreover, the vegetative meristems can give rise to a new organism, which will be a somatic clone, genetically indistinguishable from the parental organism. Since these general aspects distinguishing plant from animal development and aging have been well-reviewed [16], we will focus here on a more detailed view of peculiarities of plant telomere biology, including its latest developments. 2. Telomerase Core Components The requirement to finish the incomplete replication of chromosome ends is common for all organisms with linear chromosomes. In eukaryotes, this requirement is commonly solved by a specific nucleoprotein enzyme complex called telomerase, which is considered as an ancestral telomere maintenance system that solves the end-replication problem of linear chromosomes. In humans, telomerase activity is detected in all early developmental stages from oocytes through to blastocyst stage embryos, and increases progressively with advancing embryo stage. Telomerase reaches its highest level in morula and blastocyst stage embryos and then decreases in the inner cell mass stage. In human fetuses—when the embryonic period and organogenesis are finished—telomerase is expressed in tissue-specific stem cells. However, just after birth, telomerase activity in somatic cells is downregulated with the exception of dividing cells (e.g., proliferating cells, T-lymphocytes) [17,18] (Figure 1A). As seen in mammals, telomeres in plants are maintained by telomerase [19]. Active telomerase is detected in organs and tissues containing highly dividing meristem cells such as seedlings, root tips, young and middle-age leaves, flowers, and floral buds [20,21]. In terminally differentiated tissues (stems, mature leaves), telomerase activity is suppressed (Figure 1B). In some groups of organisms (in particular insects), telomerase has been lost and replaced by telomere-specific retrotransposons (in Drosophila) or tandem arrays of satellite repeats elongated by a gene conversion mechanism (reviewed in References [22,23]). Based on a long-term systematic search, no telomerase-independent exception has been found among vertebrates or land plants despite the variability of telomere DNA observed in land plants [24–27]. Besides the telomerase-based mechanism of telomere elongation, alternative lengthening of telomeres (ALT), which is based on homologous recombination (HR) Cells 2019, 8, 58 3 of 31 and may become active upon the loss of telomerase was described in humans as well as in plants (see below).Cells 2019, 8, x FOR PEER REVIEW 3 of 31 Figure 1. Telomerase activity in human and plant tissues. (A) During human embryonic development, high telomerase activity is detected in the blastocyst, but not in mature spermatozoa or oocytes. Highly active telomerase is detected in 16 to 20-week-old human fetuses in most somatic tissues with the exception of brain tissue [18,28]. In adults, low telomerase activity is detected in hair follicule bulbs [29], basal cells of crypt and villi or muconasal basal cells of the gastrointestinal tract, basal keratinocytes of the skin [30], lymphocytes, blood bone marrow, and stem cells [31–33], and urothelium [34]. High telomerase activity is detected in prostate tissues and endometrium [30,35]. (B) High telomerase activity is detected in plant pollen, seedling, young rosette leaves, and silliques [21,36–39]. Likewise, both apical meristems—shoot and root—show high telomerase activity [36–38]. Figures adopted from human and Arabidopsis eFP browsers [40]. In yeasts, animals, and plants, telomerase consists of the telomerase reverse transcriptase (TERT) protein subunit providing the catalytic activity, and the telomerase RNA (TR) subunit whose short region provides a template for reverse transcription [41,42]. Besides these two core subunits, the telomerase complex comprises several other accessory proteins with diverse roles in telomerase assembly, trafficking, localization, recruitment to telomeres, or the processivity of telomere synthesis [43,44]. During movement of the maturing human telomerase complex through the nucleolus to Cajal bodies and to the telomeres, the TERT catalytical subunit is associated with e.g., HSP90, p23, or pontin. Assembly of human TR, as well as other box C/D or H/ACA small nucleolar RNAs (snoRNAs), is governed by conserved scaffold proteins: dyskerin, NHP2, NOP10, NAF1 in the nucleoplasm, where NAF1 is replaced by GAR1 before the hTR RNP complex reaches the nucleolus. Several orthologues of these conserved scaffold have been identified in plants, e.g., CBF5 (dyskerin), RuvBL1 (pontin), RuvBL2a (reptin), and NAF1. The nucleolar localization of these orthologues suggests potential conservation of the trafficking pathway during telomerase maturation ([45–47]; Schorova et al., submitted). Human and plant homologues of proteins associated either with the telomerase protein subunit TERT (Table 1) or the telomerase RNA subunit (Table 2) are listed below. Figure 1. Telomerase activity in human and plant tissues. (A) During human embryonic development, high telomerase activity is detected in the blastocyst, but not in mature spermatozoa or oocytes. Highly active telomerase is detected in 16 to 20-week-old human fetuses in most somatic tissues with the exception of brain tissue [18,28]. In adults, low telomerase activity is detected in hair follicule bulbs [29], basal cells of crypt and villi or muconasal basal cells of the gastrointestinal tract, basal keratinocytes of the skin [30], lymphocytes, blood bone marrow, and stem cells [31–33], and urothelium [34]. High telomerase activity is detected in prostate tissues and endometrium [30,35]. (B) High telomerase activity is detected in plant pollen, seedling, young rosette leaves, and silliques [21,36–39]. Likewise, both apical meristems—shoot and root—show high telomerase activity [36–38]. Figures adopted from human and Arabidopsis eFP browsers [40]. In yeasts, animals, and plants, telomerase consists of the telomerase reverse transcriptase (TERT) protein subunit providing the catalytic activity, and the telomerase RNA (TR) subunit whose short region provides a template for reverse transcription [41,42]. Besides these two core subunits, the telomerase complex comprises several other accessory proteins with diverse roles in telomerase assembly, trafficking, localization, recruitment to telomeres, or the processivity of telomere synthesis [43,44]. During movement of the maturing human telomerase complex through the nucleolus to Cajal bodies and to the telomeres, the TERT catalytical subunit is associated with e.g., HSP90, p23, or pontin. Assembly of human TR, as well as other box C/D or H/ACA small nucleolar RNAs (snoRNAs), is governed by conserved scaffold proteins: dyskerin, NHP2, NOP10, NAF1 in the nucleoplasm, where NAF1 is replaced by GAR1 before the hTR RNP complex reaches the nucleolus. Several orthologues of these conserved scaffold have been identified in plants, e.g., CBF5 (dyskerin), RuvBL1 (pontin), RuvBL2a (reptin), and NAF1. The nucleolar localization of these orthologues suggests potential conservation of the trafficking pathway during telomerase maturation ([45–47]; Schorova et al., submitted). Human and plant homologues of proteins associated either with the telomerase protein subunit TERT (Table 1) or the telomerase RNA subunit (Table 2) are listed below. Cells 2019, 8, 58 4 of 31 Table 1. Comparative overview of proteins associated with the telomerase catalytic subunit TERT. Telomerase Calytic Subunit (TERT) Associated Proteins. Human TERT Associated Proteins Protein Function and Direct Interactions References Arabidopsis TERT Associated Proteins Protein Function and Direct Interactions References TERT Catalytic subunit of telomerase [48] TERT Catalytic subunit of telomerase [49] POT1 Shelterin. Int.: telomeric ssDNA, TPP1 and CTC1. [50–54] POT1a Shelterin-like. Int.: TERT, telomeric ssDNA, TER1, TRFL9, CBF5, RuvBL1, CTC1 and STN1. [47,55–58] TRF1 Shelterin. Int.: telomeric dsDNA, TIN2, TANK1, PINX1, and ATM. [59–63] TRB1-3 Shelterin-like. Int.: TERT, telomeric dsDNA, POT1b, RuvBL1 and RuvBL2a. [64–71] TRF2 Shelterin. Int.: telomeric dsDNA; TIN2, NBS1, RAD50, Apollo, Ku70, PARP1, XPF-ERCC1, BLM, FEN1, POLB, ORC, RTEL1, ATM and HP1. [61,72–80] TRP1 Possible non-telomeric functions of telomerase. Int.: TERT, telomere dsDNA in vitro, ARM, Ku70 and TRFL9. [66,69,81–83] TRFL2 Possible non-telomeric functions of telomerase. Int.: TERT, telomere dsDNA in vitro and ARM. [69,83] TRFL11 Associates with TERT. [84] KPNA1 Promotes nuclear import of the TERT. [85] ImpA4 Associates with TERT. [84] NCL Involves nucleolar localization of TERT. [86] NUC-L1 Role in telomere maintenance and telomere clustering. [87,88] pontin Telomerase assembly. Int.: TERT and dyskerin. [89] RuvBL1 Associates with TERT via TRBs, regulates telomerase activity. [84,90] reptin Telomerase assembly. Int.: dyskerin. [89] RuvBL2a Associates with TERT via TRBs, regulates telomerase activity. [84] ARMC6 Int.: TRF2, telomerase. [69,91] ARM May reflect posible non-telomeric functions of telomerase. Int.: TERT, TRP1, TRFL2, TRFL9 and CHR19. [69,92] TPP1 Shelterin, mediates telomerase recruitment. Int.: TERT, POT1, TIN2, CTC1 and STN1. [51–54,75] n.a. Cells 2019, 8, 58 5 of 31 Table 1. Cont. PINX1 Potent telomerase inhibitor. Int.: TERT and TRF1. [62] n.a. HOT1 Int.: telomeric dsDNA, active telomerase. [93] n.a. Ku70/80 Int.: TERT, TR, TRF2 and RAP1. [94,95] Ku70/80 Role in telomere length regulation, may protect blunt-ended telomeres Int.: TRP1, TER2 and TER2s. [82,96–101] Hsp90 TERT assembly. Int.: TERT. [102] Hsp90 NP_194150.1 [103] p23 TERT assembly. Int.: TERT. [102] p23 CAC16575, NP_683525 [104] Purα p.h. Unwinds dsDNA telomeric oligonucleotides. [105] PURα1 Associates with TERT. [84] SMARCAD1 p.h. SWI/SNF-like protein that presumably associates with telomeres. [106,107] CHR19 May reflect possible non-telomeric functions of telomerase. Int.: TERT, ARM, TRB1 and TRFL9. [69] PABPN1 Promotes poly(A)-dependent TR 3 end maturation. [108] RRM Associates with TERT. [92,109] MT2A p.h. Int.: HOT1. [110,111] MT2A Associates with TERT. [84,109] PA2G4 NP_006182.2 [112] G2p Associates with TERT. [84,109] The proteins depicted in grey are involved in telomere maintenance, however, their association with telomerase has not been described. The proteins in green are structural homologous to their human/plant counterparts, however, any involvement in telomere maintenance or association with telomerase has not been described so far. Direct interaction partners (Int.) of TERT-associated proteins are enumerated. Cases with not yet identified sequence homologues are denoted with n.a. ATP-dependent DNA helicase 2 subunit 1 and 2 (Ku70/80); Origin recognition complex (ORC); RuvB-like 2 (reptin); TIN2- and POT1-organizing protein (TPP1); TRF1-interacting nuclear protein 2 (TIN2); TRF1-interacting protein 1 (PINX1); 5 exonuclease Apollo (Apollo); Armadillo repeat-containing protein 6 (ARMC6); Armadillo/β-catenin-like repeat-containing protein (ARM); Ataxia telangiectasia mutated kinase (ATM); Bloom syndrome protein (BLM); Centromere-binding factor (CBF5); Conserved telomere maintenance component 1 (CTC1); DNA polymerase beta (POLB); DNA repair protein RAD50 (RAD50); Double strand DNA (dsDNA); Excision repair cross-complementation 1 (ERCC1); Flap endonuclease 1 (FEN1); H/ACA ribonucleoprotein complex subunit DKC1 (dyskerin); Heterochromatin protein 1 (HP1); Homeobox telomere-binding protein 1 (HOT1); Hsp90 co-chaperone (p23); Chromatin remodeling 19 (CHR19); Importin-α5 (KPNA1); Importin subunit alpha-4 (ImpA4); Metallothionein-like 2A (MT2A); Nijmegen breakage syndrome protein 1 (NBS1); Nucleolin (NCL); Nucleolin like 1 (NUC-L1); Heat shock protein HSP 90 (Hsp90); Poly(ADP-ribose) polymerase 1 (PARP1); Polyadenylate-binding protein (PABPN1); Proliferation-associated 2G4 (PA2G4); Proliferation-associated protein (G2p); Protection of telomeres 1 (POT1); Protection of telomeres 1a, b (POT1a, b); Pur-alpha 1 (Purα1); Regulator of telomere elongation helicase 1 (RTEL1); RNA recognition motif (RRM); RuvB-like 1 (pontin); RuvB-like 1, 2a (RuvBL1, 2a); Single strand DNA (ssDNA); Suppressor of cdc thirteen homolog (STN1); SWI/SNF-related matrix-associated actin-dependent regulator of chromatin subfamily A containing DEAD/H box 1 (SMARCAD1); Tankyrase 1 (TANK1); Telomerase reverse transcriptase (TERT); Telomerase RNA (TR); Telomerase RNA subunit 1 (TER1); Telomere repeat-binding factor 1, 2, 3 (TRB1, 2, 3); Telomere repeat-binding protein 1 (TRP1); Telomeric repeat binding factor 1-like 2, 9, 11 (TRFL 2, 9, 11); Telomeric repeat-binding factor 1, 2 (TRF1, 2); Xeroderma pigmentosum group F (XPF1); putative homolog according to NCBI blastp (p.h.). Cells 2019, 8, 58 6 of 31 Table 2. Comparative overview of proteins associated with the RNA component of telomerase. Telomerase RNA Associated Proteins Human TR Associated Proteins Protein Function and Direct Interactions References Arabidopsis TR Associated Proteins Protein Function and Direct Interactions References TR RNA subunit of telomerase [113] TER1, TER2, TER2s Putative RNA subunit of telomerase [56,100] TERT Catalytic subunit of telomerase [48,114] TERT Catalytic subunit of telomerase [100] Dyskerin H/ACA snoRNPs, associated with nucleolus. Int.: TR, GAR1, NHP2, NOP10 and TCAB1. [44,115] CBF5 H/ACA snoRNPs, Ath orthologue of Dyskerin, associated with nucleolus, subnuclear bodies and Cajal bodies, associated with telomerase RNP complex. Direct interaction with either of putative TERs not demonstrated. Int.: NAF1. [45,57] NOP10 H/ACA snoRNPs, associates with nucleolus. Int.: TR and dyskerin. [44,116] NOP10 H/ACA snoRNPs, Ath orthologue of NOP10, associates with nucleolus. [45,46] NHP2 H/ACA snoRNPs, associates with nucleolus. Int.: TR, dyskerin and TCAB1. [117,118] NHP2 H/ACA snoRNPs, Ath orthologue of NHP2, associates with nucleolus. [45,46] GAR1 H/ACA snoRNPs, associated with nucleolus. Int.: dyskerin and TCAB1. [44,118] GAR1, 2 H/ACA snoRNPs, Ath orthologues of GAR1, associate with nucleolus. [45,46] NAF1 H/ACA snoRNPs, nucleolar shuttle NAF1 is substituted by GAR1 during maturation of telomerase. Int.: dyskerin. [119] NAF1 H/ACA snoRNPs, Ath orthologue of NAF1, associates with nucleolus and Cajal bodies. Int.: CBF5. [45] Ku70/80 Int.: TR, TERT, TRF2 and RAP1. [95,120] Ku70/80 Role in telomere length regulation, may protect blunt-ended telomeres Int.: TRP1, TER2 and TER2s. [100] Cells 2019, 8, 58 7 of 31 Table 2. Cont. pontin Telomerase assembly. Int.: TERT and dyskerin. [89] RuvBL1 Associates with TERT via TRBs, regulates telomerase activity. Schorova et al., submitted reptin Telomerase assembly. Int.: dyskerin. [89] RuvBL2a Associates with TERT via TRBs, regulates telomerase activity. Schorova et al., submitted RHAU RNA helicase, unwinds a G4-quadruplex in human telomerase RNA. Int.: TR. [121] RHAU NP_850255.1, NP_175298.2, NP_680142.2, NP_178223.2 n.a. PARN Poly(A)-specific ribonuclease, 3 -end maturation of the TR. Int.: TR [122] PARN Poly(A) degradation activity, essential gene first required during early development. [123] TCAB1 H/ACA snoRNPs, driving telomerase to Cajal bodies. Int.: TR, dyskerin, NHP2 and GAR1. [124] TCAB1 NP_193883.2 n.a. The proteins in green are structural homologues to their human counterparts, however, any involvement in telomere maintenance or association with RNA component of telomerase has not been described so far. Direct interaction partners (Int.) of TR-associated proteins are enumerated. Cases when reference is not available are denoted n.a. H/ACA ribonucleoprotein complex subunit DKC1 (dyskerin); RuvB-like 2 (reptin); Arabidopsis (Ath); ATP-dependent DNA helicase 2 subunit 1 and 2 (Ku70/80); box H/ACA small nucleolar RNA-protein complexes (H/ACA snoRNPs); Centromere-binding factor (CBF5); Glycine arginine rich 1, 2 (GAR1, 2); Non-histone protein 2 (NHP2); Nuclear assembly factor 1 (NAF1); Nucleolar protein 10 (NOP10); Repressor-activator protein 1 (RAP1); RNA helicase (PARN); RNA helicase (RHAU); RuvB-like 1 (pontin); RuvB-like 1, 2a (RuvBL1, 2a); Telomerase Cajal body protein 1 (TCAB1); Telomerase reverse transcriptase (TERT); Telomere repeat-binding factors (TRBs); Telomere repeat-binding protein 1 (TRP1); Telomerase RNA subunit 1, 2, 2s (TER1, 2, 2s); Telomeric repeat-binding factor 2 (TRF2); Telomerase RNA (TR). Cells 2019, 8, 58 8 of 31 Considerable homology in TERT sequences and domain organization exists among organisms, and this homology has frequently been used to identify novel TERTs in genomic or transcriptomic data (reviewed in Reference [125]). Human TERT, as well as the plant TERTs, can be split into the N-terminal part, the central catalytic reverse transcriptase (RT) motifs, and the C-terminal extension (CTE) which is highly conserved among vertebrates as well as among plants. The N-terminal part comprises regions of both low and high similarity, e.g., the structural domains TEN (telomerase essential N-terminal domain) or TRBD (RNA-binding domain). Although most eukaryotes, including humans, harbor a single TERT gene, in the allotetraploid Nicotiana tabacum plant, three transcribed variants of the TERT gene were described, which were inherited from its diploid progenitor species [126]. Compared to the conserved structure of the TERT subunit, TRs show high sequence diversity among more distant organisms, as exemplified by the length differences of TRs in protozoa (159 nt in ciliate Tetrahymena, 2200 nt in Plasmodium), zebrafish (317 nt), mouse (397 nt), human (451 nt), and budding yeasts (1160 nt). Even within yeasts, the homology among TRs is rather low and their lengths range from 930 to more than 2000 nt [42,113,127–133]. Analogous variance of TR within the plant kingdom is still questionable, since only putative TRs have been predicted in A. thaliana so far [56]. However, several secondary structure motifs in TRs which are essential for telomerase activity are conserved in fungi and animals. Starting from the 5 -end of TR, these include a core-enclosing helix (CEH) formed by pairing the 5’-terminus of TR with the complementary internal TR region, a template boundary element (TBE)—a hairpin defining the end of the sequence recognized by TERT as a template, the template sequence itself, and a pseudoknot [133]. Except for the template sequence, none of these structural elements has been recognized in TER1 in Arabidopsis thaliana, which is the only reported candidate TR among plants so far [56]. With respect to the above-mentioned sequence diversity of plant telomere repeats, it will be interesting to learn whether and how these evolutionary changes are reflected by the corresponding TR subunits. For example, when assuming the phylogeny of Asparagales plants, telomeres switched first from Arabidopsis-like repeats (TTTAGGG)n to human-like repeats (TTAGGG)n in the divergence of the Iridaceae family, and this repeat survived all downstream speciation events until the divergence of the genus Allium, when the human-type repeat was replaced with the unusual (CTCGGTTATGGG)n repeat [24,134,135]. The molecular basis underlying these evolutionary switches in telomere DNA sequences should be sought primarily in the corresponding TRs. We can consider the following possible scenarios. (i) TR remained essentially the same across Asparagales phylogeny and the observed switches in telomere synthesis occurred either as a result of mutations in the template region of TR or in its vicinity, which could have changed the boundaries of the region used as a template, (ii) a different RNA molecule took over the TR function. Experiments are in progress in our laboratory to provide a clear answer to this question. 3. Telomere Chromatin Composition While the end-replication problem of telomeres is most commonly solved by telomerase, the other essential function of telomeres—their end-protection role (i.e., to distinguish natural chromosome ends from DNA breaks, and to eliminate unwanted repair events at telomeres)—is performed by other proteins associated with telomeres. In humans, these include proteins directly binding telomere DNA either in its double strand part (TRF1, TRF2) or at the single strand overhang (POT1). The other proteins bind telomeres via protein-protein interactions with these proteins (RAP1, TIN2, TPP1), which together form a complex termed shelterin [136,137]. Shelterin components and their interaction partners can inhibit the DNA damage response [138–141]. In addition to the end-protective function, shelterin components also play other roles as, e.g., the recruitment of telomerase to telomeres, facilitating replication fork movement through telomeres, or formation of telomere loops (t-loops) [142–149]. In particular, t-loops exist as a “closed-state” telomere conformation both in mammalians and plants [146,150]. While t-loop is considered as a structure inaccessible to telomerase, it may provide a template for telomerase-independent ALT (see below). Cells 2019, 8, 58 9 of 31 The composition of shelterin-like complexes shows differences in individual components among vertebrates, while the overall functions remain conserved. Human proteins associated with double and single strand telomeric DNA, together with their plant orthologues, are listed in Tables 3 and 4, respectively. In plants, knowledge of a shelterin-like complex is incomplete. The only proteins with confirmed in vivo telomere localization and function are members of the single-myb-histone family, telomere repeat binding (TRB) proteins, which have been characterised in Arabidopsis thaliana [66,82,151] and their orthologues were identified in other plants ([152]; Schorova et al., submitted). TRB proteins bind specifically telomeric double strand DNA through their myb-like domain of a telobox type [153,154], as well as the human core components of shelterin—TRF1 and TRF2 proteins. While the myb-like domain in TRF1 and TRF2 is localized at the C-terminus, that of TRB proteins occupies the N-terminus. Additionally, TRB proteins contain the centrally located histone-like domain (H1/5) involved in DNA sequence-unspecific DNA-protein interactions, multimerization, and interaction with POT1b (one of the plant POT1 paralogues) [65,151]. This plant-specific protein-domain organization has not been described in animals. TRB proteins bind telomeric DNA in vitro and in vivo, localize to the telomeres in vivo, interact directly with the telomerase TERT subunit, and the deregulation of telomeres was observed in mutant plants [66,68]. TRB proteins are not only components of the terminal complex associated with telomeres/telomerase, but they are also associated in vivo with promoters of translation machinery genes, which mostly contain a short telomeric sequence [67]. It seems that TRB proteins serve as epigenetic regulators that potentially affect the transcription status of thousands of genes by playing a role of recruiting subunits of multiple epigenetically active multi-protein complexes [68–71,155,156]. These findings are consistent with the observations from yeast or mammals where telomeric proteins (e.g., TRF1, TRF2, and RAP1) are able to localize outside telomeric regions and regulate the transcription of genes involved in metabolism, immunity, and differentiation [157–164]. Surprisingly, no functions in telomere maintenance were found in Arabidopsis orthologues of mammalian TRF proteins (TRFL proteins) where a myb-domain of the telobox type is located C-terminally as in human TRF1 and TRF2 [165]. However, a recent study revealed protein-protein interactions between TRFL2 and TRP1, members of the TRFL family, and TERT from A. thaliana [66,69]. Plant TRFL2 and TRP1 proteins interact with armadillo/β-catenin-like repeat-containing protein (ARM). ARM directly interacts with plant TERT [70] and might be involved in translation initiation or in regulation of recombination-related genes [69]. Moreover, ARM interacts with the chromatin remodeling protein CHR19 (Table 1). ARM, TRB1, POT1a, and CHR19 (but none of the TRFL proteins) were found among proteins that co-purified with Arabidopsis TERT using tandem affinity purification [84]. Association of TERT with proteins that are not essential for telomere maintenance may reflect possible non-telomeric functions of telomerase. A dual function for telomerase, both telomeric and non-telomeric, is not unique to plants, as mammalian telomerase is involved not only in elongation of telomeres but also non-telomeric activities have been described, including involvement in regulating cellular processes such as apoptosis, proliferation, and cell cycle progression ([166]; reviewed in Reference [167]). Human telomerase and human ARM proteins play a role in the Wnt/APC/β-catenin signaling pathway [168]. A putative human homologue of ARM, ARMC6, interacts with the shelterin protein TRF2 and immuno-precipitates telomerase activity [69]. Cells 2019, 8, 58 10 of 31 An additional telomere maintenance component is—somewhat paradoxically—Ku70/80 heterodimer, a DNA repair factor with a high affinity for DNA ends, that plays essential roles in the maintenance of genome integrity in both human and plants cells. In human cells, Ku70/80 heterodimer interacts with the RNA component of telomerase hTR [120] and with catalytic subunit hTERT [94]. In plants, Ku proteins, as well POT1b protein, are associated with TER2. This is a candidate plant TR that is not required for telomere maintenance in A. thaliana [56]. Ku70/80 is, however, important for protection of blunt-ended telomeres and for suppression of ALT (see below). An integrative updated schematic view based on these and previous studies is depicted in Figure 2. It is obvious that the number of plant telomere-associated and telomerase-associated orthologues (where they exist) is larger in comparison to their mammalian counterparts. The phenomenon of the multiplication of genes of the same family is not surprising, since in many plant families, polyploidy (i.e., whole genome duplication) resulting in retention of multiple gene paralogues may lead to their sub-functionalization, neo-functionalization, or partial or full redundancy [169,170]. In association with the previously mentioned evolutionary divergence of plant telomere DNA repeats toward human-like repeats or unusual telomeric repeats, it will be of interest to learn whether pre-existing components of plant shelterin-like complexes have adapted to the change in DNA sequence (this will be particularly interesting in proteins directly recognizing DNA sequences, such as the TRB or POT1 proteins), or whether some other proteins have replaced their function. Besides the shelterin complex in mammals and its emerging equivalents in plants, there is yet another complex termed CST (CTC1-STN1-TEN1), which is involved in telomere maintenance. This tripartite complex binds the 3 -overhang of the G-rich strand of telomeric DNA and its function in telomere maintenance is conserved in both mammals and plants, and a similar complex exists also in yeast (with Cdc13 instead of CTC1 subunit) [171]. Recently, the roles of individual components of the human CST complex in telomere maintenance were elucidated: while CTC1-STN1 limits telomerase action to prevent G-overhang over-extension, TEN1 is essential for CST function in C-strand fill-in synthesis due to its stabilizing effect on binding the whole CST complex to telomeres and DNA polymerase α engagement in telomere synthesis [172,173]. CST functions, at least in humans, are not limited only to telomeres. CST is also required to avoid replication problem at G-rich sites throughout the genome, likely resolving replication fork stalling [174]. In addition to the telomere-specific proteins, the major part of telomeres is assembled into the nucleosomal chromatin structure which shows a shorter nucleosome periodicity (spacing) than that in the other parts of the chromosomes of the same organism [175–179]. Since shorter telomeres in cultured human cells show a lower nucleosome density than that in cells with longer telomeres, a close relationship was hypothesized between histone density, heterochromatin protein associations, telomere length, and TPE [180]. Interestingly, this feature of telomeric chromatin is conserved at least in vertebrates and plants, and may reflect the specific columnar structure of telomeric chromatin with stacked nucleosomes and weak determination of nucleosome positions by telomeric DNA sequence [181]. Cells 2019, 8, 58 11 of 31 Table 3. Comparative overview of proteins associated with telomeric double strand DNA (dsDNA). Telomeric dsDNA Associated Proteins Human Telomeric dsDNA Associated Proteins Protein Function and Direct Interactions References Arabidopsis Telomeric dsDNA Associated Proteins Protein Function and Direct Interactions References TRF1 Shelterin. Int.: telomeric dsDNA, TIN2, TANK1 and PINX1. Non-telomeric: binding to ITS and chromatin and satellite DNA and modulation of their chromatin structure. Control of a common fragile site containing ITS. [59–62] [162,182] TRB1, 2, 3 Shelterin-like. Int.: telomeric dsDNA, TERT, POT1b, RuvBL1 and RuvBL2a. Non-telomeric functions - a recruitment subunit of protein complexes involved in epigenetic regulations. Binding to ITSs. [64–66]; Schorova et al., submitted [67–71] TRF2 Shelterin. Int.: telomeric dsDNA; TIN2, RAP1, NBS1, RAD50, Apollo, Ku70, PARP1, XPF-ERCC1, BLM, FEN1, POLB, ORC, RTEL1 and ATM. [61,72–80, 183–187] TRP1 Possible non-telomeric functions of telomerase. Int.: telomere dsDNA in vitro, TERT, ARM, Ku70, TRFL1 and TRFL9. [66,69,81–83] Non-telomeric function: transcriptional regulator. Binding to ITSs and satellite III. [155,163] TRFL2 Possible non-telomeric functions of telomerase. Int.: telomere dsDNA in vitro, TERT and ARM. [69,83] TRFL9 Possible non-telomeric functions of telomerase. Int.: telomere dsDNA in vitro, TRP1 and POT1a. [69,83] TBP1, TRFL1, TRFL4 Int.: telomere dsDNA in vitro. [83,188] HOT1 Int.: telomeric dsDNA, active telomerase. [93] n.a. Ku70/80 The way of association with telomeric dsDNA is not fully