Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

© 2001 Oxford University Press Human Molecular Genetics, 2001, Vol. 10, No.

20 2187–2194

Dynamic mutations: a decade of unstable expanded


repeats in human genetic disease
Robert I. Richards*

Department of Molecular Biosciences, The University of Adelaide, Adelaide, SA 5000, Australia and Centre for
Medical Genetics, Women’s and Children’s Hospital, North Adelaide, SA 5006, Australia

Received June 10, 2001; Accepted July 6, 2001

The term ‘dynamic mutation’ was introduced to distinguish the unique properties of expanding, unstable DNA
repeat sequences from other forms of mutation. The past decade has seen dynamic mutations uncovered as
the molecular basis for a growing number of human genetic diseases and for all of the characterized ‘rare’
chromosomal fragile sites. The common properties of the repeats in different diseases and fragile sites have
given insight into this unique form of DNA instability. While the dynamic mutation mechanism explains some
unusual genetic characteristics, unexpected findings have raised new questions and challenged some
assumptions about the pathways that lead from mutation to disease. This review will address the current
understanding of the molecular mechanisms involved in the dynamic mutation process and elaborate on the
pathogenic pathways that lead from expanded repeats to the diseases with which they are associated.

INTRODUCTION molecular mechanisms for both the genesis of expanded


repeats and the molecular pathways from repeat expansion to
Ten years ago the first reports appeared (1,2) of expanded
disease phenotype. These general properties include: (i) mutation
DNA repeat sequences being the molecular basis of human
manifests as a change (usually increase) in repeat copy number
genetic diseases—fragile X syndrome and spino-bulbar with mutation rate related to the initial copy number of the
muscular atrophy (Kennedy disease). In both cases a simple repeat; (ii) rare ‘founder’ events (such as loss of repeat inter-
DNA repeat sequence (CCG or CAG), which varied in copy ruption) lead to alleles with increased likelihood of undergoing
number in normal individuals, was found to be increased in changes in repeat copy number; and (iii) the diseases caused by
copy number beyond a certain threshold in affected individuals. repeat expansion exhibit a relationship between copy number
Given the idiosyncratic nature of these diseases, particularly of the repeat and the severity and/or age-at-onset of symptoms.
fragile X syndrome (3), it could have been that these expanded These properties together account for anticipation, the
repeats were manifestations of an uncommon mechanism with increasing severity/incidence and/or decreasing age-at-onset in
characteristics peculiar and unique to their associated diseases. successive generations within an affected family. However, an
Instead, however, an ever-growing list of human diseases and increasing number of exceptions and unexpected findings
chromosomal fragile sites has identified dynamic mutation as suggest that early conclusions, particularly in regard to
an important mechanism in human genetics. Consequently, possible common pathogenic mechanisms, may have been
both the molecular processes underlying this form of mutation premature.
and the pathogenic pathways leading from the mutation to its
phenotype have been the subject of intense investigation. Over
the past decade numerous reviews have updated the progress MECHANISM OF DYNAMIC MUTATION
of these investigations (4–10). Rather than reiterate this The process of dynamic mutation is affected by a variety of
progress, the purpose of this review is to identify the generally elements and factors that have been divided into those directly
accepted principles underlying dynamic mutation and to associated with the expanding repeat (cis-acting elements) and
discuss those areas where the current understanding is either those (trans-acting factors) whose interaction with the repeat
lacking or clouded by apparently conflicting observations. contributes to its instability. Examples of cis-acting factors
Comparative anatomy has long been a potent tool to develop an include the copy number and the composition of the repeat
understanding of the functional or mechanistic basis responsible (whether it is perfect or interrupted) (11). In general, higher copy
for a particular physical phenomenon. The application of number alleles that are free of interruptions (perfect repeats) are
comparative molecular anatomy to the study of different more unstable than those of lower copy number and/or contain
classes of chromosomal fragile sites and to the diseases caused interruptions (imperfect repeats). While there is indirect
by DNA repeat expansion has led to the identification of a evidence for the existence of trans-acting factors (such as
number of common properties that most likely reflect common parental gender bias in repeat instability), the identity of these

*To whom correspondence should be addressed at: Department of Molecular Biosciences, The University of Adelaide, Adelaide, SA 5000, Australia.
Tel: +618 8303 7541; Fax: +618 8303 4362; Email: robert.richards@adelaide.edu.au
2188 Human Molecular Genetics, 2001, Vol. 10, No. 20

Table 1. Founder effects in dynamic mutation diseases/fragile sites

Repeat Disease/fragile site Reported founder effect (reference)


