Dynamics and plasticity of chromosome ends: consequences in human pathologies

 

Caroline Schluth-Bolard 1, 2, Alexandre Ottaviani 1, 2, Amadou Bah 1, Amina Boussouar 1, Eric Gilson 1 and Frédérique Magdinier1, *

1 Laboratoire de Biologie Moléculaire de la Cellule, CNRS UMR5239, Ecole Normale Supérieure de Lyon, UCBL1, IFR128. 46 allée d'Italie, 69364 Lyon Cedex 07, France
2 These two authors contributed equally to this work

* To whom correspondence should be addressed. Email: Frederique.Magdinier@ens-lyon.fr; Frederique.Magdinier@univmed.fr

 

June 2009

 

 

Abstract
In eukaryotic cells, chromosome ends of linear chromosomes are particular regions of the genome formed by telomeres at the very end and subtelomeres, complex sequences that separate telomeres from chromosome-specific regions. These two regions are highly dynamic are contribute to the stability and integrity of the human genome. Furthermore, in human cells, dysregulation of these regions are implicated in a wide range of physiological events and pathological manifestations. Due to the amount of information available on the biology of telomeres, we will give an overview and discuss here what is currently known of the regulation of telomere length and homeostasis and describe the complex organization of subtelomeric regions and their implication in multiple pathologies.

Introduction
In eukaryotic cells, the genetic information is carried by linear chromosomes, which start and end with a particular structure named telomere. Telomeres are constituted by non-coding repeated DNA, usually considered of heterochromatic nature. They are bound by a specific multi-proteins complex. In particular, mammalian telomeres are constituted by TTAGGG repeats, bound to shelterin or telosome complexes containing six different polypeptides (Gilson and Geli, 2007; Palm and de Lange, 2008).
Telomeres are maintained by a ribonucleoprotein complex named telomerase. This cellular reverse transcriptase counteracts telomere shortening resulting from the incomplete telomere elongation after each round of DNA replication (Gilson and Geli, 2007) by using an RNA component complementary to the telomeric repeats. In the absence of telomerase in somatic cells, telomeres shorten at each round of replication because of the asymmetrical replication process that does not allow the entire replication of the 5' to 3' end. Over the years, accumulating evidence has strengthened the view that excessive telomere erosion is a tumor-suppressor mechanism and contributes to acquired and inherited aging processes. Thus, the balance between telomere shortening and telomerase activity and the regulation of telomere length, is a signal controlling the fate of the cell: proliferation, senescence or apoptosis.
Subtelomeres are complex sequences that separate the chromosome ends from gene specific regions. These regions are composed of repetitive sequences usually shared by different chromosome, degenerated telomeric repeats and numerous families of genes or individual ones. These regions are highly polymorphic and various haplotypes are found in the human population. The frequent rearrangements involving these regions are associated with evolution but also several unrelated human pathologies (Linardopoulou et al., 2005; Ottaviani et al., 2008; Riethman et al., 2001).
This review mainly focuses on the regulation of human telomeres and subtelomeres with a particular emphasis on their importance in pathologies.

I. Telomeres

I-A. General information
Telomeres are specialized chromatin structures at the end of linear chromosomes in eukaryotic species that prevent the chromosome ends from being recognized and processed as double strand breaks. The first clear recognition of chromosome ends and of their importance in chromosome biology came from the cytogenetic observations of Barbara Mc Clintock in maize in 1931 (McClintock, 1931). Upon X-ray irradiation, she observed an increased rate of chromosomal translocations and the appearance of ring chromosomes and envisaged that such rearrangements were caused by fusion of broken chromosome ends giving for the first time a particular role to chromosome termini in the maintenance of the genome. In the 70s, Blackburn and Gall cloned the first telomeric DNA in the ciliate Tetrahymena thermophila and described the presence of the repetitive TTGGGG sequence (Blackburn and Gall, 1978).
In many organisms, including vertebrates, plants and many protistes, the telomeres are composed of tandem repeats of simple G-rich sequences. In these cases, the G-rich strand always corresponds to the 3' end, which protrudes as a 3' overhang. Of note, the C. elegans telomeres exhibit both 3' G-rich and 5' C-rich overhangs (Raices et al., 2008). In Drosophila, the telomeres are formed by the insertion of specific retroposons. In mammals, telomeres consist in a G-rich TTAGGG hexanucleotidic sequence repeated thousands of times. The size of the telomeres varies among species from a few base pairs in ciliates to thousands of base pairs in higher eukaryotes (Fajkus et al., 1995; Kipling and Cooke, 1990; Klobutcher et al., 1981) and the size of the 3' overhang in mammals varies between 50 and 500 nucleotides (Figure 1). Telomeres are organized in a large duplex structure than can be seen by electron microscopy, called the t-loop (Telomere loop), which presumably forms through strand invasion of the duplex telomeric repeat by the 3' overhang. Since this structure hides the 3' overhang region, it has been proposed that such a conformation would protect the telomere terminus from being recognized as a damaged DNA sequence and subsequently processed by the DNA damage machinery.
Due to their particular nature, telomeres are processed and maintained by specialized pathways and protein complexes that we will describe below (Figure 1).

Figure 1. Schematic representation of chromosome ends.
Chromosome termini ends with an array of hexanucleotides repeated in tandem. The telomeric sequence found in higher eukaryotes is shown with the protruding 5' overhang ending with a G nucleotide (see text for details). Telomeric sequences are bound by a complex formed by 6 different proteins with specificity for either the single or the double strand DNA. These proteins form the shelterin or telosome complex. TRF1 and TIN2 regulate telomere length and TRF2 protects ends from fusion and prevent activation of the DNA damage response. Additional proteins able to interact with telomeric proteins and involved in DNA damage response and double-strand break repair have also been implicated in telomere length regulation and chromosome end protection (De Boeck et al., 2009; Gilson and Geli, 2007; Palm and de Lange, 2008). Some of these proteins are indicated here.

I-B. Telomere maintenance and telomere binding proteins
Telomeres are maintained by a specialized ribonucleoprotein complex named telomerase. The telomerase comprises a reverse transcriptase (TERT), a RNA component (TERC) and associates with Dyskerin. This enzymatic complex extends the 3' end of chromosomes by reverse transcription of the template region of its tightly associated RNA moiety. Telomerase expression is required for unlimited proliferation of yeast, protozoa and immortal human tumor cells, as well as for the extended proliferation in germinal, embryonic and some stem cells where it contributes to the maintenance of telomere length but is absent in most somatic cells. In the absence of telomerase, telomeres shorten at each round of replication in cells ongoing division partly because of the asymmetrical replication process that cannot allow the entire replication of parental DNA (end of replication problem, see below) (Gilson and Geli, 2007) (Figure 2).
For the TERT subunit, distantly related organisms share regions of conserved sequence motifs such as the amino-terminal, the reverse transcriptase and the carboxy-terminal domains. Each domain has a specialized function including catalysis, nucleolar localization, RNA binding, dimerization and recruitment to the telomere.
In mammals, telomeres are bound by a specialized complex formed by 6 proteins (TRF1, TRF2, POT1, TIN2, TPP1 and RAP1) and called shelterin (Figure 1). These components either bind to the double stranded DNA regions (TRF1, TRF2 and their interacting factors RAP1 (repressor activator protein 1) and TIN2 (TRF1-interacting nuclear protein 2) or the to the G-strand overhang such as the POT1-TPP1 heterodimer (Blasco, 2007a; Gilson and Geli, 2007; Palm and de Lange, 2008). This complex is implicated in the formation of the t-loop, affects the structure of telomere terminus and controls the synthesis of telomeric DNA by telomerase. The assembly of this complex relies on two bridging proteins that interact with each other, Tin2 that bridges TRF1 to TRF2 and TPP1 that bridges the TRF1-TRF2-Tin2 complex to POT1. However, the shelterin components can be found in cells as separate subcomplexes (Chen et al., 2008; Liu et al., 2004a). Moreover, photobleaching experiments revealed pools of TRF1 and TRF2 proteins with different dynamics on telomeric DNA in vivo (Mattern et al., 2004).

TRF1 was the first human telomeric binding protein isolated by biochemical methods (Bilaud et al., 1996; Chong et al., 1995). TRF1 and TRF2 share a common domain structure consisting of the TRF homology domain (TRFH) allowing homo- and heterodimerization and a C-terminal SANT/Myb DNA binding domain, which are connected through a flexible hinge domain. TRF2 N-terminal domain contains basic amino acids while TRF1 contains acidic residues. The TRFH domain of TRF1 and TRF2 contain a docking site through which they recruit other proteins to telomeres (Chen et al., 2008). In particular, TRF1 interacts with the Ku70/80 heterodimer, the BLM helicase (Hsu et al., 2000; Lillard-Wetherell et al., 2004; Opresko et al., 2004), the ATM kinase (Kishi et al., 2001), the nucleotide diphosphate kinase nm23-H2 (Nosaka et al., 1998), components of the mitotic spindle (Nakamura et al., 2001) and the transcriptional repressor SALL1 (Netzer et al., 2001).

TRF2 interacts with components of the Mre11 complex involved in non homologous end joining (NHEJ) and homologous recombination (HR) (Zhu et al., 2000), the Apollo nuclease (Lenain et al., 2006; van Overbeek and de Lange, 2006), the MDC1 DNA signaling factor (Dimitrova and de Lange, 2006), the nucleotide excision repair XPF-ERCC1 nuclease (Zhu et al., 2003) and ATM, canceling thereby the DNA damage response activated by ATM (Figure 1). In addition to binding telomeric sequences, TRF2 can fold DNA by creating higher-order structures such as the t-loop, a lasso-like structure where the end of the telomeric tract is joined to more internal telomeric sequences, or topologically-constraint DNA-protein complexes (Amiard et al., 2007). TRF2 specifically recognize the junction between double and single stranded DNA allowing the G-strand overhang to be sequestered into a protective structure (Griffith et al., 1999; Khan et al., 2007; Stansel et al., 2001). Moreover, TRF2 greatly increases the rate of Holliday junction (HJ) formation and blocks the cleavage by various types of HJ resolving activities (Poulet et al., 2009). Therefore, TRF2 could favor t-loop also by stimulating HJ formation and by preventing resolvase cleavage.
Abrogation of TRF2, or expression of a TRF2 dominant negative allele, results in loss of the G-strand overhang leading to end-to-end chromosome fusions and p53 mediated apoptosis (Denchi and de Lange, 2007; Karlseder et al., 1999; van Steensel et al., 1998). TRF1 and TRF2 are also involved in the protein counting model for telomere length regulation that controls in cis the elongation by telomerase (Ancelin et al., 2002; Loayza et al., 2004).

RAP1 is a 49 kDa protein composed of three domains: a Myb domain, an N-terminal BRCT motif and a C-terminal domain that mediates interaction with TRF2 (Li and de Lange, 2003; Li et al., 2000). Mammalian RAP1 is not able to bind directly to telomeric DNA but associates to telomere through interaction with TRF2. RAP1 has a role in telomere length regulation (Ye et al., 2004) and interacts with several components of the DNA damage machinery.

TIN2 is a 40 kDa protein that binds both TRF1 and TRF2 (Ye et al., 2004) by its central region and its amino-terminal end, respectively. It also binds to TPP1 and depletion or mutation of TIN2 has profound effect on the stability of the shelterin complex (Kim et al., 2004; Ye et al., 2004).

TPP1 (formerly PTOP, PIP1, TINT1) connects POT1 and TIN2 (Hockemeyer et al., 2005; Houghtaling et al., 2004; Liu et al., 2004b; Ye and de Lange, 2004). Impaired TPP1 function is associated with telomere deprotection and telomere loss. TPP1 binds to the amino-terminal half of TIN2 and to the carboxy-terminal region of POT1 and is important for the recruitment of POT1 to telomeres. A splice defect in the TPP1 gene results in adrenocortical dysplasia that arose spontaneously in laboratory mouse (Keegan et al., 2005).

POT1 (Protection of Telomeres 1) contains two OB folds in its N-terminus mediating the interaction with the G-strand telomeric sequence with a high affinity for the 5'-(T)TAGGGTTAG-3' sequence (Baumann and Cech, 2001; Kelleher et al., 2005; Lei et al., 2005; Loayza et al., 2004). The crystal structure of POT1 suggests a role in the protection of the 3' end (Lei et al., 2005) and depletion in POT1 after siRNA transfection results in a DNA damage response at telomeres, a reduction in the single strand and telomere fusions (Hockemeyer et al., 2005; Veldman et al., 2004; Yang et al., 2005).

Telomeres exert an effect on the different DNA transactions such as replication (Figure 2), recombination and repair. Short telomeres can replicate earlier than long telomeres (Bianchi and Shore, 2007; Gilson and Geli, 2007) and in yeast, the Sir-mediated silent chromatin emanating from telomeres might block replication initiation in the subtelomeric regions. Semi conservative DNA replication represents a problem for the full replication of linear DNA molecules and results in the loss of terminal DNA (Gilson and Geli, 2007) (Figure 2). In organisms expressing the telomerase, this loss is rapidly compensated. However, the expression of telomerase is not universal and several processing steps limit the erosion of chromosome ends in human cells. Indeed, in the absence of any counteracting mechanism, cell division will result in a loss of telomeric sequence and loss of capping eventually leading to cellular senescence or apoptosis. The replication timing differs between different species. In human cells, telomeres replicate throughout the S phase and homologous ends seem to be highly coordinated while p and q arms of a single telomere show different timing (Zou et al., 2004). The G-tail results from the processing of the 5' strand rather than from the elongation of the 3' strand. For the leading strand, the existence of a 5' resection activity converts it into a 3' overhang. Another telomere specific problem occurring on the lagging strand is the removal of the last RNA primer. During normal replication, this removal is compensated by the downstream of Okasaki fragment. In the absence of a downstream Okasaki fragment at chromosome ends, a telomere-specific event is required for the removal of the last RNA primer. In humans, the completion of telomere replication is an ATM-dependent damage signal required for the generation of the G-tail. Finally in humans and ciliates, a last step involved in the determination of the last nucleotide (either G or C) is required. The POT1 protein might modulate this activity of a nuclease responsible for this processing since POT1 deficient cells harbor random ends (Hockemeyer et al., 2005).
Genome-wide telomere length exhibits considerable variations in the human population. Several genetical and environmental factors might be involved and a few loci containing genetic determinants of telomere length have been identified by linkage analysis (Andrew et al., 2006; Baird, 2008; Gilson and Londono-Vallejo, 2007; Graakjaer et al., 2004; Vasa-Nicotera et al., 2005). Moreover, this mechanism of allele specific telomere length might be influenced by the nature or the epigenetic status of subtelomeric regions but no telomere sequences adjacent to a telomere that can modulate its length have been identified so far.
Subtelomeric domains are cold-spots for meiotic recombination in a variety of organisms (Miklos and Nankivell, 1976). However, meiotic recombination can occur at an elevated rate near some human telomeres and can have both advantageous and pathological consequences in human biology (Kipling et al., 1996). In addition, chromosome instability through the loss of telomeres has been observed in a wide range of pathologies including cancer. Loss of telomere function leading to chromosome fusion occurs through different mechanisms. Fusion can result from the loss of capping in cells deficient for some telomeric proteins or proteins involved in telomere maintenance. Chromosome instability can occur through breakage/fusion/bridge cycles (B/F/B), which are generated when a chromosome without a telomere replicates and the sister chromatids fuse at their ends. The fused chromatids form a bridge during anaphase that breaks when the two centromeres are pulled in opposite directions. Since none of these two chromatids has a telomere, the same cycle occurs at the next division and can continue for numerous generations leading to increasing rearrangements until the broken chromosome acquire a new telomere and becomes more stable (Murnane, 2006).

Figure 2. The "end of replication problem" and consequences on cell proliferation.
Schematic representation of the replication of the telomeric lagging and leading strand. During the replication of telomeres, the gap left by the degradation of the last Okasaki fragment or the stalling of the fork cannot be overcome by the replication complex that progresses for 5' to 3' leaving a single strand. The action of nucleases and the C strand degradation, regulated in part by the telomere binding proteins increase the size of this 3' overhang.

I-C. Epigenetic regulation of chromosome ends

I-C-1. Nucleosomal structures
Unlike yeast telomeres, mammalian ones contain nucleosomes. Indeed, DNA in higher eukaryotes is packaged in nucleosomes characterized by an unusual repeat length about 20-40 bp shorter than bulk nucleosome spacing. In vitro reconstitution of nucleosome formation revealed that the energy required for the wrapping of telomeric DNA and the formation of nucleosome is the highest among other sequences, suggesting that the repetitive and G-rich nature of telomeric sequence together with the periodicity limits the folding of these regions (Cacchione et al., 1997; Filesi et al., 2000; Rossetti et al., 1998; Widom, 2001). Furthermore, the reduced H1 content at telomeres is in agreement with the short nucleosome spacing (Parseghian et al., 2001; Woodcock et al., 2006) and telomeric sequences have only minor interactions with the histone tails (Cacchione et al., 2003).
Human TRF1 is able to recognize its binding site on the nucleosomal surface (Galati et al., 2006; Rossetti et al., 2001). TRF2 overexpression is associated with aberrant nucleosomal organization of distal telomeric sequences and impinges on nucleosomal density at telomeric regions (Benetti et al., 2008).