elucidated. Int.: TRF2, RAP1, TR and TERT. [95] Ku70/80 Role in telomere length regulation, may protect blunt-ended telomeres Int.: TRP1, TER2 and TER2s. [82,96–99,101] The proteins depicted in grey are involved in telomere maintenance, however, their association with telomeric dsDNA has not been fully proven yet. Direct interaction partners (Int.) interacting with telomeric dsDNA-associated proteins and concerning their telomeric functions are enumerated. No sequence homologue has been identified yet (n.a.). Double-strand DNA (dsDNA); 5 exonuclease Apollo (Apollo); Armadillo/β-catenin-like repeat-containing protein (ARM); Ataxia telangiectasia mutated kinase (ATM); ATP-dependent DNA helicase 2 subunit 1 and 2 (Ku70/80); Bloom syndrome protein (BLM); DNA polymerase beta (POLB); DNA repair protein RAD50 (RAD50); Excision repair cross-complementation 1 (ERCC1); Flap endonuclease 1 (FEN1); Homeobox telomere-binding protein 1 (HOT1); Interstitial telomeric sequences (ITSs); Nijmegen breakage syndrome protein 1 (NBS1); Origin recognition complex (ORC); Poly(ADP-Ribose); polymerase 1 (PARP1); Protection of telomeres 1b (POT1b); Regulator of telomere elongation helicase 1 (RTEL1); Repressor-activator protein 1 (RAP1); Telomerase RNA (TR); RuvB-like 1, 2a (RuvBL1, 2a); Tankyrase 1 (TANK1); Telomerase reverse transcriptase (TERT); Telomerase RNA subunit 2, 2s (TER2, TER2s); Telomere binding protein 1 (TBP1); Telomere repeat-binding factor 1, 2, 3 (TRB1, 2, 3); Telomere repeat-binding protein 1 (TRP1); Telomeric repeat binding Factor 1-like 1, 2, 4, 9 (TRFL1, 2, 4, 9); Telomeric repeat-binding factor 1 (TRF1); Telomeric repeat-binding factor 2 (TRF2); TRF1-interacting nuclear protein 2 (TIN2); TRF1-interacting protein 1 (PINX1); Xeroderma pigmentosum group F (XPF1). Cells 2019, 8, 58 12 of 31 Table 4. Comparative overview of proteins associated with telomeric single strand (ssDNA). Telomeric ssDNA Associated Proteins Human Telomeric ssDNA Associated Proteins Protein Function and Direct Interactions References Arabidopsis Telomeric ssDNA Associated Proteins Protein Function and Direct Interactions References POT1 Shelterin. Int.: telomeric ssDNA, TPP1 and CTC1. [50–54] POT1a Shelterin-like. Int.: TERT, telomeric ssDNA, TER1, TRFL9, CBF5, RuvBL1, CTC1 and STN1. [47,55–58,69,105,189] POT1b Shelterin-like. Int.: TRB1, TER2, TER2s. [56,82,100] POT1c POT1 paralogue of unknown function. [47] TERT Catalytic subunit of telomerase. [190] TERT Catalytic subunit of telomerase. STN1 CST complex subunit, prevents G-overhang overextension. Int.: CTC1, TEN1, TPP1 and POLA. [54,172,191,192] STN1 CST complex subunit, controls access of telomerase and DDR, together with POLA may be involved in C-strand synthesis. Int.: CTC1, TEN1 and POT1a. Non-telomeric function. Facilitates re-replication at non-telomeric loci. [189,193–195] TEN1 CST complex subunit, involves C-strand fill-in synthesis. Int.: STN1. [172,192] TEN1 CST complex subunit, controls access of telomerase and DDR, coordinating synthesis of the C-strand. Int.: STN1. [194] CTC1 CST complex subunit, prevents G-overhang overextension. Int.: telomeric ssDNA, STN1, TPP1 and POT1. [54,192] CTC1 CST complex subunit, controls access of the telomerase and DDR, coordinating synthesis of the C-strand. Int.: STN1, POT1a and POLA. [171,189,196] Purα p.h. Unwinds dsDNA telomeric oligonucleotides. [105] PURα1 Associates with TERT. [84] n.a. Why1 Regulates telomere-length homeostasis. Int.: telomeric ssDNA. [197] n.a. STEP1 Truncated derivative of chloroplast RNA-binding protein, role in plant telomere biogenesis. Int.: telomeric ssDNA. [198] The proteins depicted in grey are involved in telomere maintenance, however, their association with telomeric ssDNA has not been fully proven yet. The proteins in green are structural homologues of their human/plant counterparts, however, any involvement in telomere maintenance or association with telomeric sequences has not been described so far. Direct interaction partners (Int.) interacting with telomeric ssDNA associated proteins are enumerated. Cases with not yet identified sequence homologues are denoted with n.a. Single strand DNA (ssDNA); Double-strand DNA (dsDNA); Cajal bodies factor 5 (CBF5); Conserved telomere maintenance component 1 (CTC1); CST complex (CTC1, STN1 and TEN1 subunits); DNA damage response (DDR); DNA polymerase alpha (POLA); Protection of telomeres 1 (POT1); Protection of telomeres 1a, b, c (POT1a, b, c); Pur-alpha 1 (Purα1); RuvB-like 1 (RuvBL1); Single-stranded telomere-binding protein 1 (STEP1); Suppressor of cdc thirteen homolog (STN1); Telomerase reverse transcriptase (TERT); Telomerase RNA subunit 1, 2, 2s (TER1, 2, 2s); Telomeric pathways in association with STN1 (TEN1); Telomeric repeat binding factor 1 -like 9 (TRFL9); TIN2- and POT1-organizing protein (TPP1); Whirly 1 (Why1); putative homolog according to NCBI blastp (p.h.). Cells 2019, 8, 58 13 of 31 Cells 2019, 8, x FOR PEER REVIEW 14 of 31 Figure 2. An integrative schematic view of the human and plant terminal telomeric complex. (A) Human active telomerase is associated with Hsp90 and p23 chaperones as well as with TR associated conserved scaffold proteins of box H/ACA small nucleolar RNAs (dyskerin, NHP2, NOP10, GAR1). Mammalian shelterin proteins (TRF1/2, RAP1, TIN2, TPP1, and POT1) modulate access to the telomerase complex and the ATR/ATM-dependent DNA damage response pathway. The CST complex (CTC1-STN1-TEN1) affects telomerase and DNA polymerase α recruitment to the chromosomal termini, and, thus, coordinates G-overhang extension by telomerase with fill-in synthesis of the complementary C-strand (blue dashed line). G-quadruplexes, D-loops, and t-loops during telomere replication are resolved by RTEL helicase. HOT1 directly binds double strand telomere repeats and associates with the active telomerase. Telomere nucleosomes show a shorter periodicity than that in the other parts of chromosomes. For human telomere histone modifications, see Figure 3. (B) Arabidopsis telomerase is associated with TRB proteins as well as with POT1a that interacts with the dyskerin orthologue CBF5. Plants possess all orthologue proteins of conserved scaffold box H/ACA of small nucleolar RNAs (CBF5, GAR1, NOP10, NHP2). Moreover, TRB proteins interact with the telomeric sequence due to the same myb-like binding domain as that in mammalian TRF1/2. TRB proteins interact with TERT and POT1b, and, when localized at chromosomal ends, they are eligible to function as components of the plant shelterin complex. An evolutionarily conserved CST complex is suggested to coordinate the unique requirements for efficient replication of telomeric DNA in plants as well as in other organisms. In addition, plant RTEL contributes to telomere homeostasis. For the sake of clarity, only the situation in telomere with 3′ overhang is depicted. For plant telomere histone modifications, see Figure 3. 4. Telomere Epigenetics Figure 2. An integrative schematic view of the human and plant terminal telomeric complex. (A) Human active telomerase is associated with Hsp90 and p23 chaperones as well as with TR associated conserved scaffold proteins of box H/ACA small nucleolar RNAs (dyskerin, NHP2, NOP10, GAR1). Mammalian shelterin proteins (TRF1/2, RAP1, TIN2, TPP1, and POT1) modulate access to the telomerase complex and the ATR/ATM-dependent DNA damage response pathway. The CST complex (CTC1-STN1-TEN1) affects telomerase and DNA polymerase α recruitment to the chromosomal termini, and, thus, coordinates G-overhang extension by telomerase with fill-in synthesis of the complementary C-strand (blue dashed line). G-quadruplexes, D-loops, and t-loops during telomere replication are resolved by RTEL helicase. HOT1 directly binds double strand telomere repeats and associates with the active telomerase. Telomere nucleosomes show a shorter periodicity than that in the other parts of chromosomes. For human telomere histone modifications, see Figure 3. (B) Arabidopsis telomerase is associated with TRB proteins as well as with POT1a that interacts with the dyskerin orthologue CBF5. Plants possess all orthologue proteins of conserved scaffold box H/ACA of small nucleolar RNAs (CBF5, GAR1, NOP10, NHP2). Moreover, TRB proteins interact with the telomeric sequence due to the same myb-like binding domain as that in mammalian TRF1/2. TRB proteins interact with TERT and POT1b, and, when localized at chromosomal ends, they are eligible to function as components of the plant shelterin complex. An evolutionarily conserved CST complex is suggested to coordinate the unique requirements for efficient replication of telomeric DNA in plants as well as in other organisms. In addition, plant RTEL contributes to telomere homeostasis. For the sake of clarity, only the situation in telomere with 3 overhang is depicted. For plant telomere histone modifications, see Figure 3. Cells 2019, 8, 58 14 of 31 4. Telomere Epigenetics As chromatin structures, telomeres are natural targets for epigenetic modifications. At the DNA level, methylation at carbon 5 of cytosine represents the dominant mark in eukaryotic cells. Methylated cytosines (mCs) are generally enriched in heterochromatic regions of the genome and silenced promoters. Important differences in the sequence contexts, in which mCs are located, exist between animals and plants. In mammalian cells, they are predominantly located in CG doublet motifs, with the symmetry of the sequence crucial for the maintenance of the methylation pattern during DNA replication (reviewed in Reference [199]). A fraction of mCs in non-CG contexts was found in human embryonic cells. This fraction disappears after differentiation and is restored in induced pluripotent stem cells, which shows involvement of distinct methylation patterns in the regulation of gene expression [200]. Also in plants, cytosines in the CG motif are most frequently methylated, but mCs are also commonly placed in non-CG sequences, symmetrical CHG triplets (H=C or A or T), or non-symmetrical CHH motifs (reviewed in Reference [201]). In telomeres, cytosines in non-symmetrical sequence contexts are present in the telomeric C-rich strand, i.e., in CCCTAA repeats in animals and CCCTAAA repeats in plants. Using shotgun bisulfite genomic sequencing, mCs were detected in A. thaliana telomeric repeats with the inner cytosine most frequently methylated [202]. This pattern was confirmed by an independent approach, with high reliability at least in the proximal part of the telomere [203,204], and methylated telomeric cytosines were detected in cultured Nicotiana tabacum (tobacco) cells [205] and other plants [206]. Disruption of telomere homeostasis as a consequence of decreased genomic DNA methylation was observed in A. thaliana [203,207] but not in tobacco cells [205], which shows differences in the involvement of DNA methylation in regulation of telomere homeostasis between these model plants (for a more detailed review see Reference [208]). Telomeres formed by mini-satellite repeats were traditionally considered as heterochromatic regions, and, thus, associated with heterochromatin-specific histone marks. Certain differences in histone modifications in heterochromatin have been described between animals and plants. In animals, constitutive heterochromatin is defined by the presence of H3K9me3 (trimethylation of lysine 9 of histone H3) (reviewed in Reference [209]) while in plants, this mark decorates silenced euchromatic genes, and constitutive heterochromatin is associated with H3K9me2 modification [210]. Facultative heterochromatin is enriched in H3K27me3 in cells of representatives of both kingdoms. In agreement with the hypothesis of the heterochromatic character of telomeres, the importance of heterochromatin-specific epigenetic marks for telomere maintenance and genome stability was demonstrated in numerous studies using human and mouse cells as models (reviewed in Reference [211]). On the other hand, data showing a low level of heterochromatin-specific modifications and an abundance of active marks on human telomeric histones have been presented [212–214], which shows certain dynamics of the human telomeric chromatin structure. Based on these and other reports, distinct differences exist in telomeric chromatin composition between the most important mammalian models, human and mouse cells, because H3K9me3 density and HP1 enrichment were significantly higher in mouse compared to humans [215,216]. Nevertheless, according to a study utilizing quantitative locus purification [217] the heterochromatic histone modification H4K20me3 is underrepresented at mouse telomeres even though it was previously detected by others at mouse [218,219] and also human [220] telomeres in analyses based on chromatin immuno-precipitation. Further research is necessary to draw final conclusions on the epigenetic nature of mammalian telomeres, especially considering other factors mentioned below. Plant telomeric chromatin was shown to be associated with both heterochromatin-specific H3K9me2 and euchromatic H3K4me3 marks, with the latter less abundant [204,206,221]. Therefore, the plant telomeric chromatin exhibits a dual epigenetic character. Identification of the H3K27me3 modification, which is typical for facultative heterochromatin, in telomeric histones of A. thaliana [221,222] and N. tabacum [206] was rather surprising. However, it correlates with its presence at human telomeres [215], and with the recent observation that polycomb repressor complex 2-dependent Cells 2019, 8, 58 15 of 31 loading of H3K27me3 at human telomeres is essential for the proper establishment of H3K9me3 and H4K20me3 modifications [220]. Nevertheless, H3K27me3 was not detected at mouse telomeres [217]. Thus, although significantly fewer results are available on the epigenetics of telomeric chromatin in plants compared to mammals, interesting similarities as well as differences have already been described and hopefully others will be elucidated based on future studies using different model organisms, including plants with non-canonical telomere sequences [24,25,27,134]. When discussing telomeric chromatin, it is necessary to mention that analysis of epigenetic modifications may be complicated by the presence of non-terminally located telomeric repeats forming interstitial telomeric sequences (ITSs). ITSs are relatively abundant in subtelomeric, pericentromeric, and centromeric regions of most eukaryotic organisms and represent fragile parts of chromosomes, which are prone to rearrangements and recombinations. The detailed compositions of telomeres and ITSs are different. In contrast to telomeres consisting of long tracts of perfect telomeric repeats, ITSs are often degenerated and/or disrupted by non-telomeric sequences. However, ITSs may still contribute to the telomere-specific signal in epigenetic studies, mainly those based on hybridization of membrane-bound DNA. Frequently-used genome-wide sequencing analyses (ChIP-seq and bisulfite sequencing) do not completely solve this problem because telomeres, like other tandem repeats, are difficult to analyze, and even direct analysis of respective read counts (i.e., those comprising perfect telomeric repeats versus those formed by degenerated repeats and non-telomeric sequences) may be ambiguous due to the non-linearity of PCR amplification of repetitive sequences [223]. Both mammalian and plant telomeres are transcribed to long non-coding RNA called TERRA [204,224] and this transcriptional potency could reflect the relatively lower level of compactness of telomeric chromatin compared with heterochromatin. The apparent discrepancy between the association of heterochromatic marks with telomeric histones and the transcriptional activity of telomeres is weakened by the facts that a mechanistic relationship between TERRA transcription and loading of heterochromatic modifications to human telomeres has been described [220], and that in Arabidopsis a certain—maybe dominant—fraction of TERRA is transcribed from ITSs [204], which are purely heterochromatic [225]. At this stage of knowledge, it is difficult or even impossible to formulate any general conclusion on the epigenetic nature of telomeric chromatin (Figure 3). Without any doubt, the specific structure of telomeres is crucial for the maintenance of genome integrity. Telomeres are rigid enough to prevent repair and recombination at chromosome ends and to restrict telomere accessibility for telomerase, but open enough to be transcribed and, at least in a specific time window of the cell cycle, accessible to telomerase. Moreover, in disagreements about telomeric “heterochromatin” or “euchromatin”, contribution of non-histone players, mainly shelterin proteins, to the telomeric chromatin compaction should be reflected (reviewed in Reference [226]). Why not admit, that telomeric chromatin is so specific that it does not fit into the existing criteria and that these should be widened? This suggestion is strengthened by the finding that other non-genic parts of the human genome, originally thought to be uniformly heterochromatic, are associated with different combinations of histone marks [213]. It is well possible that the epigenetic state of telomeres is more dynamic than previously thought and shows tissue-specific, cell-cycle specific, and developmental stage-specific changes. This would not only explain the diverse results of the above studies, but would be consistent with our current understanding of the epigenetics of other chromosome regions. Cells 2019, 8, 58 16 of 31Cells 2019, 8, x FOR PEER REVIEW 17 of 31 Figure 3. Modifications of mammalian and plant telomere (telo.) and pericentromere (peric.) histones. The relative enrichments of selected epigenetic modifications of telomeric and pericentromeric histones in human, mouse and Arabidopsis are schematically depicted according to data presented in References [204,212,213,215,217–222,225]. 5. Telomere 3′-Overhangs, Blunt Ends, and Loops Telomeres in vertebrates, in particular humans, possess 3′-overhangs at both chromosome ends. These overhangs are of different sizes on lagging versus leading strands [227]. In human telomeres a G-overhang is prevalent whose length varies from several tens to 280 nt [228–230]. Likewise, a 5′ Crich overhang is present at the telomeres of human chromosomes, being far more prevalent in tumor cells using ALT (see below) [231]. This is not the case in Arabidopsis thaliana, Silene latifolia, and other angiosperm plants, which lack telomere overhangs or possess only short 1–3 nt overhangs at about half of their telomeres [232,233]. The telomere whose 3′- end is being synthesized in a given cell cycle by leading strand synthesis remains blunt-ended likely due to protection against end-processing by a specific exonuclease. This protection is dependent on the Ku70/80 heterodimer [233]. The role of the Ku complex in plant telomere protection was also suggested by our earlier studies, which indicated Ku as an interaction partner of AtTRP1, one of the TRF-like proteins in A. thaliana ([82]; see Reference [155] for a review). An analogous interaction between the shelterin components TRF2 and Ku70 was observed earlier in human cells [77]. Due to the asymmetry (non-equivalence) of plant telomeres, a different set of proteins may protect the telomere whose 3′-end serves as a template in “incomplete” lagging strand synthesis and can be elongated by telomerase. Protection of blunt-ended telomeres in Arabidopsis by the Ku70/80 complex seems paradoxical considering the presumed end-protective function of telomeres on one hand, and a key role of the Ku complex in non-homologous end-joining repair of double strand DNA breaks on the other hand. A possible solution of this enigma was suggested recently by a study which indicated different binding modes of the Ku complex to dsDNA breaks and to telomeres. Both functions were dissected using Ku mutants with impaired ability to translocate along DNA. While Ku sliding is not required for its association with plant telomeres, it is essential for its involvement in the non-homologous end joining pathway of DNA repair [101]. The presence of blunt-ended telomeres is, however, not common to all plants. For example, in the moss Physcomitrella patens, both telomeres of a chromosome possess overhangs and, correspondingly, lack of the Ku complex components shows no effect on telomere maintenance or end protection [234]. The Ku70/80 complex was also reported to be a negative regulator of telomerase function in Arabidopsis [99]. In addition to telomere elongation by telomerase, an extension of telomere G-strand overhangs was observed in Ku mutants, which suggests a role of Ku70/80 in C-rich telomeric strand maintenance [235]. Figure 3. Modifications of mammalian and plant telomere (telo.) and pericentromere (peric.) histones. The relative enrichments of selected epigenetic modifications of telomeric and pericentromeric histones in human, mouse and Arabidopsis are schematically depicted according to data presented in References [204,212,213,215,217–222,225]. 5. Telomere 3 -Overhangs, Blunt Ends, and Loops Telomeres in vertebrates, in particular humans, possess 3 -overhangs at both chromosome ends. These overhangs are of different sizes on lagging versus leading strands [227]. In human telomeres a G-overhang is prevalent whose length varies from several tens to 280 nt [228–230]. Likewise, a 5 C-rich overhang is present at the telomeres of human chromosomes, being far more prevalent in tumor cells using ALT (see below) [231]. This is not the case in Arabidopsis thaliana, Silene latifolia, and other angiosperm plants, which lack telomere overhangs or possess only short 1–3 nt overhangs at about half of their telomeres [232,233]. The telomere whose 3 - end is being synthesized in a given cell cycle by leading strand synthesis remains blunt-ended likely due to protection against end-processing by a specific exonuclease. This protection is dependent on the Ku70/80 heterodimer [233]. The role of the Ku complex in plant telomere protection was also suggested by our earlier studies, which indicated Ku as an interaction partner of AtTRP1, one of the TRF-like proteins in A. thaliana ([82]; see Reference [155] for a review). An analogous interaction between the shelterin components TRF2 and Ku70 was observed earlier in human cells [77]. Due to the asymmetry (non-equivalence) of plant telomeres, a different set of proteins may protect the telomere whose 3 -end serves as a template in “incomplete” lagging strand synthesis and can be elongated by telomerase. Protection of blunt-ended telomeres in Arabidopsis by the Ku70/80 complex seems paradoxical considering the presumed end-protective function of telomeres on one hand, and a key role of the Ku complex in non-homologous end-joining repair of double strand DNA breaks on the other hand. A possible solution of this enigma was suggested recently by a study which indicated different binding modes of the Ku complex to dsDNA breaks and to telomeres. Both functions were dissected using Ku mutants with impaired ability to translocate along DNA. While Ku sliding is not required for its association with plant telomeres, it is essential for its involvement in the non-homologous end joining pathway of DNA repair [101]. The presence of blunt-ended telomeres is, however, not common to all plants. For example, in the moss Physcomitrella patens, both telomeres of a chromosome possess overhangs and, correspondingly, lack of the Ku complex components shows no effect on telomere maintenance or end protection [234]. The Ku70/80 complex was also reported to be a negative regulator of telomerase function in Arabidopsis [99]. In addition to telomere elongation by telomerase, Cells 2019, 8, 58 17 of 31 an extension of telomere G-strand overhangs was observed in Ku mutants, which suggests a role of Ku70/80 in C-rich telomeric strand maintenance [235]. Besides telomerase, eukaryotic cells can also utilize a back-up mechanism of telomere maintenance—ALT—which is based on homologous recombination (HR) [236]. This telomerase-independent mechanism is activated in a number of human tumors, in human cells immortalized in culture, and also in normal somatic tissues [237]. In plants, the ALT mechanism is activated in mutants with telomerase dysfunction and possibly also during the earliest stages of normal plant development [238]. ALT relies on the formation of terminal telomeric loops (t-loops) [146], which parallels the first steps of HR. The eventual resolution of these t-loops and aberrant HR at telomeres generates not only telomeres of highly heterogeneous lengths but also extrachromosomal t-circles, which are the known hallmarks of ALT. In mutant plants that are deficient for components of the Ku70/80 complex, induction of t-circle formation was observed at telomeres but not at other regions rich in DNA repeats. Despite ongoing terminal deletions arising from excision of t-circles in mutant plants, the telomeres remain functional, which indicates an efficient telomere healing by telomerase [239]. Another interesting protein connecting telomeric loops and circles with DNA recombination and telomere replication is RTEL1. This was originally described in Caenorhabditis elegans as a functional homologue of the yeast Srs2 protein, which removes Rad51 from single strand DNA. Therefore, it prevents the homology search step of HR and helps to protect the cell from inappropriate HR (for review, see Reference [240]). Furthermore, in C. elegans, the RTEL1 helicase suppresses inappropriate recombination events by promoting disassembly of D-loop recombination intermediates, and the loss of its function results in increased genome instability [241]. In addition to its regulatory role in HR, RTEL1 acts in telomere maintenance in mammalian telomerase-positive cells [242]. This function was explained by the function of RTEL1 in opening t-loops, which blocked inappropriate excision of large telomere regions—the process known as telomere rapid deletion. To promote this t-loop unwinding, RTEL1 is recruited to telomeres in the S-phase by the telomeric protein TRF2 [186]. In addition to its role in t-loop stability, mouse RTEL1 can dissolve G4-DNA structures, which otherwise block replication fork progression and the extension of telomeres by telomerase [243]. Importantly, the role of RTEL1 in telomere dynamics was clearly confirmed by the finding that its mutation is causative for Hoyeraal-Hreidarsson syndrome, which is a severe form of dyskeratosis congenita, predisposing to bone-marrow failure and cancer. This disease is characterised by short telomeres and genome instability [244–246]. A recent report revealed that reversed replication forks occurring in telomeres of RTEL1-deficient cells is due to compromised telomere replication aberrantly recruiting telomerase, which prevents the restart of reversed replication forks at telomeres and leads to critically short telomeres [247]. In this context, telomerase paradoxically contributes to telomere shortening by stabilizing stalled replication forks at chromosome ends. In addition, the A. thaliana RTEL1 homolog suppresses HR and is involved in processing DNA replication intermediates and interstrand and intrastrand DNA cross-links. Deficiency of the Arabidopsis RTEL1 triggers a SOG1-dependent replication checkpoint in response to DNA crosslinks [248]. Similarly to the situation in mammals, the Arabidopsis RTEL1 contributes to telomere homeostasis. The concurrent loss of RTEL1 and TERT accelerates telomere shortening, which results in a developmental arrest after four generations [249] compared to 10 generations in single-mutant tert plants [250]. This observation indicates a role of RTEL1 in ALT, which otherwise partially compensates for the loss of TERT [238]. In agreement with these results, it was recently demonstrated that RAD51-dependent homologous recombination participates in ALT in A. thaliana [251]. This is not surprising when considering the essential role of RAD51 in HR, and HR as a major molecular mechanism of ALT. However, the authors further showed that this role of RAD51 is dependent on RTEL1 helicase, which possibly functions in dissolution of the D-loop after telomere replication. In P. patens, RTEL1 has been found among genes, which are up-regulated after γ-irradiation. RTEL1 knockout resulted in a severe growth deficiency, which was independent of the presence Cells 2019, 8, 58 18 of 31 of bleomycin [252], and the authors hypothesized that this growth phenotype might be the result of telomere deficiency. Thus, the functions of RTEL1 seem widely conserved. In conclusion, the requirement for RTEL1 in multiple pathways to preserve plant genome stability can be explained by its putative role in the destabilization of DNA loop structures such as D-loops and t-loops, which aligns with previous studies in mammalian systems. 