Download Telomere maintenance without telomerase

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Zinc finger nuclease wikipedia , lookup

Transposable element wikipedia , lookup

Plasmid wikipedia , lookup

Mitochondrial DNA wikipedia , lookup

DNA damage theory of aging wikipedia , lookup

Molecular cloning wikipedia , lookup

Cell-free fetal DNA wikipedia , lookup

Designer baby wikipedia , lookup

Epigenetic clock wikipedia , lookup

NEDD9 wikipedia , lookup

Gene expression programming wikipedia , lookup

Genomic library wikipedia , lookup

Mutagen wikipedia , lookup

Deoxyribozyme wikipedia , lookup

DNA vaccination wikipedia , lookup

Y chromosome wikipedia , lookup

Genome (book) wikipedia , lookup

Polycomb Group Proteins and Cancer wikipedia , lookup

Genetic engineering wikipedia , lookup

Non-coding DNA wikipedia , lookup

Therapeutic gene modulation wikipedia , lookup

DNA supercoil wikipedia , lookup

Microevolution wikipedia , lookup

History of genetic engineering wikipedia , lookup

Genome editing wikipedia , lookup

Point mutation wikipedia , lookup

Karyotype wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

X-inactivation wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Holliday junction wikipedia , lookup

Extrachromosomal DNA wikipedia , lookup

Microsatellite wikipedia , lookup

Polyploid wikipedia , lookup

Chromosome wikipedia , lookup

Replisome wikipedia , lookup

Neocentromere wikipedia , lookup

Site-specific recombinase technology wikipedia , lookup

No-SCAR (Scarless Cas9 Assisted Recombineering) Genome Editing wikipedia , lookup