CUG Myotonic dystrophy (71a,72)
CCG Fragile X syndrome (73)
CAG Huntington disease (74)a
CAG Spinocerebellar ataxia type 1 (SCA1) (75)
CAG SCA3/Machado–Joseph disease (76)
CAG Spino-bulbar muscular atrophy (77)
CAG Dentatorubral pallidoluysian atrophy (78)
AAG Friedreich ataxia (79)
AT-rich 42 bp FRA10B (19)
CAG SCA2 (80)
CAG SCA7 (81)
???? SCA4 (82)a
CAG SCA6 (83)

aThese reports appeared before the expanded repeat mutation was identified.

factors remains elusive. Candidates include the proteins cis-acting elements in and around the FRAXA repeat (21,22).
involved in DNA replication such as FEN1 (Rad27), which is One possible explanation for the inconsistency here is that
known to have a role in Okazaki fragment metabolism (12). chance has a role to play and that the differences between
populations merely reflect the different rare initial events that
What kinds of repeat sequences undergo dynamic have precipitated the dynamic mutation process. Given that
mutation?—Cis-acting factors there are multiple possibilities for generating an unstable
repeat then different populations are likely to have different
The first expanded repeats to be identified were the trinucleotides
combinations and/or frequencies of these unstable alleles.
CCG/CGG and CAG/CTG. Initially this was taken as evidence
Thus, in terms of understanding the mutation process, caution
that only trinucleotide repeats could undergo this form of
is needed when extrapolating the results observed in one
mutation (which was sometimes referred to as trinucleotide
population to the human species as a whole.
repeat expansion). Furthermore, since these two repeats could
form secondary structures, it was assumed that this was also a
What factors contribute to repeat DNA expansion?—
necessary condition for repeat instability (13). Subsequently,
Trans-acting factors
six of the ten possible trinucleotide repeats have been found to
exhibit repeat expansion in humans (14), and, more recently, 5 Gender bias in the instability of repeats during their transmission
and 6 bp microsatellite repeats (15,16) and 12, 33 and 42 bp from parent to offspring is a common property of dynamic
minisatellite repeats (17–19) have also been observed to mutations. Most cases of the congenital form of myotonic
undergo repeat copy number expansion. dystrophy are maternally inherited, while the juvenile onset
Founder effects are observed in association with dynamic Huntington disease is generally from paternal inheritance. The
mutation as a higher frequency of particular alleles in the gender of the germ line clearly contributes to these biases;
affected population, compared with the unaffected population, however, the identity of the factors involved remains obscure.
suggesting that these alleles are ‘at-risk’ for mutation. The Other factors that interact in some way with the repeat are
likelihood is that the initial mutation on these alleles is a rare also thought to contribute to repeat instability. A variety of
event that increases the instability of the repeat, setting in circumstantial evidence (particularly from studies in model
process the subsequent expansion of these few founder systems) suggests that Okazaki fragments may have a role to
mutations. Founder effects (Table 1) are a common feature of play in the expansion of DNA repeats. Several of the repeats
dynamic mutation and in many cases common flanking marker (e.g. CCG in FRAXA and CAG in DM) exhibit a bimodal
haplotypes provide evidence of a relationship between the distribution of instability with the boundary between small
expanded alleles and the longest normal (usually uninterrupted) changes and large changes in copy number approximating
alleles in the population. Therefore a common molecular Okazaki fragment length. In Escherichia coli (which has a
mechanism appears to be responsible (Fig. 1). substantially longer Okazaki fragment length), this boundary is
At the FRAXA locus, detailed haplotyping studies, together increased in copy number again to that approximating Okazaki
with sequence analysis of different repeat alleles, have provided fragment length (23). The secondary structure of CAG/CTG
evidence in support of the view that certain alleles are predisposed and CGG repeat sequences has been found to inhibit Okazaki
to instability (11,20). Similar analyses in other populations flap endonuclease (FEN1) binding and cleavage (24). In
suggest that such differences are likely to be population-specific addition, the FEN1 homologue, Rad27, has a role to play in
rather than indicative of a molecular pathway predisposed by DNA repeat instability in yeast. Yeast Rad27 mutants are
Human Molecular Genetics, 2001, Vol. 10, No. 20 2189

Figure 1. Model for common mechanism of repeat expansion. A model for the relationship between expanded alleles and the longest normal (usually perfect)
repeat alleles at a dynamic mutation locus. Interrupted alleles are blue bars, uninterrupted (perfect repeat) alleles are red bars.