6. Cellular Aging and the Immortal DNA Strand Hypothesis Cellular aging is characterized by progressive loss of physiological integrity that leads to impaired function and genomic instability and ultimately to a functional decline at the tissue and organ level. Telomere attrition during cell aging is classified as one of the several major hallmarks of aging—together with, e.g., genomic instability, epigenetic alterations, loss of proteostasis, mitochondrial dysfunction, cellular senescence, or altered intercellular communication [7]. In Metazoa, there is no universal pattern of telomere erosion [253], and, in some animals, the progressive telomere shortening with age has not been observed [254]. Nevertheless, telomere length is typically inversely correlated with lifespan, while telomerase expression co-evolved with body size [255]. A connection between cellular aging and replicative telomere shortening is widely accepted and experimentally validated in both humans and plants. Importantly, under normal conditions (in wild type plants) this type of cellular aging is prevented by telomerase activity in dividing cells [20,21,38]. The associations between telomere length and age-related disease and mortality in humans have been proven in several studies (reviewed in References [8,256,257]). However, telomere length of humans is not a determinant of aging but rather a marker able to explain life expectancy and disease risk. In animals, the distribution of cellular age varies among tissues and cell compartments, including progenitor cell compartments, depending on the influx of stem cells and the dynamics of self-renewal and differentiation of progenitor cells. In particular, the mode of cell division of progenitor cells may be: (i) symmetric self-renewal, in which progenitor cell division results in two daughter progenitor cells (one generation older) remaining in the compartment, (ii) symmetric differentiation, resulting in two differentiated cells which leave the progenitor cell compartment, or (iii) asymmetric division resulting in one progenitor and one differentiated cell. Importantly, cellular age distributions between healthy and cancerous tissues may inform dynamic changes within the hierarchical tissue structure, i.e., an acquired increased self-renewal capacity in certain tumors [258]. In this connection, it is of interest to mention the hypothesis of the immortal DNA strand [259]. This hypothesis proposes that adult stem cells segregate their template and newly synthesized DNA strands non-randomly, preferentially retaining parental DNA strands in each division. This way, adult stem cells pass mutations resulting from replication errors onto non-stem cell daughter cells that differentiate and terminate division. Adult stem cells could thus reduce the accumulation of mutations and the associated deterioration of gene functions with each cell cycle. Moreover, this strategy would also slow down replicative telomere shortening. Thus, two major factors of cellular and organismal aging could be substantially limited if immortal DNA strand segregation operates in progenitor cells. Several studies have supported this hypothesis up to now. For example, using sequential pulses of halogenated thymidine analogues, high frequencies of segregation of older and younger template strands during proliferative expansion of mouse muscle stem cells was observed [260]. Template strand co-segregation was strongly associated with asymmetric cell divisions yielding daughters with divergent fates. Daughter cells inheriting the older templates retained a more immature phenotype, whereas daughters inheriting the newer templates acquired a more differentiated phenotype. It will be of interest to learn if the validity of this hypothesis is more general, and specifically to elucidate the molecular mechanism of non-random DNA segregation in asymmetric cell division. This principle may also be functional in meristem cell division and differentiation. While replicative telomere shortening is efficiently counteracted by telomerase in wild type plants (see above), reduction of accumulation of mutations would be extremely beneficial when considering e.g., trees sustaining their growth for centuries. Low telomere loss per plant generation has been Cells 2019, 8, 58 19 of 31 found in telomerase-deficient Arabidopsis mutants [250], which indicates a possible involvement of non-random DNA strand segregation in addition to ALT [238]. Unfortunately, the application of sequential pulse labeling in planta is technically too demanding, and any direct evidence for the immortal DNA strand hypothesis is, thus, missing in plants. 7. Concluding Remarks Currently available data show remarkably conserved principles in telomere biology across eukaryotes, which is consistent with an association of telomere and telomerase emergence with the earliest steps of their evolution. At the same time, however, a number of specific features and exceptions cannot be ignored since they point to limitations of our wider understanding of these principles. Among a number of open questions to be answered, elucidation of the structure of telomeric chromatin (telochromatin), including its epigenetic and higher-order dynamics, with high spatial and temporal resolution is needed in various model systems. Furthermore, the biological relevance of non-canonical structures formed by telomeric DNA should be addressed mainly under in vivo conditions. Such studies are timely due to recent fast progress in adequate technical tools, including e.g., super-resolution and cryo-electron microscopy. Studies of repair processes at telomeres and of telomerase regulation belong to the hot topics in this field, since this knowledge can clearly be applied to promote protection of genome stability. In this respect, plants are indispensable due to the natural telomerase-competent character of their cells which allows us to examine mechanisms of repression and activation of telomerase in association with proliferation, differentiation, and dedifferentiation of plant cells. This knowledge is essential for understanding carcinogenesis and is potentially applicable to tumor therapy and cell rejuvenation. Author Contributions: P.P.S., M.F. and J.F. contributed to this paper with a literature review, drafting the paper, and approval of the final version. Funding: This work was supported by the Czech Science Foundation (projects 16-01137S and 17-09644S), by the project SYMBIT, reg. number: CZ.02.1.01/0.0/0.0/15_003/0000477 financed by the ERDF, and by the Ministry of Education, Youth and Sports of the Czech Republic under the projects CEITEC 2020 (LQ1601) and INTER-COST (LTC17077). Acknowledgments: We thank to Ladislav Dokládal and Ronald Hancock for reviewing and discussing the MS prior to submission. Conflicts of Interest: The authors declare no conflict of interest. References 1. Olovnikov, A.M. Principle of marginotomy in template synthesis of polynucleotides. Dokl. Akad. Nauk. SSSR 1971, 201, 1496–1499. 2. Muller, H.J. The remaking of chromosomes. Collect. Net. 1938, 13, 181–195. 3. McClintock, B. The fusion of broken chromosome ends of sister half-chromatids following chromatid breakage at meiotic anaphases. Mo. Agric. Exp. Stn. Res. Bull. 1938, 290, 1–48. 4. McClintock, B. The stability of broken ends of chromosomes in Zea mays. Genetics 1941, 26, 234–282. 5. Von Zglinicki, T. Oxidative stress shortens telomeres. Trends Biochem. Sci. 2002, 27, 339–344. [CrossRef] 6. Hoelzl, F.; Cornils, J.S.; Smith, S.; Moodley, Y.; Ruf, T. Telomere dynamics in free-living edible dormice (Glis glis): The impact of hibernation and food supply. J. Exp. Biol. 2016, 219, 2469–2474. [CrossRef] 7. Lopez-Otin, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The Hallmarks of Aging. Cell 2013, 153, 1194–1217. [CrossRef] 8. Simons, M.J.P. Questioning causal involvement of telomeres in aging. Ageing Res. Rev. 2015, 24, 191–196. [CrossRef] 9. Steenstrup, T.; Kark, J.D.; Verhulst, S.; Thinggaard, M.; Hjelmborg, J.V.B.; Dalgard, C.; Kyvik, K.O.; Christiansen, L.; Mangino, M.; Spector, T.D.; et al. Telomeres and the natural lifespan limit in humans. Aging US 2017, 9, 1130–1142. [CrossRef] Cells 2019, 8, 58 20 of 31 10. Factor-Litvak, P.; Susser, E.; Kezios, K.; McKeague, I.; Kark, J.D.; Hoffman, M.; Kimura, M.; Wapner, R.; Aviv, A. Leukocyte Telomere Length in Newborns: Implications for the Role of Telomeres in Human Disease. Pediatrics 2016, 137, e20153927. [CrossRef] 11. Robin, J.D.; Ludlow, A.T.; Batten, K.; Magdinier, F.; Stadler, G.; Wagner, K.R.; Shay, J.W.; Wright, W.E. Telomere position effect: Regulation of gene expression with progressive telomere shortening over long distances. Genes Dev. 2014, 28, 2464–2476. [CrossRef] 12. Victorelli, S.; Passos, J.F. Telomeres and Cell Senescence—Size Matters Not. Ebiomedicine 2017, 21, 14–20. [CrossRef] 13. Abdallah, P.; Luciano, P.; Runge, K.W.; Lisby, M.; Geli, V.; Gilson, E.; Teixeira, M.T. A two-step model for senescence triggered by a single critically short telomere. Nat. Cell Biol. 2009, 11, 988. [CrossRef] 14. Hemann, M.T.; Strong, M.A.; Hao, L.Y.; Greider, C.W. The shortest telomere, not average telomere length, is critical for cell viability and chromosome stability. Cell 2001, 107, 67–77. [CrossRef] 15. Kaul, Z.; Cesare, A.J.; Huschtscha, L.I.; Neumann, A.A.; Reddel, R.R. Five dysfunctional telomeres predict onset of senescence in human cells. Embo Rep. 2012, 13, 52–59. [CrossRef] 16. Watson, J.M.; Riha, K. Telomeres, aging, and plants: From weeds to Methuselah—A mini-review. Gerontology 2011, 57, 129–136. [CrossRef] 17. Barsov, E.V. Telomerase and primary T cells: Biology and immortalization for adoptive immunotherapy. Immunotherapy 2011, 3, 407–421. [CrossRef] 18. Shalaby, T.; Hiyama, E.; Grotzer, M.A. Telomere Maintenance as Therapeutic Target in Embryonal Tumours. Anti-Cancer Agents Med. Chem. 2010, 10, 196–212. [CrossRef] 19. Fajkus, J.; Kovarik, A.; Kralovics, R. Telomerase activity in plant cells. Febs Lett. 1996, 391, 307–309. [CrossRef] 20. Fajkus, J.; Fulneckova, J.; Hulanova, M.; Berkova, K.; Riha, K.; Matyasek, R. Plant cells express telomerase activity upon transfer to callus culture, without extensively changing telomere lengths. Mol. Gen. Genet. 1998, 260, 470–474. [CrossRef] 21. Fitzgerald, M.S.; McKnight, T.D.; Shippen, D.E. Characterization and developmental patterns of telomerase expression in plants. Proc. Natl. Acad. Sci. USA 1996, 93, 14422–14427. [CrossRef] 22. Fajkus, J.; Sykorova, E.; Leitch, A.R. Telomeres in evolution and evolution of telomeres. Chromosome Res. 2005, 13, 469–479. [CrossRef] 23. Louis, E.J. Are Drosophila telomeres an exception or the rule? Genome Biol. 2002, 3. [CrossRef] 24. Fajkus, P.; Peska, V.; Sitova, Z.; Fulneckova, J.; Dvorackova, M.; Gogela, R.; Sykorova, E.; Hapala, J.; Fajkus, J. Allium telomeres unmasked: The unusual telomeric sequence (CTCGGTTATGGG)(n) is synthesized by telomerase. Plant J. 2016, 85, 337–347. [CrossRef] 25. Peska, V.; Fajkus, P.; Fojtova, M.; Dvorackova, M.; Hapala, J.; Dvoracek, V.; Polanska, P.; Leitch, A.R.; Sykorova, E.; Fajkus, J. Characterisation of an unusual telomere motif (TTTTTTAGGG)n in the plant Cestrum elegans (Solanaceae), a species with a large genome. Plant J. 2015, 82, 644–654. [CrossRef] 26. Peska, V.; Sitova, Z.; Fajkus, P.; Fajkus, J. BAL31-NGS approach for identification of telomeres de novo in large genomes. Methods 2017, 114, 16–27. [CrossRef] 27. Tran, T.D.; Cao, H.X.; Jovtchev, G.; Neumann, P.; Novak, P.; Fojtova, M.; Vu, G.T.H.; Macas, J.; Fajkus, J.; Schubert, I.; et al. Centromere and telomere sequence alterations reflect the rapid genome evolution within the carnivorous plant genus Genlisea. Plant J. 2015, 84, 1087–1099. [CrossRef] 28. Wright, W.E.; Piatyszek, M.A.; Rainey, W.E.; Byrd, W.; Shay, J.W. Telomerase activity in human germline and embryonic tissues and cells. Dev. Genet. 1996, 18, 173–179. [CrossRef] 29. Ramirez, R.D.; Wright, W.E.; Shay, J.W.; Taylor, R.S. Telomerase activity concentrates in the mitotically active segments of human hair follicles. J. Investig. Dermatol. 1997, 108, 113–117. [CrossRef] 30. Hiyama, E.; Hiyama, K.; Yokoyama, T.; Shay, J.W. Immunohistochemical detection of telomerase (hTERT) protein in human cancer tissues and a subset of cells in normal tissues. Neoplasia 2001, 3, 17–26. [CrossRef] 31. Hiyama, E.; Hiyama, K. Telomere and telomerase in stem cells. Br. J. Cancer 2007, 96, 1020–1024. [CrossRef] 32. Hiyama, K.; Hirai, Y.; Kyoizumi, S.; Akiyama, M.; Hiyama, E.; Piatyszek, M.A.; Shay, J.W.; Ishioka, S.; Yamakido, M. Activation of Telomerase in Human-Lymphocytes and Hematopoietic Progenitor Cells. J. Immunol. 1995, 155, 3711–3715. 33. Yui, J.; Chiu, C.P.; Lansdorp, P.M. Telomerase activity in candidate stem cells from fetal liver and adult bone marrow. Blood 1998, 91, 3255–3262. Cells 2019, 8, 58 21 of 31 34. Ito, H.; Kyo, S.; Kanaya, T.; Takakura, M.; Inoue, M.; Namiki, M. Expression of human telomerase subunits and correlation with telomerase activity in urothelial cancer. Clin. Cancer Res. 1998, 4, 1603–1608. 35. Kyo, S.; Takakura, M.; Kohama, T.; Inoue, M. Telomerase activity in human endometrium. Cancer Res. 1997, 57, 610–614. 36. Jureckova, J.F.; Sykorova, E.; Hafidh, S.; Honys, D.; Fajkus, J.; Fojtova, M. Tissue-specific expression of telomerase reverse transcriptase gene variants in Nicotiana tabacum. Planta 2017, 245, 549–561. [CrossRef] 37. Ogrocka, A.; Sykorova, E.; Fajkus, J.; Fojtova, M. Developmental silencing of the AtTERT gene is associated with increased H3K27me3 loading and maintenance of its euchromatic environment. J. Exp. Bot. 2012, 63, 4233–4241. [CrossRef] 38. Riha, K.; Fajkus, J.; Siroky, J.; Vyskot, B. Developmental control of telomere lengths and telomerase activity in plants. Plant Cell 1998, 10, 1691–1698. [CrossRef] 39. Zachova, D.; Fojtova, M.; Dvorackova, M.; Mozgova, I.; Lermontova, I.; Peska, V.; Schubert, I.; Fajkus, J.; Sykorova, E. Structure-function relationships during transgenic telomerase expression in Arabidopsis. Physiol. Plant. 2013, 149, 114–126. [CrossRef] 40. Winter, D.; Vinegar, B.; Nahal, H.; Ammar, R.; Wilson, G.V.; Provart, N.J. An “Electronic Fluorescent Pictograph” Browser for Exploring and Analyzing Large-Scale Biological Data Sets. PLoS ONE 2007, 2. [CrossRef] 41. Greider, C.W.; Blackburn, E.H. Identification of a Specific Telomere Terminal Transferase-Activity in Tetrahymena Extracts. Cell 1985, 43, 405–413. [CrossRef] 42. Greider, C.W.; Blackburn, E.H. A Telomeric Sequence in the Rna of Tetrahymena Telomerase Required for Telomere Repeat Synthesis. Nature 1989, 337, 331–337. [CrossRef] 43. Chan, H.; Wang, Y.Q.; Feigon, J. Progress in Human and Tetrahymena Telomerase Structure Determination. Annu. Rev. Biophys. 2017, 46, 199–225. [CrossRef] 44. Nguyen, T.H.D.; Tam, J.; Wu, R.A.; Greber, B.J.; Toso, D.; Nogales, E.; Collins, K. Cryo-EM structure of substrate-bound human telomerase holoenzyme. Nature 2018, 557, 190. [CrossRef] 45. Lermontova, I.; Schubert, V.; Bornke, F.; Macas, J.; Schubert, I. Arabidopsis CBF5 interacts with the H/ACA snoRNP assembly factor NAF1. Plant Mol. Biol. 2007, 65, 615–626. [CrossRef] 46. Pendle, A.F.; Clark, G.P.; Boon, R.; Lewandowska, D.; Lam, Y.W.; Andersen, J.; Mann, M.; Lamond, A.I.; Brown, J.W.S.; Shaw, P.J. Proteomic analysis of the Arabidopsis nucleolus suggests novel nucleolar functions. Mol. Biol. Cell 2005, 16, 260–269. [CrossRef] 47. Rossignol, P.; Collier, S.; Bush, M.; Shaw, P.; Doonan, J.H. Arabidopsis POT1A interacts with TERT-V(18), an N-terminal splicing variant of telomerase. J. Cell Sci. 2007, 120, 3678–3687. [CrossRef] 48. Nakamura, T.M.; Morin, G.B.; Chapman, K.B.; Weinrich, S.L.; Andrews, W.H.; Lingner, J.; Harley, C.B.; Cech, T.R. Telomerase catalytic subunit homologs from fission yeast and human. Science 1997, 277, 955–959. [CrossRef] 49. Oguchi, K.; Liu, H.T.; Tamura, K.; Takahashi, H. Molecular cloning and characterization of AtTERT, a telomerase reverse transcriptase homolog in Arabidopsis thaliana. Febs Lett. 1999, 457, 465–469. [CrossRef] 50. Baumann, P.; Cech, T.R. Pot1, the putative telomere end-binding protein in fission yeast and humans. Science 2001, 292, 1171–1175. [CrossRef] 51. 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. [CrossRef] 52. Liu, D.; Safari, A.; O’Connor, M.S.; Chan, D.W.; Laegeler, A.; Qin, J.; Zhou, S.Y. PTOP interacts with POT1 and regulates its localization to telomeres. Nat. Cell Biol. 2004, 6, 673–680. [CrossRef] 53. 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. [CrossRef] 54. Chen, L.Y.; Redon, S.; Lingner, J. The human CST complex is a terminator of telomerase activity. Nature 2012, 488, 540. [CrossRef] 55. Tani, A.; Murata, M. Alternative splicing of Pot1 (Protection of telomere)-like genes in Arabidopsis thaliana. Genes Genet. Syst. 2005, 80, 41–48. [CrossRef] 56. Cifuentes-Rojas, C.; Kannan, K.; Tseng, L.; Shippen, D.E. Two RNA subunits and POT1a are components of Arabidopsis telomerase. Proc. Natl. Acad. Sci. USA 2011, 108, 73–78. [CrossRef] Cells 2019, 8, 58 22 of 31 57. Kannan, K.; Nelson, A.D.L.; Shippen, D.E. Dyskerin is a component of the Arabidopsis telomerase RNP required for telomere maintenance. Mol. Cell. Biol. 2008, 28, 2332–2341. [CrossRef] 58. Arora, A.; Beilstein, M.A.; Shippen, D.E. Evolution of Arabidopsis protection of telomeres 1 alters nucleic acid recognition and telomerase regulation. Nucleic Acids Res. 2016, 44, 9821–9830. [CrossRef] 59. Van Steensel, B.; de Lange, T. Control of telomere length by the human telomeric protein TRF1. Nature 1997, 385, 740–743. [CrossRef] 60. Kim, S.H.; Kaminker, P.; Campisi, J. TIN2, a new regulator of telomere length in human cells. Nat. Genet. 1999, 23, 405–412. [CrossRef] 61. Ye, J.Z.S.; de Lange, T. TIN2 is a tankyrase 1 PARP modulator in the TRF1 telomere length control complex. Nat. Genet. 2004, 36, 618–623. [CrossRef] 62. Zhou, X.Z.; Lu, K.P. The Pin2/TRF1-interacting: Protein PinX1 is a potent telomerase inhibitor. Cell 2001, 107, 347–359. [CrossRef] 63. Wu, Y.; Xiao, S.; Zhu, X.D. MRE11-RAD50-NBS1 and ATM function as co-mediators of TRF1 in telomere length control. Nat. Struct. Mol. Biol. 2007, 14, 832–840. [CrossRef] 64. Schrumpfova, P.; Kuchar, M.; Mikova, G.; Skrisovska, L.; Kubicarova, T.; Fajkus, J. Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. Genome 2004, 47, 316–324. [CrossRef] 65. Schrumpfova, P.P.; Kuchar, M.; Palecek, J.; Fajkus, J. Mapping of interaction domains of putative telomere-binding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. Febs Lett. 2008, 582, 1400–1406. [CrossRef] 66. Schrumpfova, P.P.; Vychodilova, I.; Dvorackova, M.; Majerska, J.; Dokladal, L.; Schorova, S.; Fajkus, J. Telomere repeat binding proteins are functional components of Arabidopsis telomeres and interact with telomerase. Plant J. 2014, 77, 770–781. [CrossRef] 67. Schrumpfova, P.P.; Vychodilova, I.; Hapala, J.; Schorova, S.; Dvoracek, V.; Fajkus, J. Telomere binding protein TRB1 is associated with promoters of translation machinery genes in vivo. Plant Mol. Biol. 2016, 90, 189–206. [CrossRef] 68. Zhou, Y.; Wang, Y.J.; Krause, K.; Yang, T.T.; Dongus, J.A.; Zhang, Y.J.; Turck, F. Telobox motifs recruit CLF/SWN-PRC2 for H3K27me3 deposition via TRB factors in Arabidopsis. Nat. Genet. 2018, 50, 638. [CrossRef] 69. Dokladal, L.; Benkova, E.; Honys, D.; Dupl’akova, N.; Lee, L.Y.; Gelvin, S.B.; Sykorova, E. An armadillo-domain protein participates in a telomerase interaction network. Plant Mol. Biol. 2018, 97, 407–420. [CrossRef] 70. Lee, W.K.; Cho, M.H. Telomere-binding protein regulates the chromosome ends through the interaction with histone deacetylases in Arabidopsis thaliana. Nucleic Acids Res. 2016, 44, 4610–4624. [CrossRef] 71. Tan, L.M.; Zhang, C.J.; Hou, X.M.; Shao, C.R.; Lu, Y.J.; Zhou, J.X.; Li, Y.Q.; Li, L.; Chen, S.; He, X.J. The PEAT protein complexes are required for histone deacetylation and heterochromatin silencing. Embo J. 2018, 37, e98770. [CrossRef] [PubMed] 72. Van Steensel, B.; Smogorzewska, A.; de Lange, T. TRF2 protects human telomeres from end-to-end fusions. Cell 1998, 92, 401–413. [CrossRef] 73. 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. [CrossRef] [PubMed] 74. Rai, R.; Hu, C.; Broton, C.; Chen, Y.; Lei, M.; Chang, S. NBS1 Phosphorylation Status Dictates Repair Choice of Dysfunctional Telomeres. Mol. Cell 2017, 65, 801. [CrossRef] 75. O’Connor, M.S.; Safari, A.; Liu, D.; Qin, J.; Zhou, S.Y. The human Rap1 protein complex and modulation of telomere length. J. Biol. Chem. 2004, 279, 28585–28591. [CrossRef] [PubMed] 76. Chen, Y.; Yang, Y.T.; van Overbeek, M.; Donigian, J.R.; Baciu, P.; de Lange, T.; Lei, M. A shared docking motif in TRF1 and TRF2 used for differential recruitment of telomeric proteins. Science 2008, 319, 1092–1096. [CrossRef] [PubMed] 77. Song, K.; Jung, D.; Jung, Y.; Lee, S.G.; Lee, I. Interaction of human Ku70 with TRF2. Febs Lett. 2000, 481, 81–85. [CrossRef] 78. Gomez, M.; Wu, J.; Schreiber, V.; Dunlap, J.; Dantzer, F.; Wang, Y.S.; Liu, Y. PARP1 is a TRF2-associated poly(ADP-ribose)polymerase and protects eroded telomere. Mol. Biol. Cell 2006, 17, 1686–1696. [CrossRef] Cells 2019, 8, 58 23 of 31 79. Dantzer, F.; Giraud-Panis, M.J.; Jaco, I.; Ame, J.C.; Schultz, I.; Blasco, M.; Koering, C.E.; Gilson, E.; Menissier-de Murcia, J.; de Murcia, G.; et al. Functional interaction between poly(ADP-ribose) polymerase 2 (PARP-2) and TRF2: PARP activity negatively regulates TRF2. Mol. Cell. Biol. 2004, 24, 1595–1607. [CrossRef] 80. Wu, Y.L.; Mitchell, T.R.H.; Zhu, X.D. Human XPF controls TRF2 and telomere length maintenance through distinctive mechanisms. Mech. Ageing Dev. 2008, 129, 602–610. [CrossRef] 81. Chen, C.M.; Wang, C.T.; Ho, C.H. A plant gene encoding a Myb-like protein that binds telomeric GGTTTAG repeats in vitro. J. Biol. Chem. 2001, 276, 16511–16519. [CrossRef] 82. Kuchar, M.; Fajkus, J. Interactions of putative telomere-binding proteins in Arabidopsis thaliana: Identification of functional TRF2 homolog in plants. Febs Lett. 2004, 578, 311–315. [CrossRef] [PubMed] 83. Karamysheva, Z.N.; Surovtseva, Y.V.; Vespa, L.; Shakirov, E.V.; Shippen, D.E. A C-terminal Myb extension domain defines a novel family of double-strand telomeric DNA-binding proteins in Arabidopsis. J. Biol. Chem. 2004, 279, 47799–47807. [CrossRef] [PubMed] 84. Majerska, J.; Schrumpfova, P.P.; Dokladal, L.; Schorova, S.; Stejskal, K.; Oboril, M.; Honys, D.; Kozakova, L.; Polanska, P.S.; Sykorova, E. Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo. Protoplasma 2017, 254, 1547–1562. [CrossRef] 85. Jeong, S.A.; Kim, K.; Lee, J.H.; Cha, J.S.; Khadka, P.; Cho, H.S.; Chung, K. Akt-mediated phosphorylation increases the binding affinity of hTERT for importin alpha to promote nuclear translocation. J. Cell Sci. 2015, 128, 2287–2301. [CrossRef] [PubMed] 86. Khurts, S.; Masutomi, K.; Delgermaa, L.; Arai, K.; Oishi, N.; Mizuno, H.; Hayashi, N.; Hahn, W.C.; Murakami, S. Nucleolin interacts with telomerase. J. Biol. Chem. 2004, 279, 51508–51515. [CrossRef] [PubMed] 87. Pontvianne, F.; Abou-Ellail, M.; Douet, J.; Comella, P.; Matia, I.; Chandrasekhara, C.; DeBures, A.; Blevins, T.; Cooke, R.; Medina, F.J.; et al. Nucleolin Is Required for DNA Methylation State and the Expression of rRNA Gene Variants in Arabidopsis thaliana. PLoS Genet. 2010, 6, e1001225. [CrossRef] 88. Pontvianne, F.; Carpentier, M.C.; Durut, N.; Pavlistova, V.; Jaske, K.; Schorova, S.; Parrinello, H.; Rohmer, M.; Pikaard, C.S.; Fojtova, M.; et al. Identification of Nucleolus-Associated Chromatin Domains Reveals a Role for the Nucleolus in 3D Organization of the A. thaliana Genome. Cell Rep. 2016, 16, 1574–1587. [CrossRef] 89. Venteicher, A.S.; Meng, Z.J.; Mason, P.J.; Veenstra, T.D.; Artandi, S.E. Identification of ATPases pontin and reptin as telomerase components essential for holoenzyme assembly. Cell 2008, 132, 945–957. [CrossRef] 90. Holt, B.F.; Boyes, D.C.; Ellerstrom, M.; Siefers, N.; Wiig, A.; Kauffman, S.; Grant, M.R.; Dangl, J.L. An evolutionarily conserved mediator of plant disease resistance gene function is required for normal Arabidopsis development. Dev. Cell 2002, 2, 807–817. [CrossRef] 91. Giannone, R.J.; McDonald, H.W.; Hurst, G.B.; Shen, R.F.; Wang, Y.S.; Liu, Y. The Protein Network Surrounding the Human Telomere Repeat Binding Factors TRF1, TRF2, and POT1. PLoS ONE 2010, 5, e12407. [CrossRef] [PubMed] 92. Lee, L.Y.; Wu, F.H.; Hsu, C.T.; Shen, S.C.; Yeh, H.Y.; Liao, D.C.; Fang, M.J.; Liu, N.T.; Yen, Y.C.; Dokladal, L.; et al. Screening a cDNA Library for Protein-Protein Interactions Directly in Planta. Plant Cell 2012, 24, 1746–1759. [CrossRef] [PubMed] 93. Kappei, D.; Butter, F.; Benda, C.; Scheibe, M.; Draskovic, I.; Stevense, M.; Novo, C.L.; Basquin, C.; Araki, M.; Araki, K.; et al. HOT1 is a mammalian direct telomere repeat-binding protein contributing to telomerase recruitment. Embo J. 2013, 32, 1681–1701. [CrossRef] [PubMed] 94. Chai, W.H.; Ford, L.P.; Lenertz, L.; Wright, W.E.; Shay, J.W. Human Ku70/80 associates physically with telomerase through interaction with hTERT. J. Biol. Chem. 2002, 277, 47242–47247. [CrossRef] 95. Fell, V.L.; Schild-Poulter, C. The Ku heterodimer: Function in DNA repair and beyond. Mutat. Res. Rev. Mutat. Res. 2015, 763, 15–29. [CrossRef] [PubMed] 96. Bundock, P.; van Attikum, H.; Hooykaas, P. Increased telomere length and hypersensitivity to DNA damaging agents in an Arabidopsis KU70 mutant. Nucleic Acids Res. 2002, 30, 3395–3400. [CrossRef] [PubMed] 97. Riha, K.; Watson, J.M.; Parkey, J.; Shippen, D.E. Telomere length deregulation and enhanced sensitivity to genotoxic stress in Arabidopsis mutants deficient in Ku70. Embo J. 2002, 21, 2819–2826. [CrossRef] 98. West, C.E.; Waterworth, W.M.; Story, G.W.; Sunderland, P.A.; Jiang, Q.; Bray, C.M. Disruption of the Arabidopsis AtKu80 gene demonstrates an essential role for AtKu80 protein in efficient repair of DNA double-strand breaks in vivo. Plant J. 2002, 31, 517–528. [CrossRef] Cells 2019, 8, 58 24 of 31 99. Gallego, M.E.; Jalut, N.; White, C.I. Telomerase dependence of telomere lengthening in Ku80 mutant Arabidopsis. Plant Cell 2003, 15, 782–789. [CrossRef] 100. Cifuentes-Rojas, C.; Nelson, A.D.L.; Boltz, K.A.; Kannan, K.; She, X.T.; Shippen, D.E. An alternative telomerase RNA in Arabidopsis modulates enzyme activity in response to DNA damage. Genes Dev. 2012, 26, 2512–2523. [CrossRef] 101. Valuchova, S.; Fulnecek, J.; Prokop, Z.; Stolt-Bergner, P.; Janouskova, E.; Hofr, C.; Riha, K. Protection of Arabidopsis blunt-ended telomeres is mediated by a physical association with the Ku heterodimer. Plant Cell 2017. [CrossRef] [PubMed] 102. Holt, S.E.; Aisner, D.L.; Baur, J.; Tesmer, V.M.; Dy, M.; Ouellette, M.; Trager, J.B.; Morin, G.B.; Toft, D.O.; Shay, J.W.; et al. Functional requirement of p23 and Hsp90 in telomerase complexes. Genes Dev. 1999, 13, 817–826. [CrossRef] 103. Chen, B.; Zhong, D.B.; Monteiro, A. Comparative genomics and evolution of the HSP90 family of genes across all kingdoms of organisms. BMC Genom. 2006, 7, 156. 104. Zhang, Z.M.; Sullivan, W.; Felts, S.J.; Prasad, B.D.; Toft, D.O.; Krishna, P. Characterization of plant p23-like proteins for their co-chaperone activities. Cell Stress Chaperones 2010, 15, 703–715. [CrossRef] [PubMed] 105. Wortman, M.J.; Johnson, E.M.; Bergemann, A.D. Mechanism of DNA binding and localized strand separation by Pur alpha and comparison with Pur family member, Pur beta. Biochim. Biophys. Acta-Mol. Cell Res. 2005, 1743, 64–78. [CrossRef] [PubMed] 106. Mermoud, J.E.; Rowbotham, S.P.; Varga-Weisz, P.D. Keeping chromatin quiet How nucleosome remodeling restores heterochromatin after replication. Cell Cycle 2011, 10, 4017–4025. [CrossRef] 107. Dona, M.; Scheid, O.M. DNA Damage Repair in the Context of Plant Chromatin. Plant Physiol. 2015, 168, 1206–1218. [CrossRef] [PubMed] 108. Nguyen, D.; St-Sauveur, V.G.; Bergeron, D.; Dupuis-Sandoval, F.; Scott, M.S.; Bachand, F. A Polyadenylation-Dependent 3 End Maturation Pathway Is Required for the Synthesis of the Human Telomerase RNA. Cell Rep. 2015, 13, 2244–2257. [CrossRef] [PubMed] 109. Dokladal, L.; Honys, D.; Rana, R.; Lee, L.Y.; Gelvin, S.B.; Sykorova, E. cDNA Library Screening Identifies Protein Interactors Potentially Involved in Non-Telomeric Roles of Arabidopsis Telomerase. Front. Plant Sci. 2015, 6, 985. [CrossRef] 110. Ma, H.L.; Su, L.; Yue, H.W.; Yin, X.L.; Zhao, J.; Zhang, S.L.; Kung, H.F.; Xu, Z.G.; Miao, J.Y. HMBOX1 interacts with MT2A to regulate autophagy and apoptosis in vascular endothelial cells. Sci. Rep. 2015, 5, 15121. [CrossRef] [PubMed] 111. Feng, X.Y.; Luo, Z.H.; Jiang, S.; Li, F.; Han, X.; Hu, Y.; Wang, D.; Zhao, Y.; Ma, W.B.; Liu, D.; et al. The telomere-associated homeobox-containing protein TAH1/HMBOX1 participates in telomere maintenance in ALT cells. J. Cell Sci. 2013, 126, 3982–3989. [CrossRef] [PubMed] 112. Lamartine, J.; Seri, M.; Cinti, R.; Heitzmann, F.; Creaven, M.; Radomski, N.; Jost, E.; Lenoir, G.M.; Romeo, G.; Sylla, B.S. Molecular cloning and mapping of a human cDNA (PA2G4) that encodes a protein highly homologous to the mouse cell cycle protein p38-2G4. Cytogenet. Cell Genet. 1997, 78, 31–35. [CrossRef] [PubMed] 113. Feng, J.L.; Funk, W.D.; Wang, S.S.; Weinrich, S.L.; Avilion, A.A.; Chiu, C.P.; Adams, R.R.; Chang, E.; Allsopp, R.C.; Yu, J.H.; et al. The Rna Component of Human Telomerase. Science 1995, 269, 1236–1241. [CrossRef] [PubMed] 114. Cohen, S.B.; Graham, M.E.; Lovrecz, G.O.; Bache, N.; Robinson, P.J.; Reddel, R.R. Protein composition of catalytically active human telomerase from immortal cells. Science 2007, 315, 1850–1853. [CrossRef] 115. Heiss, N.S.; Knight, S.W.; Vulliamy, T.J.; Klauck, S.M.; Wiemann, S.; Mason, P.J.; Poustka, A.; Dokal, I. X-linked dyskeratosis congenita is caused by mutations in a highly conserved gene with putative nucleolar functions. Nat. Genet. 1998, 19, 32–38. [CrossRef] 116. Henras, A.; Henry, Y.; Bousquet-Antonelli, C.; Noaillac-Depeyre, J.; Gelugne, J.P.; Caizergues-Ferrer, M. Nhp2p and Nop10p are essential for the function of H/ACA snoRNPs. Embo J. 1998, 17, 7078–7090. [CrossRef] 117. Saito, H.; Fujiwara, T.; Shin, S.; Okui, K.; Nakamura, Y. Cloning and mapping of a human novel cDNA (NHP2L1) that encodes a protein highly homologous to yeast nuclear protein NHP2. Cytogenet. Cell Genet. 1996, 72, 191–193. [CrossRef] [PubMed] Cells 2019, 8, 58 25 of 31 118. Watkins, N.J.; Gottschalk, A.; Neubauer, G.; Kastner, B.; Fabrizio, P.; Mann, M.; Luhrmann, R. Cbf5p, a potential pseudouridine synthase, and Nhp2p, a putative RNA-binding protein, are present together with Gar1p in all H BOX/ACA-motif snoRNPs and constitute a common bipartite structure. RNA 1998, 4, 1549–1568. [CrossRef] 119. Fatica, A.; Dlakic, M.; Tollervey, D. Naf1p is a box H/ACA snoRNP assembly factor. RNA 2002, 8, 1502–1514. 120. Ting, N.S.Y.; Yu, Y.P.; Pohorelic, B.; Lees-Miller, S.P.; Beattie, T.L. Human Ku70/80 interacts directly with hTR, the RNA component of human telomerase. Nucleic Acids Res. 2005, 33, 2090–2098. [CrossRef] 121. Sexton, A.N.; Collins, K. The 5 Guanosine Tracts of Human Telomerase RNA Are Recognized by the G-Quadruplex Binding Domain of the RNA Helicase DHX36 and Function To Increase RNA Accumulation. Mol. Cell. Biol. 2011, 31, 736–743. [CrossRef] [PubMed] 122. Moon, D.H.; Segal, M.; Boyraz, B.; Guinan, E.; Hofmann, I.; Cahan, P.; Tai, A.K.; Agarwal, S. Poly(A)-specific ribonuclease (PARN) mediates 3 -end maturation of the telomerase RNA component. Nat. Genet. 