Homologous recombination wikipedia , lookup

Helitron (biology) wikipedia , lookup

Cre-Lox recombination wikipedia , lookup

Telomerase wikipedia , lookup

Telomere wikipedia , lookup

Transcript
ã
Oncogene (2002) 21, 522 ± 531
2002 Nature Publishing Group All rights reserved 0950 ± 9232/02 $25.00
www.nature.com/onc
Telomere maintenance without telomerase
Victoria Lundblad*,1
1
Department of Molecular and Human Genetics, Baylor College of Medicine, Houston, Texas, TX 77030, USA
Recombination-dependent maintenance of telomeres, ®rst
discovered in budding yeast, has revealed an alternative
pathway for telomere maintenance that does not require
the enzyme telomerase. Experiments conducted in two
budding yeasts, S. cerevisiae and K. lactis, have shown
recombination can replenish terminal G-rich telomeric
tracts that would otherwise shorten in the absence of
telomerase, as well as disperse and amplify sub-telomeric
repeat elements. Investigation of the genetic requirements
for this process have revealed that at least two di€erent
recombination pathways, de®ned by RAD50 and
RAD51, can promote telomere maintenance. Although
critically short telomeres are very recombinogenic,
recombination among telomeres that have only partially
shortened in the absence of telomerase can also
contribute to telomerase-independent survival. These
observations provide new insights into the mechanism(s)
by which recombination can restore telomere function in
yeast, and suggest future experiments for the investigation of potentially similar pathways in human cells.
Oncogene (2002) 21, 522 ± 531. DOI: 10.1038/sj/onc/
1205079
Keywords: telomerase; EST1; ALT; recombination;
break-induced replication
Linear chromosomes must have a mechanism to ensure
that their termini are fully duplicated when the
chromosome itself is replicated. For most eukaryotes,
the ends of nuclear chromosomes are maintained by
the reverse transcriptase telomerase. Telomerase adds
species-speci®c telomeric repeats onto chromosome
termini, thereby counterbalancing sequence loss that
would ensue as a result of semi-conservative DNA
replication by the conventional DNA replication
machinery (for reviews, see Greider, 1996; McEachern
et al., 2000). This not only ensures complete replication
of the genome, but the telomeric repeats that are
synthesized by telomerase provide a binding site for the
host of telomere-speci®c proteins that prevent recognition of chromosome termini as double strand breaks
(for reviews, see Lundblad, 2000; Blackburn, 2001).
However, increasing interest has been focused on a
recombination-dependent mechanism for replenishing
telomeric repeats. This pathway, ®rst discovered in
budding yeast, is receiving attention due to the
*Correspondence: V Lundblad; E-mail: [email protected]
possibility that a potentially similar pathway operating
in human cells, referred to as ALT (for Alternative
Telomere Lengthening), might contribute to neoplastic
transformation. This review discusses the investigations
in budding yeast that have helped elucidate this
recombination-dependent process, followed by a discussion of the possible mechanisms that have been
proposed as a result of the yeast studies. A brief
summary of the ALT pathway, and a discussion about
whether ALT proceeds through a similar recombination-dependent pathway, concludes this review. The
reader is also referred to other excellent reviews on this
topic (Biessmann and Mason, 1997; Kass-Eisler and
Greider, 2000; Reddel et al., 2001; Kraus et al., 2001).
Discovery of a telomerase-independent pathway for
telomere maintenance
Several observations during the 1980's provided early
clues that recombination could potentially maintain
telomeric repeats at chromosome ends. Linear plasmids
that lacked a functional telomere, but contained
homology to sub-telomeric regions, could acquire a
functional chromosome terminus in a homologydependent process (Dunn et al., 1984). Rescue of these
linear plasmids was proposed to involve strand
invasion of the sub-telomeric region of a chromosome
by the terminus of the linear plasmid, followed by
replicative copying of the donor onto the end of the
plasmid (Figure 1a). The concept that a genetic
recombination intermediate could be converted into a
replication fork had already been proposed as a
mechanism for replication of the ends of the linear
bacteriophage T4 genome (Luder and Mosig, 1982).
Formosa and Alberts (1986) provided biochemical
support for the contribution of recombination proteins
to the formation of a replication fork, and further
suggested that homology-initiated DNA replication
might also apply to the telomere end replication
problem. Similar recombination-based models for
telomere maintenance had also been elaborated by
Bernards et al. (1983) and Walmsley et al. (1984).
Direct demonstration that recombination could maintain the natural termini of linear chromosomes, however,
awaited the analysis of telomerase-defective yeast
strains. The discovery of the S. cerevisiae EST1 gene
(Lundblad and Szostak, 1989), later shown to encode a
subunit of the telomerase holoenzyme (Hughes et al.,
2000), allowed examination of the consequences of a
Telomere maintenance without telomerase
V Lundblad
critically short
telomere
linear plasmid
resection
resection
3’
3’
strand exchange
strand exchange
extension of
invading strand
extension of
invading strand
second strand
synthesis
second strand
synthesis
final products
final products
linear healed plasmid
Figure 1 Two alternative mechanisms by which break-induced
replication can capture a telomere. (a) A linear plasmid that
contains homology to the sub-telomeric Y' repeat can initiate
strand invasion at a chromosomal Y' element. Subsequent
replication results in the nonreciprocal translocation of intact Y'
and terminal G-rich telomeric repeats onto the end of the linear
plasmid. (b) Strand invasion to initiate break-induced replication
can occur between a critically short telomere that still retains
telomeric repeat sequences and another terminus, thereby
elongating the shortened telomere. Staggered alignment of the
invading and donor strands (not depicted here) could also result
in termini that are longer than chromosome ends maintained by
telomerase
telomere replication defect in an easily experimentally
manipulated organism. A haploid yeast strain deleted for
the est1-D gene exhibited gradual telomere shortening
and an accompanying progressive decline in viability,
dubbed yeast cellular senescence. This established an
experimental paradigm for the relationship between
telomere replication and limits on replicative capacity,
which was recapitulated in human cells several years ago
(Bodnar et al., 1998). In addition, the rate of loss of a
non-essential reporter chromosome increased, in parallel
with the appearance of est17 senescence (Lundblad and
Szostak, 1989). Since chromosome loss is the most
probable outcome of a DNA double strand break
(Kramer and Haber, 1993; Sandell and Zakian, 1993),
this suggested that cell death in an est1-D cell occurs
when telomeres shorten suciently that the ends of
chromosomes resemble double strand breaks. A more
recent study has expanded the potential lethal events that
occur in an est1-D cell (Hackett et al., 2001). Their
analysis indicates that cell death may result from loss of
essential genes (as a result of extensive terminal
deletions) and/or chromosome transmission failures
resulting from end-to-end fusions that trigger a breakage-fusion-bridge cycle.
Although the majority of yeast cells die in the absence
of telomerase, Lundblad and Blackburn (1993) observed
that a small subpopulation could escape the lethal
consequences of a telomerase defect. The rare survivors
recovered from an est1-D strain not only regained the
ability to grow but also displayed dramatic changes at