prone to both DNA breakage and copy number instability of promoter; and Friedreich ataxia, in which the intronic
repeat sequences (25 and personal communication). expanded repeat sequence adopts a ‘sticky’ DNA structure that
Contradicting the proposed role of DNA replication, interferes with transcription of the frataxin gene (35). While
Kennedy and Shelbourne (26) found dramatic expansion of the these resultant diseases exhibit recessive (or X-linked) trans-
CAG repeat in the striatum of a transgenic mouse HD model. mission, the potential exists for dominant transmission where
These neurons are presumably no longer undergoing cell reduction in the level or activity of a rate limiting gene product
division, indicating that expansion is occurring independently can appear as haploinsufficiency. This appears to be the case
of DNA replication. Similarly, Kotvun and McMurray (27) for the SIX5 gene in myotonic dystrophy (36,37), as the
have found that germ-line trinucleotide repeat sequence expanded CTG repeat is located in the SIX5 promoter. The
expansion in transgenic mice also proceeded in the absence of reduction in effective ‘dosage’ of the SIX5 gene product
DNA replication via gap repair. Perhaps, given that DNA brought about by extinction of one allele is thought to
instability is occurring at the repeat sequence in the absence of contribute to some aspects of pathology (see below).
replication, some form of transcriptional healing may be
contributing to the repeat expansion process (28). Transcrip- Gain of function: the toxic polyglutamine hypothesis—old
tional healing has been invoked in the fragile site expression of concerns about a new dogma
tandemly repeated small RNA transcripts (29). In further Numerous expanded repeat sequences are located within
support of a role for breakage and repair Manley et al. (30) coding regions where they translate into expanded poly-
have found that Msh2 is required for CAG repeat somatic cell glutamine tracts in disease alleles (Table 2). This, together
instability in an HD transgenic mouse. with a substantial body of data demonstrating the toxicity of
Age-dependent somatic cell instability of the CTG repeat polyglutamine in cells, has led to the view that the expanded
sequence has been found in myotonic dystrophy (31), although polyglutamine is a common, crucial component in the patho-
here it is not confined to those cells that are affected in the genesis of these diseases. In support of this view, an antibody,
disease and in fact gives a growth advantage to lymphoblastoid 1C2, raised against the normal length polyglutamine tract of
cell lines that have undergone expansion (32). the TATA-box binding protein (TBP), was found to specifically
recognize the expanded polyglutamine in several of the
PATHOGENIC PATHWAYS expanded repeat diseases, e.g. huntingtin, ataxin 1 and others
(38). This was taken as evidence that the expanded poly-
Loss of function/haploinsufficiency glutamine adopted a different conformation that in some way
contributed to its toxicity. The context in which the repeat
Extinction of, or interference with transcription is one means sequence was located was proposed to account for why such an
by which expanded repeats cause loss of gene function. otherwise toxic conformation did not cause problems when
Examples are: fragile X syndrome, in which repeat expansion located within TBP. This theory is now somewhat flawed, as
brings about methylation of the FMR1 promoter region (33); the expanded polyglutamine tract in some alleles (n >50) of
myoclonus epilepsy (EPM1) (34), in which repeat expansion TBP has recently been associated with neurodegenerative
physically separates transcription factors in the cystatin B disease (39,40). It is possible that yet another conformation of
2190 Human Molecular Genetics, 2001, Vol. 10, No. 20

Table 2. Polyglutamine (or not) in neurodegenerative disorders

Polyglutamine diseases Non-polyglutamine/expanded repeat, dominant spino-cerebellar ataxias


SBMA a SCA8 3′-UTR expanded CUG
HD a SCA10 intronic expanded 5 bp repeat
DRPLA a SCA12 5′-UTR expanded CAG
SCA1 a

SCA2 a

SCA3 (MJD) a

SCA6 Normal polyQ 4–16


Affected polyQ 21–27
SCA7 a

TBP Normal polyQ 27–44


Affected polyQ >50

*Typical normal polyQ, n < 40; typical affected polyQ, n > 36 (some overlap because of variable intermediate alleles).