2015, 47, 1482. [CrossRef] 123. Chiba, Y.; Johnson, M.A.; Lidder, P.; Vogel, J.T.; van Erp, H.; Green, P.J. AtPARN is an essential poly(A) ribonuclease in Arabidopsis. Gene 2004, 328, 95–102. [CrossRef] [PubMed] 124. Venteicher, A.S.; Abreu, E.B.; Meng, Z.J.; McCann, K.E.; Terns, R.M.; Veenstra, T.D.; Terns, M.P.; Artandi, S.E. A Human Telomerase Holoenzyme Protein Required for Cajal Body Localization and Telomere Synthesis. Science 2009, 323, 644–648. [CrossRef] [PubMed] 125. Sykorova, E.; Fajkus, J. Structure-function relationships in telomerase genes. Biol. Cell 2009, 101, 375–392. [CrossRef] 126. Sykorova, E.; Fulneckova, J.; Mokros, P.; Fajkus, J.; Fojtova, M.; Peska, V. Three TERT genes in Nicotiana tabacum. Chromosome Res. 2012, 20, 381–394. [CrossRef] [PubMed] 127. Chakrabarti, K.; Pearson, M.; Grate, L.; Sterne-Weiler, T.; Deans, J.; Donohue, J.P.; Ares, M. Structural RNAs of known and unknown function identified in malaria parasites by comparative genomics and RNA analysis. RNA 2007, 13, 1923–1939. [CrossRef] [PubMed] 128. Webb, C.J.; Zakian, V.A. Identification and characterization of the Schizosaccharomyces pombe TER1 telomerase RNA. Nat. Struct. Mol. Biol. 2008, 15, 34–42. [CrossRef] [PubMed] 129. Leonardi, J.; Box, J.A.; Bunch, J.T.; Baumann, P. TER1, the RNA subunit of fission yeast telomerase. Nat. Struct. Mol. Biol. 2008, 15, 26–33. [CrossRef] 130. Xie, M.Y.; Mosig, A.; Qi, X.; Li, Y.; Stadler, P.F.; Chen, J.J.L. Structure and function of the smallest vertebrate telomerase RNA from teleost fish. J. Biol. Chem. 2008, 283, 2049–2059. [CrossRef] 131. Kachouri-Lafond, R.; Dujon, B.; Gilson, E.; Westhof, E.; Fairhead, C.; Teixeira, M.T. Large telomerase RNA, telomere length heterogeneity and escape from senescence in Candida glabrata. Febs Lett. 2009, 583, 3605–3610. [CrossRef] 132. Gunisova, S.; Elboher, E.; Nosek, J.; Gorkovoy, V.; Brown, Y.; Lucier, J.F.; Laterreur, N.; Wellinger, R.J.; Tzfati, Y.; Tomaska, L. Identification and comparative analysis of telomerase RNAs from Candida species reveal conservation of functional elements. RNA-A Publ. RNA Soc. 2009, 15, 546–559. [CrossRef] 133. Waldl, M.; Thiel, B.C.; Ochsenreiter, R.; Holzenleiter, A.; de Araujo Oliveira, J.V.; Walter, M.; Wolfinger, M.T.; Stadler, P.F. TERribly Difficult: Searching for Telomerase RNAs in Saccharomycetes. Genes (Basel) 2018, 9, 372. [CrossRef] [PubMed] 134. Sykorova, E.; Lim, K.Y.; Kunicka, Z.; Chase, M.W.; Bennett, M.D.; Fajkus, J.; Leitch, A.R. Telomere variability in the monocotyledonous plant order Asparagales. Proc. R. Soc. B-Biol. Sci. 2003, 270, 1893–1904. [CrossRef] [PubMed] 135. Sykorova, E.; Fajkus, J.; Meznikova, M.; Lim, K.Y.; Neplechova, K.; Blattner, F.R.; Chase, M.W.; Leitch, A.R. Minisatellite telomeres occur in the family Alliaceae but are lost in Allium. Am. J. Bot. 2006, 93, 814–823. [CrossRef] [PubMed] 136. De Lange, T. Shelterin: The protein complex that shapes and safeguards human telomeres. Genes Dev. 2005, 19, 2100–2110. [CrossRef] 137. De Lange, T. What I got wrong about shelterin. J. Biol. Chem. 2018, 293, 10453–10456. [CrossRef] 138. Palm, W.; de Lange, T. How Shelterin Protects Mammalian Telomeres. Annu. Rev. Genet. 2008, 42, 301–334. [CrossRef] 139. Sfeir, A.; de Lange, T. Removal of Shelterin Reveals the Telomere End-Protection Problem. Science 2012, 336, 593–597. [CrossRef] Cells 2019, 8, 58 26 of 31 140. Kibe, T.; Zimmermann, M.; de Lange, T. TPP1 Blocks an ATR-Mediated Resection Mechanism at Telomeres. Mol. Cell 2016, 61, 236–246. [CrossRef] 141. Zimmermann, M.; Lottersberger, F.; Buonomo, S.B.; Sfeir, A.; de Lange, T. 53BP1 Regulates DSB Repair Using Rif1 to Control 5 ‘ End Resection. Science 2013, 339, 700–704. [CrossRef] [PubMed] 142. Dalby, A.B.; Hofr, C.; Cech, T.R. Contributions of the TEL-patch Amino Acid Cluster on TPP1 to Telomeric DNA Synthesis by Human Telomerase. J. Mol. Biol. 2015, 427, 1291–1303. [CrossRef] [PubMed] 143. Latrick, C.M.; Cech, T.R. POT1-TPP1 enhances telomerase processivity by slowing primer dissociation and aiding translocation. Embo J. 2010, 29, 924–933. [CrossRef] [PubMed] 144. Nandakumar, J.; Bell, C.F.; Weidenfeld, I.; Zaug, A.J.; Leinwand, L.A.; Cech, T.R. The TEL patch of telomere protein TPP1 mediates telomerase recruitment and processivity. Nature 2012, 492, 285. [CrossRef] [PubMed] 145. Schmidt, J.C.; Cech, T.R. Human telomerase: Biogenesis, trafficking, recruitment, and activation. Genes Dev. 2015, 29, 1095–1105. [CrossRef] [PubMed] 146. 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. [CrossRef] 147. Stansel, R.M.; de Lange, T.; Griffith, J.D. T-loop assembly in vitro involves binding of TRF2 near the 3’ telomeric overhang. Embo J. 2001, 20, 5532–5540. [CrossRef] 148. Sfeir, A.; Kosiyatrakul, S.T.; Hockemeyer, D.; MacRae, S.L.; Karlseder, J.; Schildkraut, C.L.; de Lange, T. Mammalian Telomeres Resemble Fragile Sites and Require TRF1 for Efficient Replication. Cell 2009, 138, 90–103. [CrossRef] 149. Tong, A.S.; Stern, J.L.; Sfeir, A.; Kartawinata, M.; de Lange, T.; Zhu, X.D.; Bryan, T.M. ATM and ATR Signaling Regulate the Recruitment of Human Telomerase to Telomeres. Cell Rep. 2015, 13, 1633–1646. [CrossRef] 150. 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. [CrossRef] 151. Mozgova, I.; Schrumpfova, P.P.; Hofr, C.; Fajkus, J. Functional characterization of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. Phytochemistry 2008, 69, 1814–1819. [CrossRef] [PubMed] 152. Marian, C.O.; Bordoli, S.J.; Goltz, M.; Santarella, R.A.; Jackson, L.P.; Danilevskaya, O.; Beckstette, M.; Meeley, R.; Bass, H.W. The maize Single myb histone 1 gene, Smh1, belongs to a novel gene family and encodes a protein that binds telomere DNA repeats in vitro. Plant Physiol. 2003, 133, 1336–1350. [CrossRef] 153. Bilaud, T.; Koering, C.E.; BinetBrasselet, E.; Ancelin, K.; Pollice, A.; Gasser, S.M.; Gilson, E. The telobox, a Myb-related telomeric DNA binding motif found in proteins from yeast, plants and human. Nucleic Acids Res. 1996, 24, 1294–1303. [CrossRef] 154. Peska, V.; Schrumpfova, P.P.; Fajkus, J. Using the Telobox to Search for Plant Telomere Binding Proteins. Curr. Protein Pept. Sci. 2011, 12, 75–83. [CrossRef] [PubMed] 155. Schrumpfova, P.P.; Schorova, S.; Fajkus, J. Telomere- and Telomerase-Associated Proteins and Their Functions in the Plant Cell. Front. Plant Sci. 2016, 7, 851. 156. Zhou, Y.; Hartwig, B.; James, G.V.; Schneeberger, K.; Turck, F. Complementary Activities of TELOMERE REPEAT BINDING Proteins and Polycomb Group Complexes in Transcriptional Regulation of Target Genes. Plant Cell 2016, 28, 87–101. [CrossRef] 157. El Mai, M.; Wagner, K.D.; Michiels, J.F.; Ambrosetti, D.; Borderie, A.; Destree, S.; Renault, V.; Djerbi, N.; Giraud-Panis, M.J.; Gilson, E.; et al. The Telomeric Protein TRF2 Regulates Angiogenesis by Binding and Activating the PDGFR beta Promoter. Cell Rep. 2014, 9, 1047–1060. [CrossRef] 158. Krutilina, R.I.; Oei, S.L.; Buchlow, G.; Yau, P.M.; Zalensky, A.O.; Zalenskaya, I.A.; Bradbury, E.M.; Tomilin, N.V. A negative regulator of telomere-length protein TRF1 is associated with interstitial (TTAGGG)n blocks in immortal Chinese hamster ovary cells. Biochem. Biophys. Res. Commun. 2001, 280, 471–475. [CrossRef] 159. Martinez, P.; Thanasoula, M.; Carlos, A.R.; Gomez-Lopez, G.; Tejera, A.M.; Schoeftner, S.; Dominguez, O.; Pisano, D.G.; 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. [CrossRef] 160. Morse, R.H. RAP, RAP, open up! New wrinkles for RAP1 in yeast. Trends Genet. 2000, 16, 51–53. [CrossRef] 161. Rizzo, A.; Iachettini, S.; Salvati, E.; Zizza, P.; Maresca, C.; D’Angelo, C.; Benarroch-Popivker, D.; Capolupo, A.; del Gaudio, F.; Cosconati, S.; et al. SIRT6 interacts with TRF2 and promotes its degradation in response to DNA damage. Nucleic Acids Res. 2017, 45, 1820–1834. [CrossRef] Cells 2019, 8, 58 27 of 31 162. Simonet, T.; Zaragosi, L.E.; Philippe, C.; Lebrigand, K.; Schouteden, C.; Augereau, A.; Bauwens, S.; Ye, J.; Santagostino, M.; Giulotto, E.; et al. The human TTAGGG repeat factors 1 and 2 bind to a subset of interstitial telomeric sequences and satellite repeats. Cell Res. 2011, 21, 1028–1038. [CrossRef] 163. Ye, J.; Renault, V.M.; Jamet, K.; Gilson, E. Transcriptional outcome of telomere signalling. Nat. Rev. Genet. 2014, 15, 491–503. [CrossRef] [PubMed] 164. Zhang, P.; Pazin, M.J.; Schwartz, C.M.; Becker, K.G.; Wersto, R.P.; Dilley, C.M.; Mattson, M.P. Nontelomeric TRF2-REST Interaction Modulates Neuronal Gene Silencing and Fate of Tumor and Stem Cells. Curr. Biol. 2008, 18, 1489–1494. [CrossRef] [PubMed] 165. Fulcher, N.; Riha, K. Using Centromere Mediated Genome Elimination to Elucidate the Functional Redundancy of Candidate Telomere Binding Proteins in Arabidopsis thaliana. Front. Genet. 2016, 6, 349. [CrossRef] 166. Perrault, S.D.; Hornsby, P.J.; Betts, D.H. Global gene expression response to telomerase in bovine adrenocortical cells. Biochem. Biophys. Res. Commun. 2005, 335, 925–936. [CrossRef] 167. Majerska, J.; Sykorova, E.; Fajkus, J. Non-telomeric activities of telomerase. Mol. Biosyst. 2011, 7, 1013–1023. [CrossRef] [PubMed] 168. Park, J.I.; Venteicher, A.S.; Hong, J.Y.; Choi, J.; Jun, S.; Shkreli, M.; Chang, W.; Meng, Z.J.; Cheung, P.; Ji, H.; et al. Telomerase modulates Wnt signalling by association with target gene chromatin. Nature 2009, 460, 66–U77. [CrossRef] 169. Freeling, M. Bias in Plant Gene Content Following Different Sorts of Duplication: Tandem, Whole-Genome, Segmental, or by Transposition. Annu. Rev. Plant Biol. 2009, 60, 433–453. [CrossRef] 170. Mandakova, T.; Lysak, M.A. Chromosomal Phylogeny and Karyotype Evolution in x=7 Crucifer Species (Brassicaceae). Plant Cell 2008, 20, 2559–2570. [CrossRef] 171. Price, C.M.; Boltz, K.A.; Chaiken, M.F.; Stewart, J.A.; Beilstein, M.A.; Shippen, D.E. Evolution of CST function in telomere maintenance. Cell Cycle 2010, 9, 3157–3165. [CrossRef] [PubMed] 172. Feng, X.Y.; Hsu, S.J.; Bhattacharjee, A.; Wang, Y.Y.; Diao, J.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. [CrossRef] [PubMed] 173. Feng, X.Y.; Hsu, S.J.; Kasbek, C.; Chaiken, M.; Price, C.M. CTC1-mediated C-strand fill-in is an essential step in telomere length maintenance. Nucleic Acids Res. 2017, 45, 4281–4293. [CrossRef] [PubMed] 174. Stewart, J.A.; Wang, F.; Chaiken, M.F.; Kasbek, C.; Chastain, P.D.; Wright, W.E.; Price, C.M. Human CST promotes telomere duplex replication and general replication restart after fork stalling. Embo J. 2012, 31, 3537–3549. [CrossRef] [PubMed] 175. Bedoyan, J.K.; Lejnine, S.; Makarov, V.L.; Langmore, J.P. Condensation of rat telomere-specific nucleosomal arrays containing unusually short DNA repeats and histone H1. J. Biol. Chem. 1996, 271, 18485–18493. [CrossRef] [PubMed] 176. Lejnine, S.; Makarov, V.L.; Langmore, J.P. Conserved Nucleoprotein Structure at the Ends of Vertebrate and Invertebrate Chromosomes. Proc. Natl. Acad. Sci. USA 1995, 92, 2393–2397. [CrossRef] [PubMed] 177. Makarov, V.L.; Lejnine, S.; Bedoyan, J.; Langmore, J.P. Nucleosomal Organization of Telomere-Specific Chromatin in Rat. Cell 1993, 73, 775–787. [CrossRef] 178. Tommerup, H.; Dousmanis, A.; Delange, T. Unusual Chromatin in Human Telomeres. Mol. Cell. Biol. 1994, 14, 5777–5785. [CrossRef] [PubMed] 179. Fajkus, J.; Kovarik, A.; Kralovics, R.; Bezdek, M. Organization of Telomeric and Subtelomeric Chromatin in the Higher-Plant Nicotiana tabacum. Mol. Gen. Genet. 1995, 247, 633–638. [CrossRef] 180. Dejardin, J.; Kingston, R.E. Purification of Proteins Associated with Specific Genomic Loci. Cell 2009, 136, 175–186. [CrossRef] [PubMed] 181. Fajkus, J.; Trifonov, E.N. Columnar packing of telomeric nucleosomes. Biochem. Biophys. Res. Commun. 2001, 280, 961–963. [CrossRef] 182. Bosco, N.; de Lange, T. A TRF1-controlled common fragile site containing interstitial telomeric sequences. Chromosoma 2012, 121, 465–474. [CrossRef] [PubMed] 183. Sun, H.; Karow, J.K.; Hickson, I.D.; Maizels, N. The Bloom’s syndrome helicase unwinds G4 DNA. J. Biol. Chem. 1998, 273, 27587–27592. [CrossRef] [PubMed] Cells 2019, 8, 58 28 of 31 184. Muftuoglu, M.; Wong, H.K.; Imam, S.Z.; Wilson, D.M.; Bohr, V.A.; Opresko, P.L. Telomere repeat binding factor 2 interacts with base excision repair proteins and stimulates DNA synthesis by DNA polymerase beta. Cancer Res. 2006, 66, 113–124. [CrossRef] [PubMed] 185. Tatsumi, Y.; Ezura, K.; Yoshida, K.; Yugawa, T.; Narisawa-Saito, M.; Kiyono, T.; Ohta, S.; Obuse, C.; Fujita, M. Involvement of human ORC and TRF2 in pre-replication complex assembly at telomeres. Genes Cells 2008, 13, 1045–1059. [CrossRef] [PubMed] 186. Sarek, G.; Vannier, J.B.; Panier, S.; Petrini, J.H.J.; Boulton, S.J. TRF2 Recruits RTEL1 to Telomeres in S Phase to Promote T-Loop Unwinding. Mol. Cell 2015, 57, 622–635. [CrossRef] 187. Karlseder, J.; Hoke, K.; Mirzoeva, O.K.; Bakkenist, C.; Kastan, M.B.; Petrini, J.H.J.; de Lange, T. The telomeric protein TRF2 binds the ATM kinase and can inhibit the ATM-dependent DNA damage response. PLoS Biol. 2004, 2, 1150–1156. [CrossRef] 188. Hwang, M.G.; Chung, I.K.; Kang, B.G.; Cho, M.H. Sequence-specific binding property of Arabidopsis thaliana telomeric DNA binding protein 1 (AtTBP1). Febs Lett. 2001, 503, 35–40. [CrossRef] 189. Renfrew, K.B.; Song, X.Y.; Lee, J.R.; Arora, A.; Shippen, D.E. POT1a and Components of CST Engage Telomerase and Regulate Its Activity in Arabidopsis. PLoS Genet. 2014, 10, e1004738. [CrossRef] 190. Wyatt, H.D.M.; Tsang, A.R.; Lobb, D.A.; Beattie, T.L. Human Telomerase Reverse Transcriptase (hTERT) Q169 Is Essential for Telomerase Function In Vitro and In Vivo. PLoS ONE 2009, 4, e7176. [CrossRef] 191. Ganduri, S.; Lue, N.F. STN1-POLA2 interaction provides a basis for primase-pol alpha stimulation by human STN1. Nucleic Acids Res. 2017, 45, 9455–9466. [CrossRef] [PubMed] 192. Miyake, Y.; Nakamura, M.; Nabetani, A.; Shimamura, S.; Tamura, M.; Yonehara, S.; Saito, M.; Ishikawa, F. RPA-like Mammalian Ctc1-Stn1-Ten1 Complex Binds to Single-Stranded DNA and Protects Telomeres Independently of the Pot1 Pathway. Mol. Cell 2009, 36, 193–206. [CrossRef] 193. Derboven, E.; Ekker, H.; Kusenda, B.; Bulankova, P.; Riha, K. Role of STN1 and DNA Polymerase alpha in Telomere Stability and Genome-Wide Replication in Arabidopsis. PLoS Genet. 2014, 10, e1004682. [CrossRef] [PubMed] 194. Leehy, K.A.; Lee, J.R.; Song, X.Y.; Renfrew, K.B.; Shippen, D.E. MERISTEM DISORGANIZATION1 Encodes TEN1, an Essential Telomere Protein That Modulates Telomerase Processivity in Arabidopsis. Plant Cell 2013, 25, 1343–1354. [CrossRef] 195. Song, X.Y.; Leehy, K.; Warrington, R.T.; Lamb, J.C.; Surovtseva, Y.V.; Shippen, D.E. STN1 protects chromosome ends in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2008, 105, 19815–19820. [CrossRef] 196. Surovtseva, Y.V.; Churikov, D.; Boltz, K.A.; Song, X.Y.; 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. [CrossRef] [PubMed] 197. Yoo, H.H.; Kwon, C.; Lee, M.M.; Chung, I.K. Single-stranded DNA binding factor AtWHY1 modulates telomere length homeostasis in Arabidopsis. Plant J. 2007, 49, 442–451. [CrossRef] 198. Kwon, C.; Chung, I.K. Interaction of an Arabidopsis RNA-binding protein with plant single-stranded telomeric DNA modulates telomerase activity. J. Biol. Chem. 2004, 279, 12812–12818. [CrossRef] 199. Li, E.; Zhang, Y. DNA Methylation in Mammals. Cold Spring Harb. Perspect. Biol. 2014, 6, a019133. [CrossRef] 200. Lister, R.; Pelizzola, M.; Dowen, R.H.; Hawkins, R.D.; Hon, G.; Tonti-Filippini, J.; Nery, J.R.; Lee, L.; Ye, Z.; Ngo, Q.M.; et al. Human DNA methylomes at base resolution show widespread epigenomic differences. Nature 2009, 462, 315–322. [CrossRef] 201. Zhang, X. The epigenetic landscape of plants. Science 2008, 320, 489–492. [CrossRef] 202. Cokus, S.J.; Feng, S.H.; Zhang, X.Y.; Chen, Z.G.; Merriman, B.; Haudenschild, C.D.; Pradhan, S.; Nelson, S.F.; Pellegrini, M.; Jacobsen, S.E. Shotgun bisulphite sequencing of the Arabidopsis genome reveals DNA methylation patterning. Nature 2008, 452, 215–219. [CrossRef] [PubMed] 203. Ogrocka, A.; Polanska, P.; Majerova, E.; Janeba, Z.; Fajkus, J.; Fojtova, M. Compromised telomere maintenance in hypomethylated Arabidopsis thaliana plants. Nucleic Acids Res. 2014, 42, 2919–2931. [CrossRef] [PubMed] 204. Vrbsky, J.; Akimcheva, S.; Watson, J.M.; Turner, T.L.; Daxinger, L.; Vyskot, B.; Aufsatz, W.; Riha, K. siRNA-Mediated Methylation of Arabidopsis Telomeres. PLoS Genet. 2010, 6, e1000986. [CrossRef] [PubMed] 205. Majerova, E.; Fojtova, M.; Mozgova, I.; Bittova, M.; Fajkus, J. Hypomethylating drugs efficiently decrease cytosine methylation in telomeric DNA and activate telomerase without affecting telomere lengths in tobacco cells. Plant Mol. Biol. 2011, 77, 371–380. [CrossRef] [PubMed] Cells 2019, 8, 58 29 of 31 206. Majerova, E.; Mandakova, T.; Vu, G.T.H.; Fajkus, J.; Lysak, M.A.; Fojtova, M. Chromatin features of plant telomeric sequences at terminal vs. internal positions. Front. Plant Sci. 2014, 5, 593. [CrossRef] 207. Xie, X.Y.; Shippen, D.E. DDM1 guards against telomere truncation in Arabidopsis. Plant Cell Rep. 2018, 37, 501–513. [CrossRef] 208. Fojtova, M.; Fajkus, J. Epigenetic Regulation of Telomere Maintenance. Cytogenet. Genome Res. 2014, 143, 125–135. [CrossRef] 209. Fransz, P.; ten Hoopen, R.; Tessadori, F. Composition and formation of heterochromatin in Arabidopsis thaliana. Chromosome Res. 2006, 14, 71–82. [CrossRef] 210. Roudier, F.; Ahmed, I.; Berard, C.; Sarazin, A.; Mary-Huard, T.; Cortijo, S.; Bouyer, D.; Caillieux, E.; Duvernois-Berthet, E.; Al-Shikhley, L.; et al. Integrative epigenomic mapping defines four main chromatin states in Arabidopsis. Embo J. 2011, 30, 1928–1938. [CrossRef] [PubMed] 211. Schoeftner, S.; Blasco, M.A. A ‘higher order’ of telomere regulation: Telomere heterochromatin and telomeric RNAs. Embo J. 2009, 28, 2323–2336. [CrossRef] [PubMed] 212. Cubiles, M.D.; Barroso, S.; Vaquero-Sedas, M.I.; Enguix, A.; Aguilera, A.; Vega-Palas, M.A. Epigenetic features of human telomeres. Nucleic Acids Res. 2018, 46, 2347–2355. [CrossRef] [PubMed] 213. Rosenfeld, J.A.; Wang, Z.B.; Schones, D.E.; Zhao, K.; DeSalle, R.; Zhang, M.Q. Determination of enriched histone modifications in non-genic portions of the human genome. BMC Genom. 2009, 10, 143. [CrossRef] [PubMed] 214. O’Sullivan, R.J.; Kubicek, S.; Schreiber, S.L.; Karlseder, J. Reduced histone biosynthesis and chromatin changes arising from a damage signal at telomeres. Nat. Struct. Mol. Biol. 2010, 17, 1218. [CrossRef] [PubMed] 215. Arnoult, N.; Van Beneden, A.; Decottignies, A. Telomere length regulates TERRA levels through increased trimethylation of telomeric H3K9 and HP1 alpha. Nat. Struct. Mol. Biol. 2012, 19, 948–956. [CrossRef] [PubMed] 216. Garcia-Cao, M.; O’Sullivan, R.; Peters, A.H.F.M.; Jenuwein, T.; Blasco, M.A. Epigenetic regulation of telomere length in mammalian cells by the Suv39h1 and Suv39h2 histone methyltransferases. Nat. Genet. 2004, 36, 94–99. [CrossRef] 217. Saksouk, N.; Barth, T.K.; Ziegler-Birling, C.; Olova, N.; Nowak, A.; Rey, E.; Mateos-Langerak, J.; Urbach, S.; Reik, W.; Torres-Padilla, M.E.; et al. Redundant Mechanisms to Form Silent Chromatin at Pericentromeric Regions Rely on BEND3 and DNA Methylation. Mol. Cell 2014, 56, 580–594. [CrossRef] 218. Benetti, R.; Gonzalo, S.; Jaco, I.; SChotta, G.; Klatt, P.; Jenuwein, T.; Blasco, M.A. Suv4-20h deficiency results in telomere elongation and derepression of telomere recombination. J. Cell Biol. 2007, 178, 925–936. [CrossRef] 219. Gonzalo, S.; Jaco, I.; Fraga, M.F.; Chen, T.P.; Li, E.; Esteller, M.; Blasco, M.A. DNA methyltransferases control telomere length and telomere recombination in mammalian cells. Nat. Cell Biol. 2006, 8, 416. [CrossRef] 220. Montero, J.J.; Lopez-Silanes, I.; Megias, D.; Fraga, M.F.; Castells-Garcia, A.; Blasco, M.A. TERRA recruitment of polycomb to telomeres is essential for histone trymethylation marks at telomeric heterochromatin. Nat. Commun. 2018, 9, 1548. [CrossRef] 221. Sovakova, P.P.; Magdolenova, A.; Konecna, K.; Rajecka, V.; Fajkus, J.; Fojtova, M. Telomere elongation upon transfer to callus culture reflects the reprogramming of telomere stability control in Arabidopsis. Plant Mol. Biol. 2018, 98, 81–99. [CrossRef] 222. Vaquero-Sedas, M.I.; Luo, C.Y.; Vega-Palas, M.A. Analysis of the epigenetic status of telomeres by using ChIP-seq data. Nucleic Acids Res. 2012, 40, e163. [CrossRef] [PubMed] 223. Bulut-Karslioglu, A.; Perrera, V.; Scaranaro, M.; de la Rosa-Velazquez, I.A.; van de Nobelen, S.; Shukeir, N.; Popow, J.; Gerle, B.; Opravil, S.; Pagani, M.; et al. A transcription factor-based mechanism for mouse heterochromatin formation. Nat. Struct. Mol. Biol. 2012, 19, 1023. [CrossRef] [PubMed] 224. 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. [CrossRef] 225. Vaquero-Sedas, M.I.; Gamez-Arjona, F.M.; Vega-Palas, M.A. Arabidopsis thaliana telomeres exhibit euchromatic features. Nucleic Acids Res. 2011, 39, 2007–2017. [CrossRef] [PubMed] 226. Tardat, M.; Dejardin, J. Telomere chromatin establishment and its maintenance during mammalian development. Chromosoma 2018, 127, 3–18. [CrossRef] 227. Chai, W.H.; 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. [CrossRef] Cells 2019, 8, 58 30 of 31 228. Cimino-Reale, G.; Pascale, E.; Battiloro, E.; Starace, G.; Verna, R.; D’Ambrosio, E. The length of telomeric G-rich strand 3 ‘-overhang measured by oligonucleotide ligation assay. Nucleic Acids Res. 2001, 29, e35. [CrossRef] 229. 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. [CrossRef] 230. Wright, W.E.; Tesmer, V.M.; Huffman, K.E.; Levene, S.D.; Shay, J.W. Normal human chromosomes have long G-rich telomeric overhangs at one end. Genes Dev. 1997, 11, 2801–2809. [CrossRef] 231. Oganesian, L.; Karlseder, J. Mammalian 5’ C-Rich Telomeric Overhangs Are a Mark of Recombination-Dependent Telomere Maintenance. Mol. Cell 2011, 42, 224–236. [CrossRef] 232. Riha, K.; McKnight, T.D.; Fajkus, J.; Vyskot, B.; Shippen, D.E. Analysis of the G-overhang structures on plant telomeres: Evidence for two distinct telomere architectures. Plant J. 2000, 23, 633–641. [CrossRef] 233. Kazda, A.; Zellinger, B.; Rossler, M.; Derboven, E.; Kusenda, B.; Riha, K. Chromosome end protection by blunt-ended telomeres. Genes Dev. 2012, 26, 1703–1713. [CrossRef] [PubMed] 234. Fojtova, M.; Sykorova, E.; Najdekrova, L.; Polanska, P.; Zachova, D.; Vagnerova, R.; Angelis, K.J.; Fajkus, J. Telomere dynamics in the lower plant Physcomitrella patens. Plant Mol. Biol. 2015, 87, 591–601. [CrossRef] 235. Riha, K.; Shippen, D.E. Ku is required for telomeric C-rich strand maintenance but not for end-to-end chromosome fusions in Arabidopsis. Proc. Natl. Acad. Sci. USA 2003, 100, 611–615. [CrossRef] 236. Bryan, T.M.; Englezou, A.; DallaPozza, L.; Dunham, M.A.; Reddel, R.R. Evidence for an alternative mechanism for maintaining telomere length in human tumors and tumor-derived cell lines. Nat. Med. 1997, 3, 1271–1274. [CrossRef] 237. Neumann, A.A.; Watson, C.M.; Noble, J.R.; Pickett, H.A.; Tam, P.P.L.; Reddel, R.R. Alternative lengthening of telomeres in normal mammalian somatic cells. Genes Dev. 2013, 27, 18–23. [CrossRef] [PubMed] 238. Ruckova, E.; Friml, J.; Schrumpfova, P.P.; Fajkus, J. Role of alternative telomere lengthening unmasked in telomerase knock-out mutant plants. Plant Mol. Biol. 2008, 66, 637–646. [CrossRef] 239. Zellinger, B.; Akimcheva, S.; Puizina, J.; Schirato, M.; Riha, K. Ku suppresses formation of telomeric circles and alternative telomere lengthening in Arabidopsis. Mol. Cell 2007, 27, 163–169. [CrossRef] [PubMed] 240. Karpenshif, Y.; Bernstein, K.A. From yeast to mammals: Recent advances in genetic control of homologous recombination. DNA Repair 2012, 11, 781–788. [CrossRef] 241. Barber, L.J.; Youds, J.L.; Ward, J.D.; McIlwraith, M.J.; O’Neil, N.J.; Petalcorin, M.I.R.; Martin, J.S.; Collis, S.J.; Cantor, S.B.; Auclair, M.; et al. RTEL1 Maintains Genomic Stability by Suppressing Homologous Recombination. Cell 2008, 135, 261–271. [CrossRef] 242. Uringa, E.J.; Lisaingo, K.; Pickett, H.A.; Brind’Amour, J.; Rohde, J.H.; Zelensky, A.; Essers, J.; Lansdorp, P.M. RTEL1 contributes to DNA replication and repair and telomere maintenance. Mol. Biol. Cell 2012, 23, 2782–2792. [CrossRef] [PubMed] 243. Vannier, J.B.; Pavicic-Kaltenbrunner, V.; Petalcorin, M.I.R.; Ding, H.; Boulton, S.J. RTEL1 Dismantles T Loops and Counteracts Telomeric G4-DNA to Maintain Telomere Integrity. Cell 2012, 149, 795–806. [CrossRef] [PubMed] 244. Le Guen, T.; Jullien, L.; Schertzer, M.; Lefebvre, A.; Kermasson, L.; de Villartay, J.P.; Londono-Vallejo, A.; Revy, P. RTEL1 (regulator of telomere elongation helicase 1), a DNA helicase essential for genome stability. Med. Sci. 2013, 29, 1138–1144. 245. Vannier, J.B.; Sarek, G.; Boulton, S.J. RTEL1: Functions of a disease-associated helicase. Trends Cell Biol. 2014, 24, 416–425. [CrossRef] [PubMed] 246. Faure, G.; Revy, P.; Schertzer, M.; Londono-Vallejo, A.; Callebaut, I. The C-terminal extension of human RTEL1, mutated in Hoyeraal-Hreidarsson syndrome, contains Harmonin-N-like domains. Proteins-Struct. Funct. Bioinform. 2014, 82, 897–903. [CrossRef] [PubMed] 247. Margalef, P.; Kotsantis, P.; Borel, V.; Bellelli, R.; Panier, S.; Boulton, S.J. Stabilization of Reversed Replication Forks by Telomerase Drives Telomere Catastrophe. Cell 2018, 172, 439. [CrossRef] 248. Hu, Z.B.; Cools, T.; Kalhorzadeh, P.; Heyman, J.; De Veylder, L. Deficiency of the Arabidopsis Helicase RTEL1 Triggers a SOG1-Dependent Replication Checkpoint in Response to DNA Cross-Links. Plant Cell 2015, 27, 149–161. [CrossRef] 249. Recker, J.; Knoll, A.; Puchta, H. The Arabidopsis thaliana Homolog of the Helicase RTEL1 Plays Multiple Roles in Preserving Genome Stability. Plant Cell 2014, 26, 4889–4902. [CrossRef] Cells 2019, 8, 58 31 of 31 250. Riha, K.; McKnight, T.D.; Griffing, L.R.; Shippen, D.E. Living with genome instability: Plant responses to telomere dysfunction. Science 2001, 291, 1797–1800. [CrossRef] 251. Olivier, M.; Charbonnel, C.; Amiard, S.; White, C.I.; Gallego, M.E. RAD51 and RTEL1 compensate telomere loss in the absence of telomerase. Nucleic Acids Res. 2018, 46, 2432–2445. [CrossRef] [PubMed] 252. Kamisugi, Y.; Whitaker, J.W.; Cuming, A.C. The Transcriptional Response to DNA-Double-Strand Breaks in Physcomitrella patens. PLoS ONE 2016, 11, e0161204. [CrossRef] [PubMed] 253. Olsson, M.; Wapstra, E.; Friesen, C. Ectothermic telomeres: it’s time they came in from the cold. Philos. Trans. R. Soc. B-Biol. Sci. 2018, 373, 20160449. [CrossRef] [PubMed] 254. Hoelzl, F.; Smith, S.; Cornils, J.S.; Aydinonat, D.; Bieber, C.; Ruf, T. Telomeres are elongated in older individuals in a hibernating rodent, the edible dormouse (Glis glis). Sci. Rep. 2016, 6, 36856. [CrossRef] 255. Gomes, N.M.V.; Ryder, O.A.; Houck, M.L.; Charter, S.J.; Walker, W.; Forsyth, N.R.; Austad, S.N.; Venditti, C.; Pagel, M.; Shay, J.W.; et al. Comparative biology of mammalian telomeres: Hypotheses on ancestral states and the roles of telomeres in longevity determination. Aging Cell 2011, 10, 761–768. [CrossRef] [PubMed] 256. Ahmed, W.; Lingner, J. Impact of oxidative stress on telomere biology. Differentiation 2018, 99, 21–27. [CrossRef] [PubMed] 257. Zhang, J.W.; Rane, G.; Dai, X.Y.; Shanmugam, M.K.; Arfuso, F.; Samy, R.P.; Lai, M.K.P.; Kappei, D.; Kumar, A.P.; Sethi, G. Ageing and the telomere connection: An intimate relationship with inflammation. Ageing Res. Rev. 2016, 25, 55–69. [CrossRef] [PubMed] 258. Bottcher, M.A.; Dingli, D.; Werner, B.; Traulsen, A. Replicative cellular age distributions in compartmentalized tissues. J. R. Soc. Interface 2018, 15, 20180272. [CrossRef] 259. Cairns, J. Mutation Selection and Natural-History of Cancer. Nature 1975, 255, 197–200. [CrossRef] 260. Conboy, M.J.; Karasov, A.O.; Rando, T.A. High incidence of non-random template strand segregation and asymmetric fate determination in dividing stem cells and their progeny. PLoS Biol. 2007, 5, 1120–1126. © 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/). Supplement N Schořová, Š., Fajkus, J., Drábková, L.Z., Honys, D., Schrumpfová, P.P.