I-C-2. Telomere maintenance and chromatin
Much of what is currently known on the distribution and regulation of chromatin marks at telomeres and their influence on subtelomeres come from recent studies in the laboratory mouse. By chromatin immunoprecipitation experiments, it has been shown that mouse telomeres are enriched in dimethylated and trimethylated Lysine 9 on the amino terminal tail of histone H3 (H3K9) (Garcia-Cao et al., 2002; Garcia-Cao et al., 2004) and trimethylated Lysine 20 on the amino terminal tail of histone H4 (H4K20) (Gonzalo et al., 2005).
Furthermore, mammalian telomeres are enriched in all three HP1 paralogs (Koering et al., 2002; Sharma et al., 2003), which interacts with the component of the telosome, TIN2 (Kaminker et al., 2005). When HP1 isoforms are overexpressed in human cells, perturbed telomere structures lead to an increase in end-to-end fusions, an increased sensitivity to ionizing radiations, the suppression of tumorogenicity in a xenograft model and the association of hTERT to telomeres (Sharma et al., 2003). Also, loss of histone H3 methyltransferase leads to a reduction in the level of HP1 at telomeres (Garcia-Cao et al., 2004).
Recently, the existence of a histone deacetylase that specifically targets telomeric sequences has been reported. This protein, SIRT6, is a Class III histone deacetylase that belongs to the SIRT1-7 family and targets H3K9 residues (Michishita et al., 2008). SIRT6 modification of telomeric chromatin is required for the association of the WRN helicase that is essential to telomere replication by preventing loss of telomeric DNA at the lagging strand.
In mice invalidated for the telomerase (terc-/-), telomeres become shorter and the heterochromatin marks are replaced with marks characteristics of euchromatin (Benetti et al., 2007a). Moreover, the lack of the H3K9 and H4K20 methyltransferases (Suv39h1, h2 and Suv4-20h respectively) results in loss of heterochromatin marks at telomeres and aberrant telomere elongation suggesting that telomere maintenance is controlled, at least in part, by chromatin structure (Garcia-Cao et al., 2002; Garcia-Cao et al., 2004; Gonzalo et al., 2005; Gonzalo et al., 2006). Mouse embryonic fibroblasts deficient in all three members of the retinoblastoma family (Rb1, Rbl1, Rbl2) have abnormally long and heterogeneous telomeres with an increase in acetylted H3 and H4 histones, similar to what is seen in mice invalidated for the Suv39h1, h2 and Suv4-20h histone mehyltransferases (Gonzalo et al., 2005). Interstingly, the level of trimethylated H3K9 and HP1 binding are not affected. However, the level of trimethylated H4K20 and the level of the Suv4-20h histone mehyltransferase, which interacts with the Rb proteins are decreased suggesting that the assembly of chromatin at telomeres is Rb-dependent since H4K20 methylation precedes H3K9 modification (Gonzalo et al., 2005). Despite the absence of CpG dinucleotides within the telomeric track, telomeres are also sensitive to the level of DNA methylation of the subtelomeric regions. A lack of DNA methyltransferase (DNMTs) in mouse embryonic stem cells alters the DNA methylation status of subtelomeric regions, leading to telomere elongation and increased telomeric sister chromatid exchanges (Gonzalo et al., 2006). Consistent with this hypothesis, loss of heterochromatin marks at telomeres increases recombination at these regions and activation of telomerase-independent telomere elongation mechanism (ALT) (Benetti et al., 2007a; Benetti et al., 2007b; Garcia-Cao et al., 2004) suggesting a narrow link between subtelomere chromatin and telomere regulation.

I-C-3. Telomeric Position Effect (TPE)
TPE has been extensively studied in baker's yeast, although it was revealed in this model organism five years after its discovery in Drosophila melanogaster (Gehring et al., 1984; Hazelrigg et al., 1984; Levis et al., 1985). Unlike Drosophila, telomeres in Saccharomyces cerevisiae are constituted of stretches of highly repetitive telomerase-added repeats and thus resemble most of the eukaryotic telomeres therefore constituting a powerful genetic system for the study of TPE. TPE in yeast was first demonstrated by insertion of a construct containing a URA3 at the subtelomeric ADH4 locus, 1.1 kb from the VII-L telomere. Expression of the URA3 gene allows growth of the cells on plates lacking uracil. However, on plates containing a drug toxic for cells expressing URA3 (5-fluoro-orotic acid or 5-FOA), 20 to 60% of the cells were still able to grow, suggesting that the URA3 was silenced in the vicinity of the telomere (Gottschling et al., 1990). Some of the features of TPE were concomitantly described, such as the stochastic reversibility, promoter independence and expression variegation and the inward Sir-dependent heterochromatin spreading (Ottaviani et al., 2008). Increasing the length of telomeres improves TPE while telomere shortening limits the silencing of subtelomeric genes (Eugster et al., 2006; Kyrion et al., 1993; Renauld et al., 1993) but the influence of telomere length on TPE is not merely due to the length of the telomere itself but rather to the changes in the recruitment of silencing factors. Thus, the interplay between TPE and length regulation might be less direct (Ottaviani et al., 2008).

The heterochromatin nature of mammalian telomeres and their capacity to induce position effect have been controversial for many years. The first example of telomeric position effect in vivo came from the analysis of replication timing of human chromosome 22 carrying chromosomal abnormality frequently observed in pathologies such as cancer or genetic diseases. Various processes that result in the addition of a new telomere can stabilize these broken chromosome ends. One of these pathways is the process in which the broken chromosome acquires a telomeric sequence from another chromosome, homolog or sister chromatid called "telomere capture". An alternative is de novo telomere addition where the end of a broken chromosome is stabilized by telomerase-dependent addition of telomeric repeats named "telomeric healing". Telomere healing following the deletions of subtelomeric elements delays the replication timing of chromosome 22 (Ofir et al., 1999). This delayed replication is not associated with differences in DNA methylation status, condensation of the chromatin structure of the region or silencing of some subtelomeric genes located 50 kb from the telomere suggesting that the large distance between the telomere and the genes may protect from the spreading of telomeric silencing (Ofir et al., 1999). However, other studies implied that human telomeres neither modulate the expression of nearby genes nor affect the homeostasis of telomeres (Bayne et al., 1994; Sprung et al., 1996).

Compelling evidence for transcriptional silencing in the vicinity of human telomeres was provided experimentally by using transgenes inserted adjacent to telomeres, similar to the approach used with yeast after telomere fragmentation (Baur et al., 2001; Koering et al., 2002) and was fueled by additional observations in cell culture and clinical samples (Ottaviani et al., 2008). Reporter genes in the vicinity of telomeric repeats were found to be expressed on average ten-fold lower than reporter at non-telomeric sites. Overexpression of the human telomerase reverse transcriptase (hTERT) in the telomeric clones resulted in telomere extension and decrease in transgene expression (Baur et al., 2001) while overexpression of TRF1, involved in telomere length regulation, lead to the re-expression of the transgene (Koering et al., 2002) indicating the involvement of both the telomere length and architecture in TPE as observed in yeast. In addition, the treatment of cells with Trichostatin A, an inhibitor of class I and II histone deacetylases antagonizes TPE. In human cells, TPE is not sensitive to DNA methylation (Koering et al., 2002) while hypermethylation of the transgene appears as a secondary effect in TPE in mouse ES cells (Pedram et al., 2006).
In human cells, there is a correlation between HP1 delocalization and TPE alleviation by TSA treatment (Koering et al., 2002). By comparison to position effect variegation, TPE might thus be an alternative and specialized silencing process acting for instance through the interaction between the chromatin remodeling factor SALL1 and TRF1 (Netzer et al., 2001) or the telomeric shelterin component TIN2 and HP1 (Kaminker et al., 2005). Thus, in mammals, like in other simpler eukaryotic organisms, classical heterochromatin factors cooperate with telomere-associated proteins in the remodeling of the telomeric and subtelomeric regions and the propagation of the silencing at chromosome ends (Blasco, 2007b).

I-D. TERRA
Because they are constituted of highly repeated regions, are enriched with different heterochromatic marks and exert a general repressing effect on the expression of neighbouring genes, telomeres have always been considered as transcriptionally silent. Recently, Northern-Blots with telomeric probes revealed the existence of telomeric transcripts called TERRA or TelRNAs consisting of UUAGGG repeats implying that they are transcribed from the C-rich strand while antisense transcripts consisting of CCCUAA repeats are present at low to undetectable levels (Azzalin et al., 2007; Schoeftner and Blasco, 2008). TERRA has been identified in human, mouse, hamster and zebrafish, and also recently, in yeast (Luke et al., 2008) suggesting that telomere transcription is a common feature in eukaryotes. In mammals, TERRA transcripts are present in all adult tissues and cell lines but absent in the mouse embryos. These transcripts are heterogenous in size, ranging from 100 bp up to more than 9 kb. Their transcription starts in subtelomeric regions (Azzalin et al., 2007), is processed by the RNA Polymerase II and transcripts are polyadenylated (Schoeftner and Blasco, 2008). The great majority of transcripts are found in nuclear fractions of the cells where they colocalize with various telomeric components and are also found at the telomere tips in metaphases.
TERRA levels are increased in cells deficient in the Suv39h1, h2 and Suv4-20h histone methyltransferases and decreased in cells lacking the Dnmt1 DNA methyltransferase or Dicer activity suggesting that heterochromatin represses TERRA formation. At this step, the involvement of DNA methylation in TERRA synthesis is not clear (Figure 3). Indeed, the Dnmt1 effect observed by Schoeffner et al. in the laboratory mouse, differs from the recent investigation of human telomeres in cells from ICF (Immunodeficiency, centromeric region instability, facial anomalies) patients (Ng et al., 2009). This rare disorder is linked to a mutation in the de novo DNA methyltransferase DNMT3B gene and correlates with a hypomethylation of several repeats and subtelomeric regions (Kondo et al., 2000). Hypomethylation of subtelomeres was associated with an elevated level of TERRA transcript suggesting that the hypomethylation of subtelomeres and potentially TERRA promoter regions might facilitate the transcription of these repeats. Interestingly, telomerase-positive cells harbor elevated DNA methylation level at the proximal subtelomere and a reduction in telomeric transcription and TERRA production suggesting that telomere elongation negatively regulates TERRA production possibly through the methylation of the subtelomeres (Ng et al., 2009). On the opposite, telomerase-negative cancer cells maintaining their telomere via the ALT pathway, show heterogeneous methylation pattern.

Although many questions remain on the role of this transcript in telomere regulation, evidences from mammals and yeast revealed a key role in the regulation of the telomerase (Azzalin et al., 2007; Luke et al., 2008; Schoeftner and Blasco, 2008). Previous results from human cells suggested that longer telomeres produce more TERRA, which may then feed back and negatively regulate telomerase access (Azzalin et al., 2007; Schoeftner and Blasco, 2008), while in yeast, the accumulation of TERRA may inhibit telomerase action through the formation of DNA/RNA hybrids hindering thereby telomerase access (Horard and Gilson, 2008; Luke et al., 2008). This new partner in telomere biology opens a wide field of investigation for understanding the homeostasis of chromosome ends in normal and pathological human samples.

Figure 3. Distribution of the epigenetic marks in subtelomeric and telomeric regions controls the transcription of telomeres.
Telomeric DNA is packaged into nucleosomes and telomeric chromatin contains marks of heterochromatin such as HP1, trimethylated H3K9 and trimethylated H4K20. At chromosome ends, the shortening of telomeres leads to changes in chromatin condensation. In the mouse, telomere shortening is associated with loss of heterochromatin marks at telomeres and subtelomeres and increase in sister chromatid exchange. Telomeres are transcribed into non-coding transcript named TERRA or TelRNA whose production is initiated in the subtelomeric region. TERRA molecules are produced by the RNA polymerase II from the G-strand and TERRA transcripts are polyadenylated. Several factors regulate TERRA production and stability such as the mRNA decay pathways (NMD) and epigenetic mechanisms such as DNA methylation and methylation of histone tails regulated the Suv39h and Suv4-20h histone methyltransferase. Concerning DNA methylation, contradictory results have been obtained in mice or human cells. In mice, hypomethylation seems to repress TERRA while in human cells, hypermthylation blocks TERRA production possibly through hypermethylation of TERRA promoters.

I-E. Localisation of telomeres within the nuclear space
The localization of sequences within the nuclear space is of paramount importance for proper genome functions. The localization of chromosome regions at the periphery of the nucleus enriched in heterochromatin marks seems to play a central role in gene regulation, especially silencing and DNA repair (Finlan et al., 2008; Kumaran and Spector, 2008; Reddy et al., 2008; Therizols et al., 2006). The telomeres are not randomly localized within the nucleoplasm and can be found at the nuclear periphery (Gilson et al., 1993). This positioning varies greatly among organisms, cell types, cell cycle stages and individual telomeres. At the bouquet stage for instance, the clustering of all telomeres at the edge of the nucleus, is a nearly universal feature of meiosis (Scherthan, 2007). In budding yeast, the 32 telomeres are clustered into 4-6 foci which are primarily associated with the nuclear envelope (Gotta et al., 1996). This peripheral localization of telomeres is dependent on redundant pathways (Maillet et al., 2001). One acts through Ku and the second through Sir4-Esc1 (Hediger and Gasser, 2002; Taddei and Gasser, 2004). Relocation to this peripheral nuclear compartment probably does not cause repression per se (Tham et al., 2001) and silencing can be maintained without perinuclear anchoring (Gartenberg et al., 2004). All of the data on silencing and anchroring at the nuclear periphery converge toward a reservoir model where telomere clusters act as a subnuclear compartment concentrating key heterochromatin factors like the Sir proteins (Maillet et al., 1996).
A perinuclear positioning of telomeres is also observed in Plasmodium, where it favors subtelomeric gene conversion (Freitas-Junior et al., 2000) while in plants, telomeres are observed either close to the nuclear periphery (Rawlins and Shaw, 1990) or around the nucleolus (Fransz et al., 2002). In mammalian nuclei, telomeres adopt different locations (Luderus et al., 1996). While human telomeres are clustered at the nuclear periphery in sperm (Gilson et al., 1993; Zalenskaya et al., 2000), most telomeres in lymphocyte nuclei are located in the interior of the nucleoplasm (Weierich et al., 2003). Thus, it seems that by default, human telomeres are localized internally in most cell types. However, some subtelomeric elements might antagonize this internal localization and target their proximal telomere to the nuclear envelope as suggested by the presence of LADs at different subtelomeres (Guelen et al., 2008). Such an example of localization at the periphery of the nucleus is the positioning of the 4q35 subtelomeric locus, involved in the Facio-Scapulo-Humeral Dystrophy (FSHD) (Masny et al., 2004; Tam et al., 2004).

I-F. Telomeres, senescence and organismal aging
Human primary fibroblasts maintained in culture undergo many divisions before reaching the "Hayflick limit" and arresting growth (Hayflick, 1965). Later, it was shown that this limit corresponds to a critical telomere shortening and dysfunction in the absence of telomere maintenance and accumulating evidence strengthened the view that an excessive telomere erosion and subsequent activation of the senescence pathway is a tumor-suppressor mechanism (Figure 4). In mammalian somatic cells, the presence of a minimal set of very short telomeres is sufficient to trigger replicative senescence. Senescence depends on the essential phosphoinositide (PI)-3-kinase-related protein kinase ATM and ATR involved in DNA damage checkpoint. If these checkpoints fail, cells resume division, develop genomic instability and ultimately die during crisis. A few cells escape from crisis and their telomeres are maintained by recombination (Alternative telomere lengthening, ALT, see below) or telomerase reactivation. Viewing replicative senescence as one of the protective mechanisms against tumor formation, it is plausible that senescence-associated genes play significant roles in tumorogenesis repression. At this point, further studies are needed to elucidate the respective biological function of genes differentially expressed in senescent cells and cells suffering from telomere dysfunction.
A common feature among eukaryotes is the progressive decline in vitality over the time. Aging encompasses a wide spectrum of degenerative processes such as the accumulation of carbonylated proteins, damaged enzymes, protein misfolding, lipid peroxydation or activation of inflammatory response pathways. Increasing evidence supports the hypothesis of a link between modification of telomeres and senescence or aging. In S. cerevisiae, a key regulator in the aging process is the Sir2 NAD+ Histone Deacetylase (HDAC) that prevents these regions from recombination and fusion throughout cell division (Kaeberlein et al., 1999; Sinclair and Guarente, 1997).

Telomere length correlates with longevity and disease resistance in Human (Cawthon et al., 2003) whereas loss of telomerase causes accelerated aging in mice and several human syndromes (Blasco, 2007a; Chang et al., 2004). For instance, mutations reducing telomerase activity give rise to premature-ageing syndromes, which may be caused by telomere exhaustion and a reduced replicative potential of stem cells (see below). In addition, telomere shortening is observed in the Hutchinson-Gilford progeroid syndrome (HGPS) linked to a mutation in the gene encoding A-type lamins (De Sandre-Giovannoli et al., 2003; Eriksson et al., 2003) indicating a close relation between telomere erosion and aging phenotypes (Decker et al., 2009; Ding and Shen, 2008).
Deficiency in the mammalian SIRT6 protein, a member of the Sirtuin family (Dali-Youcef et al., 2007; Haigis and Guarente, 2006) increases chromosomal aberrations such as fragmented chromosomes, detached centromeres and gaps and leads to the development of a degenerative aging-like phenotype (Mostoslavsky et al., 2006). SIRT6 is H3K9 deacetylase specific for telomeres that is critical for maintaining functional telomeres, allowing WRN interaction and preventing end-to-end fusion (Michishita et al., 2008).