the expanded polyglutamine tract is adopted once the repeat diseases in which it has been identified as the disease-causing
sequence gets beyond an even greater threshold. agent. If it is, then how is the cellular specificity of this toxicity
Nuclear inclusions were found in the affected neurons of HD accounted for when at least some of these proteins are widely
and SCA1 transgenic mouse models and HD patients (41), expressed (see 52)? One possible explanation for this conundrum
prompting the assertion that these expanded polyglutamine- comes from the findings of Kennedy and Shelbourne (26),
containing nuclear inclusions cause neuronal dysfunction. which reveal affected cell-specific amplification of the
Klemment et al. (42) deleted the SCA1 self-association region unstable CAG repeat in a mouse model of Huntington’s
in transgenic mice and found that these mice still exhibited disease. Individuals affected with Huntington’s disease have
ataxia and Purkinje cell pathology; however, no evidence of also been found to exhibit somatic instability, which was most
nuclear aggregates was found. Other studies on the brains of pronounced in the affected areas of the brain (53,54). This
individuals affected with Huntington’s disease (43,44) suggests that ongoing somatic mutation may contribute to
revealed that nuclear aggregates were more commonly found diseases caused by expanded repeats. This expansion may also
in unaffected neurons than in the vulnerable striatal spiny account for the relationship between germ-line repeat copy
neurons. Consequently, it has now been proposed that rather number and age-at-onset for these diseases (7). The time taken
than being pathogenic, nuclear inclusions may even be protective for the disease to manifest could represent how long it takes for
against the toxic effects of the expanded polyglutamine (45). the disease-associated allele to reach a critical higher copy
Further conflicting evidence to a common role for poly- number threshold in the affected cell population (Fig. 2).
glutamine in pathogenesis is that disease alleles of the CAG What then is the evidence that polyglutamine has a role? The
repeat in spinocerebellar ataxia 6 are well within the normal principal data are from transgenic animal model studies,
range for other ‘polyglutamine diseases’. Functional assays on particularly those from SCA1 transgenic mice. Mutation of the
expanded CAG alleles of the SCA6 associated gene, CACNA1A, ataxin-1 nuclear localization signal severely diminishes toxicity,
demonstrate increased Ca2+ transport (46) and thus make it whereas abolition of sequences necessary for aggregation does
likely that this gain of function is a specific and unique cause of not (42). Crossing of SCA1 mice with a transgenic mouse
pathogenesis for this disorder. Taken together, these exceptions deficient in ubiquitin ligase results in more severe pathology
suggest that the notion of a common single pathogenic (55). For SCA2 and SCA6, nuclear localization does not appear
pathway involving polyglutamine is unlikely. to be necessary, suggesting different pathways than that for
As if contradictions with polyglutamine encoding genes are SCA1 (and HD) (56,57).
not enough, a growing list of repeat-associated ataxias do not In an HD transgenic mouse model where transcription of the
even encode polyglutamine (Table 2). In SCA12, the expanded transgene could be artificially turned on or off, expression of
CAG repeat is located within the 5′-untranslated region (5′-UTR) the expanded CAG-containing gene was elegantly demon-
of the PPP2R2B gene (47,48), while in SCA8 an expanded strated to be necessary for HD-like pathology (58). This animal
CUG is located within the 3′-UTR of a transcript (although model provides the first indication that HD (and hopefully
there is controversy over whether this expansion is causative of other expanded CAG repeat diseases) is potentially treatable—
neurodegenerative disease) (49–51). In addition, an intronic given that, in those patients at risk, the disease-causing gene
expanded 5 bp repeat was recently found to be associated with can either be switched off or the toxic gene product can be
SCA10 (15). cleared from the sensitive neurons.
So the question must be asked whether expanded poly- Presumably other factors, such as modification of the poly-
glutamine is a necessary and sufficient condition for those glutamine tract and/or the protein context in which the
Human Molecular Genetics, 2001, Vol. 10, No. 20 2191

Figure 2. Somatic mutation model for relationship between repeat copy number and age-at-onset. Somatic cell changes in copy number could contribute to the
relationship seen in certain neurodegenerative diseases between the copy number of the repeat locus and the age-at-onset of disease symptoms. The time taken (t)
for the repeat to reach a critical copy number threshold by means of incremental increases in repeat copy number would be determined by the inherited (germ-line)
copy number (green arrows) at birth (n). The greater is n the shorter is t. Time taken (t) to reach a disease-causing copy number is inversely proportional to ‘starting’
copy number (n), t = k/n (where k is determined by the repeat context).