*, 2019. The plant Pontin and Reptin homologues, RuvBL1 and RuvBL2a, colocalize with TERT and TRB proteins in vivo, and participate in telomerase biogenesis. Plant J. 98, 195–212 P.P.S. participated in the design of experiments, data evaluation and wrote the ms volume 98 | number 2 | April 2019 http://www.theplantjournal.com | issn 1365-313X tpj_v98_i2_issueinfo.indd 1tpj_v98_i2_issueinfo.indd 1 13-Apr-19 7:36:20 PM13-Apr-19 7:36:20 PM Lee Sweetlove Department of Plant Sciences, University of Oxford, South Parks Road, Oxford OX1 3RB, UK Tel: +44 (0) 1865 275137 Email: lee.sweetlove@plants. ox.ac.uk Research Interests: Central metabolism; Metabolic flux; Metabolic modeling; Metabolic engineering; Metabolons and metabolite channelling Previous Editors-in-Chief Dianna Bowles (Founding EiC), 1990-2001 Harry Klee, 2002-2009 Christoph Benning, 2009-2017 The Plant Journal Editorial Office, Wiley Blackwell, 9600 Garsington Road, Oxford, OX4 2DQ, UK E-mail: tpj-general@wiley.com Jim Ruddock Managing Editor E-mail: tpj-editor@wiley.com Lauren Dawson Editorial Assistant E-mail: tpj-submissions@wiley.com, tpj-manuscripts@wiley.com Rajalakshmi Sundararamanujam Production Editor E-mail: tpj@wiley.com Editor-in-Chief Editorial Office Production Office Senior Editors Editors Asaph Aharoni Weizmann Institute of Science, Department of Plant Sciences, P.O.B 26, Rehovot 76100, Israel Tel: +972 8 934 3643; Fax: +972 8 934 4181 Email: asaph.aharoni@weizmann.ac.il Research Interests Fruit metabolism and ripening control, The plant surface (epidermis and the cuticle), Secondary metabolism, Functional genomics and metabolomics, Metabolic engineering, Metabolism – development interface Jose Alonso Department of Genetics, North Carolina State University, 2548 Gardner Hall, Box 7614, Raleigh, North Carolina, 27695-7614, USA Tel: +1 919 515-5729; Fax: +1 919 515-3355 Email: jmalonso@unity.ncsu.edu Research Interests Ethylene signaling and response pathway; Hormone interactions and signal integration; Arabidopsis functional genomics and tool development; Auxin biosynthesis; Posttranscriptional control of hormonal responses Bonnie Bartel Department of Biosciences, Rice University, 6100 Main Street, MS-140, Houston,Texas 77005, USA Tel: +1 713 348 5602; Fax: +1 713 348 5154 Email:bartel@rice.edu Research Interests Cell biology; Peroxisome biology; Autophagy; Protein trafficking; Arabidopsis genetics Jörg Bohlmann Michael Smith Laboratories, University of British Columbia, 321-2185 East Mall, Vancouver V6T 1Z4 British Columbia, Canada Tel: +1 604 822 0282; Fax: +1 604 822 2114 Email: bohlmann@interchange.ubc.ca Research Interests Secondary metabolism;Terpenoid synthases, cytochrome P450; Plant-herbivore interactions; Chemical ecology; Conifers, Poplar, Grapevine Federica Brandizzi Michigan State University, Plant Research Laboratory, East Lansing, MI 48824, USA Tel: +1517 353-7872; Fax: +1 517 353-9168 Email: fb@msu.edu Research Interests Cell biology; Protein trafficking; Plant endomembranes; Endoplasmic reticulum stress; Fluorescent protein technology Jorge J. Casal Faculty of Agronomy, University of Buenos Aires, Av. San Martín 4453, 1417-Buenos Aires, Argentina Tel: +5411 4524 8070/71; Fax: +5411 4514 8730 Email: casal@ifeva.edu.ar Research Interests Light signaling circuitry, its architecture and functional implications; Functional genomics, transcriptome patterns during plant development; Photomorphogenesis in crops John Cushman University of Nevada, Department of Biochemistry, Mail Stop 330, 1664 N. Virginia St., Reno89557-0330, USA Email: jcushman@unr.edu Research Interests Abiotic stress responses; Metabolic adaptation to stress; Crassulacean acid metabolism; Biosystem engineering Brendan Davies Centre for Plant Sciences, University of Leeds, Leeds LS2 9JT, UK Tel: +44 113 343 3123; Fax: +44 113 343 3144 Email: b.h.davies@leeds.ac.uk Research Interests Plant development; Reproduction;Transcription factors; Nonsense-mediated mRNA decay Katherine Denby University ofYork, Department of BiologyYork,YO10 5DD, UK Tel: +44 1904 328 670 Research Interests Plant systems biology; Plant pathogen interaction; Functional genomics; Gene regulation; Gene regulatory networks; Crop breeding Alisdair R. Fernie Molecular Physiology, Max-Planck Institute for Molecular Plant Physiology,Golm 14476, Germany Tel: +49 331 567 8211 Email:fernie@mpimp-golm.mpg.de Research Interests Carbohydrate metabolism; Metabolomics; Primary metabolism Jaume Flexas Universitat de les Illes Balears, Biologia, Carretera de Valldemossa Km 7.5, Palma de Mallorca, 07122, Spain Tel: +34 971 172365; Fax: +34 971 173184 Email: jaume.flexas@uib.es Research Interests Photosynthesis; Ecophysiology; Abiotic stress; Mesophyll conductance; Drought Crisanto Gutierrez Centro de Biologia Molecular Severo Ochoa, Consejo Superior de Investigaciones Cientificas, Nicolas Cabrera 1, Cantoblanco, Madrid, 28049, Spain Tel: +34 911964638; Fax: +34 911964420 Email: cgutierrez@cbm.uam.es Research Interests Cell cycle regulation, DNA replication/endoreplication, Chromatin/Gene expression, Retinoblastoma/E2F, Cell fate, cell differentiation – Meristems, cell proliferation in organs Jose Gutierrez-Marcos Life Sciences, University of Warwick, Coventry, CV4 7AL, UK Tel: +44 247 6 57 5077 Email: j.f.gutierrez-marcos@warwick.ac.uk Research Interests Epigenetics; Seed development Scott Jackson Institute for Plant Breeding, Genetics and Genomics, University of Georgia , 111 Riverbend Rd, Athens, Georgia, GA 30602, USA Tel: +1 706.242.4021; Fax: +1 706-583-8120 Email: sjackson@uga.edu Research Interests Genomics; Cytogenetics; Evolutionary genomics; Legumes; Polyploidy Kerstin Kaufmann Humboldt-Universität zu Berlin, Chair for Plant Cell and Molecular Biology, Philippstr. 13, 10115 Berlin, Germany Tel: +49(0)30209349740; Fax: +49(0)30209349741 Email: kerstin.kaufmann@hu-berlin.de Research Interests Gene regulatory networks; Flower development; Plant transcription factors; Epigenetics; Transcriptional regulation; Flowering Tracy Lawson University of Essex, School of Biological Siences, Wivenhoe Park, Colchester, Essex CO4 3SQ, UK Tel: +44 1206 873327; Fax: +44 1206 872552 Email: tlawson@essex.ac.uk Research Interests Plant physiology; Plant phenotyping; Conductance; Gas exchange; Water use efficiency Olivier Loudet INRA - IJPB, Route de Saint Cyr, 78000 Versailles, France Email: olivier.loudet@versailles.inra.fr Research Interests Natural variation; Quantitative genetics; Abiotic stress tolerance; High-throughput phenotyping; Arabidopsis genetics Jianxin Ma Purdue University, 915 S. University St, West Lafayette, Indiana 47907, USA Tel: 765-496-3662; Fax: 765-496-7255; Email: maj@purdue.edu Research Interests ROS; Comparative genomics; Genome evolution;Transposable elements; Soybean genetics; Crop domestication Hiroshi Maeda University of Wisconsin – Madison, Department of Botany, 430 Lincoln Drive , Madison, Wisconsin 53706, USA Tel: +1 608-262-5833; Fax: +1 608-262-7509; Email: maeda2@wisc.edu Research Interests Amino acid metabolism; Specialized metabolism; Interaction of metabolism and plant development and physiology Ron Mittler University of NorthTexas College of Arts and Sciences, Department of Biological Sciences, College of Science, 1155 Union Circle #305220, Denton,Texas 76203-5017, USA Tel: +1 940 293 7170; Fax: +1 940 565 3821 Email: ron.mittler@unt.edu Research Interests ROS; Abiotic stress; Gene expression; Systems biology; Cancer biology Eiji Nambara University ofToronto, Cell & Systems Biology, 25 Willcocks Street,Toronto, Ontario M5S 3B2, Canada Email: eiji.nambara@utoronto.ca Research Interests Plant hormone; Biosynthesis; Catabolism; Abscisic acid; Seed; Stress response Holger Puchta Plant Molecular Biology and Biochemistry Karlsruhe Institute ofTechnology (KIT), Hertzstr. 16 Geb. 06.35, Karlsruhe D-76187, Germany Tel: +49 721 608 8894; Fax: +49 721 608 4874 Email: holger.puchta@kit.edu Research Interests DNA repair; Meiotic recombination; DNA processing enzymes; Genome evolution; Gene technology Dominique Roby Laboratory of Plant Microorganism Interactions (LIPM), UMR CNRS-INRA 2594/441, BP 52627, F-31326, Castanet-Tolosan cedex, France Tel: +33 05 61 28 55 11; Fax: +33 05 61 28 50 61 Email: dominique.roby@toulouse.inra.fr Research Interests Plant-microbe interactions; Genetic analyses of plant disease resistance pathways; Defence regulatory proteins; Programmed cell death and transcriptional regulators; Pathogen effector activities; Quantitative resistance to pathogens Ian Small University of Western Australia, ARC Centre of Excellence in Plant Energy Biology, 35 Stirling Highway, Crawley, Perth, 6009 Western Australia Tel: +61 (0)8 6488 4499; Fax: +61 (0)8 6488 4401 Email: iansmall@cyllene.uwa.edu.au Research Interests Organelle biogenesis; Gene expression in mitochondria and chloroplasts; Protein and RNA targeting; Post-transcriptional regulation of gene expression; Energy metabolism; Functional genomics, bioinformatics and systems biology Edgar Spalding Department of Botany, University of Wisconsin-Madison, 430 Lincoln Drive Madison WI 53706, USA Tel: +1 608 265 5294 Email: spalding@wisc.edu Research Interests Ion transport and electrophysiology Growth; control by light and auxin Phenotype platforms; computer vision-based Dorothea Tholl Department of Biological Sciences, Virginia Polytechnic Institute and State University,408 Latham Hall, Agquad Lane,Blacksburg, VA 24061, USA Tel: +1 540 231 4567; Fax: +1 540 231 3347 Email: tholl@vt.edu Research Interests Isoprenoid biochemistry, Secondary metabolism, Chemical diversity, Plant volatiles, Plant chemical interactions, Root biology, Metabolic pathways, Evolution, Engineering Michael F. Thomashow Michigan State University, MSU-DOE Plant Research Laboratory, Plant Biology Building, 312 Wilson Rd, East Lansing, MI 48824, USA Tel: 1 517 355 2299; Fax: 1 517 353 9168 Email: thomash6@msu.edu Research Interests Abiotic stress; Biotic stress Yves Van de Peer Department of Plant Biotechnology and Genetics, University of Gent,Technologiepark 927, 9052 Zwijnaarde, Belgium Tel: +09 331 38 07; Fax: +09 331 38 09 Email: yves.vandepeer@ugent.be Research Interests Comparative genomics, systems biology, evolutionary analysis, genome evolution, gene and genome duplication Saskia van Wees Utrecht University, Biology, Padualaan 8, Utrecht, 3584 CH, Netherlands Tel: +31-302536861; Fax: +31-302532837; Email:s.vanwees@uu.nl Research Interests Plant-microbe interactions; Plant-insect interactions; Defenses to pathogens and insects; Gene expression profiling; Gene expression regulation; Hormonal regulation of plant immunity Daoxin Xie School of Life Sciences,Tsinghua University, Beijing 100084, China Email: daoxinlab@tsinghua.edu.cn Research Interests Hormone perception; Jasmonate signal transduction; Strigolactone signaling; Hormonal crosstalk Bin Yu Center for Plant Science Innovation, & School of Biological Sciences, University of Nebraska, Lincoln, 1901 Vine St, Lincoln, NE, USA Tel: 402-472-2125; Email: byu3@unl.edu Research Interests Small RNA metabolism; Epigenetics; RNA silencing; Non-coding RNAs; Arabidopsis genetics http://www.theplantjournal.com ISSN 1365-313X (online) Federica Brandizzi Michigan State University, Plant Research Laboratory, East Lansing, MI 48824, USA Tel: (517) 353-7872; Fax: (517) 353-9168 Email: fb@msu.edu Research Interests Cell biology; Plant endomembranes; Endoplasmic reticulum stress; Fluorescent protein technology; Protein trafficking Alisdair R. Fernie Molecular Physiology, Max-Planck Institute for Molecular Plant Physiology, Golm 14476, Germany Tel: +49 331 567 8211 Email: fernie@mpimp-golm.mpg.de Research Interests Carbohydrate metabolism; Metabolomics; Primary metabolism Lyza Maron (Research highlights Editor) BoyceThompson Institute, USDA/ARS Robert W. Holley Center for Agriculture & Health, Ithaca, NY 14853, USA Research Interests Abiotic stress responses; Root biology; Natural variation; Gene expression regulation Sheila McCormick (Consultant Editor) Plant and Microbial Biology, 111 Koshland Hall, UC-Berkeley, Berkeley, CA 94720, USA Email: sheilamc@berkeley.edu Research Interests Pollen development; Plant gametes (sperm, egg, central cell); Pollen tube growth; Pollen-pistil interactions; Receptor kinase signaling (ligands); Genetics; Genomics tpj_v98_i2_issueinfo.indd 2tpj_v98_i2_issueinfo.indd 2 13-Apr-19 7:36:21 PM13-Apr-19 7:36:21 PM Subscription information The Plant Journal is published semi-monthly (four volumes per annum), with two issues published each month. Subscription prices for 2019 are: Online only: US$7274 (The Americas), US$8488 (Rest ofWorld), €4996 (Europe); £3936 (UK). Prices are exclusive of tax. Asia-Pacific GST, Canadian GST and European VAT will be applied at the appropriate rates. For more information on current tax rates, please go to www. wileyonlinelibrary.com/tax-vat. The price includes online access to the current and all online back files to January 1st 2015, where available. For other pricing options, including access information and terms and conditions, please visit www.wileyonlinelibrary.com/access. Back issues Single issues from current and prior year are available at the single issue price from cs-journals@wiley.com. Earlier issues may be obtained from the Periodicals Service Company, 11 Main Street, Germantown, NY 12526, USA. Tel: +1 518 537 4700, Fax: +1 518 537 5899, Email: psc@ periodicals.com. Publisher The Plant Journal is published by John Wiley & Sons Ltd, 9600 Garsington Road, Oxford OX4 2DQ. Tel: +44 (0)1865 776868; Fax: +44 (0)1865 714591. Journal Customer Services Fororderinginformation,claimsandanyenquiryconcerning your journal subscription please go to www.wiley customerhelp.com/ask or contact your nearest office: Americas: Email: cs-journals@wiley.com; Tel: +1 781 388 8598 or 1 800 835 6770 (Toll free in the USA & Canada). Europe, Middle East and Africa: E-mail: cs-journals@wiley. com;Tel: +44 (0) 1865 778315 Asia Pacific: E-mail: cs-journals@wiley.com; Tel: +65 6511 8000 Japan: For Japanese speaking support, e-mail: cs-japan@ wiley.com Visit our Online Customer Help available in 7 languages at www.wileycustomerhelp.com. Despatch THE PLANT JOURNAL, (ISSN 0960-7412), is published semi-monthly (four volumes per annum). US mailing agent: Mercury Media Processing, LLC 1850 Elizabeth Avenue, Suite #C, Rahway, NJ 07065, USA. Periodical postage paid at Rahway, NJ. Postmaster: Send all address changes to THE PLANT JOURNAL, JohnWiley & Sons Inc., C/OThe Sheridan Press, PO Box 465, Hanover, PA 17331. Copyright and Copying Copyright © 2019 John Wiley & Sons Ltd and the Society for Experimental Biology. All rights reserved. No part of this publication may be reproduced, stored or transmitted in any form or by any means without the prior permission in writing from the copyright holder. Authorization to copy items for internal and personal use is granted by the copyright holder for libraries and other users registered with their local Reproduction Rights Organisation (RRO), e.g. Copyright Clearance Center (CCC), 222 Rosewood Drive, Danvers, MA 01923, USA (www.copyright.com), provided the appropriate fee is paid directly to the RRO.This consent does not extend to other kinds of copying, such as copying for general distribution, for advertising and promotional purposes, for creating new collective works or for resale. Special requests should be addressed to: permissions@wiley.com. Online Open OnlineOpen is available to authors of primary research articles who wish to make their article available to nonsubscribers on publication, or whose funding agency requires grantees to archive the final version of their article. With OnlineOpen, the author, the author’s funding agency, or the author’s institution pays a fee to ensure that the article is made available to non-subscribers upon publication via Wiley Online Library, as well as deposited in the funding agency’s preferred archive. For the full list of terms and conditions, see http://wileyonlinelibrary.com/ onlineopen#OnlineOpen_Terms. Any authors wishing to send their paper OnlineOpen will be required to complete the payment form available from our website at: https://authorservices.wiley.com/ bauthor/onlineopen_order.asp. Prior to acceptance there is no requirement to inform an Editorial Office that you intend to publish your paper OnlineOpen if you do not wish to. All OnlineOpen articles are treated in the same way as any other article. They go through the journal’s standard peer-review process and will be accepted or rejected based on their own merit. The Journal is indexed by Current Contents/Life Sciences, Science Citation Index, Index Medicus, MEDLINE and BIOBASE/Current Awareness in Biological Sciences. Access to this journal is available free online within institutions in the developing world through the AGORA initiative with the FAO, the HINARI initiative with the WHO and the OARE initiative with UNEP. For information, visit www.aginternet.org, www.healthinternetwork.org, www. oarescience.org. Published by John Wiley & Sons Ltd, in association with the Society for Experimental Biology. Information on this journal can be accessed at http://www.theplantjournal. com. This journal is available online at Wiley Online Library. Visit wileyonlinelibrary.com to search the articles and register forTable of Contents email alerts. Wiley’s Corporate Citizenship initiative seeks to address the environmental, social, economic, and ethical challenges faced in our business and which are important to our diverse stakeholder groups. We have made a longterm commitment to standardize and improve our efforts around the world to reduce our carbon footprint. Follow our progress at www.wiley.com/go/citizenship. Disclaimer The Publisher, the Society for Experimental Biology and Editors cannot be held responsible for errors or any consequences arising from the use of information contained in this journal; the views and opinions expressed do not necessarily reflect those of the Publisher, the Society for Experimental Biology and editors, neither does the publication of advertisements constitute any endorsement by the Publisher, the Society for Experimental Biology and the Editors of the products advertised. Editorial correspondence Correspondence relating to editorial matters should be directed to The Plant Journal Editorial Office, John Wiley & Sons Ltd, 9600 Garsington Road, Oxford, OX4 2DQ, UK (E-mail: tpj-general@wiley.com). tpj_v98_i2_issueinfo.indd 3tpj_v98_i2_issueinfo.indd 3 13-Apr-19 7:36:21 PM13-Apr-19 7:36:21 PM RESEARCH HIGHLIGHT 193 The journey to the end of the chromosome: delivering active telomerase to telomeres in plants. L. Sweetlove and C. Gutierrez ORIGINAL ARTICLES 195 The plant Pontin and Reptin homologues, RuvBL1 and RuvBL2a, colocalize with TERT and TRB proteins in vivo, and participate in telomerase biogenesis. S. Schorˇová, J. Fajkus, L. Záveská Drábková, D. Honys and P. P. Schrumpfová 213 A newly identified cluster of glutathione S-transferase genes provides Verticillium wilt resistance in cotton. Z.-K. Li, B. Chen, X.-X. Li, J.-P. Wang, Y. Zhang, X.-F. Wang, Y.-Y. Yan, H.-F. Ke, J. Yang, J.-H. Wu, G.-N. Wang, G.-Y. Zhang, L.-Q. Wu, X.-Y. Wang and Z.-Y. Ma 228 Lotus SHAGGY-like kinase 1 is required to suppress nodulation in Lotus japonicus. C. Garagounis, D. Tsikou, P. K. Plitsi, I. S. Psarrakou, M. Avramidou, C. Stedel, M. Anagnostou, M. E. Georgopoulou and K. K. Papadopoulou 243 Extreme variation in rates of evolution in the plastid Clp protease complex. A. M. Williams, G. Friso, K. J. van Wijk and D. B. Sloan 260 Differential alternative polyadenylation contributes to the developmental divergence between two rice subspecies, japonica and indica. Q. Zhou, H. Fu, D. Yang, C. Ye, S. Zhu, J. Lin, W. Ye, G. Ji, X. Ye, X. Wu and Q. Q. Li 277 ICE1 and ZOU determine the depth of primary seed dormancy in Arabidopsis independently of their role in endosperm development. D. R. MacGregor, N. Zhang, M. Iwasaki, M. Chen, A. Dave, L. Lopez-Molina and S. Penfield 291 Comparative analysis of the reactive oxygen species-producing enzymatic activity of Arabidopsis NADPH oxidases. H. Kaya, S. Takeda, M. J. Kobayashi, S. Kimura, A. Iizuka, A. Imai, H. Hishinuma, T. Kawarazaki, K. Mori, Y. Yamamoto, Y. Murakami, A. Nakauchi, M. Abe and K. Kuchitsu 301 Targeted exome sequencing of unselected heavy-ion beam-irradiated populations reveals less-biased mutation characteristics in the rice genome. H. Ichida, R. Morita, Y. Shirakawa, Y. Hayashi and T. Abe 315 OsSHOC1 and OsPTD1 are essential for crossover formation during rice meiosis. Y. Ren, D. Chen, W. Li, D. Zhou, T. Luo, G. Yuan, J. Zeng, Y. Cao, Z. He, T. Zou, Q. Deng, S. Wang, A. Zheng, J. Zhu, Y. Liang, H. Liu, L. Wang, P. Li and S. Li 329 Suppression of tryptophan synthase activates cotton immunity by triggering cell death via promoting SA synthesis. Y. Miao, L. Xu, X. He, L. Zhang, M. Shaban, X. Zhang and L. Zhu 346 Enhancing microRNA167A expression in seed decreases the α-linolenic acid content and increases seed size in Camelina sativa. G. Na, X. Mu, P. Grabowski, J. Schmutz and C. Lu TECHNICAL ADVANCE 359 sRNA-FISH: versatile fluorescent in situ detection of small RNAs in plants. K. Huang, P. Baldrich, B. C. Meyers and J. L. Caplan 370 Rapid and reproducible phosphopeptide enrichment by tandem metal oxide affinity chromatography: application to boron deficiency induced phosphoproteomics. Y. Chen and W. Hoehenwarter CONTENTS OF VOL. 98, NO. 2, APRIL 2019 Front cover: The cover image shows an overlay and individual layers of Arabidopsis thaliana protoplast used for Bimolecular Fluorescence Complementation assay. In this case, interaction between AtCBF5 and AtTRB3 proteins was analysed. The biggest protoplast represents the final overlay of three layers, where the green one shows only chloroplasts, the red coloured protoplast symbolizes nuclear signal delivered by mRFP-VirD2NLS and finally, the yellow dot shows a specific interaction between selected proteins AtCBF5 and AtTRB3 localized in the nucleolus. For details, see article by Schorˇová et al. (pp.195–212). tpj_v98_i2_issueinfo.indd 4tpj_v98_i2_issueinfo.indd 4 13-Apr-19 7:36:21 PM13-Apr-19 7:36:21 PM RESEARCH HIGHLIGHT The journey to the end of the chromosome: delivering active telomerase to telomeres in plants Lee Sweetlove and Crisanto Gutierrez Linked article: This is a Research Highlight about Sarka Schorova et al. To view this article visit https://doi.org/10.1111/tpj. 14306. Linear chromosomes offer many advantages over circular DNA for transcription and replication of large genomes, hence their prevalence in eukaryotes. But the linear arrangement of the DNA has a massive Achilles heel: the terminal ends, or telomeres, are unstable and prone to mutation. Moreover, DNA replication cannot proceed to the end of a linear DNA molecule because the synthesis of Okazaki fragments needs RNA primers to bind ahead of the lagging strand. Eukaryotes deal with both of these problems by adding repetitive DNA sequences to the telomeres that act as a disposable buffer, protecting terminal genes from being truncated during replication and from mutation. Because the telomere is shortened during each DNA replication, it is necessary to resynthesise telomere DNA using an enzyme, telomerase reverse transcriptase. Our understanding of telomere biology is dominated by research into human telomeres. This is understandable due to the links between telomere biology and cellular mortality, ageing and a range of diseases including cancer. However, telomere biology in plants shows some specific differences to humans which may be crucial in our understanding of telomere biology in general. For example, telomerase activity in plant cells is well balanced with the cellular proliferation rate. The reversible regulation of telomerase activity is thought to be important in this context: its activity is turned off in differentiated tissues and turned on during cell periods of active cell replication, for example, during regeneration of plant tissues. Understanding the mechanism for this reversible regulation of telomerase activity could be beneficial in biomedical applications of telomere biology in humans. But where to start? From protozoans and humans, it was known that telomerase was a ribonucleoprotein, carrying its own RNA molecules that are complementary to the telomere repeats and are used as a template for telomere elongation, catalysed by the reverse transcriptase activity of the enzyme. But, in addition, a number of accessory proteins are required to deliver functional telomerase to the telomeres, to regulate its activity and to protect the elongated telomere from DNA repair enzymes. These components assemble into two distinct complexes known as shelterin and CST. Functional homologues of the CST complex have been identified in plants, but the same is not true for the shelterin complex. In plants, not all of the homologues of the six core shelterin components exist, and only some of them seem to be associated with telomeres in vivo. The goal, therefore was to identify undiscovered telomerase accessory proteins in plants and to establish how active telomerase is formed and regulated. Jirı Fajkus and his research group at Masaryk University, have been working on plant telomeres for over 20 years. A key member of his team in the hunt for plant telomeraseassociated proteins has been Petra Prochazkova Schrumpfova, first as a Ph.D. student and then through several postdoc periods. Working in Arabidopsis, the group had already established that Telomere Repeat Binding proteins (TRB) were involved in recruitment of telomerase to the telomeres. These proteins are specific to plants and contain an N–terminal Myb-like domain which is responsible for specific recognition of telomeric DNA. Attention turned to the plant homologues of two human telomere associated proteins called Pontin and Reptin after they turned up in a pull-down of TERT, the catalytic subunit of Arabidopsis telomerase, in an experiment done in collaboration with Eva Sykorova’s group at the Institute of Biophysics in Brno. The plant Pontin and Reptin homologues are encoded by RuvBL1 and RuvBL2a, respectively. But despite the fact that RuvBL proteins were isolated from plant cells as TERT-associated, Jirı and his team were not able to prove a direct interaction between TERT and RuvBL as had already been described in mammals. Serendipity then intervened. During their characterisation of RuvBL interactions with TERT, they used several proteins as negative controls. Surprisingly, one of the supposed negative controls showed reproducibly positive interaction with RuvBL proteins. It was in this way that they discovered that TRB proteins interact with RuvBL. Knowing that TRB proteins directly interact with TERT they started to closely characterise the trimeric complex TERT-TRB-RuvBL and that is © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd 193 The Plant Journal (2019) 98, 193–194 doi: 10.1111/tpj.14328 the focus of the highlighted paper which is drawn from the MSc and PhD research of Sarka Schorova with Petra Prochazkova Schrumpfova and Jirı Fajkus as joint corresponding authors. The work also involved Lenka Zaveska Drabkova, a postdoc from David Honys’s group at the Institute of Experimental Botany in Prague who did phylogenetic analysis of the RuvBL family in plants. That collaboration started late one afternoon during a shortterm visit of Petra Prochazkova Schrumpfova to David Honys’s lab that was focused on a completely different scientific topic. Such is the nature of science and scientists! In this highlighted paper (Schorova et al., 2019), a combination of BiFC, yeast-two hybrid and pull-down assays confirmed that there is no direct interaction between RuvBLs and TERT, but that the interaction is mediated by TRBs as an intermediary. It was also shown that RuvBL proteins form hetero- and homo-oligomers in vivo. Proof of the importance of RuvBL1/2 for telomerase biogenesis was provided by analysis of Arabidopsis knockout lines which had substantially reduced telomerase activity in flower buds (a rapidly proliferating tissue with high telomerase requirement). This crucial experiment turned out to be the hardest part of the research, with identification of knockout alleles a real struggle. Jirı and Petra say that they had to genotype hundreds of individual plants from several lines and were only able to identify a few heterozygous individuals of each gene with homozygous mutants being lethal. Further protein interaction experiments identified another protein in the complex: CBF5, a homologue of mammalian dyskerin, a known telomerase-associated protein. Cell biological analyses were able to place all of these proteins in the nucleolus and some of them in Cajal bodies and, combined with previous studies, the authors were able to put together the most complete picture of the plant telomerase complex to date, as shown in Figure 1. One of the most interesting facets of this picture is the similarities and divergence between plants and humans. On the one hand, identification of Reptin and Pontin in Arabidopsis and their conservation in humans, shows that the factors involved in telomerase biogenesis and function are evolutionary ancient. On the other, the interactions and mechanism of action of plant Reptin and Pontin is different than in human cells. The TERT subunit of Arabidopsis telomerase does not interact directly with Reptin and Pontin but through TRBs which in human cells are telomereassociated proteins but not TERT-accessory factors. This reveals that different mechanisms have evolved although using basically the same set of factors, a finding that would justify a similar study in other eukaryotic lineages to define the evolutionary history of complex formation between telomeric repeats, TERT, accessory factors and shelterin proteins. One possible reason to explain the variety of mechanisms suggested by this study is the specific organization, and possibly the 3D structure, of TERT RNA (TER) molecules which may limit the ability of TERT to interact directly with them or require other bridging proteins, as it occurs in Arabidopsis. Differences in the subnuclear localization of telomeric sequences may be also important. For Jirı and his team, work will continue to unpick the regulation of synthesis of both basic subunits of telomerase, TER and TERT, their intracellular trafficking and assembly into the holoenzyme complex, together with a number of associated factors. REFERENCE Schorova, S., Fajkus, J., Zaveska Drabkova, L., Honys, D. and Prochazkova Schrumpfova, P. (2019) The plant Pontin and Reptin homologues, RuvBL1 and RuvBL2a, colocalize with TERT and TRB proteins in vivo, and participate in telomerase biogenesis. Plant J. 98, 195–212. Figure 1. RuvBL1 and RuvBL2a, are orthologues of human Pontin and Reptin, respectively, in Arabidopsis. Besides their mutual interactions, RuvBL1 associates with the catalytic subunit of telomerase (TERT) in the nucleolus in vivo. In contrast to mammals, interactions between TERT and RuvBL proteins in Arabidopsis are not direct but are mediated by one of the Telomere Repeat Binding (TRB) proteins. The plant orthologue of human dyskerin, named CBF5, is indirectly associated with TRB proteins but not with the RuvBL proteins in the plant nucleus/nucleolus, and interacts with the Protection of telomere 1 (POT1a) in the nucleolus or cytoplasmic foci. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 193–194 194 Lee Sweetlove and Crisanto Gutierrez The plant Pontin and Reptin homologues, RuvBL1 and RuvBL2a, colocalize with TERT and TRB proteins in vivo, and participate in telomerase biogenesis Sarka Schorova1 , Jirı Fajkus1,2,3, * , Lenka Zaveska Drabkova4 , David Honys4 and Petra Prochazkova Schrumpfova1, * 1 Laboratory of Functional Genomics and Proteomics, National Centre for Biomolecular Research, Faculty of Science, Masaryk University, Brno, Czech Republic, 2 Mendel Centre for Plant Genomics and Proteomics, Central European Institute of Technology, Masaryk University, Brno, Czech Republic, 3 Institute of Biophysics of the Czech Academy of Sciences, v.v.i., Brno, Czech Republic, and 4 Laboratory of Pollen Biology, Institute of Experimental Botany of the Czech Academy of Sciences, v.v.i., Prague, Czech Republic Received 26 November 2018; revised 8 February 2019; accepted 26 February 2019; published online 4 March 2019. *For correspondence (e-mails petra.proch.schrumpfova@gmail.com; fajkus@sci.muni.cz). SUMMARY Telomerase maturation and recruitment to telomeres is regulated by several telomerase- and telomereassociated proteins. Among a number of proteins, human Pontin and Reptin play critical roles in telomerase biogenesis. Here we characterized plant orthologues of Pontin and Reptin, RuvBL1 and RuvBL2a, respectively, and show association of Arabidopsis thaliana RuvBL1 (AtRuvBL1) with the catalytic subunit of telomerase (AtTERT) in the nucleolus in vivo. In contrast to mammals, interactions between AtTERT and AtRuvBL proteins in A. thaliana are not direct and they are rather mediated by one of the Arabidopsis thaliana Telomere Repeat Binding (AtTRB) proteins. We further show that plant orthologue of dyskerin, named AtCBF5, is indirectly associated with AtTRB proteins but not with the AtRuvBL proteins in the plant nucleus/nucleolus, and interacts with the Protection of telomere 1 (AtPOT1a) in the nucleolus or cytoplasmic foci. Our genomewide phylogenetic analyses identify orthologues in RuvBL protein family within the plant kingdom. Dysfunction of AtRuvBL genes in heterozygous T-DNA insertion A. thaliana mutants results in reduced telomerase activity and indicate the involvement of AtRuvBL in plant telomerase biogenesis. Keywords: telomerase assembly, Pontin, Reptin, AtTERT, AtTRB, AtRuvBL, AtPOT1a, nucleolus, Arabidopsis. Linked article: This paper is the subject of a Research Highlight article. To view this Research Highlight article visit https://doi.org/10.1111/tpj.14328. INTRODUCTION Telomeres are nucleoprotein structures at the ends of eukaryotic chromosomes that protect linear chromosomes. Telomeric structures are formed by telomeric DNA, RNA, histones, and a number of other proteins that bind telomeric DNA, either directly or indirectly, together forming the protein telomere cap (Fajkus and Trifonov, 2001; de Lange, 2005; Schrumpfova et al., 2016a). The core component of the telomere cap in mammals is a six-protein complex called shelterin. The specific telomeric double-stranded DNA binding of the shelterin is mediated by its TRF1 and TRF2 (Telomere Repeat Binding Factors 1 and 2) components through their Myb-like domain of a telobox type (Bilaud et al., 1996; Peska et al., 2011). In Arabidopsis thaliana Telomere Repeat Binding (AtTRB) proteins, that contain Myb-like domain of a telobox type and bind plant telomeric repeats in vitro (Schrumpfova et al., 2004; Mozgova et al., 2008), were found to colocalize with telomeres in situ and in vivo (Dvorackova et al., 2010; Schrumpfova et al., 2014; Dreissig et al., 2017), directly interacted with the telomerase reverse transcriptase (AtTERT) (Schrumpfova et al., 2014) and physically interacted with AtPOT1b (Protection Of Telomeres 1) (Schrumpfova et al., 2008). Moreover, shortening of telomeres was observed in attrb knockout mutants (Schrumpfova et al., 2014, 2019; Zhou et al., 2018). © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd 195 The Plant Journal (2019) 98, 195–212 doi: 10.1111/tpj.14306 Telomere- or telomerase-associated proteins can regulate lengths of telomere tracts by affecting the assembly of active telomerase complex or by modulation of the accessibility of telomeres to telomerase. The process of maturation and recruitment of human telomerase is partially understood (Schmidt and Cech, 2015; MacNeil et al., 2016; Schmidt et al., 2016). However, a similar description of telomerase assembly and recruitment to the telomeres in plants is still missing which would allow to distinguish between general and specific features of these processes. Among a number of proteins, which were identified as associated with human telomerase, proteins named RuvBL (RuvB-like), that share limited sequence similarity to the bacterial RuvB helicase, were also identified. RuvBL proteins belong to the evolutionarily highly conserved AAA+ family (ATPase Associated with various cellular Activities) that are involved in ATP binding and hydrolysis (Matias et al., 2006). Eukaryotic RuvBL1 (Pontin, TIP49a, Rvb1, TAP54a) and RuvBL2 (Reptin, TIP48, TIP49b, Rvb2, TAP54b) participate in many diverse cellular activities like chromatin remodeling (Jha et al., 2008), transcriptional regulation (Ohdate et al., 2003; Gallant, 2007), oncogenic transformation (Osaki et al., 2013), epigenetic regulations (Gallant, 2007) or DNA-damage signaling (Rosenbaum et al., 2013). RuvBL1 and RuvBL2 can also play a role in the assembly of box C/D or H/ACA of small nucleolar RNAs (snoRNAs) with specific proteins to form functional ribonucleoprotein particles (RNPs) (Watkins et al., 2004; McKeegan et al., 2007; Boulon et al., 2008; Zhao et al., 2008). Participation of RuvBL1 and RuvBL2 proteins in diverse cellular processes, as well as their association with specific interactors, can vary among cytoplasm, nucleus and nucleolus (Izumi et al., 2012). RuvBL1 and also RuvBL2 monomers can assemble into different oligomeric forms, including hexameric structure with a central channel, or dodecamer composed of two hetero-hexameric rings with alternating RuvBL1 and RuvBL2 monomers (Torreira et al., 2008; Niewiarowski et al., 2010). RuvBL structure suggests that these proteins can act as scaffolding proteins, which explains their appearance in various cellular protein complexes (Matias et al., 2006; Mao and Houry, 2017). Mammalian RuvBL1 and RuvBL2, also termed as Pontin and Reptin, respectively, were found to play a critical role in telomerase biogenesis. Telomerase is a ribonucleoprotein enzyme complex composed of two core subunits: the catalytic telomerase reverse transcriptase (TERT) protein subunit and the telomerase RNA (TR) subunit (containing a box H/ACA motif). It performs the addition of telomeric DNA repeats onto the telomeres (Greider, 1996; Zhang et al., 2011). Proper assembly of TERT with TR into a functional complex is a stepwise regulated process governed also by multiple associated proteins (Schmidt and Cech, 2015; MacNeil et al., 2016). Human TR (hTR), as well as other box H/ACA snoRNPs, is associated with conserved scaffold proteins: dyskerin, NHP2, NOP10, NAF1 in the nucleoplasm, where NAF1 is replaced by GAR1 before the hTR RNP complex reaches the nucleolus. Association of hTR RNP with hTERT is proceeded in the nucleolus and the subsequent formation of catalytically active telomerase holoenzyme is localized into the Cajal bodies (CBs) (MacNeil et al., 2016) that are evolutionary conserved mobile nuclear substructures involved in the RNA modification and the RNP assembly processes (Cioce and Lamond, 2005). Venteicher et al. (2008) demonstrated that hRuvBL1 (Pontin) and hRuvBL2 (Reptin) are interdependent proteins and are recruited to hTERT complexes through the association between hTERT and hRuvBL1. Additionally, they showed that both hRuvBL1 and hRuvBL2 directly interact with dyskerin and may help to assemble or remodel a nascent hTERT/hTR/ dyskerin complex. The scaffold proteins, including dyskerin, together with hRuvBL1 and hRuvBL2, are required for a proper assembly of hTR RNP and are involved in the biogenesis of telomerase. A homologue of human RuvBL1 from A. thaliana has been already described by Holt et al. (2002). They observed that plants with reduced AtRuvBL1 (AT5G22330) mRNA levels had morphological defects and suggested that AtRuvBL1 was required in meristem development. Moreover, they observed that T-DNA insertion mutation in AtRuvBL1 gene was lethal. In our laboratory, AtRuvBL1 protein and also one of two AtRuvBL2 homologues, named AtRuvBL2a (AT5G67630), were purified together with AtTERT using Tandem Affinity Purification (TAP) from A. thaliana suspension cultures (Majerska et al., 2017). In this study, we examined a mutual interaction of AtRuvBL1-AtRuvBL2a proteins and demonstrated that AtRuvBL proteins are associated with AtTERT in the nucleolus in vivo. However, in contrast to mammalian counterparts, interactions between AtTERT and AtRuvBL proteins are not direct and are likely to be mediated by one of the AtTRB proteins. We prove that AtTRB3 protein physically interacts with AtRuvBL1 and simultaneously with AtTERT. We further show that in plants, similarly to mammals, telomerase assembly is a dynamic process, as is supported by our observation that AtCBF5, a plant orthologue of dyskerin, is in the plant nucleus/nucleolus indirectly associated with three of AtTRB proteins, but not with the AtRuvBL proteins, and interacts with the AtPOT1a in the cytoplasmic or nucleolus foci. Heterozygous T-DNA insertion mutants in AtRuvBL1 or AtRuvBL2a genes show reduced telomerase activity indicating the potential involvement of AtRuvBL proteins in telomerase assembly in A. thaliana. To identify new homologues of RuvBL protein family and elucidate their evolutionary relationships, we performed a survey of 83 plant species (80 angiosperms, one gymnosperm and two bryophytes). © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 196 Sarka Schorova et al. RESULTS AtRuvBL proteins form homomers or mutually interact RuvBL1 and RuvBL2 proteins from mammals and yeast can co-exist in different monomeric or oligomeric complexes comprising dimers, trimers, hexamers or doublehexamers that can be composed as mixed multimers (Torreira et al., 2008; Niewiarowski et al., 2010; Queval et al., 2014). Each RuvBL monomer contains three basic domains (DI, DII, DIII) (Figure 1a). Domain I (DI) together with domain III (DIII) represent the AAA+ core and are sufficient to form hexameric rings. In the AAA+ domain, the Walker A/B motifs are responsible for ATP binding and hydrolysis, while sensor I/II motifs sense whether the protein is bound to di- or triphosphates. Domain II (DII) corresponds to an insertion that is unique to RuvBL in comparison with other AAA+ family members (Silva-Martin et al., 2016). Figure 1. AtRuvBL proteins can form homo- or hetero-oligomers. (a) Schematic representation of the conserved motifs of the RuvBL proteins from Arabidopsis thaliana. DI, DII, DIII, Domain I, II, III; Walker A/B, Walker motifs; Sensor I/II, sensors; Arg finger, arginine finger. AtRuvBL2a and AtRuvBL2b form closely related sequence pairs. (b) Y2H system is used to assess homo- or heteromerization of AtRuvBL proteins. Two sets of plasmids carrying the indicated protein fused to either the GAL4 DNA-binding domain (BD) or the GAL4 activation domain (AD) are constructed and introduced into yeast strain PJ69-4a carrying reporter genes His3 and Ade2. Clear AtRuvBL1 and also AtRuvBL2a homomerization is detected on histidine-deficient plates. Mutual interaction between AtRuvBL1 and AtRuvBL2a is detected not only on histidine-deficient plates but also under stringent adenine selection. Co-transformation with an empty vector (AD, BD) serves as a negative control. (c) Co-IP is performed with the TNT-RRL expressed AtRuvBL1* and AtRuvBL2a* (35 S-labelled*, prey) mixed with their protein counterparts AtRuvBL1 and AtRuvBL2a, fused with Myc-tag (anchor) and incubated with anti-Myc antibody. In the control experiment, the AtRuvBL* proteins are incubated with Myc-antibody and protein G-coupled magnetic beads in the absence of partner protein. Input (I), Unbound (U) and Bound (B) fractions are collected and run in 12% SDS-PAGE gels. Mutual AtRuvBL1 and AtRuvBL2a interactions appear to be stronger than entirely homo-interactions between AtRuvBL proteins. (d) BiFC confirms homo- and also mutual heteromerization of AtRuvBL1 and AtRuvBL2a proteins. A. thaliana leaf protoplasts are co-transfected with 10 lg of each of the plasmids encoding nYFP-tagged or cYFPÀtagged AtRuvBL1, AtRuvBL2a or AtGAUT10 clones (as negative control) and simultaneously with mRFPÀVirD2NLS clone. Bright Field (left); RFP, mRFPÀVirD2NLS (red fluorescent protein fused with nuclear localization signal) labels cell nuclei and determines transfection efficiency; YFP, yellow fluorescent protein signals indicate specific protein–protein interactions (PPI) also marked by white arrows; Chl, chloroplast autofluorescence is marked by green pseudocolor, chloroplast autofluorescence is also visible in the YFP channel. Scale bars = 10 lm. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 Characterization of plant Pontin and Reptin 197 To examine whether plant homologues of RuvBL proteins form homomers or mutual heteromers as their mammalian counterparts, or exist only as monomers, we performed several assays for protein–protein interactions (PPIs): yeast two-hybrid system (Y2H), co-immunoprecipitation (Co-IP) and bimolecular fluorescence complementation (BiFC). First, we tested AtRuvBL1 and AtRuvBL2a homo-interactions. BiFC assay performed in A. thaliana leaf protoplasts, which enables direct visualization of protein interactions in living cells, demonstrated that AtRuvBL1 and AtRuvBL2a form homodimers or homomultimers in vivo. These results were confirmed using a GAL4 based Y2H assay, in which interactions took place inside the nucleus. We observed a clear homomeric interaction of AtRuvBL1 proteins as well as of AtRuvBL2a proteins in Y2H mating assay. The homomerization was further verified by Co-IP experiments in which proteins were expressed in the Coupled Transcription/Translation Rabbit Reticulocyte Lysate (TNT-RRL) System using the same vectors as in Y2H (Figure 1b, c). Additionally, we expanded our BiFC study (Majerska et al., 2017) and tested heteromerization of AtRuvBL1 and AtRuvBL2a not only in Nicotiana tabacum BY-2 protoplasts, but also in A. thaliana leaf protoplasts (Figure 1d). Analysis of subcellular localization of the AtRuvBL1-AtRuvBL2a interactions further showed that one reciprocal interaction of nYFP/AtRuvBL1 and cYFP/AtRuvBL2a was negative, and cYFP/AtRuvBL1 and nYFP/AtRuvBL2a showed nuclear, but not nucleolar localization, maybe due to the presence of a tag that may induce conformational changes of the AtRuvBL proteins (Cheung et al., 2010). Using Y2H assay, we confirmed clear interaction not only on histidine-deficient (ÀHis) plates but also under stringent adenine (ÀAde) selection. Both Y2H and Co-IP experiments revealed that mutual AtRuvBL1ÀAtRuvBL2a interaction seemed stronger than pure homomerization of Figure 2. AtRuvBL1 interacts indirectly with N-terminal part of Arabidopsis thaliana catalytic subunit AtTERT. The analyses were performed as described in Figure 1. (a) Schematic depiction of the plant catalytic subunit of telomerase (AtTERT) showing functional motifs. The regions of structural domains TEN (telomerase essential N-terminal domain), TRBD (RNA-binding domain), RT (reverse transcriptase domain) and CTE (C-terminal extension) are depicted above the conserved RT motifs (1, 2, A, B0 , C, D and E), telomerase-specific motifs (T2, CP, QFP and T) and a NLS (nucleus localisation-like signal). All the depicted AtTERT fragments were used in protein–protein interaction analysis (amino acid numbering is shown). All AtTERT fragments were fused with activation domains (AD/BD or nYFP/ cYFP) and used for further BiFC, Y2H and Co-IP analysis. (b) BiFC in A. thaliana leaf protoplasts were used to detect the interaction between AtRuvBL1 and all AtTERT fragments from schematic depiction. Here we show PPI interaction (white arrows) of two N-terminal fragments of AtTERT (AtTERT 1À233 and AtTERT 1À271) and AtRuvBL1 located in the nucleolus. AtGAUT10, negative control; RFP, nucleus marker; YFP, detects PPI; Chl, Chloroplast autofluorescence. Scale bars = 10 lm. (c) Y2H system fails to detect the interactions between AtRuvBL1 protein and N-terminal fragments of AtTERT (AtTERT 1À233 and AtTERT 1À271). BD, GAL4 DNA-binding domain; AD, GAL4 activation domain. (d) Co-IP analysis does not detect interactions between AtTERT fragments and AtRuvBL1 protein which were demonstrated by BiFC. I, Input; U, Unbound; B, Bound fractions; asterisks*, 35 S-labelling. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 198 Sarka Schorova et al. AtRuvBL proteins. These results showed that RuvBL1 and RuvBL2a proteins from A. thaliana are able to form both homo- and heteromers, as well as their homologues in diverse organisms, although they preferably form hetero- mers. AtRuvBL1 and AtTERT colocalize in the nucleus but contrary to mammalian homologues do not interact directly Human RuvBL proteins are involved in the biogenesis and maturation of human telomerase complex. Human hRuvBL1 directly interacts with hTERT catalytic subunit. hRuvBL2 does not exhibit direct interaction with hTERT and seems to be recruited to an hTERT complex through bridging hRuvBL1 molecules (Venteicher et al., 2008). To gain a deeper insight whether direct RuvBL-TERT interaction is conserved throughout the higher eukaryotes, we applied the above described Y2H, Co-IP and BiFC techniques. As TERT is a high-molecular-weight protein (approximately 130 kDa), we used the Gateway-compatible donor vectors carrying the AtTERT fragments that were described in Lee et al. (2012) and Zachova et al. (2013) (Figure 2a). We observed a clear nuclear interaction between AtRuvBL1 protein and AtTERT N-terminal fragments covering AtTERT domains localized in positions 1- 233 and 1-271 in the A. thaliana leaf protoplasts using BiFC (Figure 2b). These results supported the observation from tobacco BY2 culture protoplasts where N-terminal fragments of AtTERT interact with AtRuvBL1 (Majerska et al., 2017). As the central reverse transcriptase (RT) domain of hTERT is implicated in hRuvBL1 binding (Venteicher et al., 2008), we expanded our interest to the other AtTERT fragments. However, no interactions were detected between AtRuvBL1 protein and AtTERT fragments localized in positions 229À582, 597À987 and 972À1123, therefore covering RT or C-terminal domains of AtTERT. Likewise, no interaction was observed between any of AtTERT fragments and AtRuvBL2a protein (Figure S1). Notably, interactions of the N-terminal fragments between AtTERT domains and AtRuvBL1 were not confirmed in Y2H or Co-IP (Figure 2c, d). This discrepancy can be caused by the fact that the BiFC analysis detects the presence of proteins within the same macromolecular complex even in the absence of a direct contact between the proteins fused to the cYFP and nYFP fragments. The presence of proteins within the visualized macromolecular complex generally indicates that they participate in the same biological process (Kerppola, 2009). Our data show the interaction between AtRuvBL1 and AtTERT is localized in the nucleus and supports the suggestion of Majerska et al., that AtRuvBL1-AtTERT interaction is mediated by an unknown partner and occurs in plant cells but not in RRL lysate or yeast system. AtRuvBL proteins physically interact with AtTRB proteins Previously, we have described that members of plant-specific group of AtTRB proteins physically interact with the Nterminal part of AtTERT and colocalized with telomeres in situ (Schrumpfova et al., 2004, 2014; Mozgova et al., 2008; Dreissig et al., 2017; Zhou et al., 2018). AtTRB1 interaction with double-stranded telomeric DNA is mediated by the Myb-like domain, while the H1/5 domain is involved in DNA sequence-non-specific DNA-protein interactions, interaction with AtPOT1b (Schrumpfova et al., 2008) and in the multimerization of AtTRB1 (Mozgova et al., 2008) (Figure 3a). According to these findings, AtTRB proteins might be components of a putative shelterin-like complex in plants that modulates access of the telomerase to telomeres (Schrumpfova et al., 2016a, 2019). Our BiFC analysis revealed the AtTRB3 protein interaction with both AtRuvBL1 and AtRuvBL2a proteins in the nucleus (Figure 3b). These interactions were confirmed by Y2H (Figure 3c) and also by Co-IP (Figure 3d), in which AtTRB3 Figure 3. AtTRB3 proteins directly interact with AtRuvBL1 and AtRuvBL2a proteins. The methods are performed as is described in Figure 1. (a) Schematic representation of the conserved motifs of the AtTRB3 protein from Arabidopsis thaliana. Myb-like, Telobox-containing Myb domain; H1/H5, histone-like domain; coiled-coil, C-terminal domain. (b) BiFC shows interaction between AtTRB3 and both AtRuvBL proteins. PPIs marked by white arrows are localized in the nucleus. AtGAUT10, negative control; RFP, nucleus marker; YFP, detects PPI; Chl, Chloroplast autofluorescence. Scale bars = 10 lm. (c) Y2H results show interactions between AtTRB3 and both AtRuvBL proteins on His- deficient plates. BD, GAL4 DNA-binding domain; AD, GAL4 activation domain; asterisks*, 5 mM 3-aminotriazol. (d) Co-IP results confirm direct interactions between radioactively labelled AtTRB3 and Myc-tagged AtRuvBL proteins. I, Input; U, Unbound; B, Bound fractions; asterisks*, 35S-labelling. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 Characterization of plant Pontin and Reptin 199 protein was radioactively labelled by 35 S-methionine and mixed with its putative protein partners AtRuvBL1 or AtRuvBL2a fused with Myc-tag, and incubated with antiMyc antibody. Clear nuclear interaction of AtTRB2 and AtRuvBL2a, but not of AtTRB2 with AtRuvBL1, was detected by BiFC and verified by Y2H and Co-IP. However, the nuclear interaction of AtTRB1 with AtRuvBL1 observed in BiFC seems to be indirect, as it was not proven by Y2H or Co-IP assays, but indicates that both proteins are present in the same macromolecular complex (Figure S2). Collectively, direct interactions of AtTRBs with AtTERT, as well as with AtRuvBL proteins, imply the role of AtTRB proteins as mediators of the interactions between AtRuvBL proteins and AtTERT telomerase subunit in vivo. AtTRB3 protein mediates interaction between AtRuvBL1 and AtTERT Our data, showing the indirect interaction between the Nterminal part of AtTERT and AtRuvBL1, suggested that this interaction could be mediated by AtTRB3 protein. We performed Co-IP assay with all three proteins of interest (Figure 4). Two prey proteins AtRuvBL1 and AtTRB3, were labelled with 35 S-methionine during the expression in TNTÀRRL system. N-terminal fragment of AtTERT (AtTERT 1-271), fused with Myc-tag as an anchor, was expressed in TNTÀRRL system in non-radioactive form ensuring a better resolution of the prey proteins in the 12% SDS-PAGE separation. Radioactively labelled AtTERT fragment was expressed in parallel tube to affirm the proper AtTERT 1- 271 expression. The complex was captured with anti-Mycantibody and protein G-coupled magnetic beads. Several negative controls were performed, where some of the monitored proteins were not present, to ensure specificity of the AtRuvBL1ÀAtTRB3ÀAtTERT complex. From these negative controls, it is evident that AtRuvBL1 protein neither directly interacts with the AtTERT 1À271 fragment nor is non-specifically bound to the magnetic beads. Conversely, the presence of AtTRB3 in immunoprecipitation mixture resulted in reproducible and significant increase of the AtRuvBL1 in the immunoprecipitated complex. So, it is evident that AtRuvBL1 is recruited to the AtTERT complex through an interaction with AtTRB3 protein, which mediates interaction of both proteins, AtTERT and AtRuvBL1. Plant homologue of mammalian dyskerin AtCBF5 associates with AtTRB proteins in the plant nucleus Mammalian protein dyskerin is a core component of mature and functional telomerase complex (He et al., 2002; Schmidt and Cech, 2015; MacNeil et al., 2016). Dyskerin binds the H/ACA box of small nuclear and nucleolar RNAs (sn- and sno-RNAs) and belongs to conserved scaffold proteins of human hTR (MacNeil et al., 2016). Plant homologue AtCBF5 (also named AtNAP57) is localized within nucleoli and CBs (Lermontova et al., 2007) and associates with enzymatically active telomerase RNP particles in an RNA-dependent manner (Kannan et al., 2008). Here we observed a clear indirect interaction of AtCBF5, fused with cYFP, with all three examined nYFP/AtTRB proteins using BiFC technique (Figure 5). As has already been discussed above, BiFC analysis can detect the presence of proteins within the same macromolecular complex even without a direct contact between the proteins fused with cYFP/nYFP (Kerppola, 2009). We assume that the interactions between AtCBF5 and AtTRBs are indirect because we Figure 4. AtTRB3 protein is mediator of AtRuvBL1 and AtTERT interaction. (a) Co-Immunoprecipitation of the three proteins of interest. Two proteins AtRuvBL1 and AtTRB3 are radioactively labelled by 35 S-methionine (marked with asterisks) during the expression in TNT-RRL lysate and subsequently incubated with non-radioactive Myc-tagged AtTERT 1À271 fragment and anti-Myc antibody. In the control experiments, the proteins are incubated with Myc-antibody and protein G-coupled magnetic beads in the absence of one or both partner proteins. Radioactively labeled AtTERT fragment is expressed in parallel tube as a control of the expression. From penult column it is evident that the presence of AtTRB3 results in significant increase of the AtRuvBL1 in the immunoprecipitated complex. I, Input; U, Unbound; B, Bound fractions were collected and run in 12% SDS-PAGE gels. (b) Schematic depiction of the putative protein complex formed by proteins AtRuvBL1, AtTRB3 and AtTERT. AtRuvBL1 is depicted in its presumed hexameric form and AtTRB3 in its dimeric form. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 200 Sarka Schorova et al. were not able to confirm the AtCBF5ÀAtTRBs interactions observed by BiFC in Y2H mating assay. Also Co-IP did not reveal any iteraction between proteins expressed in TNTÀRRL system, fused with Myc-tag (AtRuvBL1, AtRuvBL2a, AtTRB1, AtTRB2 and AtTRB3) and with radioactively labelled AtCBF5 as a prey. Additionally, no interaction was detected between AtCBF5 and any of AtRuvBL proteins neither in BiFC nor in Y2H or Co-IP. As a positive control we used the interaction between AtCBF5 and AtPOT1a. Here we show that the AtCBF5 interacts with AtPOT1a not only in Y2H and Co-IP, as was shown in Kannan et al. (2008), but also in the plant nucleus using BiFC assay. In addition to the nucleolar localization of AtPOT1a–AtCBF5 interactions, we also observed this interaction in several Figure 5. AtCBF5 associates with AtTRB proteins indirectly. The methods are done in the same manner as in Figure 1. (a) BiFC assay shows indirect interaction between AtCBF5 and three proteins from AtTRB family (AtTRB1-3). AtCBF5 interacts also with AtPOT1a protein. PPIs are marked by white arrows. AtGAUT10, negative control; RFP, nucleus marker; YFP, detects PPI; Chl, Chloroplast autofluorescence. Scale bars represent 10 lm. (b) Y2H assay analysis does not detect the interaction between AtCBF5 and AtTRB proteins which was found by BiFC. AtCBF5 protein interacts only with AtPOT1a on histidine deficient plate (-His). BD, GAL4 DNA-binding domain; AD, GAL4 activation domain. (c) Co-IP analysis shows interaction only between AtCBF5 and AtPOT1a protein, fused with Myc-tag and incubated with Myc-antibody and protein G-coupled magnetic beads. I, Input; U, Unbound; B, Bound fractions; asterisks*, 35 S-labelling. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 Characterization of plant Pontin and Reptin 201 nuclear and cytoplasmic foci (Figure S3). Further, we observed a weak interaction between AtPOT1aÀAtRuvBL1 proteins in Y2H and Co-IP assays but not in BiFC system (Figure S4). As a negative control in BiFC assay, we cotransfected protoplasts with cYFP/AtGAUT10. AtGAUT10 protein did not interact with any of the proteins of interest fused with nYFP: AtRuvBL1, AtRuvBL2a, AtTRB1, AtTRB2, AtTRB3 or AtPOT1a. Co-transformation with an empty vector (AD, BD) served as a negative control in Y2H experiments. In Co-IP experiment, the AtCBF5 proteins were incubated with Myc-antibody and protein G-coupled magnetic beads in the absence of partner protein as negative control. Together, we conclude that AtTRB proteins are associated in very close proximity with AtCBF5, the plant homologue of mammalian dyskerin, in the plant nucleus. However, at the same time, AtCBF5 is not localized in the nearby complex with the AtRuvBL proteins in vivo. Association of AtRuvBLs, AtTRBs and AtTERT indicates the formation of their complex in the nucleolus During the assembly of a fully functional complex of the human telomerase, the mature hTR gets recruited to the nucleolus where it binds the hTERT complex. Both of the core telomerase components, hTR and also hTERT, are previously processed by several proteins, including hRuvBL1 and hRuvBL2. It has already been published that in the interphase, the AtTRB proteins showed preferential localization to the nucleus and specially to the nucleolus (Dvorackova et al., 2010). In comparison with the mammalian nucleoli, plant nucleoli are larger, more frequently undergo fusions, and sometimes have a central clear region, often called the nucleolar vacuole, the size of which depends on nucleoli transcriptional activity (Shaw and Brown, 2012; Stepinski, 2014). We analyzed the subcellular localizations of the interactions of our proteins of interest: AtTERT 1-271, AtRuvBL1, AtRuvBL2a, AtTRB3 and AtCBF5 fused with nYFP- or cYFPtag in routinely performed BiFC experiments. The nucleoli were marked by control plasmid mRFPÀAtFibrillarin 1 (Pih et al., 2000). Figure 6 shows interactions between AtRuvBL1ÀAtTERT, AtTRB3ÀAtTERT, AtRuvBL1ÀAtTRB3 and AtRuvBL2aÀAtTRB3, which occupy distinct areas within the plant nucleus that match to the plant nucleolus. The number of the PPIs foci localized exclusively in the nucleolus is listed in the Table S1. Similar patterns of nuclear or nucleolar PPI localization is visible also in Figure S5 where the whole nucleus was marked by mRFPÀVirD2NLS. However, the AtCBF5ÀAtTRB3 interaction showed different localization pattern than the other examined PPIs. AtCBF5ÀAtTRB3 interaction seems to be localized in nucleoli and sometimes in additional nuclear bodies at the periphery or outside the nucleoli, which is consistent with localization of free AtCBF5 (Lermontova et al., 2007). Together, our data indicate formation of AtRuvBLsÀAtTRBsÀAtTERT complex in the nucleolus. Dysfunction of AtRuvBL genes reduces telomerase activity In human cells, the hRuvBL1 and hRuvBL2 proteins associate with a significant population of hTERT molecules that do not yield high-level telomerase activity, measured by Telomere Repeat Amplification Protocol (TRAP). The depletion of hRuvBL1 and hRuvBL2 markedly impaired telomerase RNP accumulation and diminished human telomerase activity (Venteicher et al., 2008). To assess whether mutations in AtRuvBL genes have any impact on telomerase activity in A. thaliana, we set to perform TRAP assay on telomerase extracts isolated from T-DNA insertion mutant lines. Extensive search of several T-DNA Figure 6. Association of AtRuvBLs, AtTRBs and AtTERT in the nucleolus in A. thaliana leaf protoplasts. Protoplasts are co-transfected with mRFPÀAtFibrillarin 1 encoding RFP that labels nucleolus and simultaneously with each of the plasmids encoding nYFP-tagged or cYFP-tagged AtRuvBL1, AtRuvBL2a, AtTERT 1À271, AtTRB3 or AtCBF5 to determine PPI localization. AtRuvBL1ÀAtTERT, AtTRB3ÀAtTERT, AtRuvBL1ÀAtTRB3 or AtRuvBL2aÀAtTRB3 interactions show nucleolar localization. Plant homologue of mammalian dyskerin, AtCBF5, is associated with AtTRB3 in the nucleolus and in additional nuclear bodies at the periphery of the nucleolus. RFP, marked nucleus; YFP, detects PPI; Scale bars = 10 lm. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 202 Sarka Schorova et al. insertion lines, which are available from several plant databases, revealed only two suitable plant lines with a limited number of heterozygous mutant plants but with no homozygous mutant plants: SAIL_397_C11 (AtRuvBL1) and GK-543F01 (AtRuvBL2a). In an additional seven tested TDNA insertion plant lines we did not detect any viable mutant or heterozygous plants for AtRuvBL1 or AtRuvBL2a genes (Table S2). Furthermore, genotypic ratios of offspring of individual heterozygous plants did not follow the expected Mendelian genotypic ratio. The observed ratio for AtRuvBL1+/À and for AtRuvBL2a+/À plants was (51:21:0) and (91:10:0), respectively, instead of (1:2:1) (Figure 7a) and the cause of this phenomenon will be further investigated. Quantitative TRAP assay performed with telomerase extract isolated from flower buds of individual AtRuvBL2a heterozygous plants demonstrated that relative telomerase activity showed apparent reduction in comparison with telomerase extract from wild-type A. thaliana (Col-0+/+ ) buds (Figure 7b). T-DNA insertion mutation in AtRuvBL1 gene was lethal (Holt et al., 2002) but we detected viable heterozygous AtRuvBL1 plants. These plants showed a milder reduction of telomerase activity than AtRuvBL2a+/À plants, which supports the assumption that AtRuvBL1 protein is essential for meristem development (Holt et al., 2002). Human RuvBL proteins are direct interactors of transcription factor MYC that is required for expressing many genes involved in cell-cycle transition events and proliferation (Wood et al., 2000). hRuvBL2 regulates MYC-dependent transcription of TERT via targeting the hTERT promoter (Wood et al., 2000; Li et al., 2010; Flavin et al., 2011; Zhao et al., 2014). We analyzed the levels of AtTERT transcripts in AtRuvBL1 and AtRuvBL2a heterozygous plants to detect whether the decrease of telomerase activity was caused by the negative regulation of AtTERT promoter i.e. the decrease of the abundance of AtTERT transcripts. We did not observe significant changes in transcripts of AtTERT gene in AtRuvBL1 heterozygous mutant plants compared with the wild-type A. thaliana. Instead, we observed very slight, though significant, increase in AtTERT transcripts in AtRuvBL2a heterozygous mutant plant lines (Figure S6). Due to the difficulties in maintaining the heterozygous AtRuvBL plant lines for several subsequent generations, we were not able to analyze the transgenerational effects of reduced telomerase activity on telomere lengths in plants heterozygous in AtRuvBL1 and AtRuvBL2a genes. However, in the analyzed generation of AtRuvBL1+/À and AtRuvBL2a+/À plants that were descendants of heterozygous predecessors, we did not detect any significant changes in telomere lengths compared with the wild-type plants using Terminal Restriction Fragment analysis (TRF) (Figure S7). Together, we conclude that the depletion of AtRuvBL1 and especially of AtRuvBL2a proteins reduces telomerase activity which suggests a conserved role of AtRuvBL proteins in maturation of functional telomerase complex across the mammals and also plants. Identification and phylogenetic analysis of the RuvBL family in plants RuvBL proteins, showing association with TERT in human cells, represent a group of proteins well conserved across all eukaryotic kingdoms, including Fungi, Animalia or Plantae. Here, we present a genome-wide analysis of RuvBL proteins in 80 vascular plant species, one gymnosperm and two bryophytes, totally 83 taxa, that were analyzed for the presence of all three basic domains (DI, DII, DIII). The evolutionary relationships among the RuvBL proteins were Figure 7. Reduction of relative telomerase activity in heterozygous AtRuvBL mutant plants. (a) Genotypic ratio of the offspring of heterozygous AtRuvBL1 and AtRuvBL2a T-DNA insertion mutant plants. Homozygous mutant plants in AtRuvBL genes are fully absent and even the number of heterozygous plants does not follow the Mendelian genotypic ratio. (b) Samples isolated from AtRuvBL1+/À and AtRuvBL2a+/À buds are analyzed in three technical replicates by quantitative TRAP. Data are related to wildtype Col-0 sample (telomerase activity in Col-0 buds are arbitrarily chosen as 1). Relative telomerase activity is reduced in both AtRuvBL1+/À and especially in AtRuvBL2a+/À samples. P < 0.05 are considered as significant. Single asterisk denotes 0.01 < P < 0.05. Three asterisks denote 0.01 < P < 0.001. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 Characterization of plant Pontin and Reptin 203 determined using maximum likelihood analyses based on multiple alignments producing a phylogenetic tree depicting the relationships among all currently accessible RuvBL sequences. The evolutionary hypotheses from these analyses were highly congruent. RuvBL protein family was divided in two distinct groups based on the similarity of sequences and branch length. Sequence similarity between RuvBL1 and RuvBL2 is generally low, about 35– 40% while the sequence similarity within RuvBL subfamilies is about 80%. For instance, in A. thaliana AtRuvBL2a and AtRuvBL2b share 82% similarity. On the other hand, AtRuvBL1 with AtRuvBL2a or AtRuvBL1 with AtRuvBL2b share 37.5 and 38.8% similarity, respectively. However, only a subset of RuvBL1 was clearly separated (100% BS; blue branch in Figure 8). Surprisingly, based on BLAST search, RuvBL1 was found only in dicots and basal angiosperms (Amborella trichopoda) up to now, RuvBL2 was represented in both, dicots and monocots from angiosperms, but also in gymnosperms (Picea sitchensis) and bryophytes (Physcomitrella patens and Marchantia polymorpha). The number of the homologues varied from 1 to 8 (Data S1 and S2). Figure 8. Phylogenetic analysis of the RuvBL family in plants. Unrooted phylogenetic tree of 190 proteins sequences of RuvBL family with enumerated plant species. Numbers above branches means bootstrap support values. Orthologues from Arabidopsis thaliana and Nicotiana tabacum are in bold letters. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 204 Sarka Schorova et al. DISCUSSION The formation of functional and enzymatically active telomerase, a multisubunit RNP complex, is a dynamic process governed by number of cofactors. In mammals, hRuvBL1 and hRuvBL2 proteins, Pontin and Reptin, respectively, are present in early steps of telomerase RNP biogenesis. We characterized plant homologues of RuvBL proteins: AtRuvBL1 and AtRuvBL2a, previously co-purified together with telomerase protein subunit AtTERT from A. thaliana suspension cultures (Majerska et al., 2017). Here we show that AtRuvBL1 protein colocalizes with Nterminal part of AtTERT subunit of plant telomerase also in vivo. However, in contrast with the AtRuvBL mammalian counterparts, their interaction in plants seems to be indirect. Association of AtRuvBL proteins with AtTERT in the plant nucleolus appears to be bridged by telomeric AtTRB proteins. Requirement of AtRuvBL proteins for a proper telomerase assembly is endorsed by the fact that depletion of AtRuvBL1 and especially of AtRuvBL2a protein, reduces telomerase activity in plants heterozygous for AtRuvBL1 or AtRuvBL2a genes. Moreover, AtTRB proteins are associated in the plant cell with a homologue of mammalian dyskerin, AtCBF5, that plays a role in telomerase RNP biogenesis and directly interacts with AtPOT1a protein. AtTRB proteins thus play a role of interaction hubs not only in telomere chromatin structure but also in telomerase biogenesis. AtRuvBL proteins are able to multimerize, which is analogous to the situation in mammalian cells, and our data show preference to form mutual heteromers. Detailed summary of protein–protein interactions between AtRuvBLs, AtTRBs, AtTERT fragments, AtPOT1s and AtCBF5 proteins, that have been detected using BiFC, Y2H or Co-IP assays in this and other relevant publications, are given in the Table 1. Our detailed phylogeny proved that RuvBL proteins are evolutionarily conserved in land plants and implied plausible functional conservation of the RuvBL proteins. However, further biochemical validation of the possible conservation of mutual RuvBLÀTRB interaction across the plant kingdom can be limited by the fact that the number of paralogues varies from 1 to 8 members in between RuvBL proteins. The multiplication of genes of the same family is not surprising as, in many plant families, the polyploidy (i.e. whole-genome duplication, WGD), resulting in retention of multiple gene paralogs may lead to their sub-functionalization, neo-functionalization or partial or full redundancy (Mandakova and Lysak, 2008; Freeling, 2009). These limitations might be deteriorated by the fact that the AtRuvBL proteins can be involved in a similar biochemical pathway but their interaction partners might slightly differ (this paper; Venteicher et al., 2008). RuvBL proteins are involved in various cellular processes The exact function even of mammalian RuvBL proteins is still quite unknown as they interact with many molecular complexes with vastly different downstream effectors (Mao and Houry, 2017). Among others, hRuvBL2 was shown to regulate hTERT promoter likely through the regulation of MYC (c-myc), the transcription factor for TERT (Wood et al., 2000; Li et al., 2010; Flavin et al., 2011; Zhao et al., 2014). We observed no significant changes in transcripts of AtTERT gene in AtRuvBL1 heterozygous mutant plants, however we detected a very slight increase in transcripts of AtTERT gene in AtRuvBL2a heterozygous plants. Although the transcript levels of AtTERT gene were slightly increased in AtRuvBL2a heterozygous plant lines, we observed a very significant reduction of telomerase activity Table 1 A summary table of protein–protein interactions Summary table shows all interactions between AtRuvBLs, AtTRBs, AtTERT fragments, AtPOT1s and AtCBF5 proteins that are detected using BiFC, Y2H or Co-IP assays. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 Characterization of plant Pontin and Reptin 205 in these plants. Telomerase activity was reduced also in AtRuvBL1 heterozygous T-DNA insertion plant lines. To verify whether the regulation of telomerase activity was affected due to the compromised assembly of telomerase complex rather than due to regulation of transcript levels of AtTERT gene in AtRuvBL-dependent manner, however, needs further investigation. The participation of RuvBL proteins in heterogeneous cellular process as well as their association with specific interactors can vary between cytoplasm, nucleus and nucleolus (Mao and Houry, 2017). It seems that, also in A. thaliana, the function of AtRuvBL proteins is not specific only to the telomerase assembly, as they were suggested as regulators of disease resistance (R) genes (Holt et al., 2002). It has already been published that AtRuvBL1 is essential in meristem development (Holt et al., 2002), the function consistent with its function in telomerase assembly observed in this work. Our extensive, but unsuccessful, Figure 9. Comparative model of telomerase assembly in human and Arabidopsis. (a) Human TR binds dyskerin, NHP2, NOP10, and GAR1 and human TERT associates with the chaperones Hsp90 and p23. The telomerase RNP is retained into the nucleoli through the interaction between TERT and nucleolin. Assembly of TR and TERT into catalytically active telomerase is aided by Pontin (hRuvBL1) and Reptin (hRuvBL2) AAA+ ATPases. Telomerase is recruited to Cajal bodies (CBs) by its interaction with TCAB1. The CBs will colocalize with telomeres, and telomerase is recruited to telomeres by the interaction with the shelterin component TPP1 (MacNeil et al., 2016; Lim et al., 2017). (b) Arabidopsis CBF5, GAR1, NOP10, NHP2, but in contrast with human cells also NAF1, were localized into the plant nucleolus (Pendle et al., 2005; Lermontova et al., 2007). In the plant nucleolus, we observe colocalization of TERT with RuvBL AAA+ ATPases complex bridged by telomeric TRB proteins, as well as the interaction of telomeric protein POT1a with CBF5. Arabidopsis telomeres cluster at the periphery of the nucleolus which is mediated by the presence of nucleolin. Recruitment of the mature telomerase complex to telomeres with or without commitment of CBs in Arabidopsis needs further investigation. Proteins that were already proven as associated with CBs are highlighted in color in Cajal bodies. Proteins that have not yet been experimentally proven as CBs associated are marked with black and white. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 206 Sarka Schorova et al. effort to receive homozygous mutant plants in several T-DNA insertion lines mutant in the AtRuvBL1 and AtRuvBL2a genes also indicated the essentiality of AtRuvBL proteins in various cellular processes in plants. Furthermore, genotypic ratio of offspring of individual heterozygous plants does not follow the Mendelian genotypic ratio, indicating that both AtRuvBL proteins are essential regulators of plant development. Therefore, we suggest to investigate the function of AtRuvBLs in plant sporophyte or female gametophyte development in future studies. Nucleolus localization of telomerase assembly complex Telomere maintenance requires a proper assembly of the TERT and TR components of telomerase into RNP as well as a number of cofactors involved in maturation, stability and subcellular localization of telomerase. In mammals, the association of hTR RNP with hTERT proceeds in the nucleolus during the early S-phase (Lee et al., 2014). Assembled and catalytically active telomerase RNP separates from the nucleoli and is transported to CBs during the S-phase for subsequent recruitment to telomeric chromatin and telomere extension (Figure 9a) (MacNeil et al., 2016). Association of hTERT with human RuvBL proteins, Pontin and Reptin, peaks in S-phase, which may reflect cell-cycle regulation of total TERT and/or assembly of telomerase on telomeres (Venteicher et al., 2008). RuvBL1 and RuvBL2a proteins, together with, for example, Fibrillarin 1 and many other proteins, were purified and identified in nucleoli isolated from A. thaliana cell culture protoplasts (Pendle et al., 2005). Our data indicated that plant homologues of human Pontin and Reptin, the AtRuvBL proteins, are associated in the plant nucleolus with AtTERT, together with AtTRB proteins (Figure 9b). AtTRB proteins are highly dynamic and during the interphase, they are preferentially localized to the nucleolus or nuclear bodies of different size (Dvorackova, 2010). AtTRBs behave as typical nucleolar resident proteins, being largely dispersed at prophase, coinciding with nucleolar disassembly. However, a small but detectable amount of the protein remains associated with the chromatin throughout mitosis (Azum-Gelade et al., 1994; Dvorackova et al., 2010). Similarly, to the AtTRB proteins, also the N-terminal part of AtTERT was detected in the nucleoli in A. thaliana (Rossignol et al., 2007; Zachova et al., 2013). In mammals, the telomerase RNP is retained in nucleoli through the interaction between hTERT and nucleolin in the dense fibrillar component (Khurts et al., 2004; Lee et al., 2014). In A. thaliana, null mutants for the nucleolar protein NUCLEOLIN 1 cause telomere shortening on all chromosome arms (Pontvianne et al., 2016). Telomeres in A. thaliana do not form a Rabl conformation, as in some other species, but telomeres and their flanking regions strongly associated with the nucleolus in a rosette-like organization (Armstrong et al., 2001; Fransz et al., 2002; Roberts et al., 2009; Pontvianne et al., 2016; Schrumpfova et al., 2016a). Our data indicated the presence of AtTERTÀAtTRBÀAtRuvBL complex in the nucleolus. Nucleolar localization of the AtTERTÀAtTRBÀAtRuvBL complex together with the close proximity of telomeres to the nucleolus, suggested the conservation of the recruitment of the maturating telomerase to the nucleolus during the telomerase assembly. Figure 9 shows a comparative model of the assembly and localization of telomerase in mammalian and plant cells. Plausible conservation of the telomerase trafficking pathway Cajal bodies are spherical suborganelles localized in the nucleoplasm either in the vicinity of the nucleolus and/or they are present free. The function of CBs is not completely understood, but they were implicated mainly in snRNAs synthesis and processing. CBs also contribute to the biogenesis of telomerase. In S-phase, CBs colocalize with telomeres and facilitate recruitment of the mature mammalian telomerase complex to the telomeres. Human dyskerin, hNHP2, hNOP10 and hGAR1, that displaces hNAF1 in the hTR RNP, belong to conserved scaffold proteins, which colocalize with CBs and are involved in hTR RNP assembly (MacNeil et al., 2016). Expression of putative AtGAR1, AtNOP10, AtNHP2 genes encoding protein components of the H/ACA box snoRNP complex correlate with that of AtCBF5, a plant homologue of dyskerin (Lermontova et al., 2007). AtCBF5 directly interacts with AtNAF1 (Lermontova et al., 2007) and has been identified as a component of the enzymatically active A. thaliana telomerase RNP (Kannan et al., 2008). AtCBF5 localizes in nucleoli and sometimes in additional nuclear bodies at the periphery or outside the nucleoli, but AtCBF5 also colocalizes with TMG-capped snRNA, a marker for CBs (Lermontova et al., 2007). Here we show that plant dyskerin, AtCBF5, indirectly interacts with AtTRB proteins in the plant nucleolus or in other nuclear bodies. It has already been published that AtTRBs are located not only in the nucleolus but also in nuclear bodies of different size, some of which might be CBs (visualized by a marker protein Coilin) (Dvorackova, 2010). Dvorackova detected significant colocalization of AtTRB1 with Coilin present in the CBs adjacent to the nucleolus. However, no colocalization was detected between signals corresponding to the AtTRB1 and free CBs in the nucleoplasm. Presence of AtTRB1 protein entirely in the CBs adjacent to the nucleoli implies a potential conservation of the trafficking pathway during the telomerase maturation, which comprises movement of maturating telomerase complex through nucleolus to CBs and finally to the telomeres. Notably, not all the organisms (e.g. budding yeast and ciliates) rely on the CBs trafficking since telomerase RNAs from these species do not have H/ACA or CAB box motifs, and further studies © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 Characterization of plant Pontin and Reptin 207 are needed to prove this hypothesis. We observe that interaction between AtCBF5 and AtPOT1a is localized mostly in the nucleolus but in few cases also in cytoplasmic foci. The cytoplasmic localization is not surprising as it has already been shown that plant AtPOT1a and AtPOT1b, as well as their human homologue hPot1, are localized in the nucleus, as well in the cytoplasm (Chen et al., 2007; Rossignol et al., 2007). The assembly of hTR RNP to the telomerase holoenzyme is not fully elucidated and it is highly complex multistep process. Therefore, the absence of the interaction between AtCBF5 and AtRuvBLs in the plant nucleus in our experiments is also not surprising. For example, Machado-Pinilla et al. (2012) showed that dyskerin was sandwiched between two hSHQ1 domains in the first steps of the biogenesis of telomerase. C-terminal tail of hCBF5 was essential for hSHQ1 release mediated by hRuvBLs. However, a stable interaction with the tails is not a part of the process because hRuvBLs bind to hCBF5 in a pull-down assay, even in the absence of its tail. Assembly of functional AtTER RNP, as well as the assembly of mammalian hTR RNP, is certainly a multistep process that may include AtTER, AtCBF5, AtTRBs, AtRuvBLs, AtPOT1a and many other factors, whose presence/participation/mutual interactions will be the subjects of our future research. Dynamics and complexity of mutual interactions can be demonstrated by the fact that we detect the interacting complex of AtCBF5ÀAtPOT1a in the nucleolus or in the cytoplasmic and nuclear foci using BiFC assay, while AtCBF5ÀAtTRBs interactions are localized entirely to the nucleoli and additional nuclear bodies. Moreover, association of AtTRB3 with AtTERT and AtRuvBLs is entirely localized to the nucleolus. Concluding remarks Homologues of the mammalian Pontin and Reptin, named RuvBL proteins, as well as TRB proteins, might be involved in diverse processes in the plant cell. AtTRB proteins are not only components of terminal complex associated with telomeres and catalytic subunit of telomerase, AtTERT (Schrumpfova et al., 2016a, 2019), but they also serve as epigenetic regulators that potentially impact the transcription status of thousands of genes as subunits of epigenetically active multiprotein complexes (Lee and Cho, 2016; Schrumpfova et al., 2016b; Zhou et al., 2016; Dokladal et al., 2018; Tan et al., 2018). AtRuvBL1 protein has been assumed as a regulator of R genes so far and is essential in meristem development (Holt et al., 2002). Here we suggest involvement of AtRuvBL proteins in telomerase assembly pathway in A. thaliana. We detected new interactions of AtTRB proteins with AtRuvBL proteins, localized the AtTERTÀAtTRBÀAtRuvBL complex exclusively in the nucleolus and observed that heterozygous T-DNA insertion mutants in AtRuvBL1 or AtRuvBL2a genes showed reduced telomerase activity. Further, our results showed interactions of AtCBF5, the plant orthologue of dyskerin, with AtTRB and AtPOT1, but not with the AtRuvBL proteins, which expanded our knowledge on the telomerase assembly process. Indispensability of the AtRuvBL proteins for the plant development was supported by our finding that homozygous atruvbl1 and atruvbl2a mutant plants were not viable. Furthermore, we identified new homologues RuvBL proteins and analyzed their evolutionary relationships in plants. Altogether, our data show that the plant homologues of Pontin and Reptin, AtRuvBLs, and also AtTRB are involved in telomerase assembly and suggest conservation of telomerase trafficking pathway via the nucleolus to the telomeres in plants. EXPERIMENTAL PROCEDURES Searching transcriptomes and genomes for RuvBL homologues RuvBL homologues were identified by BLASTP searches using A. thaliana proteins from the TAIR database (https://www.arabid opsis.org/) to query NCBI protein databases (http://www.ncbi.nlm. nih.gov/). The BLASTP searches used default parameters, adjusted to the lowest E-value. The duplicates from all searches were eliminated. We conducted an iterative search of the UniProt database (http://www.uniprot.org/) and the Phytozome version 11 database (https://phytozome.jgi.doe.gov) was next searched for proteins not found by BLASTP. We analyzed all sequences independently of their annotations, with no prior assumptions. Information summary of accession numbers for RuvBL are in Data S1 and S2. Sequence alignment Amino acid sequences were aligned using the Clustal Omega algorithm (Sievers et al., 2011) in the Mobyle platform (Neron et al., 2009), with homology detection by HMM-HMM comparisons (Soding, 2005). Protein isoforms with the same length were also used, because the differential expression patterns producing protein isoforms from various tissues suggested that isoforms could have different biological functions in vivo (Chen et al., 2014). Phylogenetic reconstruction Maximum likelihood (ML) analyses of the matrices were performed in RAxML 8.2.4 (Stamatakis, 2014) to examine differences in optimality between alternative topologies. Using the Akaike information criterion as implemented in Modeltest 3.8 (Posada and Crandall, 1998), a GTR+I+Γ model was chosen as the best-fitting model, and 1000 replications were run for bootstrap values. The final data set for RuvBL contained 190 proteins of different species and length 576 bp. Phylogenetic trees were constructed and modified with iTOL v3.4 (Letunic and Bork, 2016). Transgenic constructs The Gateway-compatible donor and destination vectors carrying the AtTERT (AtTERT 1-233, 1-271, 229-582, 597-987, 958-1123) fragments were described in Zachova et al. (2013). The Gateway-compatible donor vectors carrying AtRuvBL1, AtRuvBL2a, AtPOT1a, AtPOT1b and AtGAUT10 were described in Majerska et al. (2017). The AtTRB1, 2 and 3 constructs have described previously (Schrumpfova et al., 2014). © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 208 Sarka Schorova et al. The cloned cDNA sequence of AtCBF5 (GC105080 from Arabidopsis Information Resource (http://www.arabidopsis.org/)) in pENTR223 was used as entry vector. For preparation of yeast twohybrid (Y2H) and/or BiFC constructs, DNA fragments were introduced into the destination Gateway vectors pGBKT7-DEST, pGADT7-DEST (Horak et al., 2008) and/or the pSAT5-DEST-c(175end)EYFPÀC1(B), pSAT4-DEST-n(174)EYFPÀCl (Lee et al., 2012) using the LR recombinase reaction (Invitrogen, Carlsbad, CA, USA). PCR-based genotyping of plant lines Plants with annotated T-DNA insertion within AtRuvBL1 gene (SAIL_397_C11, WiscDsLoxHs027_03G, WiscDsLoxHs117_06F, WiscDsLoxHs168_06D) and AtRuvBL2a gene (GK-543F01, SALK_071103, SALK_144539, SALK_144540, SAIL_500_C04) in the Col-0 background were used (Figure S8). To distinguish between wild-type plants and those that were heterozygous for the T-DNA insertion in the AtRuvBL1 or AtRuvBL2a genes, we isolated genomic DNA from leaves by the standard protocol of Dellaporta et al. (1983). The genomic DNA was used for PCR analysis using MyTaq DNA polymerase (Bioline, http://www.bioline.com). The conditions used were in accordance with the manufacturer’s instructions. The primers used were specific for T-DNA and AtRuvBL1 or AtRuvBL2a genes (Table S3, Figure S9). Thermal conditions were 95°C for 1 min (initial denaturation), followed by 30 cycles of 95°C for 30 sec, 60°C for 30 sec and 72°C for 1 min 20 sec, with a final extension at 72°C for 10 min. RT-PCR Total RNA was extracted from approximately 100 mg of frozen young leaves using an RNeasy plant mini kit (Qiagen, Venlo, Netherlands) and RNA samples were treated with TURBO DNA-free (Applied Biosystems/Ambion, http://www.lifetechnologies.com TURBO DNA-free). The quality and quantity of RNA were determined by electrophoresis on 1% w/v agarose gels and by measurement of absorbance using NanoDropTM 2000/2000c spectrophotometer (https://www.thermofisher.com/). Reverse transcription was performed using random nonamers (Sigma-Aldrich, http://www.sigmaaldrich.com) with 1 lg RNA and Mu-MLV RT (New England Biolabs, https://www.neb.com/). Quantification of transcript levels of the AtRuvBL1, AtRuvBL2a (Figure S10) and AtTERT genes (Fojtova et al., 2011) was carried out by FastStart l SYBR Green Master (Roche, Basel, Switzerland), a Rotorgene 6000 cycler (Qiagen) and using the Ubiquitin-10 gene as suitable references for quantitative analyses in A. thaliana. A 2 ll aliquot of cDNA, from two biological replicates, were added to the 20 ll reaction mix; the final concentration of each forward and reverse primer (sequences are given in Table S3) was 0.25 lM. Three technical replicates were done for each reaction that was measured in triplicates; the PCR cycle consisted of 15 min of initial denaturation followed by 40 cycles of 15 sec at 95°C, 20 sec at 56°C and 30 sec at 72°C. SYBR Green I fluorescence was monitored consecutively after the extension step (Fojtova et al., 2011) sequences of primers are given in Table S3. Statistical analysis was performed using unpaired Student’s t-test. Quantitative TRAP assay Protein extracts from buds were prepared as described by Fitzgerald et al. (1996). qTRAP analysis was performed as described in Herbert et al. (2006) using FastStart SYBR Green Master (Roche) and TS21 and TEL-PR primers. Samples were analyzed in triplicate. A 1 ll aliquot of extract diluted to 50 ng llÀ1 protein concentration was added to the 20 ll reaction mix. Ct values were determined using the Rotorgene 3000 (Qiagen) machine software, and relative telomerase activity was calculated by the DCt method (Pfaffl, 2004). TRF analysis TRF analysis was performed as described previously (Ruckova et al., 2008) using 500 ng genomic DNA isolated from 5 to 7 weeks old rosette leaves using NucleoSpin Plant II (Machery Nagel). Hybridized samples (Hybond XL, GE Healthcare, Chicago, IL, USA) by Southern hybridization method were radioactively marked by random priming, in which the telomeric probe