their chromosomal termini, due to global ampli®cation
of both telomeric and sub-telomeric repeat sequences.
These extensive rearrangements arose as a result of
recombination, as the appearance of survivors was
blocked if the est1-D strain was also defective for
RAD52, which is responsible for the majority of
homologous recombination events in yeast. Survivors
could be grouped into two general classes based on the
pattern of ampli®cation of telomeric and/or subtelomeric sequences. Although survivors were recovered
as healthy growing clones arising from a culture of
mostly inviable cells, continued single colony propagation of both classes of survivors revealed variable growth
patterns: sub-clones could be isolated in which a
senescence phenotype re-appeared, although survivors
readily reappeared again from these senescing subclones. Highly dynamic changes in telomere structure,
with sequences rapidly lost or acquired at individual
telomeres of survivor strains, was also observed,
providing a possible explanation for the growth
variability. Finally, neither est17 strains early in
senescence nor subsequently isolated survivors displayed
a genome-wide hyper-recombination phenotype. However, this analysis did not exclude the possibility that in a
cell that lacked telomerase, unreplicated telomeres could
themselves become hyper-recombinogenic, a point that
will be re-addressed later in this review.
Although all survivors recovered from an est1-D
strain shared common features, such as RAD52dependence and highly dynamic changes at their
telomeres, they could be grouped into two classes
based on notable di€erences in their telomeric
structure. One class of survivors was distinguished by
extensive ampli®cation of sub-telomeric repeats (called
Y' elements). In most wild type strains of S. cerevisiae,
approximately 2/3 of the telomeres contain tandem
arrays of one to four Y' elements, separated by short
stretches of *50 to 100 bp of G-rich telomeric repeats
(Figure 2a). In a telomerase-pro®cient strain, exchanges
of Y' elements between telomeres had been shown to
occur at low frequency, resulting in the occasional
dispersal of these repeats among chromosome ends
(Horowitz and Haber, 1985). However, in one class of
survivors, Y' copy number increased dramatically
(5200-fold in some survivor strains), forming extended
tandem arrays at many telomeres (Figure 2a). In
addition, telomeres that did not have any Y' elements
in the pre-survivor est17 strain acquired Y' elements.
Sub-telomeric repeat ampli®cation was also accompanied by a correspondingly substantial increase in the
total amount of telomeric DNA (due to the short
523
Oncogene
Telomere maintenance without telomerase
V Lundblad
524
further suggested that conversion to a telomeraseindependent survivor was a multi-step process that
involved multiple rounds of recombination, occurring
during the growth of an est17 culture, with each
bene®cial recombination event providing a slight
selective advantage. This model implied that there
was a gradient of telomere function, such that any
reduction in telomere length could reduce the replicative capacity of the est17 culture, and any recombination event that restored function to even a single
telomere could contribute to improved growth. Subsequent work showed that a strain with a partially
defective telomerase, resulting in stably short telomeres
but no senescence, had a severe growth disadvantage
relative to a telomerase-pro®cient strain, as assessed by
liquid competition experiments (Morris and Lundblad,
1997). Alternative models, which propose that critically
short telomeres are primarily responsible for triggering
senescence and serving as substrates for recombination,
are considered later in this review.
Figure 2 Telomeric structure in telomerase-defective survivors.
(a) S. cerevisiae chromosome ends consist of telomeres that have
one to four copies of the sub-telomeric repeat element Y' (which
are bracketed by short stretches of telomeric repeats) as well as
telomeres that lack Y' elements. One class of telomerase-defective
survivors (type I) is characterized by extensive ampli®cation and
dispersal of Y' elements to most telomeres, although the terminal
G-rich tract remains short. In a second class of survivors (type
II), the G-rich terminal tract is itself elongated by recombination.
(b) Survivors of a telomerase defect in K. lactis exhibit extended
telomeres, similar to the type II survivors of S. cerevisiae. Note
that K. lactis telomeres do not have sub-telomeric repeats
bracketed by telomeric repeats
stretches of telomeric tracts bracketing each Y'
element), with up to 4% of the genome composed of
telomeric repeat DNA in some survivors. However,
despite the overall increase in telomeric DNA, the
terminal G-rich repeat tract remained short (and did
not appear to undergo further shortening) in this class
of survivors.
This analysis also identi®ed a second class of est17
survivors, with a distinctively di€erent telomeric restriction pattern: survivors in this group were characterized
by changes in the multiple sizes of telomeric restriction
bands, relative to the parental strain. Survivors in this
second class did not exhibit extensive sub-telomeric
repeat ampli®cation, although many, if not all, telomeres
appeared to acquire at least partial Y' sequences. The full
characterization of the nature of the telomeric changes
occurring in this class of survivors came from subsequent
work by Teng and Zakian (1999), discussed in more
detail in the next section.
Based on the above observations, Lundblad and
Blackburn (1993) proposed that, in the absence of
telomerase, recombination could maintain G-rich
telomeric repeats and restore telomere function. We
Oncogene
Figure 3 Models for recombination-mediated exchanges between telomeres. (a) Recombination between the terminal G-rich
tract of a telomere that lacks Y' elements and the internal G-rich
tract of a Y'-containing telomere results in conversion of the nonY'-containing telomere to one that now has both sub-telomeric
repeats and additional telomeric G-rich repeats. (b) Integration of
an extrachromosomal Y' element restores a terminally short
telomere, and also converts it to a Y'-containing telomere. (c)
Exchanges between G-rich terminal tracts, analogous to the
nonreciprocal translocation in part A, could also maintain
telomeres
Telomere maintenance without telomerase
V Lundblad
For the class of survivors that displayed Y'
ampli®cation, Y' elements were proposed to be
acquired by either a nonreciprocal translocation event
(Figure 3a), or by integration of extrachromosomal Y'
circles into a critically short telomere (Figure 3b). The
latter model was based on the demonstration that Y'
elements, which contained a weak origin of replication
(Chan and Tye, 1983), could be propagated as
extrachromosomal plasmids (Horowitz and Haber,
1985). Expansion of sub-telomeric Y' elements to
multiple telomeres in this class of est17 survivors
would have two consequences. First, this would
increase the extent of homology between telomeres,
thereby enhancing subsequent recombination events
and continuing to drive the process. In addition, the
ampli®ed arrays of Y'-associated telomeric repeats
would increase the amount of G-rich telomeric repeat
DNA present at each chromosome, which provides a
reservoir of G-rich telomeric repeat DNA for recombination onto the extreme terminus of chromosomes.
Recombination between terminal G-rich telomeric
repeats, without Y' ampli®cation, also appeared to be
capable of maintaining the terminal G-rich tract
(Figure 3c). This was inferred from the observation
that even survivors that exhibited extensive ampli®cation of Y' elements still possessed individual telomeres
composed only of G-rich repeats, with no newly
acquired Y' element. These G-rich tracts therefore
appeared to be maintained by telomere ± telomere