Also several environmental factors that affect telomere length such as stress, smoking, obesity and socio-economic status might also accelerate aging (Canela et al,. 2007; Cherkas et al., 2006; Epel et al., 2004; Valdes et al., 2005).

Figure 4. Telomere attrition controls the proliferative capacity of somatic cells.
Telomere length is central to the process of aging and tumorigenesis in human cells. As result of the "end of replication problem", telomeres shorten after each cell division. When telomeres reach a critical length, the recruitment of telomere binding proteins and telomere capping are insufficient. Uncapped chromosome end resembles a double stranded DNA break that is highly unstable and can give rise to chromosome rearrangements. Since telomere erosion limits the proliferative capacity of cells, premalignant transformed clones cease to expend when their telomeres shorten. Critically shortened telomeres elicit a potent DNA damage response and accumulate several DNA damage markers such as phosphorylated gH2AX, 53 BP1, CHK2 for instance (d'Adda di Fagagna et al., 2003; Takai et al., 2003). Many of these proteins localize directly to dysfunctional telomeres to form dysfunctional telomere-induced foci (TIFs). However, entry into senescence or apoptosis can be bypassed by dysfunctional cell cycle controls and lead to the accumulation of chromosomal rearrangement and a global instability of the genome. Telomeres are tightly linked to tumorigenesis and the reactivation of telomerase circumvents telomere shortening and enhances the proliferative capacity of a subset of cells.

I-G. Telomeres and pathologies

I-G-1. Telomeres and cancer

I-G-1a. Reactivation of telomerase
Telomere dysfunction has a dual role in neoplasia either by initiating or suppressing tumorogenesis. As mentioned earlier, telomeres shorten after each division until they reach a critical length incompatible with normal capping (Figure 4). The induction of a DNA damage pathway is triggered, followed by growth arrest. In the presence of a functional p53 pathway, telomere shortening and inherent genomic instability promote cellular senescence, a potent tumor suppressor mechanism. However, in a few cells with disabled checkpoints, short dysfunctional telomeres escape this control step and accumulate rearranged DNA perpetuated through recurrent breakage-fusion-bridge cycles. Generalized chromosomal instability leads to cell death, however a few cells may escape and be maintained after reactivation of a telomere maintenance mechanism.
Human cancers and some human somatic cells are able to maintain or extend their telomere length. Indeed, in approximately 90-95% of human cancers and 60-70 % of immortalized cell lines, telomerase is upregulated. Futhermore, hTERT together with the large T and H-ras oncogenes result in the tumorigenic conversion of normal epithelial and fibroblast cells (Hahn et al., 1999). Chromosomal aberrations resulting from telomere-driven chromosomal instability contributes to the acquisition of the tumoral phenotype and early cancerous lesions seem to accumulate more breaks toward chromosome ends whereas more advanced lesions accumulate breaks along chromosomes (Gisselsson et al., 2000). At later stage, further telomere shortening and interstitial breaks amplify the global instability and heterogeneity observed in tumoral samples. In cellular models, the absence of TRF2 contributes to tumor development in cells expressing the telomerase and an ER-SV40 antigen suggesting that telomeric proteins also contribute to cancer (Brunori et al., 2006). However, in another cellular model, the dominant negative form of TRF2 introduced in melanoma cell lines is associated with a loss of tumorogenicity in at least one cell line (Biroccio et al., 2006). In human tumors, altered levels of telomeric proteins have been observed and might also been associated with tumor phenotype (Bellon et al., 2006; Lin et al., 2006; Ning et al., 2006; Oh et al., 2005; Poncet et al., 2008; Yamada et al., 2002; Yamada et al., 2000).
The targeting of telomere may thus be a promising target in the cure of a wide range of cancer types. However, only a few potent small molecules have been discovered so far and the main targets of these bioactive compounds are either the components of the telomerase or the ligands of G-quadruplexes formed at chromosome ends (Gomez et al., 2006a; Gomez et al., 2006b; Salvati et al., 2007; Tahara et al., 2006).

I-G-1b. The ALT pathway
Approximately 5% of human cancer cells maintain their telomere by ALT, Alternative Lengthening of Telomeres. ALT cells are characterized by the presence of long and heterogeneous telomeres. ALT appears more common in tumors derived from tissues of mesenchymal and neuroepithelial origin. The onset of the ALT pathway might be associated with loss of p53 or p16INK4A or activation of the Cyclin D1. The ALT mechanism likely involves recombination-mediated DNA replication, however, the exact mechanism is still unclear. A mechanism of replication/recombination using a circular telomeric DNA generated by homologous recombination might be implicated, allowing the extension of a telomere end using a sister telomere as a template (Dunham et al., 2000; Yeager et al., 1999). Cellular characteristics of the ALT pathway include the heterogeneous and rapid change in individual telomere length and the presence of intra-nuclear aggregates defined as ALT-asociated promyelocytic leukemia (PML) protein nuclear bodies (APBs) (Henson et al., 2002; Yeager et al., 1999). These APBs contain extra-chromosomal telomeric DNA, TRF1 and TRF2 and several proteins involved in DNA recombination and replication. Decreased DNA methylation of the subtelomeric region in DNA methyltransferase deficient mice is associated with features of ALT including long and heterogeneously sized telomeres, increased telomeric recombination and the presence of APBs (Gonzalo et al., 2006). However, in a subset of cells that resembles typical ALT cells, the absence of APBs in telomerase-negative immortalized cells has been reported (Marciniak et al., 2005; Wu et al., 2000) suggesting the existence of alternative pathways for ALT but the mechanisms underlying either the onset of ALT or telomerase reactivation remain elusive.

I-G-2. Telomere dysfunction in genetic diseases
Recent experimental evidences revealed the existence of different genetic diseases linked to mutation in the telomerase or telomerase associated proteins (Garcia et al., 2007; Kirwan and Dokal, 2009).
Dyskeratosis congenitae (DKC) or Zinsser-Engman-Cole syndrome is a rare inherited pathology with an incidence of 1/1 000 000 individuals. The pathology is characterized by the mucocutaneous triad of abnormal skin pigmentation, nail dystrophy and mucosal leukoplakia. Additional symptoms such as dental, gastrointestinal, pulmonary, immunological or neurological abnormalities have also been reported. Usually, patients appear healthy at birth and develop the mucocutaneous features later in life (around the age of 10) with death occurring around the median age of 16. For some patients, malignancy usually occurs in the third decade and includes carcinomas, leukemias and lymphomas.
Three genetic forms have been described for the disease. The first form is recessive, linked to the X chromosome and mainly occurs in male (MIM 305000) but autosomal dominant (MIM 127550) and autosomal recessive forms (MIM 224230) have also been identified.
In these genetic diseases, mutations in the gene encoding components of the telomerase result in telomerase deficiency, telomere shortening, end-to-end fusions, increased chromosomal instability and the subsequent development of multisymptomatic syndromes.

The X-linked form of DKC was identified by linkage analysis and positional cloning and involves mutations spread across the DKC1 gene (encoding Dyskerin) on Xq28 (Heiss et al., 1998; Knight et al., 1998). Dyskerin has been predicted to have a pseudouridylation activity and DC pathogenesis might be in part linked to defective ribosome biogenesis. Furthermore, Dyskerin and other associated proteins (GAR1, NH2, NPO10) are part of the telomerase complex and interact with the telomerase RNA component (TERC). Dyskerin mutations lead to reduction in TERC and experimental evidences converge toward abnormalities in telomere maintenance in patients suffering DC.

The autosomal dominant form of DKC is associated with mutations in the TERC gene on chromosome 3q. The pathology results in reduced telomerase activity through either impaired RNA accumulation/stability or catalytic defect of the telomerase complex. The genetic alteration underlying the autosomal recessive form of DKC is currently unknown (Mitchell et al., 1999; Vulliamy et al., 2001; Vulliamy et al., 2004). In around 11% of cases of AD-DC, mutations in the TINF2 gene have been reported (Savage et al., 2008; Walne et al., 2008). This gene encodes the TIN2 shelterin component and TIN2 mutation might lead to a more direct degradation of telomere by preventing TRF1 binding and possibly by leaving telomere ends unprotected although the mechanism has not been fully determined yet (Heiss et al., 1998; Knight et al., 1998).
Despite the heterogeneity of symptoms in patients with DKC, some features seem to overlap with the Hoyerhaal-Hreidarsson syndrome (HH, OMIM 300240), a severe multisystem disorder occurring in the neonatal period and infancy. As observed in the X-linked form of DKC, several DKC1 mutations have now been identified in patients affected with this syndrome (Marrone et al., 2007).

Like in patients with DKC, individuals with aplastic anemia and myelodysplasia have short telomeres and increased chromosomal instability compared to age-matched controls. These patients are believed to suffer from a defect at the level of stem cells, especially affecting the renewal and proliferative capacity of haematopoietic and epithelial tissues (Vulliamy et al., 2002; Yamaguchi et al., 2005). Idiopathic pulmonary fibrosis (IPF) is a progressive fatal lung disease characterized by lung scarring and abnormal gas exchange that usually arise around the age of 50. However, a familial form of IPF with early onset has been described in occidental countries. In a significant percentage of cases, this pathology has been associated with mutations in TERC or HTERT (Tsakiri et al., 2007), which are different from those usually observed in patients with DC.

In general, the mutations of the telomerase complex or proteins involved in telomere maintenance might affect the recruitment of telomerase, the activity of the enzyme and in turn lead to a dysfunctional telomere maintenance, especially in stem cells whose renewal potential become limited.

II. Subtelomeres

II-A. General information
Subtelomeres are DNA sequences placed between chromosome-specific regions and chromosome ends with features that distinguish them from the rest of the genome (Mefford and Trask, 2002; Riethman, 2008; Riethman et al., 2001). Human subtelomeres vary in size from 10 to up to 500 kb in human cells. They contain repetitive sequences of different types and numerous genes but very little are known on their function in the regulation of cellular homeostasis (Mefford and Trask, 2002; Riethman, 2008; Riethman et al., 2001). However, these regions, prone to recombination and rearrangements, are associated with genome evolution, human disorders but also aging possibly through TPE (Ottaviani et al., 2008). In the human population the subtelomeric regions are highly polymorphic and the rate of recombination at chromosome ends is higher than in the rest of the genome. Such rearrangements participate in the genome variability and the length of variation may be up to hundreds of kilobases among different haplotypes (Figure 5). Although the coverage of chromosome ends has not been fully achieved, available sequences allowed the representation of a detailed paralogy map showing that several blocks of sequences are shared by different human subtelomeres (Flory et al., 2004; Linardopoulou et al., 2005; Riethman et al., 2001). Various tandemly repeated units called Telomere-Associated Repeats (TAR1), short native telomeric arrays and numerous degenerate telomere-like repeats are also located at variable distance from the telomere and subtelomeres contain members of 25 small families of genes encoding potentially functional proteins (Flory et al., 2004; Linardopoulou et al., 2005; Riethman et al., 2001). Interestingly, many of them are involved in the adaptation to the environmental changes like in other species suggesting that the plasticity of chromosome ends is likely to play a key role in genome evolution and that abnormalities or dysregulation of these genes may have phenotypical consequences (see below for examples).

Figure 5. Representation of the human telomeric and subtelomeric regions.
In eukaryotes, the subtelomeres are patchworks of genes (pink rectangles) interspersed within repeated elements (blue rectangles) and degenerated telomeric repeats (ITS, internal Telomeric Sequences). In human, large polymorphic blocks of repeated sequences are distributed between the different chromosomes and subtelomeres contain genes.

II-B. Subtelomeric sequences

II-B-1. Families of genes
Human subtelomeric genes vary in copy number and chromosomal distribution. Large families of genes are present at subtelomeres such as odorant and cytokine receptors, homeodomain proteins, secretoglobins together with several genes of unknown function.

II-B-1a. The WASH genes
The most terminally located human subtelomeric genes were recently identified and encode a third class of the Wiskott-Aldrich Syndrome protein (WASP) family (Linardopoulou et al., 2007). Five WASP family members are known in mammals and are involved in cell motility, phagocytosis, cytokinesis and in processes such as angiogenesis, embryogenesis, inflammatory immune response, microbial infection and cancer metastasis (Millard et al., 2004; Takenawa and Suetsugu, 2007; Yamaguchi and Condeelis, 2007). Thus, the recent characterization of the human subtelomeric MGC52000 genes allowed the identification of these new members of the WASP family, named WASH for Wiskott Aldrich Syndrome protein and Scar Homolog. Human genome harbors mutltiple functional WASH paralogs at subtelomeres with the coding sequence ending within 5 kb of the telomere. Several functional WASH variants and multiple pseudogenes have been identified per genome and subtelomeric dynamic might contribute to the variation and diversification of the WASH family. The WASH genes are found in numerous species suggesting conservation during evolution with extensive duplication and dispersal to chromosome ends in primates.
Their colocalization with actin in vivo suggests a role in actin polymerization and cytoskeleton reorganization. The drosophila WASH protein was recently identified as a component of a nuclear complex containing various transcriptional factors and chromatin modifiers and is essential in development. Interestingly, the WASP gene defective in the Wiskott-Aldrich syndrome causes eczema, thrombocytopenia and immunodeficiency (Ochs and Thrasher, 2006) and by analogy, loss of the WASH genes in human through subtelomeric rearrangements for instance might have consequences in pathologies.

II-B-1b. The α and β Defensin loci
The human genome is rich in genomic regions repeated several times. These regions named CNVs (Copy Number Variations) are highly polymorphic but also involved in predisposition to diseases. Among these CNVs, the subtelomeric α and β Defensin loci located on chromosome 20 and at the 8p23.1 subtelomere and containing at least 8 Defensin genes, are associated with different pathologies mainly characterized by an altered immune response. Defensin are small cationic secreted peptides (3-5 kDa) with antimicrobial activity against gram-positive and gram-negative bacteria as well as fungi and enveloped viruses acting by disrupting membrane integrity and function (Ganz, 2003). Some Defensin are produced constitutively while some are synthetized in response to microbial products or pro-inflammatory cytokines amplifying subsequent innate and adaptative immune response (Ganz, 2003; Lehrer and Ganz, 2002). α Defensin are mainly expressed in neutrophils and paneth cells of the intestine while β Defensin are expressed by epithelial tissues.
This locus is the fastest changing CNV currently known (Abu Bakar et al., 2009) and probably the most clinically relevant CNV involved in lupus for the α Defensin (Bennett et al., 2003; Ishii et al., 2005), the Crohn's disease (Fellerman et al., 2006), psoriasis (Hollox et al., 2008), but also potentially to other inflammatory disorders for the β Defensin genes.

II-B-1c. Olfactory Receptors
Olfactory receptors (ORs) comprise one of the largest gene families in the genome of mammals with over 400 olfactory receptor genes and pseudogenes in human (Glusman et al., 2001; Niimura and Nei, 2007). Each neuron in the olfactory epithelium expresses a single allele of a single OR gene and axon neurons expressing the same gene converge in the olfactory bulb of the brain. A single odorant can be recognized by a number of different receptor types and the differential affinity of these different receptors together with the combinatorial use of other OR allow the detection of millions of chemicals by the olfactory system. Human ORs are often clustered and might represent > 0.1% of the human genome. The numerous OR clusters arose through tandem duplications and several ORs clusters are localized in subtelomeric or centromeric regions.

Over 60% of human OR genes bear one or several sequence disruptions likely resulting in the inactivation of the corresponding protein. Interestingly, the human OR repertoire is interspersed with numerous repeated elements suggesting a high capacity to recombine. In some cases, recurrent 8p rearrangements might occur as a consequence of an inversion polymorphism mediated by two ORs genes between the subtelomeric 8p23.1 and 4p16 loci (Wieczorek et al., 2000a; Wieczorek et al., 2000b). In addition, unequal crossovers between two OR genes in 8p clusters are responsible for the formation of three recurrent chromosome rearrangements (inv dup(8p), +der(8p) and inv(8p)) associated with distinct phenotypes (Ciccone et al., 2006; Giglio et al., 2001). Furthermore, a much faster functional deterioration of the large OR gene superfamily occurred in the human lineage compared to the great apes and old world monkeys.
It is thus tempting to speculate that a lesser need for the sense of smell in humans compared to other species involved a rapid evolution of the OR gene repertoire, possibly through subtelomeric rearrangements (Gilad et al., 2003).