was prepared according to a modified protocol from Ijdo et al. (1991). Telomeric signals were visualized using an FLA7000 imager (Fujifilm, Tokyo, Japan). Evaluation of fragment lengths was performed using a Gene Ruler 1 kb DNA ladder (Fermentas, http://www.thermoscien tificbio.com/fermentas/) as the standard. Mean telomere lengths were calculated as described by Grant et al. (2001). Yeast two-hybrid assay Yeast two-hybrid experiments were performed using the Matchmaker TM GAL4-based two-hybrid system (Clontech, Kyoto, Japan) as described in Schrumpfova et al. (2014). AtRuvBL1 and AtRuvBL2a constructs from pDONR/221 entry clones were subcloned into the Gateway-compatible destination vector pGBKT7DEST (bait vector) and pGADT7-DEST (prey vector). cDNA sequences encoding AtTERT fragments from pDONR/221 entry clones and AtCBF5 from PENTR223 entry clone were subcloned into the Gateway-compatible destination vector pGBKT7-DEST (bait vector). AtPOT1a constructs were subcloned from pDONR/ 221 entry clones into the Gateway-compatible destination vector pGADT7-DEST (prey vector). The pGADT7 prey vectors (Clontech) carrying AtTRB1-3 and AtPOT1a have been described previously (Schrumpfova et al., 2008). Successful co-transformation of each bait/prey combination into Saccharomyces cerevisiae PJ69-4a was confirmed on SD plates lacking Leu and Trp, and positive interactions were selected on SD medium lacking Leu, Trp and His (with or without 3-aminotriazol (3-AT)) or SD medium lacking Leu, Trp and Ade. Co-transformation with an empty vector and homodimerization of the AtTRB1 protein served as negative and positive control, respectively (Schrumpfova et al., 2014). Protein expression was verified by immunoblotting in equal amounts of protein extracts separated by SDS-PAGE (12%), blotted onto nitrocellulose membrane, and probed with mouse anti-Myc (1:1000; SigmaAldrich) and mouse anti-HA (1:1000) primary antibodies binding to specific protein epitope tags of AD- and BD-fusion proteins, followed by an anti-mouse HRP-conjugated secondary antibodies (1:8000; Sigma-Aldrich) for chemiluminescence detection. In vitro translation and co-immunoprecipitation Additionally, the Y2H constructs were employed for verification in assay as described in Schrumpfova et al. (2008). Briefly, radioactively (35 S-Met) labelled proteins with hemagglutinin tag (HA) (pGADT7, pGADT7-DEST), as well as non-radioactively labelled protein partners with a Myc-tag (pGBKT7, pGBKT7-DEST) were separately expressed in the TNT Quick Coupled Transcription/ Translation System (TNT-RRL) (Promega, Fitchburg, WI, USA) in 50 ll of each reaction according to the manufacturer’s instructions. The co-immunoprecipitation procedure was performed as described by Schrumpfova et al. (2008) with 1 lg anti-Myc-tag polyclonal antibody (Sigma-Aldrich, St. Louis, MO, USA) and incubated overnight at 4°C with 10 ll protein G magnetic particles (Dynabeads, Invitrogen-Dynal). © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 Characterization of plant Pontin and Reptin 209 During the co-immunoprecipitation with three proteins of interest, two radioactively labeled proteins with HA-tag (AtRuvBL1, AtTRB3) and one non-radioactively labeled AtTERT 1-271 fragment with Myc-tag were expressed separately in TNT-RRL and incubated in the same manner as previous Co-IP together with protein G magnetic particles (Dynabeads, Invitrogen-Dynal) and 1 lg antiMyc-tag polyclonal antibody (Sigma). Input, Unbound and Bound fractions were separated by 12% SDS-PAGE and analyzed by FLA7000 imager (Fujifilm). Bimolecular fluorescence complementation Arabidopsis thaliana leaf protoplasts were prepared and co-transfected with DNA (10 lg of each construct) as was described in Lee et al. (2012). The same entry vectors (pDONR/221, PENTR223), already used for AtRuvBL1, AtRuvBL2a, AtTERT fragments, AtCBF5 and AtPOT1a Y2H constructs cloning (Majerska et al., 2017) or entry vectors used for cloning AtTRB1-3 (Schrumpfova et al., 2008) were ligated into pSAT5-DEST-c(175-end)EYFPÀC1(B), pSAT4-DEST-n(174)EYFPÀCl vectors. As a negative control, we used the cYFP/AtGAUT10 construct. To control transformation efficiency and to label cell nuclei, we co-transfected a plasmid expressing mRFP fused to the nuclear localization signal of the VirD2 protein of A. tumefaciens (mRFPÀVirD2NLS; Citovsky et al., 2006). To label nucleolus we co-transfected a plasmid expressing mRFP fused to the to AtFibrillarin 1 (Pih et al., 2000). Transfected protoplasts were incubated in the light, at room temperature overnight, and then observed for fluorescence using a Zeiss AxioImager Z1 epifluorescence microscope equipped with filters for YFP (Alexa Fluor 488), RFP (Texas Red) and CY5 (chloroplast autofluorescence). The mRFPÀVirD2NLS and AtGAUT10ÀcEYFP constructs for BiFC experiments were kindly provided by Prof. Stanton B. Gelvin (Purdue University, USA). ACCESSION NUMBERS AtRuvBL1 (AT5G22330); AtRuvBL2a (AT5G67630); AtTERT (AT5G16850); AtTRB1 (AT1G49950); AtTRB2, formerly TBP3 (AT5G67580); AtTRB3, formerly TBP2 (AT3G49850); AtCBF5 (AT3G57150); AtPOT1a (AT2G05210); GAUT10 (AT2G20810); AtFibrillarin 1 (AT5G52470). ACKNOWLEDGEMENTS This work was supported by the Czech Science Foundation (16- 1137S), by the project SYMBIT, reg. number: CZ.02.1.01/0.0/0.0/ 15_003/0000477 financed by the ERDF, and by the Ministry of Education, Youth and Sports of the Czech Republic under the project CEITEC 2020 (LQ1601). We thank Inna Lermontova (IPK Gatersleben) for AtCBF5 constructs. We further thank Eva Sykorova and Vratislav Peska for providing us with the collection of TERT fragments and Fibrillarin 1-RFP vectors. We thank to Miloslava Fojtova for help with qTRAP assay. Jennifer McMahon is gratefully acknowledged for proofreading. The access to the computing and storage facilities owned by parties and projects contributing to the National Grid Infrastructure MetaCentrum provided under the program ‘Projects of Large Infrastructure for Research, Development, and Innovations’ (LM2010005) was highly appreciated, as was the access to the CERIT-SC computing and storage facilities provided under the program ‘Center CERIT Scientific Cloud’, part of the Operational Program Research and Development for Innovations, reg. no. CZ.1.05/3.2.00/08.0144. Core Facility Plant Sciences of CEITEC MU is acknowledged for the cultivation of experimental plants used in this paper. CONFLICTS OF INTEREST The authors declare no conflicts of interest. AUTHORS’ CONTRIBUTIONS PPS designed the study, supervised the project and wrote the manuscript with support from SS, JF, LZD and DH.SS performed the experiments. LZD designed the phylogeny analysis. JF helped to supervise the project and wrote the manuscript. All authors discussed the results and contributed to the final manuscript. SUPPORTING INFORMATION Additional Supporting Information may be found in the online version of this article. Figure S1. AtRuvBL1 does not interact with either the RT-domain or the C-terminus of AtTERT. Figure S2. AtTRB2 proteins directly interact with AtRuvBL2a protein and AtTRB1 is associated with AtRuvBL1 in BiFC assay. Figure S3. Nucleolar or cytoplasmic localization of AtPOT1a– AtCBF5 interactions. Figure S4. Weak interaction between AtPOT1a and AtRuvBL1 pro- teins. Figure S5. Association of AtRuvBLs, AtTRBs and AtTERT in the nucleolus in A. thaliana leaf protoplasts where the whole nucleus is marked. Figure S6. Relative transcript levels of AtTERT gene in AtRuvBL1 and AtRuvBL2a heterozygous mutant plants. Figure S7. Terminal restriction fragment analysis. Figure S8. Schematic illustration of specific primers and T-DNA insertion location within the AtRuvBL1 and AtRuvBL2a genes. Figure S9. Example of PCR analysis of genomic DNA isolated from wild-type (Wt) plants and heterozygous AtRuvBL1 and AtRuvBL2a plants. Figure S10. Relative AtRuvBL1 and AtRuvBL2a transcript levels in heterozygous AtRuvBL1 and AtRuvBL2a plants. Table S1. The number of PPIs foci with exclusively nucleolar local- ization. Table S2. List of T-DNA insertion lines. Table S3. List of primers. Data S1. List of the analyzed plant species sorted by phylogenetic system with number of homologues. Data S2. List of the analyzed plant species for AtRuvBL homologues and their accession numbers. REFERENCES Armstrong, S.J., Franklin, F.C. and Jones, G.H. (2001) Nucleolus-associated telomere clustering and pairing precede meiotic chromosome synapsis in Arabidopsis thaliana. J. Cell Sci. 114, 4207–4217. Azum-Gelade, M.C., Noaillac-Depeyre, J., Caizergues-Ferrer, M. and Gas, N. (1994) Cell cycle redistribution of U3 snRNA and fibrillarin. Presence in the cytoplasmic nucleolus remnant and in the prenucleolar bodies at telophase. J. Cell Sci. 107 (Pt 2), 463–475. Bilaud, T., Koering, C.E., Binet-Brasselet, E., Ancelin, K., Pollice, A., Gasser, S.M. and Gilson, E. (1996) The telobox, a Myb-related telomeric DNA binding motif found in proteins from yeast, plants and human. Nucleic Acids Res. 24, 1294–1303. Boulon, S., Marmier-Gourrier, N., Pradet-Balade, B. et al. (2008) The Hsp90 chaperone controls the biogenesis of L7Ae RNPs through conserved machinery. J. Cell Biol. 180, 579–595. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 210 Sarka Schorova et al. Chen, L.Y., Liu, D. and Songyang, Z. (2007) Telomere maintenance through spatial control of telomeric proteins. Mol. Cell. Biol. 27, 5898–5909. Chen, Y., Zou, M. and Cao, Y. (2014) Transcriptome analysis of the Arabidopsis semi-in vivo pollen tube guidance system uncovers a distinct gene expression profile. J. Plant Biol. 57, 93–105. Cheung, K.L., Huen, J., Houry, W.A. and Ortega, J. (2010) Comparison of the multiple oligomeric structures observed for the Rvb1 and Rvb2 proteins. Biochem. Cell Biol. 88, 77–88. Cioce, M. and Lamond, A.I. (2005) Cajal bodies: a long history of discovery. Annu. Rev. Cell Dev. Biol. 21, 105–131. Citovsky, V., Lee, L.Y., Vyas, S., Glick, E., Chen, M.H., Vainstein, A., Gafni, Y., Gelvin, S.B. and Tzfira, T. (2006) Subcellular localization of interacting proteins by bimolecular fluorescence complementation in planta. J. Mol. Biol. 362, 1120–1131. Dellaporta, S.L., Wood, J. and Hicks, J.B. (1983) A Plant DNA minipreparation. Version II. Plant. Mol. Biol. Rep. 1, 19–21. Dokladal, L., Benkova, E., Honys, D., Duplˇakova, N., Lee, L.Y., Gelvin, S.B. and Sykorova, E. (2018) An armadillo-domain protein participates in a telomerase interaction network. Plant Mol. Biol. 97, 407–420. Dreissig, S., Schiml, S., Schindele, P., Weiss, O., Rutten, T., Schubert, V., Gladilin, E., Mette, M.F., Puchta, H. and Houben, A. (2017) Live-cell CRISPR imaging in plants reveals dynamic telomere movements. Plant J. 91, 565–573. Dvorackova, M. (2010) Analysis of Arabidopsis telomere-associated proteins in vivo. PhD thesis, Masaryk University, Brno. Dvorackova, M., Rossignol, P., Shaw, P.J., Koroleva, O.A., Doonan, J.H. and Fajkus, J. (2010) AtTRB1, a telomeric DNA-binding protein from Arabidopsis, is concentrated in the nucleolus and shows highly dynamic association with chromatin. Plant J. 61, 637–649. Fajkus, J. and Trifonov, E.N. (2001) Columnar packing of telomeric nucleosomes. Biochem. Biophys. Res. Commun. 280, 961–963. Fitzgerald, M.S., McKnight, T.D. and Shippen, D.E. (1996) Characterization and developmental patterns of telomerase expression in plants. Proc. Natl Acad. Sci. USA, 93, 14422–14427. Flavin, P., Redmond, A., McBryan, J. et al. (2011) RuvBl2 cooperates with Ets2 to transcriptionally regulate hTERT in colon cancer. FEBS Lett. 585, 2537–2544. Fojtova, M., Peska, V., Dobsakova, Z., Mozgova, I., Fajkus, J. and Sykorova, E. (2011) Molecular analysis of T-DNA insertion mutants identified putative regulatory elements in the AtTERT gene. J. Exp. Bot. 62, 5531–5545. Fransz, P., De Jong, J.H., Lysak, M., Castiglione, M.R. and Schubert, I. (2002) Interphase chromosomes in Arabidopsis are organized as well defined chromocenters from which euchromatin loops emanate. Proc. Natl Acad. Sci. USA, 99, 14584–14589. Freeling, M. (2009) Bias in plant gene content following different sorts of duplication: tandem, whole-genome, segmental, or by transposition. Annu. Rev. Plant Biol. 60, 433–453. Gallant, P. (2007) Control of transcription by Pontin and Reptin. Trends Cell Biol. 17, 187–192. Grant, J.D., Broccoli, D., Muquit, M., Manion, F.J., Tisdall, J. and Ochs, M.F. (2001) Telometric: a tool providing simplified, reproducible measurements of telomeric DNA from constant field agarose gels. Biotechniques, 31, 1314–1318. Greider, C.W. (1996) Telomere length regulation. Annu. Rev. Biochem. 65, 337–365. He, J., Navarrete, S., Jasinski, M., Vulliamy, T., Dokal, I., Bessler, M. and Mason, P.J. (2002) Targeted disruption of Dkc1, the gene mutated in Xlinked dyskeratosis congenita, causes embryonic lethality in mice. Oncogene, 21, 7740–7744. Herbert, B.S., Hochreiter, A.E., Wright, W.E. and Shay, J.W. (2006) Nonradioactive detection of telomerase activity using the telomeric repeat amplification protocol. Nat. Protoc. 1, 1583–1590. Holt, B.F., Boyes, D.C., Ellerstr€om, M., Siefers, N., Wiig, A., Kauffman, S., Grant, M.R. and Dangl, J.L. (2002) An evolutionarily conserved mediator of plant disease resistance gene function is required for normal Arabidopsis development. Dev. Cell, 2, 807–817. Horak, J., Grefen, C., Berendzen, K.W., Hahn, A., Stierhof, Y.D., Stadelhofer, B., Stahl, M., Koncz, C. and Harter, K. (2008) The Arabidopsis thaliana response regulator ARR22 is a putative AHP phospho-histidine phosphatase expressed in the chalaza of developing seeds. BMC Plant Biol. 8, 77. Ijdo, J.W., Wells, R.A., Baldini, A. and Reeders, S.T. (1991) Improved telomere detection using a telomere repeat probe (TTAGGG)n generated by PCR. Nucleic Acids Res. 19, 4780. Izumi, N., Yamashita, A. and Ohno, S. (2012) Integrated regulation of PIKKmediated stress responses by AAA+ proteins RUVBL1 and RUVBL2. Nucleus, 3, 29–43. Jha, S., Shibata, E. and Dutta, A. (2008) Human Rvb1/Tip49 is required for the histone acetyltransferase activity of Tip60/NuA4 and for the downregulation of phosphorylation on H2AX after DNA damage. Mol. Cell. Biol. 28, 2690–2700. Kannan, K., Nelson, A.D. and Shippen, D.E. (2008) Dyskerin is a component of the Arabidopsis telomerase RNP required for telomere maintenance. Mol. Cell. Biol. 28, 2332–2341. Kerppola, T.K. (2009) Visualization of molecular interactions using bimolecular fluorescence complementation analysis: characteristics of protein fragment complementation. Chem. Soc. Rev. 38, 2876–2886. Khurts, S., Masutomi, K., Delgermaa, L., Arai, K., Oishi, N., Mizuno, H., Hayashi, N., Hahn, W.C. and Murakami, S. (2004) Nucleolin interacts with telomerase. J. Biol. Chem. 279, 51508–51515. de Lange, T. (2005) Shelterin: the protein complex that shapes and safeguards human telomeres. Genes Dev. 19, 2100–2110. Lee, W.K. and Cho, M.H. (2016) Telomere-binding protein regulates the chromosome ends through the interaction with histone deacetylases in Arabidopsis thaliana. Nucleic Acids Res. 44, 4610–4624. Lee, L.Y., Wu, F.H., Hsu, C.T. et al. (2012) Screening a cDNA library for protein–protein interactions directly in planta. Plant Cell, 24, 1746–1759. Lee, J.H., Lee, Y.S., Jeong, S.A., Khadka, P., Roth, J. and Chung, I.K. (2014) Catalytically active telomerase holoenzyme is assembled in the dense fibrillar component of the nucleolus during S phase. Histochem. Cell Biol. 141, 137–152. Lermontova, I., Schubert, V., Bornke, F., Macas, J. and Schubert, I. (2007) Arabidopsis CBF5 interacts with the H/ACA snoRNP assembly factor NAF1. Plant Mol. Biol. 65, 615–626. Letunic, I. and Bork, P. (2016) Interactive tree of life (iTOL) v3: an online tool for the display and annotation of phylogenetic and other trees. Nucleic Acids Res. 44, W242–W245. Li, W., Zeng, J., Li, Q., Zhao, L., Liu, T., Bjorkholm, M., Jia, J. and Xu, D. (2010) Reptin is required for the transcription of telomerase reverse transcriptase and over-expressed in gastric cancer. Mol. Cancer, 9, 132. Lim, C.J., Zaug, A.J., Kim, H.J. and Cech, T.R. (2017) Reconstitution of human shelterin complexes reveals unexpected stoichiometry and dual pathways to enhance telomerase processivity. Nat. Commun. 8, 1075. Machado-Pinilla, R., Liger, D., Leulliot, N. and Meier, U.T. (2012) Mechanism of the AAA+ ATPases pontin and reptin in the biogenesis of H/ACA RNPs. RNA, 18, 1833–1845. MacNeil, D.E., Bensoussan, H.J. and Autexier, C. (2016) Telomerase regulation from beginning to the end. Genes (Basel), 7, 1–33. Majerska, J., Schrumpfova, P.P., Dokladal, L., Schorova, S., Stejskal, K., Oboril, M., Honys, D., Kozakova, L., Polanska, P.S. and Sykorova, E. (2017) Tandem affinity purification of AtTERT reveals putative interaction partners of plant telomerase in vivo. Protoplasma, 254, 1547–1562. Mandakova, T. and Lysak, M.A. (2008) Chromosomal phylogeny and karyotype evolution in x=7 crucifer species (Brassicaceae). Plant Cell, 20, 2559–2570. Mao, Y.Q. and Houry, W.A. (2017) The role of Pontin and Reptin in cellular physiology and cancer etiology. Front. Mol. Biosci. 4, 58. Matias, P.M., Gorynia, S., Donner, P. and Carrondo, M.A. (2006) Crystal structure of the human AAA+ protein RuvBL1. J. Biol. Chem. 281, 38918–38929. McKeegan, K.S., Debieux, C.M., Boulon, S., Bertrand, E. and Watkins, N.J. (2007) A dynamic scaffold of pre-snoRNP factors facilitates human box C/D snoRNP assembly. Mol. Cell. Biol. 27, 6782–6793. Mozgova, I., Schrumpfova, P.P., Hofr, C. and Fajkus, J. (2008) Functional characterization of domains in AtTRB1, a putative telomere-binding protein in Arabidopsis thaliana. Phytochemistry, 69, 1814–1819. Neron, B., Menager, H., Maufrais, C., Joly, N., Maupetit, J., Letort, S., Carrere, S., Tuffery, P. and Letondal, C. (2009) Mobyle: a new full web bioinformatics framework. Bioinformatics, 25, 3005–3011. Niewiarowski, A., Bradley, A.S., Gor, J., McKay, A.R., Perkins, S.J. and Tsaneva, I.R. (2010) Oligomeric assembly and interactions within the human RuvB-like RuvBL1 and RuvBL2 complexes. Biochem. J. 429, 113–125. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 Characterization of plant Pontin and Reptin 211 Ohdate, H., Lim, C.R., Kokubo, T., Matsubara, K., Kimata, Y. and Kohno, K. (2003) Impairment of the DNA binding activity of the TATA-binding protein renders the transcriptional function of Rvb2p/Tih2p, the yeast RuvBlike protein, essential for cell growth. J. Biol. Chem. 278, 14647–14656. Osaki, H., Walf-Vorderwulbecke, V., Mangolini, M., Zhao, L., Horton, S.J., Morrone, G., Schuringa, J.J., de Boer, J. and Williams, O. (2013) The AAA+ ATPase RUVBL2 is a critical mediator of MLL-AF9 oncogenesis. Leukemia, 27, 1461–1468. Pendle, A.F., Clark, G.P., Boon, R., Lewandowska, D., Lam, Y.W., Andersen, J., Mann, M., Lamond, A.I., Brown, J.W. and Shaw, P.J. (2005) Proteomic analysis of the Arabidopsis nucleolus suggests novel nucleolar functions. Mol. Biol. Cell, 16, 260–269. Peska, V., Schrumpfova, P.P. and Fajkus, J. (2011) Using the telobox to search for plant telomere binding proteins. Curr. Protein Pept. Sci. 12, 75–83. Pfaffl, M.W. (2004) Quantification strategies in real-time PCR. In A-Z of Quantitative PCR (Bustin, S.A., ed). La Jolla, CA: International University Line, pp 87–112. Pih, K.T., Yi, M.J., Liang, Y.S., Shin, B.J., Cho, M.J., Hwang, I. and Son, D. (2000) Molecular cloning and targeting of a fibrillarin homolog from Arabidopsis. Plant Physiol. 123, 51–58. Pontvianne, F., Carpentier, M.C., Durut, N. et al. (2016) Identification of nucleolus-associated chromatin domains reveals a role for the nucleolus in 3D organization of the A. thaliana genome. Cell Rep. 16, 1574– 1587. Posada, D. and Crandall, K.A. (1998) MODELTEST: testing the model of DNA substitution. Bioinformatics, 14, 817–818. Queval, R., Papin, C., Dalvai, M., Bystricky, K. and Humbert, O. (2014) Reptin and Pontin oligomerization and activity are modulated through histone H3 N-terminal tail interaction. J. Biol. Chem. 289, 33999–34012. Roberts, N.Y., Osman, K. and Armstrong, S.J. (2009) Telomere distribution and dynamics in somatic and meiotic nuclei of Arabidopsis thaliana. Cytogenet. Genome Res. 124, 193–201. Rosenbaum, J., Baek, S.H., Dutta, A., Houry, W.A., Huber, O., Hupp, T.R. and Matias, P.M. (2013) The emergence of the conserved AAA+ ATPases Pontin and Reptin on the signaling landscape. Sci. Signal. 6, mr1. Rossignol, P., Collier, S., Bush, M., Shaw, P. and Doonan, J.H. (2007) Arabidopsis POT1A interacts with TERT-V(I8), an N-terminal splicing variant of telomerase. J. Cell Sci. 120, 3678–3687. Ruckova, E., Friml, J., Prochazkova Schrumpfova, P. and Fajkus, J. (2008) Role of alternative telomere lengthening unmasked in telomerase knockout mutant plants. Plant Mol. Biol. 66, 637–646. Schmidt, J.C. and Cech, T.R. (2015) Human telomerase: biogenesis, trafficking, recruitment, and activation. Genes Dev. 29, 1095–1105. Schmidt, J.C., Zaug, A.J. and Cech, T.R. (2016) Live cell imaging reveals the dynamics of telomerase recruitment to telomeres. Cell, 166, 1188–1197 e1189. Schrumpfova, P., Kuchar, M., Mikova, G., Skrısovska, L., Kubicarova, T. and Fajkus, J. (2004) Characterization of two Arabidopsis thaliana myb-like proteins showing affinity to telomeric DNA sequence. Genome, 47, 316–324. Schrumpfova, P.P., Kuchar, M., Palecek, J. and Fajkus, J. (2008) Mapping of interaction domains of putative telomere-binding proteins AtTRB1 and AtPOT1b from Arabidopsis thaliana. FEBS Lett. 582, 1400–1406. Schrumpfova, P.P., Vychodilova, I., Dvorackova, M., Majerska, J., Dokladal, L., Schorova, S. and Fajkus, J. (2014) Telomere repeat binding proteins are functional components of Arabidopsis telomeres and interact with telomerase. Plant J. 77, 770–781. Schrumpfova, P., Schorova, S. and Fajkus, J. (2016a) Telomere- and telomerase-associated proteins and their functions in the plant cell. Front. Plant Sci. 7, 851. Schrumpfova, P.P., Vychodilova, I., Hapala, J., Schorova, S., Dvoracek, V. and Fajkus, J. (2016b) Telomere binding protein TRB1 is associated with promoters of translation machinery genes in vivo. Plant Mol. Biol. 90, 189–206. Schrumpfova, P.P., Fojtova, M. and Fajkus, J. (2019) Telomeres in plants and humans: not so different, not so similar. Cells, 8, 1–31. Shaw, P. and Brown, J. (2012) Nucleoli: composition, function, and dynamics. Plant Physiol. 158, 44–51. Sievers, F., Wilm, A., Dineen, D. et al. (2011) Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539. Silva-Martin, N., Dauden, M.I., Glatt, S., Hoffmann, N.A., Kastritis, P., Bork, P., Beck, M. and Muller, C.W. (2016) The combination of x-ray crystallography and cryo-electron microscopy provides insight into the overall architecture of the dodecameric Rvb1/Rvb2 complex. PLoS One, 11, e0146457. Soding, J. (2005) Protein homology detection by HMM-HMM comparison. Bioinformatics, 21, 951–960. Stamatakis, A. (2014) RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics, 30, 1312–1313. Stepinski, D. (2014) Functional ultrastructure of the plant nucleolus. Protoplasma, 251, 1285–1306. Tan, L.M., Zhang, C.J., Hou, X.M., Shao, C.R., Lu, Y.J., Zhou, J.X., Li, Y.Q., Li, L., Chen, S. and He, X.J. (2018) The PEAT protein complexes are required for histone deacetylation and heterochromatin silencing. EMBO J. 37, 1–21. Torreira, E., Jha, S., Lopez-Blanco, J.R., Arias-Palomo, E., Chacon, P., Canas, C., Ayora, S., Dutta, A. and Llorca, O. (2008) Architecture of the pontin/ reptin complex, essential in the assembly of several macromolecular complexes. Structure, 16, 1511–1520. Venteicher, A.S., Meng, Z., Mason, P.J., Veenstra, T.D. and Artandi, S.E. (2008) Identification of ATPases pontin and reptin as telomerase components essential for holoenzyme assembly. Cell, 132, 945–957. Watkins, N.J., Lemm, I., Ingelfinger, D., Schneider, C., Hossbach, M., Urlaub, H. and Luhrmann, R. (2004) Assembly and maturation of the U3 snoRNP in the nucleoplasm in a large dynamic multiprotein complex. Mol. Cell, 16, 789–798. Wood, M.A., McMahon, S.B. and Cole, M.D. (2000) An ATPase/helicase complex is an essential cofactor for oncogenic transformation by c-Myc. Mol. Cell, 5, 321–330. Zachova, D., Fojtova, M., Dvorackova, M., Mozgova, I., Lermontova, I., Peska, V., Schubert, I., Fajkus, J. and Sykorova, E. (2013) Structure-function relationships during transgenic telomerase expression in Arabidopsis. Physiol. Plant. 149, 114–126. Zhang, Q., Kim, N.K. and Feigon, J. (2011) Architecture of human telomerase RNA. Proc. Natl Acad. Sci. USA, 108, 20325–20332. Zhao, R., Kakihara, Y., Gribun, A. et al. (2008) Molecular chaperone Hsp90 stabilizes Pih1/Nop17 to maintain R2TP complex activity that regulates snoRNA accumulation. J. Cell Biol. 180, 563–578. Zhao, Y., Cheng, D., Wang, S. and Zhu, J. (2014) Dual roles of c-Myc in the regulation of hTERT gene. Nucleic Acids Res. 42, 10385– 10398. Zhou, Y., Hartwig, B., James, G.V., Schneeberger, K. and Turck, F. (2016) Complementary activities of TELOMERE REPEAT BINDING proteins and polycomb group complexes in transcriptional regulation of target genes. Plant Cell, 28, 87–101. Zhou, Y., Wang, Y., Krause, K., Yang, T., Dongus, J.A., Zhang, Y. and Turck, F. (2018) Telobox motifs recruit CLF/SWN-PRC2 for H3K27me3 deposition via TRB factors in Arabidopsis. Nat. Genet. 50, 638–644. © 2019 The Authors The Plant Journal © 2019 John Wiley & Sons Ltd, The Plant Journal, (2019), 98, 195–212 212 Sarka Schorova et al. SUPPORTING INFORMATION Figure S1. AtRuvBL1 Does Not Interact with either the RT-domain or the C-terminus of AtTERT. The analyses are performed as described in Figure 1. (a) Schematic depiction of the catalytic subunit of plant telomerase - AtTERT. (b) BiFC assay detects interaction between cYFP/AtRuvBL1 and the N-terminal part of AtTERT (nYFP/AtTERT 1-271). No interactions are detected between AtRuvBL1 protein and AtTERT fragments covering AtTERT domains localized in regions 229-582, 597-987 and 972-1123. No interactions are observed between any of AtTERT fragments and AtRuvBL2a protein in A. thaliana leaf protoplasts. PPIs are marked with white arrows. AtGAUT10, negative control; RFP, nucleus marker; YFP, detects PPI; Chl, Chloroplast autofluorescence. Scale bars = 10 μm. (c) Y2H system results show no interaction between AD/AtRuvBL1 nor AD/AtRuvBL2a proteins and fragments of AtTERT fused with GAL4 DNA-binding domain (BD). Vectors with AD or BD without gene of interest are used as negative control. (d) Co-IP analysis does not detect interaction between AtTERT fragments and AtRuvBL1 or AtRuvBL2a protein as is demonstrated by BiFC. I, Input; U, Unbound; B, Bound fractions; asterisks*,35 S-labelling. Figure S2. AtTRB2 Proteins Directly Interact with AtRuvBL2a Protein and AtTRB1 is Associated with AtRuvBL1 in BiFC Assay. The methods are performed as is described in Figure 1. (a) Schematic representation of the conserved motifs of the AtTRB1 and AtTRB2 proteins from A. thaliana. Myb-like, Telobox-containing Myb domain; H1/H5, histone-like domain; coiled-coil, Cterminal domain. (b) BiFC assay reveals interactions between nYFP/AtTRB1 and cYFP/AtRuvBL1 and between nYFP/AtTRB2 and cYFP/AtRuvBL2a. PPIs are marked with white arrows. AtGAUT10, negative control; RFP, Red Fluorescent Protein, positive control; YFP, Yellow Fluorescent Protein, PPI; Chl, Chloroplast autofluorescence, control. Scale bars = 10 μm. (c) Y2H results confirm the interaction between AD/AtTRB2 and BD/AtRuvBL2 on His- deficient plates, but not the interaction between AD/AtTRB1 and BD/AtRuvBL1. BD, GAL4 DNA-binding domain; AD, GAL4 activation domain; asterisks*, 5mM 3-aminotriazol. (d) Co-IP results confirm interactions between radioactively labelled AtTRB2 and Myc-tagged AtRuvBL2 protein. I, Input; U, Unbound; B, Bound fractions; asterisks*,35S-labelling. Figure S3. Nucleolar or Cytoplasmic Localization of AtPOT1a-AtCBF5 Interactions. BiFC performed in A. thaliana leaf protoplasts show either clear nucleolar AtPOT1a-AtCBF5 foci or cytoplasmic and nuclear foci. The examples of AtPOT1a-AtCBF5 interactions localized in the nucleolus are given in the first four columns. The following four columns show AtPOT1a-AtCBF5 interactions localized in the cytoplasmic and nucleus foci. The whole nucleus was marked by mRFP- VirD2NLS. Figure S4. Weak Interaction Between AtPOT1a and AtRuvBL1 Proteins. The analyses are performed as is described in Figure 1. (a) Y2H results show weak but reproducible interactions between AtPOT1a and AtRuvBL1 protein on His- deficient plates. BD, GAL4 DNA-binding domain; AD, GAL4 activation domain. (b) Co-IP results confirm direct interactions between radioactively labelled AtRuvBL1 and Myctagged AtPOT1a protein. I, Input; U, Unbound; B, Bound fractions; asterisks*, 35S-labelling (c) BiFC does not detect any interaction between AtPOT1a and AtRuvBL1 proteins. AtGAUT10, negative control; RFP, nucleus marker; YFP, detects PPI; Chl, Chloroplast autofluorescence. Scale bars = 10 μm Figure S5. Association of AtRuvBLs, AtTRBs and AtTERT in the nucleolus in A. thaliana Leaf Protoplasts where the Whole Nucleus is Marked. A. thaliana leaf protoplast are co-transfected with mRFP-VirD2NLS encoding RFP that labels the whole nucleus and simultaneously with each of the plasmids encoding nYFP-tagged or cYFP-tagged AtRuvBL1, AtRuvBL2a, AtTERT 1-271, AtTRB3 or AtCBF5 to determine PPI localization. AtRuvBL1AtTERT, AtTRB3-AtTERT, AtRuvBL1-AtTRB3 or AtRuvBL2a-AtTRB3 interactions show nucleolar localization. Plant homologue of mammalian dyskerin, AtCBF5, is associated with AtTRB3 in nucleolus and in additional nuclear bodies at the periphery of nucleolus. RFP, marked nucleus; YFP, detects PPI; Scale bars = 10 μm. Figure S6. Relative Transcript Levels of AtTERT Gene in AtRuvBL1 and AtRuvBL2a Heterozygous Mutant Plants. No significant differences are detected in the transcript levels of AtTERT gene in the heterozygous mutant plants in AtRuvBL1 gene. We detect a slight increase of transcripts of AtTERT gene in the plants heterozygous in AtRuvBL2a genes compared to the wild-type. Levels or AtTERT transcripts from WT Col-0 are arbitrarily set as 1. P values<0.05 are considered significant. Single star denotes 0.01