recombination. Thus, the key feature of the overall
model was that the acquisition of G-rich telomeric
repeats through recombination was responsible for
restoring telomere function and viability.
Subsequent work from several laboratories has
shown that the appearance of survivors is not speci®c
to yeast strains that are defective for EST1 function.
Yeast strains lacking any telomerase subunit (Est1p,
Est2p, Est3p or TLC1), or defective in the telomerase
recruitment function of Cdc13p, all produce survivors
in a recombination-dependent manner that is indistinguishable from that observed for est17 survivors
(Singer and Gottschling, 1994; Lendvay et al., 1996;
Teng and Zakian, 1999).
Recombination-mediated telomere maintenance can
elongate terminal telomeric G-rich repeats
Analysis of the consequences of a telomerase defect in a
related budding yeast, Kluveromyces lactis, revealed
additional insights regarding how recombination could
maintain telomeres. In K. lactis, loss of telomerase
function had the same consequences as in S. cerevisiae:
gradual attrition of telomeric repeats, accompanied by a
senescence phenotype (McEachern and Blackburn,
1995). Survivors could also be recovered from senescent
cultures of telomerase-de®cient K. lactis strains, through
a mechanism that was again dependent on RAD52
function (McEachern and Blackburn, 1996). However,
analysis of telomere structure in these survivors uncovered some notable di€erences. Unlike S. cerevisiae, K.
lactis telomeres do not have tandem arrays of subtelomeric repeats interspersed with G-rich telomeric
repeat sequences (note that, unlike telomeric repeats,
sub-telomeric repeat elements show little conservation
from species to species). The absence of sub-telomeric
repeats meant that a class of survivors analogous to the
S. cerevisiae survivors that exhibited extensive Y'
ampli®cation was not recovered in K. lactis. Instead,
the terminal G-rich telomeric tract was greatly elongated
in telomerase-defective K. lactis survivors (Figure 2b).
Di€erent telomeres in post-senescent survivors exhibited
variable lengths, often with tract lengths longer than
telomeres of a telomerase-pro®cient strain. However,
these elongated telomeres still exhibited the same gradual
shortening observed during the initial senescence of a
newly generated telomerase-defective strain, suggesting
that the recombination mechanism that was replenishing
telomeric repeats was not as ecient as that provided by
telomerase. As with S. cerevisiae survivors, K. lactis
survivors also displayed an unstable growth phenotype,
manifested as secondary episodes of transient senescence.
These observations led McEachern and Blackburn
(1996) to present a model for telomere repeat array
lengthening (Figure 1b) similar to that proposed by
Dunn et al. (1984) for the healing of linear plasmids
(Figure 1a). They proposed that the telomere shortening that occurs in the absence of telomerase results in
loss of the essential `capping' function of a telomere, so
that chromosome ends are now perceived as double
strand breaks. Processing of these exposed termini by
DNA repair enzymes could produce a 3' single
stranded overhang that could invade another chromosomal telomere. Extension of this invading strand,
followed by second strand synthesis, would result in a
non-reciprocal translocation, with sequences from the
donor chromosome transferred to the `uncapped'
chromosome. These gene conversion events could be
the result of invasion of a 3' strand that still retains
homology with a telomeric repeat tract on another
chromosome; staggered alignment of the invading and
donor strands would generate a telomeric tract longer
than that observed in telomerase-plus cells.
Although the primary structural change observed in
K. lactis survivors was extension of the telomeric tract
array, gene conversion events in sub-telomeric regions
of K. lactis survivors were also observed. Since 11 of
the 12 K. lactis chromosome termini share homology in
the sub-telomeric region, this indicates that strand
invasions could initiate well into sub-telomeric regions.
However, McEachern and Blackburn pointed out that
the majority of the recombination events that they
observed must depend on retention of at least a few
telomeric repeats at the time that chromosome termini
undergo recombination-mediated elongation.
Through a careful molecular analysis, Teng and
Zakian (1999) showed that the second class of S.
cerevisiae survivors identi®ed by Lundblad and Blackburn (1993) exhibited many of the properties displayed
by survivors recovered from K. lactis. Telomeres in this
second class of S. cerevisiae survivors, which they re-
525
Oncogene
Telomere maintenance without telomerase
V Lundblad
526
named type II survivors (while survivors with extensive
Y' ampli®cation were labeled type I survivors), were
shown to terminate with very long, heterogeneous
tracts of G-rich telomeric repeats (Figure 2a). As had
been previously observed in K. lactis, these extended
telomeric tracts were not stably maintained: they
displayed the slow attrition of a telomerase-defective
strain, as well as more abrupt changes that were
reminiscent of previous observations showing that
elongated telomeric tracts could undergo rapid truncation (Li and Lustig, 1996; Bucholc et al., 2001). Thus,
although type I and type II survivors both retain
terminal G-rich repeat tracts, terminal sequences are
greatly elongated in type II survivors. Notably, the
short tracts that are retained on the telomeres of type I
survivors are apparently more stable than the continually shortening type II telomeric tracts (although it
cannot be ruled out that additional shortening of a
type I telomere might be lethal and thus would not
contribute to the signal on a telomeric Southern blot).
Teng and Zakian also studied in detail the growth
properties of the two classes of survivors. In their strain
background, all type II survivors exhibited a stable
growth phenotype, and when propagated in culture, type
II survivors also had a selective advantage over type I
survivors. The growth disadvantage of type I survivors
might be due ± at least in part ± to the burden of
replicating a genome increased by up to 10%, due to the
extensive ampli®cation of Y' and telomeric repeat
sequences. However, type I survivors also showed a high
propensity for conversion to type II survivors, suggesting
that telomeres in the absence of telomerase may be
substrates for several di€erent types of recombination
events. The eventual outcome may be dictated by a
balance between the susceptibility of an under-replicated
telomere to di€erent types of recombination events,
versus the selective advantage of di€erent recombinant
telomeric structures. In other words, telomeres in budding
yeast may be more recombinogenic for the types of events
that result in type I survivors, but the selective growth
advantages of type II survivors allows this class of
survivors to dominate. This may have implications for
the analysis of potential intermediates for similar
telomere maintenance pathways in human cells.
senescence is observed (McEachern and Iyer, 2001;
Rizki and Lundblad, 2001). However, additional
observations indicate critically short telomeres may
also be preferred substrates for at least one type of
recombination (Teng et al., 2000). Thus, both variations of this model are likely to contribute to
telomerase-independent telomere maintenance.
An initial clue that recombination-mediated exchanges were important for cell survival even early in
the propagation of a telomerase-de®cient strain came
from observations of the consequence of loss of
RAD52 function. A rad527 mutation not only blocked
the appearance of survivors, but in addition, an est1-D
rad52-D strain also displayed a greatly accelerated
senescence phenotype (Lundblad and Blackburn, 1993).