II-C. Regulation of subtelomeric sequences

II-C-1. Does telomere shortening modulate expression of subtelomeric genes?
In the human population, subtelomeric regions are highly polymorphic and length variation can be up to hundreds of kilobases among the different haplotypes. As described above, telomere shortening affects the epigenetic regulation of subtelomeres and might also impact on their recombination rate. Thus, transcriptional regulation of natural subtelomeric genes in human cells likely depends on telomere length, the structure of the telomeric chromatin but also on the composition of the subtelomeric regions and the spatial organization of chromosome ends.
The effect of telomere shortening on the expression of subtelomeric genes was recently investigated during senescence in human fibroblasts maintained in culture for an extended period of time (Ning et al., 2003). A total of 34 subtelomeric genes and the length of the corresponding telomeres were analyzed in young and senescent cells. Despite a differential expression for 17 out of these 34 genes, telomere length alone is not sufficient to determine the expression status of telomeric genes (Ning et al., 2003) suggesting a complex interplay between telomere and subtelomere regulation.
Age-dependent telomere erosion might also be a key player in the regulation of subtelomeric genes in elders as it was observed experimentally in artificial systems (Baur et al., 2001; Koering et al., 2002). In mammals, aging is associated with a multitude of gene expression changes and increasing evidence supports the hypothesis of a link between senescence or aging and modification of chromatin since the architecture of the telomeric and subtelomeric regions is also remodelled during these two processes and a number of factors that can influence directly or indirectly telomere structure may alter the expression of subtelomeric genes by changing telomere conformation and maintenance and vice versa although clear demonstration in higher eukaryotes are still lacking.

II-C-2. Telomeres and subtelomeres: adaptation to environment?
Different mechanisms could be designed to accommodate the evolution of environmental conditions throughout life. One way might be the rapid regulation of subtelomeric genes by Telomeric Position Effect. Indeed, a subtelomeric enrichment of genes related to stress response and metabolism in non-optimal growth conditions appears to be a conserved feature in many yeast species (Robyr et al., 2002) and clustering stress response genes at subtelomeres might be an evolutionary conserved strategy allowing their reversible silencing and a fast response to changes in environmental conditions (Barry et al., 2003; Borst and Ulbert, 2001; Dreesen et al., 2007; Ottaviani et al., 2008). Aging is characterized by an increasing susceptibility to environmental stress and a wide range of diseases. Interestingly, older subjects are more susceptible than younger ones to pathogenic stimuli (Krabbe et al., 2004) and the IgH genes cluster localized at subtelomeres is up-regulated in aging hematopoietic stem cells. Also, in human hepatic stellate cells undergoing senescence, increased expression of genes mediating inflammatory response has been observed (Schnabl et al., 2003). Interestingly, numerous genes encoding cytokines are located at subtelomeric loci and might be also influenced by telomere length.
Impressively, sensory perception may also affect life span in higher animals (Libert et al., 2007; Lindemann, 2001) and some gustatory and olfactory neurons either promote or inhibit longevity (Alcedo and Kenyon, 2004). Olfactory genes are preferentially positioned at subtelomeric positions, it would be interesting to correlate changes in olfactory stimuli in aged individuals and control of expression of subtelomeric clusters of genes by telomeres.

II-D. Subtelomeres and pathologies

II-D-1. General information
As described above, subtelomeres are highly polymorphic and a broad range of expression level of natural subtelomeric genes is likely to be found from individual to individual as described in yeast (Pryde and Louis, 1999). Thus, telomere length-mediated transcriptional regulation of natural subtelomeric genes in human cells is likely to operate through the telomeric heterochromatin structure, involving long and variable stretches of subtelomeric sequences and renders analysis of telomeres and subtelomeres challenging in human pathologies. The only naturally occurring situations wherein telomeric repeats are adjacent to unique sequences are those that occur in patients with truncated chromosomes ends that have been repaired by the process of telomere healing or that lead to the formation of ring chromosomes. However, the molecular pathogeneses associated with these rearrangements have never been investigated.
Moreover, TPE may play a direct role in human diseases as a result of repositioning of active genes near telomeres or subtelomeric sequences following such chromosome rearrangements and subtelomeric element may either participate in the spreading of silencing in the vicinity of a telomere or shelter genes from this silencing.

II-D-2. Idiopathic mental retardation
Mental retardation (MR) affects 3 % of the general population (Tyson et al., 2004) and approximately 15 % of cases are explained by chromosome aberrations (Stankiewicz and Beaudet, 2007). Since subtelomeric regions are among the most gene rich regions of the genome and are particularly prone to recombination, it was logical to think that subtelomeric imbalances would account for MR. However, subtelomeric regions are difficult to explore with standard cytogenetic techniques and new strategies had to be developed in order to confirm their role in MR. In 1995, Flint et al. found 6% of cryptic subtelomeric imbalances in MR patients using polymorphic microsatellite markers (Flint et al., 1995). Since then, faster techniques have been developed such as fluorescent in situ hybridization (FISH) (Knight and Flint, 2000), Multiplex Ligation dependent Probe Amplification (Rooms et al., 2004) or array comparative genomic hybridization (CGH) (Ballif et al., 2007a; Ballif et al., 2007b; Veltman et al., 2002). They allowed the rapid screening of children with unexplained MR. The largest study on 12,000 MR patients investigated by FISH demonstrated that 2.5 % of the clinical cases with severe to mild mental retardation display relatively small subtelomeric abnormalities of all the chromosomes arms with exception of the p-arm of the acrocentric chromosomes. Subtelomeric imbalances include deletions, duplications, unbalanced translocations and complex rearrangements (Shao et al., 2008). They are terminal as well as interstitial. Their size is extremely variable. However, these rearrangements are quite large since 40% of them are over 5 Mb in size (Ballif et al., 2007a; Ballif et al., 2007b). Fifty percent of the subtelomeric imbalances are familial cases. In a series of 56 families, Adeyinka and colleagues demonstrated that 65% of the derivative chromosomes were inherited from a parent carrier of a balanced translocation, and that 32% of the subtelomeric deletions were inherited from a parent presented with normal clinical features or a milder phenotype than the affected children (Adeyinka et al., 2005). Among the benign subtelomeric copy number variations, it can be distinguished common telomeric polymorphisms and transmitted subtelomeric imbalances without phenotypic effect (Ledbetter and Martin, 2007). Common telomeric polymorphisms are present in at least 1 % of the population. They are well known for telomeres 2q, 4q, 7q, 9p, 10q, Xp and Yq. Diagnostic assays now avoid the detection of these clinical insignificant variations. Transmitted subtelomeric imbalances without phenotypic effect are less common (Barber, 2008). They have a wide range of sizes from 150 kb to 10 Mb. They have now been detected at 24 of the 41 telomeres (Balikova et al., 2007). Several mechanisms may explain the absence of abnormal phenotype in carriers of the subtelomeric imbalances: variable expressivity, unmasking of recessive allele, somatic mosaicism in the normal parent and epigenetic modifications.
During the recent years, the development of high resolution genetic analysis techniques allowed a better characterization of the genotype of patients affected with such developmental delays and lead to the identification of different genes disrupted by these subtle terminal deletions (Kleefstra et al., 2006; Lamb et al., 1993; Walter et al., 2004). Nevertheless, gene deletion does not always explain the pathological manifestations. Among the hundreds of patients analyzed, the size of the subtelomeric region disrupted may be accompanied by variable degrees of chromatin condensation and explain the penetrance of the clinical manifestations in patients (Walter et al., 2004). Surprisingly, rather few subtelomeric imbalances are associated with a distinct, recognizable phenotype. Genotype-phenotype correlations for these syndromes are not well established. If one or few genes could have been identified for some syndromes, molecular mechanisms are still unclear for others. It appears that gene dosage alone cannot account for the phenotype in many cases and those alternative mechanisms, in particular epigenetic modifications, may be involved. We will take three examples to illustrate this point.
Thus, it is conceivable that genes residing in close proximity to healed telomeres become epigenetically inactivated contributing to the phenotype. However, characterization of the rearrangement's effect on gene expression is still needed to prove that mental retardation is caused by modification of the chromatin architecture at telomeric and subtelomeric loci.

II-D-3. The 22qter syndrome
The 22qter deletion syndrome is characterized by severe neonatal hypotonia, global development delay, autistic-like behavior, normal to accelerated growth, absent to severely delayed speech and minor dysmorphic features (Balikova et al., 2007). The deletion is usually terminal although cases of interstitial deletions sparing the telomere have been described. Deletions range from 130 kb to 9 Mb in size (Wilson et al., 2003). However, there is little correlation between the size of the deletion and the severity of phenotype (Koolen et al., 2005). It has been demonstrated that all the patients shared a deletion of SHANK3, a gene encoding a scaffolding protein found in excitatory synapses, and that a recurrent breakpoint was in this gene (Bonaglia et al., 2006; Wilson et al., 2003). Moreover, patients with SHANK3 disruption or point mutations present the same phenotype than the 22qter deletion syndrome (Bonaglia et al., 2001; Durand et al., 2007). So, it appears that SHANK3 is the major gene responsible for at least the neuro-behavioral phenotype of 22qter deletion syndrome.

II-D-4. The cri-du-chat syndrome
Cri-du-chat syndrome is due to the terminal deletion of the short arm of chromosome 5 (5pter). The phenotype is characterized by microcephaly, facial dysmorphism, high-pitched cat-like cry, severe mental retardation and speech delay (Cerruti Mainardi, 2006). The range of deletion size is also very wide, since deletion can be visible on standard karyogram. Genotype-phenotype correlation studies delineated three critical regions corresponding to cry (5p15.31), speech delay (5p15.32-15.33) and facial dysmorphism (5p15.31-15.2) (Zhang et al., 2005) but no gene has been identified for each specific feature yet. The severity of mental retardation seems to be correlated with the deletion size although some patient presented with a disproportionately severe retardation regarding to the deletion size.

II-D-5. The 1p36 mental retardation syndrome
The 1p36 monosomy syndrome is the most frequent subtelomeric microdeletion syndrome with a frequency of 1/5,000. It is characterized by mental retardation, developmental delay, hearing impairment, seizures, growth impairment, hypotonia, heart defect and distinctive dysmorphic features (Gajecka et al., 2007). Two thirds of de novo rearrangements are apparently simple terminal truncation. The remaining third corresponds to complex structures including deletions with inverted duplication, large duplications and triplications with small deletions and interstitial deletions. Attempts have been made to demonstrate that monosomy 1p36 was a contiguous gene syndrome. Candidate genes for seizures and facial features have been proposed (Heilstedt et al., 2003a; Heilstedt et al., 2003b). However, arguments are in favor of alternative mechanisms. No correlation between the deletion size and the number of clinical features could be observed (Gajecka et al., 2007). There is neither common breakpoint nor common deletion interval in monosomy 1p36 patients. Redon et al., studied 6 patients using tiling path array CGH, two of the six patients presented with very similar features but had non-overlapping 1p36 deletions (Redon et al., 2005). Thus it was proposed that the 1p36 monosomy syndrome might be due to a positional effect rather than haploinsufficiency of contiguous genes.

II-D-6. Facio-Scapulo-Humeral Dystrophy
One of the best-characterized human genetic diseases potentially linked to TPE is the Facio-Scapulo-Humeral Dystrophy (FHSD). This puzzling pathology is associated with the deletion of repeated elements at the 4q35 locus. Normal 4q35 chromosome termini carry from 11 up to 150 copies of a 3.3 kb repeated element named D4Z4 while in FSHD patients the pathogenic allele has only 1-10 repeats. This autosomal dominant disorder is the first most common myopathy clinically described by a progressive and asymmetric weakening of the muscles of the face, scapular girdle and upper limb. The pathogenic alteration does not reside within the gene responsible for the disease but is rather related to an epigenetic mechanism. Several hypotheses have been proposed to explain this enigmatic pathology (Gabellini et al., 2004; van der Maarel and Frants, 2005). Evidence for the binding of a repressor complex to D4Z4 that might regulate the expression of the nearby genes was provided (Gabellini et al., 2002) but remains controversial (Jiang et al., 2003; Winokur et al., 2003). However, the most popular hypothesis to explain this dystrophy is the involvement of PEV or TPE (van Deutekom et al., 1996). D4Z4 shares some of the properties of heterochromatic sequences such as DNA methylation and it was postulated that D4Z4 and surrounding sequences would be packed as heterochromatin leading to the silencing of nearby genes. In patients, the partial loss of the D4Z4 repeat would lead to local chromatin relaxation and to the transcriptional upregulation of genes (Hewitt et al., 1994; Winokur et al., 1994). However, the analysis of the chromatin structure of this locus either in normal individuals or in FSHD patients does not fully support this hypothesis (Jiang et al., 2003). Alternatively D4Z4 may act as an insulator, separating heterochromatic telomeric sequences distal to D4Z4 from euchromatic sequences upstream (van Deutekom et al., 1996). We recently showed that FSHD might be associated with a gain of function of CTCF (Ottaviani et al., 2009).
Interestingly, different allelic variants might also be linked to the pathology and the 4q35 subtelomeric region appear as a mosaic of regulatory elements. The understanding of the cross talks between D4Z4, the 4q35 subtelomere and the telomere would provide insights in the deciphering of this complex epigenetic disease and the involvement of the D4Z4 subtelomeric element in transcriptional activity or replication timing of the 4q35 chromosome end.

II-D-7. Ring chromosomes
Hundreds of patients have been reported with various combinations of malformations, minor abnormalities and growth retardation usually associated with mental retardation linked to the formation of a ring chromosome (Cote et al., 1981; Kosztolanyi, 1987). Ring chromosomes are thought to be formed by deletion near the end(s) of chromosomes followed by fusion at breakage points and have been described for all human chromosomes. The resulting phenotypes vary greatly depending on the size and the nature of the deleted segments. Most ring chromosomes are formed by fusion of the deleted ends of both chromosome arms coupled with the loss of genetic material. However, in a few cases, the rings are formed by telomere-telomere fusion with little or no loss of chromosomal material and have intact subtelomeric and telomeric sequences suggesting that the "ring syndrome" might be associated with the silencing of genes in the vicinity of a longer telomere. The formation of intact ring caused by telomere-telomere fusion and associated with putative telomeric position effects has been reported for different autosomes (Pezzolo et al., 1993; Sigurdardottir et al., 1999; Vermeesch et al., 2002). For instance, a severe seizure disorder with features of non-convulsive epilepsy is a characteristic of ring chromosome 20. In this pathology, the formation of ring chromosome is generally associated with a breakage in each chromosome arm and the subsequent fusion of the broken ends with the loss of the telomere, subtelomeric regions or CHRNA4 and KCNQ2, two well-known epilepsy genes. In a patient with a typical severe epilepsy, classical cytogenetic methods, chromosome and quantitative FISH showed that the ring had a longer telomere than either of the 20p or 20q telomere ends suggesting that telomeric position effect silences the CHRNA4 and KCNQ2 genes (Zou et al., 2006).

Strikingly, most of these genetic diseases associated with mental retardation and different malformations are either linked to terminal deletion or fusion raising the hypothesis of a major contribution for telomere and subtelomere integrity in development. However, attempts to make genotype-phenotype correlations with specific anomalies have been difficult because of the paucity of reported cases and the variability in the size of the terminal deletion. In addition these chromosomal abnormalities are often mosaic and the occurrence of sister chromatid exchange complicates the description of these heterogeneous developmental disorders and the precise classification of the genetic alterations.

Conclusions
Telomeres on natural chromosomes are dynamic regions involved in numerous cellular pathways, which in turn controls cell fate. Next to telomeres, the nature and structure of subtelomeric regions might directly act on the specific topology of chromatin at chromosome ends. Consequently, a number of factors that can influence directly or indirectly the telomere length would likely affect the expression of subtelomeric sequences by changing telomere conformation and maintenance and vice versa. Furthermore, chromosome ends are associated with multiple pathologies and a more complete knowledge of telomere and subtelomere regulation, especially those involving epigenetic mechanisms would likely provide important insights into the role of chromosome ends in cancer, age-related diseases or response to environmental stress and infection but also numerous pathologies such as developmental disorders, mental retardation, infertility and spontaneous recurrent miscarriages. In addition the deciphering of the molecular mechanism sustaining these pathologies would provide a new avenue for the development of therapeutic approaches aimed at correcting the molecular defects caused by inappropriate modification of telomeric silencing and rearrangements of chromosome ends.

Acknowledgements
The work in Gilson lab is supported by the Ligue Nationale contre le Cancer (Equipe labellisée) and by the Association Française contre les Myopathies (AFM).

Bibliography

Cytological observations of deficiencies involving known genes translocations and an inversion in Zea mays.
McClintock B.
Mo. Agric. Exp. Res. Stn. Res Bull. 1931;163:4-30.
 