polyglutamine is located, contribute to the cellular specificity mediating the pathogenic effects of repeat expansion (for
of toxicity. Examples of where this apparently is (59) or is not review see 10). Three genes (DMPK, SIX5 and DMWD) are all
(60) the case have been reported. The relative contribution of located in the immediate vicinity of the DM expanded CTG
various intrinsic and extrinsic factors is therefore likely to repeat. At least three distinct pathogenic pathways are thought
differ between the different disease genes (and experimental to contribute to DM. (i) The expanded repeat affects the
model systems). Genetic screens for factors that modify poly- splicing of the DMPK gene in which it is located (64). (ii) The
glutamine-generated pathology in Drosophila have produced a RNA transcript containing the expanded repeat is both
large number and variety of candidates (61,62), and it will be inappropriately compartmentalized and titrates a crucial
of great interest to see which (if any) of these has a role in protein (muscleblind) (65). Expression of the CTG repeat in
human polyglutamine disease. another muscle-specific transcript is able to cause at least some
Bias of ascertainment was invoked by Penrose to account for of the DM phenotype (myotonia and myopathy) (66). (iii) The
(and discredit) the observation of anticipation seen by clinical DM CTG repeat is also located in the promoter region of the
geneticists in myotonic dystrophy families (63). While SIX5 gene and its expansion also interferes with the expression
expanded DNA repeats now give a molecular basis (and of this gene. The SIX5 gene encodes a transcription factor
validation) for the phenomenon of anticipation, the relative whose concentration is critical for eye, muscle and testicular
ease with which expanded CAG repeats can be screened for as development (67). Heterozygous deletion of Six5 is sufficient
a cause of neurodegenerative disease may well have led to an to cause ocular cataracts in transgenic mice (36,37), and
overemphasis of the relative importance of this form of therefore the SIX5 haploinsufficiency brought about by the
mutation and perhaps contributed to the view that expanded DM CTG repeat expansion is likely to contribute to the DM
CAG repeats necessarily have a common pathogenic pathway phenotype.
(via their encoded expanded polyglutamines). While there is a The FRAXE chromosomal fragile site is associated with a
substantial body of evidence in support of a role for poly- mild form of mental retardation (68). The expanded CCG
glutamine in some repeat expansion diseases there is a growing repeat was first initially located in the 5′-UTR of a gene,
list of exceptions and inconsistencies which suggest (at the referred to as FMR2. The expression of FMR2 is extinguished
least) that this is not necessarily a common pathogenic by repeat expansion and subsequent methylation of the CpG
pathway exhibited by all neurodegenerative diseases in which island promoter region (69), in a manner analogous to the
an expanded repeat (let alone an expanded CAG) is the silencing of the FMR1 gene at the FRAXA locus in fragile X
disease-causing mutation. syndrome (33). It was therefore thought that the loss of FMR2
was responsible for the phenotype associated with FRAXE.
Further evidence for alternative pathways—multiple gene However, an additional gene, FMR3, has now been identified
effects that shares the same methylated CpG island (70). Expression
In myotonic dystrophy there is good evidence to support not of FMR3 is also silenced by FRAXE full mutation and there-
only a role for multiple genes, but also multiple pathways in fore this gene may also contribute to the phenotype.
2192 Human Molecular Genetics, 2001, Vol. 10, No. 20

Figure 3. Multiple pathogenic pathways for dynamic mutation diseases. Alternative and sometimes multiple pathways contribute to the complex pathology brought
about by expansion of the repeat sequences within or near genes.