Whereas an est1-D strain at the 25 generation time
point was indistinguishable from an EST1 strain, an
est1-D rad52-D strain displayed an obvious growth
defect by 25 generations and was completely inviable
by *50 generations (depicted schematically in Figure
4). This strongly suggested that recombination was
contributing to telomere function even in a newly
generated telomerase-defective strain, during the initial
period of telomere shortening.
As a reciprocal test of this premise, Rizki and
Lundblad (2001) examined the consequences of genetic
alterations that increased the frequency of genetic
exchanges between telomeric repeat DNA. This work
showed that defects in mismatch repair ± which relieve
a block to recombination between imperfectly matched
sequences ± enhanced the proliferative capacity of
telomerase-defective strains in both S. cerevisiae and K.
lactis. Using liquid competition assays, the e€ects of a
mismatch repair mutation were detected during the
very early stages of growth of a telomerase-defective
yeast strain, even before a senescence phenotype was
evident (Figure 4). Increased survival rates were shown
to be due to elimination of the anti-recombination
activity of the mismatch repair pathway, rather than an
indirect e€ect of increased mutation rates. Notably,
enhancement of survival of telomerase-defective K.
Do critically short telomeres promote the
recombination-dependent pathway?
The original model by Lundblad and Blackburn
proposed a stochastic process of recombination events
at telomeres, with selection pressure for viability
ensuring that there would be a cumulative acquisition
of telomeric sequences. A variant of this model is that
telomeres become more recombinogenic in the absence
of telomerase. A further extension of the model is that
recombination occurs in a burst, triggered by critically
short telomeres, rather than a succession of cumulative
exchanges. Recent studies from several di€erent
laboratories suggest that short telomeres are in fact
recombinogenic, even before critically shortening or
Oncogene
Figure 4 Schematic representation of the relative proliferative
capacity of a telomerase-defective strain (est27), a telomerasedefective strain that is also eliminated for the major pathway for
recombination (est27 rad527) or a telomerase-defective strain in
which homeologous recombination between chromosome ends is
increased, due to a defect in mismatch repair (est27 msh27)
Telomere maintenance without telomerase
V Lundblad
lactis cells was also dependent on elimination of
mismatch repair. Since K. lactis telomeres are composed of perfect telomeric G-rich repeats, this strongly
suggested that recombination between partially mismatched sub-telomeric regions could contribute to
survival in the absence of telomerase.
More direct support for the potential contribution of
recombination between sub-telomeric repeats to telomerase-independent survival came from studies by
McEachern and Iyer (2001). Recombination between
telomeres was monitored by following phenotypically
silent changes in the telomeric repeat tract or the
URA3 gene placed in the sub-telomeric region of one
telomere. In a telomerase-defective strain, recombination-mediated exchanges were greatly elevated, by up
to 1000-fold. Notably, the URA3 gene exhibited rapid
dispersal to most, or even all, of the 12 telomeres,
reminiscent of the ampli®cation of Y' elements
observed in S. cerevisiae survivors (although note that
in these experiments, URA3 was not bracketed by
telomeric repeats). Even in cells in which telomerase
was still partially active (i.e. telomeres were shortened,
but not suciently to confer a senescence phenotype),
gene conversion rates were increased by 100 to 200fold, again arguing that recombination at telomeres
can be substantially enhanced before chromosome ends
become terminally short. The authors proposed that
the enhanced recombination rates of stably short
telomere mutants re¯ects a weakened ability of a short
telomere to provide a telomeric `cap' that protects
chromosome ends from DNA repair activities that
normally act on double strand breaks.
However, a strikingly di€erent conclusion was reached
when the telomeric changes occurring during the
appearance of type II survivors were analysed in detail
(Teng et al., 2000). Monitoring of telomere length at
multiple time points during the outgrowth of a tlc17
strain in culture showed that telomeres gradually
shortened during the pre-survivor period, as previously
observed for a telomerase defect, and then abruptly
lengthened; this correlated with a sudden switch to an
improved growth rate of the culture. When an
individually marked telomere in an already established
type II survivor was monitored, the previously observed
continuous shortening of type II telomeres was interspersed with abrupt elongation events, increasing in size
in 1 to 2 kb segments of presumably telomeric repeat
tract DNA. Based on these observations, Zakian and
colleagues argued against an incremental process of
telomere lengthening and instead proposed that a onestep event was responsible. They speculated that the
substrate for this event could be extra-chromosomal
telomeric repeat DNA, generated by intra-chromatid
recombination.
One simple way to reconcile these two sets of
observations is to consider a model in which there is a
graded increase in recombination that correlates with the
severity of the telomeric length defect. In other words,
shortened telomeres exhibit enhanced susceptibility to
recombination-mediated exchanges, but critically short
telomeres are even more recombinogenic. Notably, in the
study by McEachern and Iyer (2001), greatly elongated
telomeric tracts were not observed in short telomere
strains with partial loss of telomerase function, despite
the increased telomeric gene conversion rate. This
contrasts with the greatly elongated telomeric tracts
displayed by survivors of a telomerase null mutation
(McEachern and Blackburn, 1996). However, one
alternative to a graded recombination response is that
partially uncapped telomeres are healed by a di€erent
recombination mechanism than that utilized by critically
short telomeres.
527
A summary of potential molecular mechanisms
Work from several di€erent labs in the last several years
has expanded our understanding of the genetic requirements for telomerase-independent telomere maintenance
in yeast. The ®rst step in this direction came from
collaborative work from the Greider and Haber
laboratories that investigated the e€ects of a large set
of DNA recombination mutations on the ability of a
tlc17 strain to form survivors (Le et al., 1999). Their
results demonstrated that at least two di€erent recombination pathways, de®ned by RAD50 and RAD51, were
involved in telomerase-independent survival. Both
rad507 tlc17 and rad517 tlc17 strains gave rise to
survivors, but a triply mutant rad507 rad517 tlc17 strain
was as defective as a rad527 tlc17 strain at generating
post-senescent survivors. These two pathways correlated
with the two classes of survivors recovered from S.
cerevisiae: type I survivors required RAD51 function
whereas type II survivors were dependent on RAD50
function (Teng et al., 2000; Chen et al., 2001). However,
the correlation between recombination pathway and
survivor class may not be absolute, as type II survivors
could form in a rad507 mutant background, at least at
low frequency, but could not be stably maintained (Chen
et al., 2001). Additional analysis has shown that other
members of the RAD51 epistasis group (RAD54, RAD55
and RAD57) were also required for the appearance of
type I survivors (Le et al., 1999; Chen et al., 2001). Based
both on information about genetic requirements and
observations of the structural alterations that occur at
telomeres, four general models have been proposed for