The Limited in Vitro Lifetime of Human Diploid Cell Strains.
Hayflick L.
Exp Cell Res. 1965 Mar;37:614-36.
PMID 14315085
 
Telomeric satellite DNA functions in regulating recombination.
Miklos GL, Nankivell RN.
Chromosoma. 1976 Jun 30;56(2):143-67.
PMID 976019
 
A tandemly repeated sequence at the termini of the extrachromosomal ribosomal RNA genes in Tetrahymena.
Blackburn EH, Gall JG.
J Mol Biol. 1978 Mar 25;120(1):33-53.
PMID 642006
 
The cytogenetic and clinical implications of a ring chromosome 2.
Cote GB, Katsantoni A, Deligeorgis D.
Ann Genet. 1981;24(4):231-5.
PMID 6977305
 
All gene-sized DNA molecules in four species of hypotrichs have the same terminal sequence and an unusual 3' terminus.
Klobutcher LA, Swanton MT, Donini P, Prescott DM.
Proc Natl Acad Sci U S A. 1981 May;78(5):3015-9.
PMID 6265931
 
Functional analysis of the white gene of Drosophila by P-factor-mediated transformation.
Gehring WJ, Klemenz R, Weber U, Kloter U.
Embo J. 1984 Sep;3(9):2077-85.
PMID 16453549
 
Transformation of white locus DNA in drosophila: dosage compensation, zeste interaction, and position effects.
Hazelrigg T, Levis R, Rubin GM.
Cell. 1984 Feb;36(2):469-81.
PMID 6319027
 
Effects of genomic position on the expression of transduced copies of the white gene of Drosophila.
Levis R, Hazelrigg T, Rubin GM.
Science. 1985 Aug 9;229(4713):558-61.
PMID 2992080
 
Does "ring syndrome" exist? An analysis of 207 case reports on patients with a ring autosome.
Kosztolanyi G.
Hum Genet. 1987 Feb;75(2):174-9.
PMID 3817812
 
Position effect at S. cerevisiae telomeres: reversible repression of Pol II transcription.
Gottschling DE, Aparicio OM, Billington BL, Zakian VA.
Cell. 1990 Nov 16;63(4):751-62.
PMID 2225075
 
Hypervariable ultra-long telomeres in mice.
Kipling D, Cooke HJ.
Nature. 1990 Sep 27;347(6291):400-2.
PMID 2170845
 
Localization of ribosomal and telomeric DNA sequences in intact plant nuclei by in-situ hybridization and three-dimensional optical microscopy.
Rawlins DJ, Shaw PJ.
J Microsc. 1990 Jan;157(Pt 1):83-9.
PMID 2299663
 
Telomeres and the functional architecture of the nucleus.
Gilson E, Laroche T, Gasser SM.
Trends Cell Biol. 1993 Apr;3(4):128-34.
PMID 14731767
 
RAP1 and telomere structure regulate telomere position effects in Saccharomyces cerevisiae.
Kyrion G, Liu K, Liu C, Lustig AJ.
Genes Dev. 1993 Jul;7(7A):1146-59.
PMID 8319907
 
De novo truncation of chromosome 16p and healing with (TTAGGG)n in the alpha-thalassemia/mental retardation syndrome (ATR-16).
Lamb J, Harris PC, Wilkie AO, Wood WG, Dauwerse JG, Higgs DR.
Am J Hum Genet. 1993 Apr;52(4):668-76.
PMID 8460633
 
Presence of telomeric and subtelomeric sequences at the fusion points of ring chromosomes indicates that the ring syndrome is caused by ring instability.
Pezzolo A, Gimelli G, Cohen A, Lavaggetto A, Romano C, Fogu G, Zuffardi O.
Hum Genet. 1993 Aug;92(1):23-7.
PMID 8365723
 
Silent domains are assembled continuously from the telomere and are defined by promoter distance and strength, and by SIR3 dosage.
Renauld H, Aparicio OM, Zierath PD, Billington BL, Chhablani SK, Gottschling DE.
Genes Dev. 1993 Jul;7(7A):1133-45.
PMID 8319906
 
Sandwiching of a gene within 12 kb of a functional telomere and alpha satellite does not result in silencing.
Bayne RA, Broccoli D, Taggart MH, Thomson EJ, Farr CJ, Cooke HJ.
Hum Mol Genet. 1994 Apr;3(4):539-46.
PMID 8069295
 
Analysis of the tandem repeat locus D4Z4 associated with facioscapulohumeral muscular dystrophy.
Hewitt JE, Lyle R, Clark LN, Valleley EM, Wright TJ, Wijmenga C, van Deutekom JC, Francis F, Sharpe PT, Hofker M, et al.
Hum Mol Genet. 1994 Aug;3(8):1287-95.
PMID 7987304
 
The DNA rearrangement associated with facioscapulohumeral muscular dystrophy involves a heterochromatin-associated repetitive element: implications for a role of chromatin structure in the pathogenesis of the disease.
Winokur ST, Bengtsson U, Feddersen J, Mathews KD, Weiffenbach B, Bailey H, Markovich RP, Murray JC, Wasmuth JJ, Altherr MR, et al.
Chromosome Res. 1994 May;2(3):225-34.
PMID 8069466
 
A human telomeric protein.
Chong L, van Steensel B, Broccoli D, Erdjument-Bromage H, Hanish J, Tempst P, de Lange T.
Science. 1995 Dec 8;270(5242):1663-7.
PMID 7502076
 
Organization of telomeric and subtelomeric chromatin in the higher plant Nicotiana tabacum.
Fajkus J, Kovarik A, Kralovics R, Bezdek M.
Mol Gen Genet. 1995 Jun 10;247(5):633-8.
PMID 7603443
 
The detection of subtelomeric chromosomal rearrangements in idiopathic mental retardation.
Flint J, Wilkie AO, Buckle VJ, Winter RM, Holland AJ, McDermid HE.
Nat Genet. 1995 Feb;9(2):132-40.
PMID 7719339
 
The telobox, a Myb-related telomeric DNA binding motif found in proteins from yeast, plants and human.
Bilaud T, Koering CE, Binet-Brasselet E, Ancelin K, Pollice A, Gasser SM, Gilson E.
Nucleic Acids Res. 1996 Apr 1;24(7):1294-303.
PMID 8614633
 
The clustering of telomeres and colocalization with Rap1, Sir3, and Sir4 proteins in wild-type Saccharomyces cerevisiae.
Gotta M, Laroche T, Formenton A, Maillet L, Scherthan H, Gasser SM.
J Cell Biol. 1996 Sep;134(6):1349-63.
PMID 8830766
 
Structural variation of the pseudoautosomal region between and within inbred mouse strains.
Kipling D, Wilson HE, Thomson EJ, Lee M, Perry J, Palmer S, Ashworth A, Cooke HJ.
Proc Natl Acad Sci U S A. 1996 Jan 9;93(1):171-5.
PMID 8552598
 
Structure, subnuclear distribution, and nuclear matrix association of the mammalian telomeric complex.
Luderus ME, van Steensel B, Chong L, Sibon OC, Cremers FF, de Lange T.
J Cell Biol. 1996 Nov;135(4):867-81.
PMID 8922373
 
Evidence for silencing compartments within the yeast nucleus: a role for telomere proximity and Sir protein concentration in silencer-mediated repression.
Maillet L, Boscheron C, Gotta M, Marcand S, Gilson E, Gasser SM.
Genes Dev. 1996 Jul 15;10(14):1796-811.
PMID 8698239
 
Effect of telomere length on telomeric gene expression.
Sprung CN, Sabatier L, Murnane JP.
Nucleic Acids Res. 1996 Nov 1;24(21):4336-40.
PMID 8932391
 
Identification of the first gene (FRG1) from the FSHD region on human chromosome 4q35.
van Deutekom JC, Lemmers RJ, Grewal PK, van Geel M, Romberg S, Dauwerse HG, Wright TJ, Padberg GW, Hofker MH, Hewitt JE, Frants RR.
Hum Mol Genet. 1996 May;5(5):581-90.
PMID 8733123
 
In vitro low propensity to form nucleosomes of four telomeric sequences.
Cacchione S, Cerone MA, Savino M.
FEBS Lett. 1997 Jan 2;400(1):37-41.
PMID 9000509
 
Extrachromosomal rDNA circles--a cause of aging in yeast.
Sinclair DA, Guarente L.
Cell. 1997 Dec 26;91(7):1033-42.
PMID 9428525
 
X-linked dyskeratosis congenita is caused by mutations in a highly conserved gene with putative nucleolar functions.
Heiss NS, Knight SW, Vulliamy TJ, Klauck SM, Wiemann S, Mason PJ, Poustka A, Dokal I.
Nat Genet. 1998 May;19(1):32-8.
PMID 9590285
 
1.4 Mb candidate gene region for X linked dyskeratosis congenita defined by combined haplotype and X chromosome inactivation analysis.
Knight SW, Vulliamy TJ, Heiss NS, Matthijs G, Devriendt K, Connor JM, D'Urso M, Poustka A, Mason PJ, Dokal I.
J Med Genet. 1998 Dec;35(12):993-6.
PMID 9863595
 
Association of nucleoside diphosphate kinase nm23-H2 with human telomeres.
Nosaka K, Kawahara M, Masuda M, Satomi Y, Nishino H.
Biochem Biophys Res Commun. 1998 Feb 13;243(2):342-8.
PMID 9480811
 
Nucleosome assembly on telomeric sequences.
Rossetti L, Cacchione S, Fua M, Savino M.
Biochemistry. 1998 May 12;37(19):6727-37.
PMID 9578556
 
TRF2 protects human telomeres from end-to-end fusions.
van Steensel B, Smogorzewska A, de Lange T.
Cell. 1998 Feb 6;92(3):401-13.
PMID 9476899
 
Mammalian telomeres end in a large duplex loop.
Griffith JD, Comeau L, Rosenfield S, Stansel RM, Bianchi A, Moss H, de Lange T.
Cell. 1999 May 14;97(4):503-14.
PMID 10338214
 
Creation of human tumour cells with defined genetic elements.
Hahn WC, Counter CM, Lundberg AS, Beijersbergen RL, Brooks MW, Weinberg RA.
Nature. 1999 Jul 29;400(6743):464-8.
PMID 10440377
 
The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms.
Kaeberlein M, McVey M, Guarente L.
Genes Dev. 1999 Oct 1;13(19):2570-80.
PMID 10521401
 
p53- and ATM-dependent apoptosis induced by telomeres lacking TRF2.
Karlseder J, Broccoli D, Dai Y, Hardy S, de Lange T.
Science. 1999 Feb 26;283(5406):1321-5.
PMID 10037601
 
A telomerase component is defective in the human disease dyskeratosis congenita.
Mitchell JR, Wood E, Collins K.
Nature. 1999 Dec 2;402(6761):551-5.
PMID 10591218
 
Position effect of human telomeric repeats on replication timing.
Ofir R, Wong AC, McDermid HE, Skorecki KL, Selig S.
Proc Natl Acad Sci U S A. 1999 Sep 28;96(20):11434-9.
PMID 10500194
 
Limitations of silencing at native yeast telomeres.
Pryde FE, Louis EJ.
Embo J. 1999 May 4;18(9):2538-50.
PMID 10228167
 
Clinical, cytogenetic, and fluorescence in situ hybridization findings in two cases of "complete ring" syndrome.
Sigurdardottir S, Goodman BK, Rutberg J, Thomas GH, Jabs EW, Geraghty MT.
Am J Med Genet. 1999 Dec 22;87(5):384-90.
PMID 10594875
 
Telomerase-negative immortalized human cells contain a novel type of promyelocytic leukemia (PML) body.
Yeager TR, Neumann AA, Englezou A, Huschtscha LI, Noble JR, Reddel RR.
Cancer Res. 1999 Sep 1;59(17):4175-9.
PMID 10485449
 
Telomere maintenance by recombination in human cells.
Dunham MA, Neumann AA, Fasching CL, Reddel RR.
Nat Genet. 2000 Dec;26(4):447-50.
PMID 11101843
 
The main role of the sequence-dependent DNA elasticity in determining the free energy of nucleosome formation on telomeric DNAs.
Filesi I, Cacchione S, De Santis P, Rossetti L, Savino M.
Biophys Chem. 2000 Jan 24;83(3):223-37.
PMID 10647852
 
Frequent ectopic recombination of virulence factor genes in telomeric chromosome clusters of P. falciparum.
Freitas-Junior LH, Bottius E, Pirrit LA, Deitsch KW, Scheidig C, Guinet F, Nehrbass U, Wellems TE, Scherf A.
Nature. 2000 Oct 26;407(6807):1018-22.
PMID 11069183
 
Chromosomal breakage-fusion-bridge events cause genetic intratumor heterogeneity.
Gisselsson D, Pettersson L, Hoglund M, Heidenblad M, Gorunova L, Wiegant J, Mertens F, Dal Cin P, Mitelman F, Mandahl N.
Proc Natl Acad Sci U S A. 2000 May 9;97(10):5357-62.
PMID 10805796
 
Ku acts in a unique way at the mammalian telomere to prevent end joining.
Hsu HL, Gilley D, Galande SA, Hande MP, Allen B, Kim SH, Li GC, Campisi J, Kohwi-Shigematsu T, Chen DJ.
Genes Dev. 2000 Nov 15;14(22):2807-12.
PMID 11090128
 
Perfect endings: a review of subtelomeric probes and their use in clinical diagnosis.
Knight SJ, Flint J.
J Med Genet. 2000 Jun;37(6):401-9.
PMID 10851249
 
Whole-genome methylation scan in ICF syndrome: hypomethylation of non-satellite DNA repeats D4Z4 and NBL2.
Kondo T, Bobek MP, Kuick R, Lamb B, Zhu X, Narayan A, Bourc'his D, Viegas-Pequignot E, Ehrlich M, Hanash SM.
Hum Mol Genet. 2000 Mar 1;9(4):597-604.
PMID 10699183
 
Identification of human Rap1: implications for telomere evolution.
Li B, Oestreich S, de Lange T.
Cell. 2000 May 26;101(5):471-83.
PMID 10850490
 
Effect of the size of the deletion and clinical manifestation in Wolf-Hirschhorn syndrome: analysis of 13 patients with a de novo deletion.
Wieczorek D, Krause M, Majewski F, Albrecht B, Horn D, Riess O, Gillessen-Kaesbach G.
Eur J Hum Genet. 2000a Jul;8(7):519-26.
PMID 10909852
 
Unexpected high frequency of de novo unbalanced translocations in patients with Wolf-Hirschhorn syndrome (WHS).
Wieczorek D, Krause M, Majewski F, Albrecht B, Meinecke P, Riess O, Gillessen-Kaesbach G.
J Med Genet. 2000b Oct;37(10):798-804.
PMID 11183188
 
NBS1 and TRF1 colocalize at promyelocytic leukemia bodies during late S/G2 phases in immortalized telomerase-negative cells. Implication of NBS1 in alternative lengthening of telomeres.
Wu G, Lee WH, Chen PL.
J Biol Chem. 2000 Sep 29;275(39):30618-22.
PMID 10913111
 
Role of human telomerase reverse transcriptase and telomeric-repeat binding factor proteins 1 and 2 in human hematopoietic cells.
Yamada K, Yajima T, Yagihashi A, Kobayashi D, Koyanagi Y, Asanuma K, Yamada M, Moriai R, Kameshima H, Watanabe N.
Jpn J Cancer Res. 2000 Dec;91(12):1278-84.
PMID 11123427
 
Chromatin structure of telomere domain in human sperm.
Zalenskaya IA, Bradbury EM, Zalensky AO.
Biochem Biophys Res Commun. 2000 Dec 9;279(1):213-8.
PMID 11112441
 
Cell-cycle-regulated association of RAD50/MRE11/NBS1 with TRF2 and human telomeres.
Zhu XD, Kuster B, Mann M, Petrini JH, de Lange T.
Nat Genet. 2000 Jul;25(3):347-52.
PMID 10888888
 
Pot1, the putative telomere end-binding protein in fission yeast and humans.
Baumann P, Cech TR.
Science. 2001 May 11;292(5519):1171-5.
PMID 11349150
 
Telomere position effect in human cells.
Baur JA, Zou Y, Shay JW, Wright WE.
Science. 2001 Jun 15;292(5524):2075-7.
PMID 11408657
 
Disruption of the ProSAP2 gene in a t(12;22)(q24.1;q13.3) is associated with the 22q13.3 deletion syndrome.
Bonaglia MC, Giorda R, Borgatti R, Felisari G, Gagliardi C, Selicorni A, Zuffardi O.
Am J Hum Genet. 2001 Aug;69(2):261-8.
PMID 11431708
 
Control of VSG gene expression sites.
Borst P, Ulbert S.
Mol Biochem Parasitol. 2001 Apr 25;114(1):17-27.
PMID 11356510
 
Olfactory receptor-gene clusters, genomic-inversion polymorphisms, and common chromosome rearrangements.
Giglio S, Broman KW, Matsumoto N, Calvari V, Gimelli G, Neumann T, Ohashi H, Voullaire L, Larizza D, Giorda R, Weber JL, Ledbetter DH, Zuffardi O.
Am J Hum Genet. 2001 Apr;68(4):874-83.
PMID 11231899
 
The complete human olfactory subgenome.
Glusman G, Yanai I, Rubin I, Lancet D.
Genome Res. 2001 May;11(5):685-702.
PMID 11337468
 
Telomeric protein Pin2/TRF1 as an important ATM target in response to double strand DNA breaks.
Kishi S, Zhou XZ, Ziv Y, Khoo C, Hill DE, Shiloh Y, Lu KP.
J Biol Chem. 2001 Aug 3;276(31):29282-91.
PMID 11375976
 
Receptors and transduction in taste.
Lindemann B.
Nature. 2001 Sep 13;413(6852):219-25.
PMID 11557991
 
Ku-deficient yeast strains exhibit alternative states of silencing competence.
Maillet L, Gaden F, Brevet V, Fourel G, Martin SG, Dubrana K, Gasser SM, Gilson E.
EMBO Rep. 2001 Mar;2(3):203-10.
PMID 11266361
 