CONCLUSIONS 2. La Spada, A.R., Wilson, E.M., Lubahn, D.B., Harding, A.E. and
Fischbeck, K.H. (1991) Androgen receptor gene mutations in X-linked
Given that increasing the copy number of existing DNA repeat spinal and bulbar atrophy. Nature, 352, 77–79.
sequence would appear to be a relatively simple process, in 3. Sherman, S.L., Jacobs, P.A., Morton, N.E., Froster-Iskenius, U.,
Howard-Peebles, P.N., Nielsen, K.B., Partington, M.W., Sutherland, G.R.,
reality the biological causes and consequences of such Turner, G. and Watson, M. (1985) Further segregation analysis of the
phenomena can be remarkably complex. In fact, a major issue fragile X syndrome with special reference to transmitting males.
in understanding both the process of dynamic mutation and the Hum. Genet., 69, 289–299.
pathway from genotype to phenotype is being able to distinguish 4. Richards, R.I. and Sutherland, G.R. (1992) Dynamic mutation: a new
class of mutations causing human disease. Cell, 70, 709–712.
between cause and consequence. It certainly appears as though 5. Richards, R.I. and Sutherland, G.R. (1997) Dynamic mutation: possible
there are multiple components to the mutation process and it is mechanisms and significance in human disease. Trends Biochem. Sci., 22,
clear that there can be multiple pathways to the resultant 432–436.
6. Richards, R.I. and Sutherland, G.R. (1994) Simple repeat DNA is not
disease (Fig. 3). So far dynamic mutations have been detected
replicated simply. Nat. Genet., 6, 114–116.
in diseases (and at fragile site loci) with very high penetrance. 7. Zoghbi, H.Y. (1996) The expanding world of ataxins. Nat. Genet., 14,
It will be intriguing to see whether DNA repeats also 237–238.
contribute to multifactorial diseases and disease susceptibility. 8. Wells, R.D. and Warren, S.T. (eds) (1998) Genetic Instabilities and
Hereditary Neurological Diseases. Academic Press, San Diego, CA.
9. Cummings, C.J. and Zoghbi, H.Y. (2000) Fourteen and counting:
ACKNOWLEDGEMENTS unravelling trinucleotide repeat diseases. Hum. Mol. Genet., 9, 909–916.
10. Tapscott, S.J. (2000) Deconstructing myotonic dystrophy. Science, 289,
I thank Keith Johnson, Catherine Freudenreich and Alec 1701–1702.
Jeffreys for communicating data prior to publication and 11. Gunter, C., Paradee, W., Crawford, D.C., Meadows, K.A., Newman, J.,
Kunst, C.B., Nelson, D.L., Schwartz, C., Murray, A., Macpherson, J.N.
Kathryn Friend, Sonia Dayan, Merran Finnis and Lynne et al. (1998) Re-examination of factors associated with expansion of CGG
Hobson for constructive comments on drafts of this review. I repeats using a single nucleotide polymorphism in FMR1.
am grateful to Shelley Richards for her support and encourage- Hum. Mol. Genet., 7, 1935–1946.
12. Gordenin, D.A., Kunkel, T.A. and Resnick, M.A. (1997) Repeat
ment. The preparation of this review was supported by grants expansion—all in a flap? Nat. Genet., 16, 116–118.
from the Anti-Cancer Foundation of South Australia and the 13. Gacy, A.M., Goellner, G.M., Spiro, C., Chen, X., Gupta, G., Bradbury, E.M.,
National Health and Medical Research Council of Australia. Dyer, R.B., Mikesell, M.J., Yao, J.Z., Johnson, A.J. et al. (1998) GAA
instability in Friedreich’s ataxia shares a common, DNA-directed and
intraallelic mechanism with other trinucleotide repeats. Mol. Cell, 1, 583–593.
REFERENCES 14. Lindblad, K., Zander, C., Schalling, M. and Hudson, T. (1994) Growing
triplet repeats. Nat. Genet., 7, 124.
1. Kremer, E., Pritchard, M., Lynch, M., Yu, S., Holman, K., Warren, S.T., 15. Matsuura, T., Yamagata, T., Burgess, D.L., Rasmussen, A., Grewal, R.P.,
Schlessinger, D., Sutherland, G.R. and Richards, R.I. (1991) DNA Watase, K., Khajavi, M., McCall, A.E., Davis, C.F., Zu, L. et al. (2000)
instability at the fragile X maps to a trinucleotide repeat sequence Large expansion of the ATTCT pentanucleotide repeat in spinocerebellar
p(CCG)n. Science, 252, 1711–1714. ataxia. Nat. Genet., 26, 191–194.
Human Molecular Genetics, 2001, Vol. 10, No. 20 2193