how recombination can maintain telomeres. The next
four sections discuss these models.
Break-induced replication (BIR)
BIR is a one-ended recombination event that initiates a
replication fork (Figure 1). Strand invasion of an intact
chromosome by the broken end (via regions of
homology shared between the reporter and donor
chromosomes) is followed by replicative copying of
donor sequences. If replication continues to the end of
the donor chromosome, this will result in the
nonreciprocal transfer of chromosomal DNA ± as
well as terminal telomeric DNA sequences ± to the site
of the double strand break. Thus, through BIR, a
Oncogene
Telomere maintenance without telomerase
V Lundblad
528
Oncogene
broken chromosome can be healed by capturing a
telomere from the donor (Bosco and Haber, 1998).
Note that a key feature of this replicative mechanism is
that the broken chromosome is not repaired at the
expense of the donor chromosome, as would be the
case if telomere capture occurred as the result of a
nonreciprocal crossover. For a more complete discussion of BIR, see Kraus et al. (2001).
Direct demonstration that BIR can be utilized as a
response to a telomerase defect has come from the
recent work by Greider and colleagues (Hackett et al.,
2001). Their analysis of the molecular events occurring
in late generation est17 cultures showed that not only
telomeric repeat DNA was lost: much more substantial
terminal deletions, extending for many kilobases, also
occurred. If left unrepaired, these extensive terminal
deletions would presumably lead to lethality, either due
to the loss of essential genes or deleterious e€ects on
chromosome transmission. However, these deletions
could be healed by break-induced replication, via
exchanges with other chromosomes through short
stretches of homology shared by the donor and
terminally deleted chromosomes. This indicates that a
primary mechanism for the generation of telomeraseindependent survivors relies on BIR. Although the BIR
events recovered by Hackett et al. (2001) involved
exchanges between non-homologous chromosomes
(because of their experimental design, which dictated
what they recovered), healing by BIR can also occur
through intrachromosomal translocations (Bosco and
Haber, 1998) or unequal sister chromatid exchange.
Multiple studies by the Haber laboratory, using HOinduced double strand breaks, have de®ned the genetic
requirements for BIR, as well as enhancer sites that
can facilitate BIR events (Signon et al., 2001; Malkova
et al., 2001; Kraus et al., 2001). These studies have
demonstrated that a RAD51-independent pathway can
promote BIR, using non-random sites to initiate the
process. However, RAD51-dependent pathway(s) for
BIR are also available (discussed in Kraus et al., 2001).
This suggests that survivors that exhibit extensive
ampli®cation of sub-telomeric elements (type I)
proceed by a RAD51-dependent BIR pathway (Figure
1a), whereas survivors that expand the terminal
telomeric tract (type II) are promoted by a RAD51independent, RAD50-dependent process (Figure 1b).
Thus, both classes of survivors may establish and/or
maintain their telomeres through BIR.
However, one genetic observation about the requirements for BIR that does not correlate well with the
requirements of the survivor pathway is the role of the
Sgs1p helicase. Although Sgs1p is not required for
conventional BIR (Signon et al., 2001), the appearance
of type II survivors is eliminated in telomerasedefective strains that also lack SGS1 (Huang et al.,
2001; Cohen and Sinclair, 2001; Johnson et al., 2001).
One possible explanation, proposed by Louis and
colleagues (Huang, et al. 2001) for the telomere-speci®c
SGS1 requirement may be due to a need to disrupt Gquartets and other secondary structures that singlestranded G-rich telomeric DNA can form.
Integration of extrachromosomal DNA
Integration of extra chromosomal DNA could provide an
alternative mechanism for sudden alterations in the size of
telomeric and sub-telomeric tracts, via excision and
reintegration of chromosomal DNA, as an interchromosomal recombination process (Figure 3b). If the excised
DNA also contains an origin of replication, as has been
demonstrated for Y' elements (Chan and Tye, 1983), this
provides a rapid means of expansion and dispersal of both
sub-telomeric and telomeric DNA. This may provide
either an alternative, or even an additional, route to the
production or maintenance of type I survivors, in
addition to the BIR pathway discussed above. Reintegration by recombination between telomeric DNA present
on the Y' circle and terminal telomeric tracts on the
recipient chromosome, as depicted in Figure 1b, would
also provide a means of replenishing the chromosome
terminus. Notably, in an S. cerevisiae strain that has few
Y' elements, with no tandem arrays or bracketing internal
telomeric repeat tracts, type I survivors are rare (Huang et
al., 2001), possibly due to the inability to generate
extrachromosomal Y' circles.
Whether excised DNA consisting of simple telomeric
DNA is also capable of participating in recombinationmediated reintegration events is unknown. However,
Lustig and colleagues (Bucholc et al., 2001) have
proposed a model to explain rapid shortening of
elongated telomeric repeat tracts, termed TRD (Telomeric Rapid Deletion), such that a potential product of
this model is excised telomeric DNA. Since rapid
truncations are observed in type II survivors (Teng and
Zakian, 1999), this argues that TRD is probably
operative in type II survivors strains. If reintegration of
telomeric DNA can contribute to survivor formation,
this may be a mechanism employed by type II survivors,
possibly as a maintenance mechanism of this survivor
state once formed. Intriguingly, human ALT cells, which
appear to maintain telomeres by a non-telomerase
mechanism, contain a novel variant of promyelocytic
Figure 5 Rolling circle replication (a) or t-loop mediated
extension (b) provide two mechanisms for rapid expansion of
the G-rich tract of a telomere
Telomere maintenance without telomerase
V Lundblad
leukemia (PML) bodies that is composed of telomeric
DNA, as well as telomere binding proteins and proteins
involved in DNA repair and replication (including
Rad51p and Rad52p) (Yeager et al., 1999). However,
whether ALT-associated PML bodies are an obligatory
intermediate in either the establishment or maintenance
steps of the ALT pathway, rather than a byproduct of
elongated telomeres, has not yet been resolved.
Rolling circle replication
Rolling circle DNA replication has received increasing
attention as a potential mechanism for rapid expansion
of the G-rich telomeric tract (Figure 5a). Invasion of a
telomeric tract by an extrachromosomal telomeric
circle could potentially establish a replication fork that
initiates rolling circle replication, thereby allowing
extensive elongation of the telomeric tract. Such a
model is particularly appealing when considering
mechanisms that would promote the abrupt elongation
observed during the initial appearance of type II
survivors. As pointed out by Teng et al. (2000),
conventional gene conversion events between critically
short telomeres does not provide an easy explanation
for the sudden elongation observed at the termini of
type II telomeres, whereas rolling circle replication
could provide one potential route for rapid elongation.
Rolling circle replication has been well characterized
as a replicative mechanism used by small genomes such
as single-stranded bacteriophages and bacterial plasmids (for a review, see Novick, 1998). These replicons
depend on speci®c initiator proteins which both melt
open and nick the initiator region, thereby providing a
3' OH that can serve as a primer for the initiation of
DNA synthesis. If rolling circle replication does in fact
occur at telomeres, the 3' terminus of the G-rich strand