A specific interaction between the telomeric protein Pin2/TRF1 and the mitotic spindle.
Nakamura M, Zhou XZ, Kishi S, Kosugi I, Tsutsui Y, Lu KP.
Curr Biol. 2001 Oct 2;11(19):1512-6.
PMID 11591318
 
SALL1, the gene mutated in Townes-Brocks syndrome, encodes a transcriptional repressor which interacts with TRF1/PIN2 and localizes to pericentromeric heterochromatin.
Netzer C, Rieger L, Brero A, Zhang CD, Hinzke M, Kohlhase J, Bohlander SK.
Hum Mol Genet. 2001 Dec 15;10(26):3017-24.
PMID 11751684
 
Distribution of somatic H1 subtypes is non-random on active vs. inactive chromatin II: distribution in human adult fibroblasts.
Parseghian MH, Newcomb RL, Hamkalo BA.
J Cell Biochem. 2001;83(4):643-59.
PMID 11746507
 
Integration of telomere sequences with the draft human genome sequence.
Riethman HC, Xiang Z, Paul S, Morse E, Hu XL, Flint J, Chi HC, Grady DL, Moyzis RK.
Nature. 2001 Feb 15;409(6822):948-51.
PMID 11237019
 
Specific interactions of the telomeric protein Rap1p with nucleosomal binding sites.
Rossetti L, Cacchione S, De Menna A, Chapman L, Rhodes D, Savino M.
J Mol Biol. 2001 Mar 9;306(5):903-13.
PMID 11237607
 
T-loop assembly in vitro involves binding of TRF2 near the 3' telomeric overhang.
Stansel RM, de Lange T, Griffith JD.
EMBO J. 2001 Oct 1;20(19):5532-40.
PMID 11574485
 
Localization of yeast telomeres to the nuclear periphery is separable from transcriptional repression and telomere stability functions.
Tham WH, Wyithe JS, Ferrigno PK, Silver PA, Zakian VA.
Mol Cell. 2001 Jul;8(1):189-99.
PMID 11511372
 
The RNA component of telomerase is mutated in autosomal dominant dyskeratosis congenita.
Vulliamy T, Marrone A, Goldman F, Dearlove A, Bessler M, Mason PJ, Dokal I.
Nature. 2001 Sep 27;413(6854):432-5.
PMID 11574891
 
Role of DNA sequence in nucleosome stability and dynamics.
Widom J.
Q Rev Biophys. 2001 Aug;34(3):269-324.
PMID 11838235
 
Targeting assay to study the cis functions of human telomeric proteins: evidence for inhibition of telomerase by TRF1 and for activation of telomere degradation by TRF2.
Ancelin K, Brunori M, Bauwens S, Koering CE, Brun C, Ricoul M, Pommier JP, Sabatier L, Gilson E.
Mol Cell Biol. 2002 May;22(10):3474-87.
PMID 11971978
 
Interphase chromosomes in Arabidopsis are organized as well defined chromocenters from which euchromatin loops emanate.
Fransz P, De Jong JH, Lysak M, Castiglione MR, Schubert I.
Proc Natl Acad Sci U S A. 2002 Oct 29;99(22):14584-9.
PMID 12384572
 
Inappropriate gene activation in FSHD: a repressor complex binds a chromosomal repeat deleted in dystrophic muscle.
Gabellini D, Green MR, Tupler R.
Cell. 2002 Aug 9;110(3):339-48.
PMID 12176321
 
A role for the Rb family of proteins in controlling telomere length.
Garcia-Cao M, Gonzalo S, Dean D, Blasco MA.
Nat Genet. 2002 Nov;32(3):415-9.
PMID 12379853
 
Nuclear organization and silencing: putting things in their place.
Hediger F, Gasser SM.
Nat Cell Biol. 2002 Mar;4(3):E53-5.
PMID 11875445
 
Alternative lengthening of telomeres in mammalian cells.
Henson JD, Neumann AA, Yeager TR, Reddel RR.
Oncogene. 2002 Jan 21;21(4):598-610.
PMID 11850785
 
Human telomeric position effect is determined by chromosomal context and telomeric chromatin integrity.
Koering CE, Pollice A, Zibella MP, Bauwens S, Puisieux A, Brunori M, Brun C, Martins L, Sabatier L, Pulitzer JF, Gilson E.
EMBO Rep. 2002 Nov;3(11):1055-61.
PMID 12393752
 
Defensins of vertebrate animals.
Lehrer RI, Ganz T.
Curr Opin Immunol. 2002 Feb;14(1):96-102.
PMID 11790538
 
The complex structure and dynamic evolution of human subtelomeres.
Mefford HC, Trask BJ.
Nat Rev Genet. 2002 Feb;3(2):91-102.
PMID 11836503
 
Microarray deacetylation maps determine genome-wide functions for yeast histone deacetylases.
Robyr D, Suka Y, Xenarios I, Kurdistani SK, Wang A, Suka N, Grunstein M.
Cell. 2002 May 17;109(4):437-46.
PMID 12086601
 
High-throughput analysis of subtelomeric chromosome rearrangements by use of array-based comparative genomic hybridization.
Veltman JA, Schoenmakers EF, Eussen BH, Janssen I, Merkx G, van Cleef B, van Ravenswaaij CM, Brunner HG, Smeets D, van Kessel AG.
Am J Hum Genet. 2002 May;70(5):1269-76.
PMID 11951177
 
Ring syndrome caused by ring chromosome 7 without loss of subtelomeric sequences.
Vermeesch JR, Baten E, Fryns JP, Devriendt K.
Clin Genet. 2002 Nov;62(5):415-7.
PMID 12431259
 
Association between aplastic anaemia and mutations in telomerase RNA.
Vulliamy T, Marrone A, Dokal I, Mason PJ.
Lancet. 2002 Jun 22;359(9324):2168-70.
PMID 12090986
 
Decreased gene expression for telomeric-repeat binding factors and TIN2 in malignant hematopoietic cells.
Yamada K, Yagihashi A, Yamada M, Asanuma K, Moriai R, Kobayashi D, Tsuji N, Watanabe N.
Anticancer Res. 2002 Mar-Apr;22(2B):1315-20.
PMID 12168944
 
Why are parasite contingency genes often associated with telomeres?
Barry JD, Ginger ML, Burton P, McCulloch R.
Int J Parasitol. 2003 Jan;33(1):29-45.
PMID 12547344
 
Interferon and granulopoiesis signatures in systemic lupus erythematosus blood.
Bennett L, Palucka AK, Arce E, Cantrell V, Borvak J, Banchereau J, Pascual V.
J Exp Med. 2003 Mar 17;197(6):711-23.
PMID 12642603
 
Acetylated nucleosome assembly on telomeric DNAs.
Cacchione S, Luis Rodriguez J, Mechelli R, Franco L, Savino M.
Biophys Chem. 2003 Jun 1;104(2):381-92.
PMID 12878307
 
Association between telomere length in blood and mortality in people aged 60 years or older.
Cawthon RM, Smith KR, O'Brien E, Sivatchenko A, Kerber RA.
Lancet. 2003 Feb 1;361(9355):393-5.
PMID 12573379
 
A DNA damage checkpoint response in telomere-initiated senescence.
d'Adda di Fagagna F, Reaper PM, Clay-Farrace L, Fiegler H, Carr P, Von Zglinicki T, Saretzki G, Carter NP, Jackson SP.
Nature. 2003 Nov 13;426(6963):194-8.
PMID 14608368
 
Lamin a truncation in Hutchinson-Gilford progeria.
De Sandre-Giovannoli A, Bernard R, Cau P, Navarro C, Amiel J, Boccaccio I, Lyonnet S, Stewart CL, Munnich A, Le Merrer M, Levy N.
Science. 2003 Jun 27;300(5628):2055.
PMID 12702809
 
Recurrent de novo point mutations in lamin A cause Hutchinson-Gilford progeria syndrome.
Eriksson M, Brown WT, Gordon LB, Glynn MW, Singer J, Scott L, Erdos MR, Robbins CM, Moses TY, Berglund P, Dutra A, Pak E, Durkin S, Csoka AB, Boehnke M, Glover TW, Collins FS.
Nature. 2003 May 15;423(6937):293-8.
PMID 12714972
 
Defensins: antimicrobial peptides of innate immunity.
Ganz T.
Nat Rev Immunol. 2003 Sep;3(9):710-20.
PMID 12949495
 
Human specific loss of olfactory receptor genes.
Gilad Y, Man O, Paabo S, Lancet D.
Proc Natl Acad Sci U S A. 2003 Mar 18;100(6):3324-7.
PMID 12612342
 
Population data suggest that deletions of 1p36 are a relatively common chromosome abnormality.
Heilstedt HA, Ballif BC, Howard LA, Kashork CD, Shaffer LG.
Clin Genet. 2003a Oct;64(4):310-6.
PMID 12974736
 
Physical map of 1p36, placement of breakpoints in monosomy 1p36, and clinical characterization of the syndrome.
Heilstedt HA, Ballif BC, Howard LA, Lewis RA, Stal S, Kashork CD, Bacino CA, Shapira SK, Shaffer LG.
Am J Hum Genet. 2003b May;72(5):1200-12.
PMID 12687501
 
Testing the position-effect variegation hypothesis for facioscapulohumeral muscular dystrophy by analysis of histone modification and gene expression in subtelomeric 4q.
Jiang G, Yang F, van Overveld PG, Vedanarayanan V, van der Maarel S, Ehrlich M.
Hum Mol Genet. 2003 Nov 15;12(22):2909-21.
PMID 14506132
 
Rap1 affects the length and heterogeneity of human telomeres.
Li B, de Lange T.
Mol Biol Cell. 2003 Dec;14(12):5060-8.
PMID 14565979
 
Telomere length and the expression of natural telomeric genes in human fibroblasts.
Ning Y, Xu JF, Li Y, Chavez L, Riethman HC, Lansdorp PM, Weng NP.
Hum Mol Genet. 2003 Jun 1;12(11):1329-36.
PMID 12761048
 
Replicative senescence of activated human hepatic stellate cells is accompanied by a pronounced inflammatory but less fibrogenic phenotype.
Schnabl B, Purbeck CA, Choi YH, Hagedorn CH, Brenner D.
Hepatology. 2003 Mar;37(3):653-64.
PMID 12601363
 
Human heterochromatin protein 1 isoforms HP1(Hsalpha) and HP1(Hsbeta) interfere with hTERT-telomere interactions and correlate with changes in cell growth and response to ionizing radiation.
Sharma GG, Hwang KK, Pandita RK, Gupta A, Dhar S, Parenteau J, Agarwal M, Worman HJ, Wellinger RJ, Pandita TK.
Mol Cell Biol. 2003 Nov;23(22):8363-76.
PMID 14585993
 
DNA damage foci at dysfunctional telomeres.
Takai H, Smogorzewska A, de Lange T.
Curr Biol. 2003 Sep 2;13(17):1549-56.
PMID 12956959
 
Three-dimensional arrangements of centromeres and telomeres in nuclei of human and murine lymphocytes.
Weierich C, Brero A, Stein S, von Hase J, Cremer C, Cremer T, Solovei I.
Chromosome Res. 2003;11(5):485-502.
PMID 12971724
 
Molecular characterisation of the 22q13 deletion syndrome supports the role of haploinsufficiency of SHANK3/PROSAP2 in the major neurological symptoms.
Wilson HL, Wong AC, Shaw SR, Tse WY, Stapleton GA, Phelan MC, Hu S, Marshall J, McDermid HE.
J Med Genet. 2003 Aug;40(8):575-84.
PMID 12920066
 
Expression profiling of FSHD muscle supports a defect in specific stages of myogenic differentiation.
Winokur ST, Chen YW, Masny PS, Martin JH, Ehmsen JT, Tapscott SJ, van der Maarel SM, Hayashi Y, Flanigan KM.
Hum Mol Genet. 2003 Nov 15;12(22):2895-907.
PMID 14519683
 
ERCC1/XPF removes the 3' overhang from uncapped telomeres and represses formation of telomeric DNA-containing double minute chromosomes.
Zhu XD, Niedernhofer L, Kuster B, Mann M, Hoeijmakers JH, de Lange T.
Mol Cell. 2003 Dec;12(6):1489-98.
PMID 14690602
 
Regulation of C. elegans longevity by specific gustatory and olfactory neurons.
Alcedo J, Kenyon C.
Neuron. 2004 Jan 8;41(1):45-55.
PMID 14715134
 
Essential role of limiting telomeres in the pathogenesis of Werner syndrome.
Chang S, Multani AS, Cabrera NG, Naylor ML, Laud P, Lombard D, Pathak S, Guarente L, DePinho RA.
Nat Genet. 2004 Aug;36(8):877-82.
PMID 15235603
 
Accelerated telomere shortening in response to life stress.
Epel ES, Blackburn EH, Lin J, Dhabhar FS, Adler NE, Morrow JD, Cawthon RM.
Proc Natl Acad Sci U S A. 2004 Dec 7;101(49):17312-5.
PMID 15574496
 
An SMC-domain protein in fission yeast links telomeres to the meiotic centrosome.
Flory MR, Carson AR, Muller EG, Aebersold R.
Mol Cell. 2004 Nov 19;16(4):619-30.
PMID 15546621
 
When enough is enough: genetic diseases associated with transcriptional derepression.
Gabellini D, Green MR, Tupler R.
Curr Opin Genet Dev. 2004 Jun;14(3):301-7.
PMID 15172674
 
Epigenetic regulation of telomere length in mammalian cells by the Suv39h1 and Suv39h2 histone methyltransferases.
Garcia-Cao M, O'Sullivan R, Peters AH, Jenuwein T, Blasco MA.
Nat Genet. 2004 Jan;36(1):94-9.
PMID 14702045
 
Sir-mediated repression can occur independently of chromosomal and subnuclear contexts.
Gartenberg MR, Neumann FR, Laroche T, Blaszczyk M, Gasser SM.
Cell. 2004 Dec 29;119(7):955-67.
PMID 15620354
 
The relative lengths of individual telomeres are defined in the zygote and strictly maintained during life.
Graakjaer J, Pascoe L, Der-Sarkissian H, Thomas G, Kolvraa S, Christensen K, Londono-Vallejo JA.
Aging Cell. 2004 Jun;3(3):97-102.
PMID 15153177
 
A dynamic molecular link between the telomere length regulator TRF1 and the chromosome end protector TRF2.
Houghtaling BR, Cuttonaro L, Chang W, Smith S.
Curr Biol. 2004 Sep 21;14(18):1621-31.
PMID 15380063
 
TIN2 mediates functions of TRF2 at human telomeres.
Kim SH, Beausejour C, Davalos AR, Kaminker P, Heo SJ, Campisi J.
J Biol Chem. 2004 Oct 15;279(42):43799-804.
PMID 15292264
 
Inflammatory mediators in the elderly.
Krabbe KS, Pedersen M, Bruunsgaard H.
Exp Gerontol. 2004 May;39(5):687-99.
PMID 15130663
 
Association and regulation of the BLM helicase by the telomere proteins TRF1 and TRF2.
Lillard-Wetherell K, Machwe A, Langland GT, Combs KA, Behbehani GK, Schonberg SA, German J, Turchi JJ, Orren DK, Groden J.
Hum Mol Genet. 2004 Sep 1;13(17):1919-32.
PMID 15229185
 
Telosome, a mammalian telomere-associated complex formed by multiple telomeric proteins.
Liu D, O'Connor MS, Qin J, Songyang Z.
J Biol Chem. 2004a Dec 3;279(49):51338-42.
PMID 15383534
 
PTOP interacts with POT1 and regulates its localization to telomeres.
Liu D, Safari A, O'Connor MS, Chan DW, Laegeler A, Qin J, Songyang Z.
Nat Cell Biol. 2004b Jul;6(7):673-80.
PMID 15181449
 
DNA binding features of human POT1: a nonamer 5'-TAGGGTTAG-3' minimal binding site, sequence specificity, and internal binding to multimeric sites.
Loayza D, Parsons H, Donigian J, Hoke K, de Lange T.
J Biol Chem. 2004 Mar 26;279(13):13241-8.
PMID 14715659
 
Localization of 4q35.2 to the nuclear periphery: is FSHD a nuclear envelope disease?
Masny PS, Bengtsson U, Chung SA, Martin JH, van Engelen B, van der Maarel SM, Winokur ST.
Hum Mol Genet. 2004 Sep 1;13(17):1857-71.
PMID 15238509
 
Dynamics of protein binding to telomeres in living cells: implications for telomere structure and function.
Mattern KA, Swiggers SJ, Nigg AL, Lowenberg B, Houtsmuller AB, Zijlmans JM.
Mol Cell Biol. 2004 Jun;24(12):5587-94.
PMID 15169917
 
Signalling to actin assembly via the WASP (Wiskott-Aldrich syndrome protein)-family proteins and the Arp2/3 complex.
Millard TH, Sharp SJ, Machesky LM.
Biochem J. 2004 May 15;380(Pt 1):1-17.
PMID 15040784
 
The Werner syndrome helicase and exonuclease cooperate to resolve telomeric D loops in a manner regulated by TRF1 and TRF2.
Opresko PL, Otterlei M, Graakjaer J, Bruheim P, Dawut L, Kolvraa S, May A, Seidman MM, Bohr VA.
Mol Cell. 2004 Jun 18;14(6):763-74.
PMID 15200954
 