16. Bois, P.R.J., Southgate, L. and Jeffreys, A.J. (2001) Length of develop cataracts: implications for myotonic dystrophy. Nat. Genet., 25,
uninterrupted repeat determines instability at the unstable mouse 105–109.
expanded simple tandem repeat family MMS10 derived from independent 38. Trottier, Y., Lutz, Y., Stevanin, G., Imbert, G., Devys, D., Cancel, G.,
SINE B1 elements. Mamm. Genome, 12, 104–111. Saudou, F., Weber, C., David, G., Tora, L. et al. (1995) Polyglutamine
17. Lalioti, M.D., Scott, H.S., Buresi, C., Rossier, C., Bottani, A., expansion as a pathological epitope in Huntington’s disease and four
Morris, M.A., Malafosse, A. and Antonarakis, S.E. (1997) Dodecamer dominant cerebellar ataxias. Nature, 378, 403–406.
repeat expansion in cystatin B gene in progressive myoclonus epilepsy. 39. Koide, R., Kobayashi, S., Shimohata, T., Ikeuchi, T., Maruyama, M.,
Nature, 386, 847–851. Saito, M., Yamada, M., Takahashi, H. and Tsuji, S. (1999) A neurological
18. Yu, S., Mangelsdorf, M., Hewett, D., Hobson, L., Baker, E., Eyre, H., disease caused by an expanded CAG trinucleotide repeat in the TATA-binding
Lapsys, N., Le Paslier, D., Doggett, N., Sutherland, G.R. and Richards, R.I. protein gene: a new polyglutamine disease? Hum. Mol. Genet., 8, 2047–2053.
(1997) Human chromosomal fragile site FRA16B is an amplified AT-rich 40. Zuhlke, C., Hellenbroich, Y., Dalski, A., Kononowa, N., Hagenah, J.,
minisatellite repeat. Cell, 88, 367–374. Vieregge, P., Riess, O., Klein, C. and Schwinger, E. (2001) Different
19. Hewett, D.R., Handt, O., Mangelsdorf, M., Hobson, L., Eyre, H., Baker, E., types of repeat expansion in the TATA-binding protein are associated
Sutherland, G.R., Schuffenhauer, S., Mao, J. and Richards, R.I. (1998) with a new form of inherited ataxia. Eur. J. Hum. Genet., 9, 160–164.
Structure of FRA10B reveals common elements in repeat expansion and 41. Lunkes, A. and Mandel, J.-L. (1997) Polyglutamines, nuclear inclusions
chromosomal fragile genesis. Mol. Cell, 1, 773–781. and neurodegeneration. Nat. Med., 3, 1201–1202.
20. Kunst, C.B. and Warren, S.T. (1994) Cryptic and polar variation of the fragile 42. Klement, I.A., Skinner, P.J., Kaytor, M.D., Yi, H., Hersch, S.M.,
X repeat could result in predisposing normal alleles. Cell, 77, 853–861. Clark, H.B., Zoghbi, Y.H. and Orr, H.T (1998) Ataxin-1 nuclear
21. Crawford, D.C., Zhang, F., Wilson, B., Warren, S.T. and Sherman, S.L. localisation and aggregation: role in polyglutamine-induced disease in
(2000) Survey of the Fragile X Syndrome CGG repeat and the short-tandem SCA1 transgenic mice. Cell, 95, 41–53.
repeat and single-nucleotide-polymorphism haplotypes in an African 43. Gutekunst, C.A., Li, S.-H., Yi, H., Mulroy, J.S., Kuermmerle, S.,
American population. Am. J. Hum. Genet., 66, 480–493. Jones, R., Rye, D., Ferrante, R.J., Hersch, S.M. and Li, X.J. (1999)
22. Larsen, L.A., Armstrong, J.S., Gronskov, K., Hjalgrim, H., Macpherson, J.N., Nuclear and neutrophil aggregates in Huntington’s disease: relationship to
Brondum-Nielsen, K., Hasholt, L., Norgaard-Pedersen, B. and Vuust, J. neuropathology. J. Neurosci., 19, 2522–2534.
(2000) Haplotype and AGG-interspersion analysis of FMR1 (CGG)n 44. Kuermmerle, S., Gutekunst, C.A., Klein, A.M., Li, X.J., Li, S.H.,
alleles in the Danish population: Implications for multiple mutational Beal, M.F., Hersch, S.M. and Ferrante, R.J. (1999) Huntington aggregates
pathways towards Fragile X alleles. Am. J. Med. Genet., 93, 99–106. may not predict neuronal death in Huntington’s disease. Ann. Neurol., 46,
23. Sarkar, P.S., Chang, H.C., Boudi, F.B. and Reddy, S. (1998) CTG repeats 842–849.
show bimodal amplification in E. coli. Cell, 95, 531–540. 45. Sisodia, S.S. (1998) Nuclear inclusions in glutamine repeat disorders: are
24. Spiro, C., Pelletire, R., Rolfsmeier, M.L., Dixon, M.J., Lahue, R.S., they pernicious, coincidental or beneficial? Cell, 95, 1–4.
Gupta, G., Park, M.S., Chen, X., Mariappan, S.V. and McMurray, C.T. 46. Restituito, S., Thompson, R.M., Eliet, J., Raike, R.S., Riedl, M.,
(1999) Inhibition of FEN-1 processing by DNA secondary structure at Charnet, P. and Gomez, C.M. (2000) The polyglutamine expansion in
trinucleotide repeats. Mol. Cell, 4, 1079–1085. spinocerebellar ataxia type 6 causes a β subunit-specific enhanced
25. Freudenreich, C.H., Kantrow, S.M. and Zakian, V.A. (1998) Expansion and activation of P/Q-type calcium channels in Xenopus oocytes. J. Neurosci.,
length-dependent fragility of CTG repeats in yeast. Science, 279, 853–856. 20, 6394–6403.
47. Holmes, S.E., O’Hearn, E.E., McInnes, M.G., Gorelick-Feldman, D.A.,
26. Kennedy, L. and Shelbourne, P.F. (2000) Dramatic mutation instability in
Kleiderlein, J.J., Callahan, C., Kwak, N.G., Ingersoll-Ashworth, R.G.,