of the telomere could provide priming activity.
However, this model leaves open many questions, such
as what activities are required to initiate the process
(which involves melting of both the duplex telomeric
tract and the invading duplex extrachromosomal DNA,
regulation of assembly of the replisome, etc.). In
addition, proving that recombination-mediated elongation is templated by a single extrachromosomal
telomeric DNA circle (Figure 5a), as opposed to
integration of multiple extrachromosomal circles
(Figure 3b), will be experimentally challenging.
Elongation of a t-loop
t-loops provide an alternative to the rolling circle
replication model (Figure 5b). t-loops are intrachromosomal structures that are formed when the 3' single-strand
G-rich extension at the end of the chromosome invades
the duplex telomeric region (Grith et al., 1999; MunozJordan et al., 2001). These structures have been proposed
to provide an architectural means of protecting chromosome ends from degradative and/or DNA repair activities, by masking the telomeric single strand terminus. In
addition, such structures potentially regulate elongation
by telomerase, since a sequestered 3' terminus would not
be accessible to telomerase. The 3' terminus of the
invading G-rich telomeric strand could also potentially
initiate DNA replication, thereby providing a telomeraseindependent means for elongation of this terminus.
Formation of t-loops does in fact resemble a recombination intermediate, and bears substantial resemblance to
the interchromosomal BIR process depicted in Figure 1b.
Furthermore, Rad50p (along with other components of
the Rad50/Mre11/Nbs1 protein complex) have been
invoked as players in either the formation and/or
maintenance of t-loop structures (Zhu et al., 2000),
suggesting that t-loops might be the initiating structure
for type II survivor formation. However, although the
structure of a t-loop is certainly suggestive of a replicative
model for extension of the chromosome terminus, this
presents somewhat of a paradox: t-loops have been
proposed as a mechanism to provide chromosome end
protection (Grith et al., 1999), whereas recombinationmediated telomere maintenance is thought to occur in
response to a loss of end protection.
Although t-loops have not been identi®ed so far in
yeast, detection of these telomeric structures in
evolutionary unrelated organisms such as mammals,
ciliates and trypanosomes suggests that this is a
conserved feature of eukaryotic telomeres. However,
although trypanosome t-loops are small (*1 kb)
relative to the average size of mammalian t-loops, this
still substantially exceeds the size of telomeric tracts in
yeast, particularly the even shorter tracts that occur
when telomerase is lacking. Therefore, a potential role
for t-loops in rescuing yeast senescence and promoting
survivors remains speculative at this time.
529
Potential parallels to an ALT (Alternative Telomere
Lengthening) pathway in human cells
As mentioned at the start of this review, interest in the
yeast survivor pathway has been enhanced by the
possibility that recombination may contribute to telomere maintenance in human cells as well. If so, this may
have particular signi®cance to processes that promote
neoplasia. In fact, although most immortalized human
cells, including tumor cells, have substantially upregulated telomerase (Shay and Bacchetti, 1997), a subset
of immortalized cells do not display high levels of
telomerase activity (Bryan et al., 1995, 1997). Furthermore, telomeres in these cells show a strikingly di€erent
pro®le, displaying heterogeneous telomeres that range
from very short to lengths that are elongated by up to
50 kb. The absence of detectable telomerase enzyme
activity, coupled with the striking telomeric pro®le, has
led to considerable speculation that ALT may proceed by
the same mechanism(s) that promote telomere elongation
in K. lactis survivors or type II survivors in S. cerevisiae.
The speci®c genetic requirements that promote the
ALT pathway have not yet been elucidated, due to the
diculties of genetic analysis in human cells, nor has an
inducible system for ALT, where the intermediate steps
Oncogene
Telomere maintenance without telomerase
V Lundblad
530
in the pathway can be analysed, been established. In
addition, the steady attrition of telomeric DNA observed
in K. lactis and type II S. cerevisiae survivors have not
been observed. However, numerous studies have shown
that human chromosome termini are subject to enhanced
levels of recombination, as monitored by sequence
polymorphisms in sub-telomeric regions (see Wilkie et
al., 1991; Baird et al., 2000 for several examples).
Additional work indicates that terminal deletions found
on human chromosomes can be healed by capture of subtelomeric and telomeric sequences from another chromosome (Meltzer et al., 1993). A study by Reddel's
laboratory (Dunham et al., 2000) was designed to ask
speci®cally whether ALT telomeres participated in
recombination, by monitoring the transmission of a
selectable marker inserted into telomeric sequences in
either ALT or telomerase-plus immortalized cells. This
marker spread to other chromosome ends much more
frequently in ALT cells than in telomerase-plus cells,
suggesting that the telomeres in ALT cells (or a subset,
such as the shortest telomeres?) were more recombinogenic than telomeres maintained by telomerase. Notably,
when the same marker was placed in a sub-telomeric
location, dispersal to other chromosomes was not
observed, in contrast to the results obtained in K. lactis
by McEachern and Iyer (2001).
It is also worth noting that recombination may
contribute to telomere maintenance and potentially
tumor formation, even without conversion to the full
ALT pathway. In particular, any genetic alteration that
enhances recombination rates at telomeres ± such as
defects in mismatch repair ± may partially relieve the
telomere maintenance barrier to proliferation, thereby
contributing to the mechanism of oncogenesis.
Can the ALT pathway and telomerase-mediated telomere
maintenance co-exist in the same cell?
A number of recent papers have been directed at the
above question, by examining the consequences of
reconstitution of telomerase activity in ALT cells as a
result of ectopic expression of telomerase subunits
(Perrem et al., 2001; Cerone et al., 2001; Ford et al.,
2001; Grobelny et al., 2001) or by fusion of ALT cells
with telomerase-positive tumor cell lines (Perrem, et al.
2001). However, the conclusions from these four
groups were somewhat contradictory. Two groups
reported that when telomerase was ectopically expressed, short telomeres were preferentially elongated,
but the ALT pathway was not repressed (Grobelny et
al., 2001; Cerone et al., 2001). In contrast, Shay and
Wright and colleagues, using the same ectopic
expression approach, observed rapid reduction of the
longest telomeres and concluded that ALT was
inhibited (Ford et al., 2001). Reddel and co-workers
did not observe inhibition of ALT when telomerase
was ectopically expressed, but when ALT cells were
fused with telomerase-positive tumor cells, telomeres
showed rapid initial decreases in length, followed by
more gradual shortening (Perrem et al., 2001).
Oncogene
Observations from yeast may provide some insight into
the potential complexities of these experiments. The
genomic alterations that occur in telomerase-defective
survivors can be very stable, even when telomerase is
restored. Introduction of EST1 back into an est1-D
survivor that exhibits extensive Y' ampli®cation results in
elongation of the terminal telomeric tract back to wild
type length, as well as a substantial reduction in Y' copy
number (Lundblad and Blackburn, 1993; Teng and
Zakian, 1999). However, even with extensive propagation, the distribution of Y' elements does not return to its