Subtelomeric deletions detected in patients with idiopathic mental retardation using multiplex ligation-dependent probe amplification (MLPA).
Rooms L, Reyniers E, van Luijk R, Scheers S, Wauters J, Ceulemans B, Van Den Ende J, Van Bever Y, Kooy RF.
Hum Mutat. 2004 Jan;23(1):17-21.
PMID 14695528
 
Multiple pathways for telomere tethering: functional implications of subnuclear position for heterochromatin formation.
Taddei A, Gasser SM.
Biochim Biophys Acta. 2004 Mar 15;1677(1-3):120-8.
PMID 15020053
 
The 4q subtelomere harboring the FSHD locus is specifically anchored with peripheral heterochromatin unlike most human telomeres.
Tam R, Smith KP, Lawrence JB.
J Cell Biol. 2004 Oct 25;167(2):269-79.
PMID 15504910
 
Elucidation of a cryptic interstitial 7q31.3 deletion in a patient with a language disorder and mild mental retardation by array-CGH.
Tyson C, McGillivray B, Chijiwa C, Rajcan-Separovic E.
Am J Med Genet A. 2004 Sep 1;129A(3):254-60.
PMID 15326624
 
Loss of hPot1 function leads to telomere instability and a cut-like phenotype.
Veldman T, Etheridge KT, Counter CM.
Curr Biol. 2004 Dec 29;14(24):2264-70.
PMID 15620654
 
Disease anticipation is associated with progressive telomere shortening in families with dyskeratosis congenita due to mutations in TERC.
Vulliamy T, Marrone A, Szydlo R, Walne A, Mason PJ, Dokal I.
Nat Genet. 2004 May;36(5):447-9.
PMID 15098033
 
Subtelomere FISH in 50 children with mental retardation and minor anomalies, identified by a checklist, detects 10 rearrangements including a de novo balanced translocation of chromosomes 17p13.3 and 20q13.33.
Walter S, Sandig K, Hinkel GK, Mitulla B, Ounap K, Sims G, Sitska M, Utermann B, Viertel P, Kalscheuer V, Bartsch O.
Am J Med Genet A. 2004 Aug 1;128(4):364-73.
PMID 15264281
 
TIN2 binds TRF1 and TRF2 simultaneously and stabilizes the TRF2 complex on telomeres.
Ye JZ, Donigian JR, van Overbeek M, Loayza D, Luo Y, Krutchinsky AN, Chait BT, de Lange T.
J Biol Chem. 2004 Nov 5;279(45):47264-71.
PMID 15316005
 
TIN2 is a tankyrase 1 PARP modulator in the TRF1 telomere length control complex.
Ye JZ, de Lange T.
Nat Genet. 2004 Jun;36(6):618-23.
PMID 15133513
 
Asynchronous replication timing of telomeres at opposite arms of mammalian chromosomes.
Zou Y, Gryaznov SM, Shay JW, Wright WE, Cornforth MN.
Proc Natl Acad Sci U S A. 2004 Aug 31;101(35):12928-33.
PMID 15322275
 
Subtelomere deletions and translocations are frequently familial.
Adeyinka A, Adams SA, Lorentz CP, Van Dyke DL, Jalal SM.
Am J Med Genet A. 2005 May 15;135(1):28-35.
PMID 15810004
 
Role of the RB1 family in stabilizing histone methylation at constitutive heterochromatin.
Gonzalo S, Garcia-Cao M, Fraga MF, Schotta G, Peters AH, Cotter SE, Eguia R, Dean DC, Esteller M, Jenuwein T, Blasco MA.
Nat Cell Biol. 2005 Apr;7(4):420-8.
PMID 15750587
 
POT1 protects telomeres from a transient DNA damage response and determines how human chromosomes end.
Hockemeyer D, Sfeir AJ, Shay JW, Wright WE, de Lange T.
EMBO J. 2005 Jul 20;24(14):2667-78.
PMID 15973431
 
Isolation and expression profiling of genes upregulated in the peripheral blood cells of systemic lupus erythematosus patients.
Ishii T, Onda H, Tanigawa A, Ohshima S, Fujiwara H, Mima T, Katada Y, Deguchi H, Suemura M, Miyake T, Miyatake K, Kawase I, Zhao H, Tomiyama Y, Saeki Y, Nojima H.
DNA Res. 2005;12(6):429-39.
PMID 16769699
 
Higher-order nuclear organization in growth arrest of human mammary epithelial cells: a novel role for telomere-associated protein TIN2.
Kaminker P, Plachot C, Kim SH, Chung P, Crippen D, Petersen OW, Bissell MJ, Campisi J, Lelievre SA.
J Cell Sci. 2005 Mar 15;118(Pt 6):1321-30.
PMID 15741234
 
Urogenital and caudal dysgenesis in adrenocortical dysplasia (acd) mice is caused by a splicing mutation in a novel telomeric regulator.
Keegan CE, Hutz JE, Else T, Adamska M, Shah SP, Kent AE, Howes JM, Beamer WG, Hammer GD.
Hum Mol Genet. 2005 Jan 1;14(1):113-23.
PMID 15537664
 
Human protection of telomeres 1 (POT1) is a negative regulator of telomerase activity in vitro.
Kelleher C, Kurth I, Lingner J.
Mol Cell Biol. 2005 Jan;25(2):808-18.
PMID 15632080
 
Molecular characterisation of patients with subtelomeric 22q abnormalities using chromosome specific array-based comparative genomic hybridisation.
Koolen DA, Reardon W, Rosser EM, Lacombe D, Hurst JA, Law CJ, Bongers EM, van Ravenswaaij-Arts CM, Leisink MA, van Kessel AG, Veltman JA, de Vries BB.
Eur J Hum Genet. 2005 Sep;13(9):1019-24.
PMID 15986041
 
Switching human telomerase on and off with hPOT1 protein in vitro.
Lei M, Zaug AJ, Podell ER, Cech TR.
J Biol Chem. 2005 May 27;280(21):20449-56.
PMID 15792951
 
Human subtelomeres are hot spots of interchromosomal recombination and segmental duplication.
Linardopoulou EV, Williams EM, Fan Y, Friedman C, Young JM, Trask BJ.
Nature. 2005 Sep 1;437(7055):94-100.
PMID 16136133
 
A novel telomere structure in a human alternative lengthening of telomeres cell line.
Marciniak RA, Cavazos D, Montellano R, Chen Q, Guarente L, Johnson FB.
Cancer Res. 2005 Apr 1;65(7):2730-7.
PMID 15805272
 
Up-regulation of telomere-binding proteins, TRF1, TRF2, and TIN2 is related to telomere shortening during human multistep hepatocarcinogenesis.
Oh BK, Kim YJ, Park C, Park YN.
Am J Pathol. 2005 Jan;166(1):73-80.
PMID 15632001
 
Tiling path resolution mapping of constitutional 1p36 deletions by array-CGH: contiguous gene deletion or "deletion with positional effect" syndrome?
Redon R, Rio M, Gregory SG, Cooper RA, Fiegler H, Sanlaville D, Banerjee R, Scott C, Carr P, Langford C, Cormier-Daire V, Munnich A, Carter NP, Colleaux L.
J Med Genet. 2005 Feb;42(2):166-71.
PMID 15689456
 
Obesity, cigarette smoking, and telomere length in women.
Valdes AM, Andrew T, Gardner JP, Kimura M, Oelsner E, Cherkas LF, Aviv A, Spector TD.
Lancet. 2005 Aug 20-26;366(9486):662-4.
PMID 16112303
 
The D4Z4 repeat-mediated pathogenesis of facioscapulohumeral muscular dystrophy.
van der Maarel SM, Frants RR.
Am J Hum Genet. 2005 Mar;76(3):375-86.
PMID 15674778
 
Mapping of a major locus that determines telomere length in humans.
Vasa-Nicotera M, Brouilette S, Mangino M, Thompson JR, Braund P, Clemitson JR, Mason A, Bodycote CL, Raleigh SM, Louis E, Samani NJ.
Am J Hum Genet. 2005 Jan;76(1):147-51.
PMID 15520935
 
Mutations in TERT, the gene for telomerase reverse transcriptase, in aplastic anemia.
Yamaguchi H, Calado RT, Ly H, Kajigaya S, Baerlocher GM, Chanock SJ, Lansdorp PM, Young NS.
N Engl J Med. 2005 Apr 7;352(14):1413-24.
PMID 15814878
 
POT1 and TRF2 cooperate to maintain telomeric integrity.
Yang Q, Zheng YL, Harris CC.
Mol Cell Biol. 2005 Feb;25(3):1070-80.
PMID 15657433
 
High-resolution mapping of genotype-phenotype relationships in cri du chat syndrome using array comparative genomic hybridization.
Zhang X, Snijders A, Segraves R, Niebuhr A, Albertson D, Yang H, Gray J, Niebuhr E, Bolund L, Pinkel D.
Am J Hum Genet. 2005 Feb;76(2):312-26.
PMID 15635506
 
Mapping genetic loci that determine leukocyte telomere length in a large sample of unselected female sibling pairs.
Andrew T, Aviv A, Falchi M, Surdulescu GL, Gardner JP, Lu X, Kimura M, Kato BS, Valdes AM, Spector TD.
Am J Hum Genet. 2006 Mar;78(3):480-6.
PMID 16400618
 
Increased expression of telomere length regulating factors TRF1, TRF2 and TIN2 in patients with adult T-cell leukemia.
Bellon M, Datta A, Brown M, Pouliquen JF, Couppie P, Kazanji M, Nicot C.
Int J Cancer. 2006 Nov 1;119(9):2090-7.
PMID 16786598
 
TRF2 inhibition triggers apoptosis and reduces tumourigenicity of human melanoma cells.
Biroccio A, Rizzo A, Elli R, Koering CE, Belleville A, Benassi B, Leonetti C, Stevens MF, D'Incalci M, Zupi G, Gilson E.
Eur J Cancer. 2006 Aug;42(12):1881-8.
PMID 16750909
 
Identification of a recurrent breakpoint within the SHANK3 gene in the 22q13.3 deletion syndrome.
Bonaglia MC, Giorda R, Mani E, Aceti G, Anderlid BM, Baroncini A, Pramparo T, Zuffardi O.
J Med Genet. 2006 Oct;43(10):822-8.
PMID 16284256
 
TRF2 inhibition promotes anchorage-independent growth of telomerase-positive human fibroblasts.
Brunori M, Mathieu N, Ricoul M, Bauwens S, Koering CE, Roborel de Climens A, Belleville A, Wang Q, Puisieux I, Decimo D, Puisieux A, Sabatier L, Gilson E.
Oncogene. 2006 Feb 16;25(7):990-7.
PMID 16205637
 
Cri du Chat syndrome.
Cerruti Mainardi P.
Orphanet J Rare Dis. 2006 Sep 5;1:33.
PMID 16953888
 
The effects of social status on biological aging as measured by white-blood-cell telomere length.
Cherkas LF, Aviv A, Valdes AM, Hunkin JL, Gardner JP, Surdulescu GL, Kimura M, Spector TD.
Aging Cell. 2006 Oct;5(5):361-5.
PMID 16856882
 
Inversion polymorphisms and non-contiguous terminal deletions: the cause and the (unpredicted) effect of our genome architecture.
Ciccone R, Mattina T, Giorda R, Bonaglia MC, Rocchi M, Pramparo T, Zuffardi O.
J Med Genet. 2006 May;43(5):e19.
PMID 16648372
 
MDC1 accelerates nonhomologous end-joining of dysfunctional telomeres.
Dimitrova N, de Lange T.
Genes Dev. 2006 Dec 1;20(23):3238-43.
PMID 17158742
 
The finger subdomain of yeast telomerase cooperates with Pif1p to limit telomere elongation.
Eugster A, Lanzuolo C, Bonneton M, Luciano P, Pollice A, Pulitzer JF, Stegberg E, Berthiau AS, Forstemann K, Corda Y, Lingner J, Geli V, Gilson E.
Nat Struct Mol Biol. 2006 Aug;13(8):734-9.
PMID 16878131
 
The human telomeric protein TRF1 specifically recognizes nucleosomal binding sites and alters nucleosome structure.
Galati A, Rossetti L, Pisano S, Chapman L, Rhodes D, Savino M, Cacchione S.
J Mol Biol. 2006 Jul 7;360(2):377-85.
PMID 16756990
 
The G-quadruplex ligand telomestatin inhibits POT1 binding to telomeric sequences in vitro and induces GFP-POT1 dissociation from telomeres in human cells.
Gomez D, O'Donohue MF, Wenner T, Douarre C, Macadre J, Koebel P, Giraud-Panis MJ, Kaplan H, Kolkes A, Shin-ya K, Riou JF.
Cancer Res. 2006a Jul 15;66(14):6908-12.
PMID 16849533
 
Telomestatin-induced telomere uncapping is modulated by POT1 through G-overhang extension in HT1080 human tumor cells.
Gomez D, Wenner T, Brassart B, Douarre C, O'Donohue MF, El Khoury V, Shin-Ya K, Morjani H, Trentesaux C, Riou JF.
J Biol Chem. 2006b Dec 15;281(50):38721-9.
PMID 17050546
 
DNA methyltransferases control telomere length and telomere recombination in mammalian cells.
Gonzalo S, Jaco I, Fraga MF, Chen T, Li E, Esteller M, Blasco MA.
Nat Cell Biol. 2006 Apr;8(4):416-24.
PMID 16565708
 
Mammalian sirtuins--emerging roles in physiology, aging, and calorie restriction.
Haigis MC, Guarente LP.
Genes Dev. 2006 Nov 1;20(21):2913-21.
PMID 17079682
 
Loss-of-function mutations in euchromatin histone methyl transferase 1 (EHMT1) cause the 9q34 subtelomeric deletion syndrome.
Kleefstra T, Brunner HG, Amiel J, Oudakker AR, Nillesen WM, Magee A, Genevieve D, Cormier-Daire V, van Esch H, Fryns JP, Hamel BC, Sistermans EA, de Vries BB, van Bokhoven H.
Am J Hum Genet. 2006 Aug;79(2):370-7.
PMID 16826528
 
The Apollo 5' exonuclease functions together with TRF2 to protect telomeres from DNA repair.
Lenain C, Bauwens S, Amiard S, Brunori M, Giraud-Panis MJ, Gilson E.
Curr Biol. 2006 Jul 11;16(13):1303-10.
PMID 16730175
 
Expression of telomere-associated genes as prognostic markers for overall survival in patients with non-small cell lung cancer.
Lin X, Gu J, Lu C, Spitz MR, Wu X.
Clin Cancer Res. 2006 Oct 1;12(19):5720-5.
PMID 17020976
 
Genomic instability and aging-like phenotype in the absence of mammalian SIRT6.
Mostoslavsky R, Chua KF, Lombard DB, Pang WW, Fischer MR, Gellon L, Liu P, Mostoslavsky G, Franco S, Murphy MM, Mills KD, Patel P, Hsu JT, Hong AL, Ford E, Cheng HL, Kennedy C, Nunez N, Bronson R, Frendewey D, Auerbach W, Valenzuela D, Karow M, Hottiger MO, Hursting S, Barrett JC, Guarente L, Mulligan R, Demple B, Yancopoulos GD, Alt FW.
Cell. 2006 Jan 27;124(2):315-29.
PMID 16439206
 
Telomeres and chromosome instability.
Murnane JP.
DNA Repair (Amst). 2006 Sep 8;5(9-10):1082-92.
PMID 16784900
 
TRF2 promotes multidrug resistance in gastric cancer cells.
Ning H, Li T, Zhao L, Li J, Liu J, Liu Z, Fan D.
Cancer Biol Ther. 2006 Aug;5(8):950-6.
PMID 16760674
 
The Wiskott-Aldrich syndrome.
Ochs HD, Thrasher AJ.
J Allergy Clin Immunol. 2006 Apr;117(4):725-38; quiz 39.
PMID 16630926
 
Telomere position effect and silencing of transgenes near telomeres in the mouse.
Pedram M, Sprung CN, Gao Q, Lo AW, Reynolds GE, Murnane JP.
Mol Cell Biol. 2006 Mar;26(5):1865-78.
PMID 16479005
 
G-Quadruplex stabilization by telomestatin induces TRF2 protein dissociation from telomeres and anaphase bridge formation accompanied by loss of the 3' telomeric overhang in cancer cells.
Tahara H, Shin-Ya K, Seimiya H, Yamada H, Tsuruo T, Ide T.
Oncogene. 2006 Mar 23;25(13):1955-66.
PMID 16302000
 
Telomere tethering at the nuclear periphery is essential for efficient DNA double strand break repair in subtelomeric region.
Therizols P, Fairhead C, Cabal GG, Genovesio A, Olivo-Marin JC, Dujon B, Fabre E.
J Cell Biol. 2006 Jan 16;172(2):189-99.
PMID 16418532
 
Apollo, an Artemis-related nuclease, interacts with TRF2 and protects human telomeres in S phase.
van Overbeek M, de Lange T.
Curr Biol. 2006 Jul 11;16(13):1295-302.
PMID 16730176
 
Role of linker histone in chromatin structure and function: H1 stoichiometry and nucleosome repeat length.
Woodcock CL, Skoultchi AI, Fan Y.
Chromosome Res. 2006;14(1):17-25.
PMID 16506093
 