starting point: each telomere retains at least one Y'
element in an EST1 strain that has a previous history as a
telomerase-defective survivor (V Lundblad, unpublished
results). When this strain is converted back to est1-D, it
again exhibits telomere shortening and senescence, but in
this newly senescent strain, survivors appear more readily
and at earlier time points. This suggests that the retained
single Y' elements prime the re-appearance of survivors.
Similarly, following reintroduction of telomerase into a
type II survivor, telomeres return to a roughly wild type
telomeric restriction pattern only very slowly (Teng and
Zakian, 1999). This suggest that even if telomeres are no
longer highly recombinogenic after the reintroduction of
telomerase, the stable extended telomeric tracts still decay
only very slowly.
Extending these observations to human cells, this
suggests that using telomere length as a monitor of
whether a recombination-promoted pathway for telomere maintenance is still operating in the presence of
telomerase may not be ideal. This monitors the product,
rather than the presence of an active mechanism. The
solution may be to develop assays that directly monitor
the frequency of recombination events at telomeres. One
such assay, using a reporter gene that marks a telomeric
tract, has been described above (Dunham et al., 2000). An
alternative approach relies on the presence of natural
variants in the telomeric repeat tract. Human telomerase
synthesizes TTAGGG repeats onto chromosome termini
(Morin, 1989). However, the proximal end of this repeat
tract contains variant telomeric repeats such as
TGAGGG and TCAGGG, which can be monitored by
a PCR-based assay (Baird et al., 1995). Since the
distribution of these variant repeats is highly variable,
this assay may provide a semi-quantitative means of
monitoring the extent of recombination at human
chromosome ends.
Summary and future possibilities
Experiments using budding yeasts have provided both
speci®c models and genetic requirements that dictate the
frequency of recombination-dependent exchanges between telomeres in cells that lack telomerase. These
observations have implicated speci®c recombination
proteins and also highlighted other alterations, such as
defects in mismatch repair, that may similarly in¯uence
the survival of human cells that fail to express telomerase.
Additional studies, not discussed in this review, suggest
that alterations in components of telomeric chromatin
Telomere maintenance without telomerase
V Lundblad
can also in¯uence either the types of recombination events
that occur at the telomere (Teng et al., 2000) or the
frequency with which the survival pathway appears
(Chandra and Lundblad, unpublished observations).
All of these examples in yeast provide a possible roadmap
for experiments that may help elucidate the ALT
pathway.
Acknowledgments
I thank Jim Haber and Jenny Hackett for explaining the
subtleties of the genetic requirements of BIR to me, Sara
Evans for excellent editorial assistance and ®gure preparation, and Lou Zumstein, Rachel Cervantes and Erin
Pennock for critical reading of the manuscript. Work in
the author's laboratory is supported by grants from the
NIH and the Ellison Medical Foundation.
531
References
Baird DM, Coleman J, Rosser ZH and Royle NJ. (2000).
Am. J. Hum. Genet., 66, 235 ± 250.
Baird DM, Je€reys AJ and Royle NJ. (1995). EMBO J., 14,
5433 ± 5443.
Bernards A, Michels PA, Lincke CR and Borst P. (1983).
Nature, 303, 592 ± 597.
Biessmann H and Mason JM. (1997). Chromosoma, 106, 63 ±
69.
Blackburn EH. (2001). Cell, 106, 661 ± 673.
Bodnar AG, Ouellette M, Frolkis M, Holt SE, Chiu C-P,
Morin GB, Harley CB, Shay JW, Lichtsteiner S and
Wright WE. (1998). Science, 279, 349 ± 352.
Bosco G and Haber JE. (1998). Genetics, 150, 1037 ± 1047.
Bryan TM, Englezou A, Dalla-Pozza L, Dunham MA and
Reddel RR. (1997). Nat. Med., 3, 1271 ± 1274.
Bryan TM, Englezou A, Gupta J, Bacchetti S and Reddel
RR. (1995). EMBO J., 14, 4240 ± 4248.
Bucholc M, Park Y and Lustig AJ. (2001). Mol. Cell. Biol.,
21, 6559 ± 6573.
Cerone MA, Londono-Vallejo JA and Bacchetti S. (2001).
Hum. Mol. Genet., 10, 1945 ± 1952.
Chan CS and Tye BK. (1983). J. Mol. Biol., 168, 505 ± 523.
Chen Q, Ijpma A and Greider CW. (2001). Mol. Cell. Biol.,
21, 1819 ± 1827.
Cohen H and Sinclair DA. (2001). Proc. Natl. Acad. Sci.
USA, 98, 3174 ± 3179.
Dunham MA, Neumann AA, Fasching CL and Reddel RR.
(2000). Nat. Genet., 26, 447 ± 450.
Dunn B, Szauter P, Pardue ML and Szostak JW. (1984). Cell,
39, 191 ± 201.
Ford LP, Zou Y, Pongracz K, Gryaznov SM, Shay JW and
Wright WE. (2001). J. Biol. Chem., 276, 32198 ± 32203.
Formosa T and Alberts BM. (1986). J. Biol. Chem., 261,
6107 ± 6118.
Greider CW. (1996). Annu. Rev. Biochem., 65, 337 ± 365.
Grith JD, Comeau L, Rosen®eld S, Stansel RM, Bianchi A,
Moss H and de Lange T. (1999). Cell, 97, 503 ± 514.
Grobelny JV, Kulp-McEliece M and Broccoli D. (2001).
Hum. Mol. Genet., 10, 1953 ± 1961.
Hackett JA, Feldser DM and Greider CW. (2001). Cell, 106,
275 ± 286.
Horowitz H and Haber JE. (1985). Mol. Cell. Biol., 5, 2369 ±
2380.
Huang P, Pryde FE, Lester D, Maddison RL, Borts RH,
Hickson ID and Louis EJ. (2001). Curr. Biol., 11, 125 ± 129.
Hughes TR, Evans SK, Weilbaecher RG and Lundblad V.
(2000). Curr. Biol., 10, 809 ± 812.
Johnson FB, Marciniak RA, McVey M, Stewart SA, Hahn
WC and Guarente L. (2001). EMBO J., 20, 905 ± 913.
Kass-Eisler A and Greider CW. (2000). Trends Biochem. Sci.,
25, 200 ± 204.
Kramer KM and Haber JE. (1993). Genes Dev., 7, 2345 ± 2356.
Kraus E, Leung WY and Haber JE. (2001). Proc. Natl. Acad.
Sci. USA, 98, 8255 ± 8262.
Le S, Moore JK, Haber JE and Greider CW. (1999).
Genetics, 152, 143 ± 152.
Lendvay TS, Morris DK, Sah J, Balasubramanian B and
Lundblad V. (1996). Genetics, 144, 1399 ± 1412.
Li B and Lustig AJ. (1996). Genes Dev., 10, 1310 ± 1326.
Luder A and Mosig G. (1982). Proc. Natl. Acad. Sci. USA,
79, 1101 ± 1105.
Lundblad V. (2000). Mutat. Res., 451, 227 ± 240.
Lundblad V and Blackburn EH. (1993). Cell, 73, 347 ± 360.
Lundblad V and Szostak JW. (1989). Cell, 57, 633 ± 643.
Malkova A, Signon L, Schaefer CB, Naylor ML, Theis JF,
Newlon CS and Haber JE. (2001). Genes Dev., 15, 1055 ±
1060.
McEachern MJ and Blackburn EH. (1995). Nature, 376,
403 ± 409.
McEachern MJ and Blackburn EH. (1996). Genes Dev., 10,
1822 ± 1834.
McEachern MJ and Iyer S. (2001). Mol. Cell., 7, 695 ± 704.
McEachern MJ, Krauskopf A and Blackburn EH. (2000).
Annu. Rev. Genet., 34, 331 ± 358.
Meltzer PS, Guan XY and Trent JM. (1993). Nature Genet.,
4, 252 ± 255.
Morin GB. (1989). Cell, 59, 521 ± 529.
Morris DK and Lundblad V. (1997). Curr. Biol., 7, 969 ± 976.
Munoz-Jordan JL, Cross GA, de Lange T and Grith JD.
(2001). EMBO J., 20, 579 ± 588.
Novick RP. (1998). Trends Biochem. Sci., 23, 434 ± 438.
Perrem K, Colgin LM, Neumann AA, Yeager TR and
Reddel RR. (2001). Mol. Cell. Biol., 21, 3862 ± 3875.
Reddel RR, Bryan TM, Colgin LM, Perrem KT and Yeager
TR. (2001). Radiat. Res., 155, 194 ± 200.
Rizki A and Lundblad V. (2001). Nature, 411, 713 ± 716.
Sandell LL and Zakian VA. (1993). Cell, 75, 729 ± 739.
Shay JW and Bacchetti S. (1997). Eur. J. Cancer, 33, 787 ±
791.
Signon L, Malkova A, Naylor ML, Klein H and Haber JE.
(2001). Mol. Cell. Biol., 21, 2048 ± 2056.
Singer MS and Gottschling DE. (1994). Science, 266, 404 ±
409.
Teng CS, Chang J, McCowan B and Zakian AV. (2000). Mol.
Cell., 6, 947 ± 952.
Teng SC and Zakian VA. (1999). Mol. Cell. Biol., 19, 8083 ±
8093.
Walmsley RW, Chan CS, Tye BK and Petes TD. (1984).
Nature, 310, 157 ± 160.
Wilkie AO, Higgs DR, Rack KA, Buckle VJ, Spurr NK,
Fischel-Ghodsian N, Ceccherini I, Brown WR and Harris
PC. (1991). Cell, 64, 595 ± 606.
Yeager TR, Neumann AA, Englezou A, Huschtscha LI,
Noble JR and Reddel RR. (1999). Cancer Res., 59, 4175 ±
4179.
Zhu XD, Kuster B, Mann M, Petrini JH and Lange T.
(2000). Nat. Genet., 25, 347 ± 352.
Oncogene