Mosaic ring 20 with no detectable deletion by FISH analysis: Characteristic seizure disorder and literature review.
Zou YS, Van Dyke DL, Thorland EC, Chhabra HS, Michels VV, Keefe JG, Lega MA, Feely MA, Uphoff TS, Jalal SM.
Am J Med Genet A. 2006 Aug 1;140(15):1696-706.
PMID 16835934
 
A topological mechanism for TRF2-enhanced strand invasion.
Amiard S, Doudeau M, Pinte S, Poulet A, Lenain C, Faivre-Moskalenko C, Angelov D, Hug N, Vindigni A, Bouvet P, Paoletti J, Gilson E, Giraud-Panis MJ.
Nat Struct Mol Biol. 2007 Feb;14(2):147-54.
PMID 17220898
 
Telomeric repeat containing RNA and RNA surveillance factors at mammalian chromosome ends.
Azzalin CM, Reichenbach P, Khoriauli L, Giulotto E, Lingner J.
Science. 2007 Nov 2;318(5851):798-801.
PMID 17916692
 
Subtelomeric imbalances in phenotypically normal individuals.
Balikova I, Menten B, de Ravel T, Le Caignec C, Thienpont B, Urbina M, Doco-Fenzy M, de Rademaeker M, Mortier G, Kooy F, van den Ende J, Devriendt K, Fryns JP, Speleman F, Vermeesch JR.
Hum Mutat. 2007 Oct;28(10):958-67.
PMID 17492636
 
The clinical utility of enhanced subtelomeric coverage in array CGH.
Ballif BC, Sulpizio SG, Lloyd RM, Minier SL, Theisen A, Bejjani BA, Shaffer LG.
Am J Med Genet A. 2007a Aug 15;143A(16):1850-7.
PMID 17632771
 
Development of a high-density pericentromeric region BAC clone set for the detection and characterization of small supernumerary marker chromosomes by array CGH.
Ballif BC, Hornor SA, Sulpizio SG, Lloyd RM, Minier SL, Rorem EA, Theisen A, Bejjani BA, Shaffer LG.
Genet Med. 2007b Mar;9(3):150-62.
PMID 17413419
 
Telomere length regulates the epigenetic status of mammalian telomeres and subtelomeres.
Benetti R, Garcia-Cao M, Blasco MA.
Nat Genet. 2007a Feb;39(2):243-50.
PMID 17237781
 
Suv4-20h deficiency results in telomere elongation and derepression of telomere recombination.
Benetti R, Gonzalo S, Jaco I, Schotta G, Klatt P, Jenuwein T, Blasco MA.
J Cell Biol. 2007b Sep 10;178(6):925-36.
PMID 17846168
 
Early replication of short telomeres in budding yeast.
Bianchi A, Shore D.
Cell. 2007 Mar 23;128(6):1051-62.
PMID 17382879
 
Telomere length, stem cells and aging.
Blasco MA.
Nat Chem Biol. 2007a Oct;3(10):640-9.
PMID 17876321
 
The epigenetic regulation of mammalian telomeres.
Blasco MA.
Nat Rev Genet. 2007b Apr;8(4):299-309.
PMID 17363977
 
High-throughput telomere length quantification by FISH and its application to human population studies.
Canela A, Vera E, Klatt P, Blasco MA.
Proc Natl Acad Sci U S A. 2007 Mar 27;104(13):5300-5.
PMID 17369361
 
Sirtuins: the 'magnificent seven', function, metabolism and longevity.
Dali-Youcef N, Lagouge M, Froelich S, Koehl C, Schoonjans K, Auwerx J.
Ann Med. 2007;39(5):335-45.
PMID 17701476
 
Protection of telomeres through independent control of ATM and ATR by TRF2 and POT1.
Denchi EL, de Lange T.
Nature. 2007 Aug 30;448(7157):1068-71.
PMID 17687332
 
Telomere structure and function in trypanosomes: a proposal.
Dreesen O, Li B, Cross GA.
Nat Rev Microbiol. 2007 Jan;5(1):70-5.
PMID 17160000
 
Mutations in the gene encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders.
Durand CM, Betancur C, Boeckers TM, Bockmann J, Chaste P, Fauchereau F, Nygren G, Rastam M, Gillberg IC, Anckarsater H, Sponheim E, Goubran-Botros H, Delorme R, Chabane N, Mouren-Simeoni MC, de Mas P, Bieth E, Roge B, Heron D, Burglen L, Gillberg C, Leboyer M, Bourgeron T.
Nat Genet. 2007 Jan;39(1):25-7.
PMID 17173049
 
Monosomy 1p36 deletion syndrome.
Gajecka M, Mackay KL, Shaffer LG.
Am J Med Genet C Semin Med Genet. 2007 Nov 15;145C(4):346-56.
PMID 17918734
 
Human diseases of telomerase dysfunction: insights into tissue aging.
Garcia CK, Wright WE, Shay JW.
Nucleic Acids Res. 2007;35(22):7406-16.
PMID 17913752
 
Telomere length profiles in humans: all ends are not equal.
Gilson E, Londono-Vallejo A.
Cell Cycle. 2007 Oct 15;6(20):2486-94.
PMID 17726375
 
How telomeres are replicated.
Gilson E, Geli V.
Nat Rev Mol Cell Biol. 2007 Oct;8(10):825-38.
PMID 17885666
 
Interactions of TRF2 with model telomeric ends.
Khan SJ, Yanez G, Seldeen K, Wang H, Lindsay SM, Fletcher TM.
Biochem Biophys Res Commun. 2007 Nov 9;363(1):44-50.
PMID 17850765
 
Cryptic telomere imbalance: a 15-year update.
Ledbetter DH, Martin CL.
Am J Med Genet C Semin Med Genet. 2007 Nov 15;145C(4):327-34.
PMID 17910073
 
Regulation of Drosophila life span by olfaction and food-derived odors.
Libert S, Zwiener J, Chu X, Vanvoorhies W, Roman G, Pletcher SD.
Science. 2007 Feb 23;315(5815):1133-7.
PMID 17272684
 
Human subtelomeric WASH genes encode a new subclass of the WASP family.
Linardopoulou EV, Parghi SS, Friedman C, Osborn GE, Parkhurst SM, Trask BJ.
PLoS Genet. 2007 Dec;3(12):e237.
PMID 18159949
 
Telomerase reverse-transcriptase homozygous mutations in autosomal recessive dyskeratosis congenita and Hoyeraal-Hreidarsson syndrome.
Marrone A, Walne A, Tamary H, Masunari Y, Kirwan M, Beswick R, Vulliamy T, Dokal I.
Blood. 2007 Dec 15;110(13):4198-205.
PMID 17785587
 
Extensive gains and losses of olfactory receptor genes in Mammalian evolution.
Niimura Y, Nei M.
PLoS ONE. 2007;2(1):e708.
PMID 17684554
 
Telomere damage induced by the G-quadruplex ligand RHPS4 has an antitumor effect.
Salvati E, Leonetti C, Rizzo A, Scarsella M, Mottolese M, Galati R, Sperduti I, Stevens MF, D'Incalci M, Blasco M, Chiorino G, Bauwens S, Horard B, Gilson E, Stoppacciaro A, Zupi G, Biroccio A.
J Clin Invest. 2007 Nov;117(11):3236-47.
PMID 17932567
 
Telomere attachment and clustering during meiosis.
Scherthan H.
Cell Mol Life Sci. 2007 Jan;64(2):117-24.
PMID 17219025
 
Use of array CGH in the evaluation of dysmorphology, malformations, developmental delay, and idiopathic mental retardation.
Stankiewicz P, Beaudet AL.
Curr Opin Genet Dev. 2007 Jun;17(3):182-92.
PMID 17467974
 
The WASP-WAVE protein network: connecting the membrane to the cytoskeleton.
Takenawa T, Suetsugu S.
Nat Rev Mol Cell Biol. 2007 Jan;8(1):37-48.
PMID 17183359
 
Adult-onset pulmonary fibrosis caused by mutations in telomerase.
Tsakiri KD, Cronkhite JT, Kuan PJ, Xing C, Raghu G, Weissler JC, Rosenblatt RL, Shay JW, Garcia CK.
Proc Natl Acad Sci U S A. 2007 May 1;104(18):7552-7.
PMID 17460043
 
Regulation of the actin cytoskeleton in cancer cell migration and invasion.
Yamaguchi H, Condeelis J.
Biochim Biophys Acta. 2007 May;1773(5):642-52.
PMID 16926057
 
Telomere dynamics in human cells.
Baird DM.
Biochimie. 2008 Jan;90(1):116-21.
PMID 17854970
 
Terminal 3p deletions: phenotypic variability, chromosomal non-penetrance, or gene modification?
Barber JC.
Am J Med Genet A. 2008 Jul 15;146A(14):1899-901.
PMID 18553547
 
Role of TRF2 in the assembly of telomeric chromatin.
Benetti R, Schoeftner S, Munoz P, Blasco MA.
Cell Cycle. 2008 Nov 1;7(21):3461-8.
PMID 18971622
 
A shared docking motif in TRF1 and TRF2 used for differential recruitment of telomeric proteins.
Chen Y, Yang Y, van Overbeek M, Donigian JR, Baciu P, de Lange T, Lei M.
Science. 2008 Feb 22;319(5866):1092-6.
PMID 18202258
 
Model of human aging: recent findings on Werner's and Hutchinson-Gilford progeria syndromes.
Ding SL, Shen CY.
Clin Interv Aging. 2008;3(3):431-44.
PMID 18982914
 
Recruitment to the nuclear periphery can alter expression of genes in human cells.
Finlan LE, Sproul D, Thomson I, Boyle S, Kerr E, Perry P, Ylstra B, Chubb JR, Bickmore WA.
PLoS Genet. 2008 Mar;4(3):e1000039.
PMID 18369458
 
Domain organization of human chromosomes revealed by mapping of nuclear lamina interactions.
Guelen L, Pagie L, Brasset E, Meuleman W, Faza MB, Talhout W, Eussen BH, de Klein A, Wessels L, de Laat W, van Steensel B.
Nature. 2008 Jun 12;453(7197):948-51.
PMID 18463634
 
Psoriasis is associated with increased beta-defensin genomic copy number.
Hollox EJ, Huffmeier U, Zeeuwen PL, Palla R, Lascorz J, Rodijk-Olthuis D, van de Kerkhof PC, Traupe H, de Jongh G, den Heijer M, Reis A, Armour JA, Schalkwijk J.
Nat Genet. 2008 Jan;40(1):23-5.
PMID 18059266
 
Telomeric RNA enters the game.
Horard B, Gilson E.
Nat Cell Biol. 2008 Feb;10(2):113-5.
PMID 18246034
 
A genetic locus targeted to the nuclear periphery in living cells maintains its transcriptional competence.
Kumaran RI, Spector DL.
J Cell Biol. 2008 Jan 14;180(1):51-65.
PMID 18195101
 
The Rat1p 5' to 3' exonuclease degrades telomeric repeat-containing RNA and promotes telomere elongation in Saccharomyces cerevisiae.
Luke B, Panza A, Redon S, Iglesias N, Li Z, Lingner J.
Mol Cell. 2008 Nov 21;32(4):465-77.
PMID 19026778
 
SIRT6 is a histone H3 lysine 9 deacetylase that modulates telomeric chromatin.
Michishita E, McCord RA, Berber E, Kioi M, Padilla-Nash H, Damian M, Cheung P, Kusumoto R, Kawahara TL, Barrett JC, Chang HY, Bohr VA, Ried T, Gozani O, Chua KF.
Nature. 2008 Mar 27;452(7186):492-6.
PMID 18337721
 
Telomeric position effect: from the yeast paradigm to human pathologies?
Ottaviani A, Gilson E, Magdinier F.
Biochimie. 2008 Jan;90(1):93-107.
PMID 17868970
 
How shelterin protects mammalian telomeres.
Palm W, de Lange T.
Annu Rev Genet. 2008;42:301-34.
PMID 18680434
 
Changes in the expression of telomere maintenance genes suggest global telomere dysfunction in B-chronic lymphocytic leukemia.
Poncet D, Belleville A, t'kint de Roodenbeke C, Roborel de Climens A, Ben Simon E, Merle-Beral H, Callet-Bauchu E, Salles G, Sabatier L, Delic J, Gilson E.
Blood. 2008 Feb 15;111(4):2388-91.
PMID 18077792
 
C. elegans telomeres contain G-strand and C-strand overhangs that are bound by distinct proteins.
Raices M, Verdun RE, Compton SA, Haggblom CI, Griffith JD, Dillin A, Karlseder J.
Cell. 2008 Mar 7;132(5):745-57.
PMID 18329362
 
Transcriptional repression mediated by repositioning of genes to the nuclear lamina.
Reddy KL, Zullo JM, Bertolino E, Singh H.
Nature. 2008 Mar 13;452(7184):243-7.
PMID 18272965
 
Human subtelomeric copy number variations.
Riethman H.
Cytogenet Genome Res. 2008;123(1-4):244-52.
PMID 19287161
 
TINF2, a component of the shelterin telomere protection complex, is mutated in dyskeratosis congenita.
Savage SA, Giri N, Baerlocher GM, Orr N, Lansdorp PM, Alter BP.
Am J Hum Genet. 2008 Feb;82(2):501-9.
PMID 18252230
 
Developmentally regulated transcription of mammalian telomeres by DNA-dependent RNA polymerase II.
Schoeftner S, Blasco MA.
Nat Cell Biol. 2008 Feb;10(2):228-36.
PMID 18157120
 
Identification of chromosome abnormalities in subtelomeric regions by microarray analysis: a study of 5,380 cases.
Shao L, Shaw CA, Lu XY, Sahoo T, Bacino CA, Lalani SR, Stankiewicz P, Yatsenko SA, Li Y, Neill S, Pursley AN, Chinault AC, Patel A, Beaudet AL, Lupski JR, Cheung SW.
Am J Med Genet A. 2008 Sep 1;146A(17):2242-51.
PMID 18663743
 
TINF2 mutations result in very short telomeres: analysis of a large cohort of patients with dyskeratosis congenita and related bone marrow failure syndromes.
Walne AJ, Vulliamy T, Beswick R, Kirwan M, Dokal I.
Blood. 2008 Nov 1;112(9):3594-600.
PMID 18669893
 
Allelic recombination between distinct genomic locations generates copy number diversity in human beta-defensins.
Abu Bakar S, Hollox EJ, Armour JA.
Proc Natl Acad Sci U S A. 2009 Jan 20;106(3):853-8.
PMID 19131514
 
Telomere-associated proteins: cross-talk between telomere maintenance and telomere-lengthening mechanisms.
De Boeck G, Forsyth RG, Praet M, Hogendoorn PC.
J Pathol. 2009 Feb;217(3):327-44.
PMID 19142887
 
Telomere length in Hutchinson-Gilford progeria syndrome.
Decker ML, Chavez E, Vulto I, Lansdorp PM.
Mech Ageing Dev. 2009 Jun;130(6):377-83.
PMID 19428457
 
Dyskeratosis congenita, stem cells and telomeres.
Kirwan M, Dokal I.
Biochim Biophys Acta. 2009 Apr;1792(4):371-9.
PMID 19419704
 
Telomerase activity is associated with an increase in DNA methylation at the proximal subtelomere and a reduction in telomeric transcription.
Ng LJ, Cropley JE, Pickett HA, Reddel RR, Suter CM.
Nucleic Acids Res. 2009 Mar;37(4):1152-9.
PMID 19129228
 
The D4Z4 macrosatellite repeat acts as a CTCF and A-type lamins-dependent insulator in facio-scapulo-humeral dystrophy.
Ottaviani A, Rival-Gervier S, Boussouar A, Foerster AM, Rondier D, Sacconi S, Desnuelle C, Gilson E, Magdinier F.
PLoS Genet. 2009 Feb;5(2):e1000394.
PMID 19247430
 
TRF2 promotes, remodels and protects telomeric Holliday junctions.
Poulet A, Buisson R, Faivre-Moskalenko C, Koelblen M, Amiard S, Montel F, Cuesta-Lopez S, Bornet O, Guerlesquin F, Godet T, Moukhtar J, Argoul F, Declais AC, Lilley DM, Ip SC, West SC, Gilson E, Giraud-Panis MJ.
EMBO J. 2009 Mar 18;28(6):641-51.
PMID 19197240
 
Written2009-06Caroline Schluth-Bolard, Alexandre Ottaviani, Amadou Bah, Amina Boussouar, Eric Gilson, Frédérique Magdinier
de Biologie Moléculaire de la Cellule, CNRS UMR5239, Ecole Normale Supérieure de Lyon, UCBL1, IFR128. 46 allée d'Italie, 69364 Lyon Cedex 07, France

Citation

This paper should be referenced as such :
Schluth-Bolard, C ; Ottaviani, A ; Bah, A ; Boussouar, A ; Gilson, E ; Magdinier, F
Dynamics, plasticity of chromosome ends: consequences in human pathologies
Atlas Genet Cytogenet Oncol Haematol. 2010;14(5):501-524.
Free journal version : [ pdf ]   [ DOI ]
On line version : http://AtlasGeneticsOncology.org/Deep/SubTelomereID20025.htm