HD mouse striatum: does polyglutamine load contribute to cell-specific
Sherr, M., Sumner, A.J. et al. (1999) Expansion of a novel CAG
vulnerability in Huntington’s disease? Hum. Mol. Genet., 9, 2539–2544.
trinucleotide repeat in the 5′ region of PPP2R2B is associated with
27. Kotvun, I.V. and McMurray, C.T. (2001) Trinucleotide expansion in
SCA12. Nat. Genet., 23, 391–392.
haploid germ cells by gap repair. Nat. Genet., 27, 407–411.
48. Fujigasaki, H., Verma, I.C., Camuzat, A., Margolis, R.L., Zander, C.,
28. Citterio, E., Vermeulen, W. and Hoeijmakers, J.H. (2000) Transcriptional Lebre, A.S., Jamot, L., Saxena, R., Anand, I., Holmes, S.E. et al. (2001)
healing. Cell, 26, 447–450. SCA12 is a rare locus for autosomal dominant cerebellar ataxia: a study of
29. Yu, A., Fan, H.Y., Liao, D., Bailey, A.D. and Weiner, A.M. (2000) an Indian family. Ann. Neurol., 49, 117–121.
Activation of p53 or loss of the Cockayne syndrome group B repair 49. Stevanin, G., Herman, A., Durr, A., Jodice, C., Frontali, M., Agid, Y. and
protein causes metaphase fragility of human U1, U2 and 5S genes. Brice, A. (2000) Are (CTG)n expansions at the SCA8 locus rare
Mol. Cell, 5, 801–810. polymorphisms? Nat. Genet., 24, 213.
30. Manley, K., Shirley, T.L., Flaherty, L. and Messer, A. (1999) Msh2 50. Worth, P.F., Houlden, H., Giunti, P., Davis, M.B. and Wood, N.W. (2000)
deficiency prevents in vivo somatic instability of the CAG repeat in Large, expanded repeats in SCA8 are not confined to patients with
Huntington disease transgenic mice. Nat. Genet., 23, 471–473. cerebellar ataxia. Nat. Genet., 24, 214–215.
31. Martorell, L., Monckton, D.G., Garnez, J., Johnson, K.J., Gich, I., 51. Mosely, M.L., Schut, L.J., Bird, T.D., Day, J.W. and Ranum, L.P.W.
Lopez de Munain, A. and Baiget, M. (1998) Progression of somatic CTG (2000) Reply. Nat. Genet., 24, 215.
repeat length heterogeneity in the blood cells of myotonic dystrophy 52. Sieradzan, K.A. and Mann, D.M.A. (2001) The selective vulnerability of
patients. Hum. Mol. Genet., 7, 307–312. nerve cells in Huntington’s disease. Neuropathol. Appl. Neurobiol., 27, 1–21.
32. Khajavi, M., Tari, A.M., Patel, N.B., Tsuji, K., Siwak, D.R., 53. Telenius, H., Kremer, B., Goldberg, Y.P., Theilmann, J., Andrew, S.E.,
Meistrich, M.L., Terry, N.H.A. and Ashizawa, T. (2001) ‘Mitotic drive’ Zeisler, J., Adam, S., Greenberg, C., Ives, E.J., Clarke, L.A. et al. (1994)
of expanded CTG repeats in myotonic dystrophy type 1 (DM1). Somatic and gonadal mosaicism of the Huntington disease gene CAG
Hum. Mol. Genet., 10, 855–863. repeat in brain and sperm. Nat. Genet., 6, 409–414.
33. Pieretti, M., Zhang, F.P., Fu, Y.-H., Warren, S.T., Oostra, B.A., 54. Aronin, N., Chase, K., Young, C., Sapp, E., Schwarz, C., Matta, N.,
Caskey, C.T. and Nelson, D.T. (1991) Absence of expression of FMR-1 Kornreich, R., Landwehrmeyer, B., Bird, E., Beal, M.F. et al. (1995) CAG
gene in fragile X syndrome. Cell, 66, 817–822. expansion affects the expression of mutant Huntingtin in the Huntington’s
34. Lalioti, M., Scott, H.S. and Antonarakis, S.E. (1999) Altered spacing of disease brain. Neuron, 15, 1193–1201.
promoter elements due to the dodecamer repeat expansion contributes to 55. Cummings, C.J., Reinstein, E., Sun, Y., Antalffy, B., Jiang, Y., Ciechanover, A.,
reduced expression of the cystatin B gene in EPM1. Hum. Mol. Genet., 8, Orr, H.T., Beaudet, A.L. and Zoghbi, H.Y. (1999) Mutation of the E6-AP
1791–1798. ubiquitin ligase reduces nuclear inclusion frequency while accelerating
35. Sakamoto, N., Ohshima, K., Montermini, L., Pandolfo, M. and polyglutamine-induced pathology in SCA1 mice. Neuron, 24, 879–892.
Wells, R.D. (2001) Sticky DNA, a self-associated complex formed at long 56. Ishikawa, K., Fujigasaki, H., Saegusa, H., Ohwada, K., Fujita, T.,
GAA.TTC repeats in intron 1 of the frataxin gene, inhibits transcription. Iwamoto, H., Komatsuzaki, Y., Toru, S., Toriyama, H., Watanabe, M.
J. Biol. Chem., 276, 27171–27177. et al. (1999) Abundant expression and cytoplasmic aggregations of α1A
36. Sarkar, P.S., Appukuttan, B., Han, J., Ito, Y., Ai, C., Tsai, W., Chai, Y., voltage-dependent calcium channel protein associated with neurodegeneration
Stout, J.T. and Reddy, S. (2000) Heterozygous loss of Six5 in mice is in spinocerebellar ataxia type 6. Hum. Mol. Genet., 8, 1185–1193.
sufficient to cause ocular cataracts. Nat. Genet., 25, 110–114. 57. Huynh, D.P., Figueroa, K., Hoang, N. and Pulst, S.-M. (2000) Nuclear
37. Klesert, T.R., Cho, D.R., Clark, J.I., Maylie, J., Adelman, J., Snider, L., localisation or inclusion body formation of ataxin-2 are not necessary for
Yuen, E.C., Soriano, P. and Tapscott, S.J. (2000) Mice deficient in Six5 SCA2 pathogenesis in mouse or human. Nat. Genet., 26, 44–50.
2194 Human Molecular Genetics, 2001, Vol. 10, No. 20

58. Yamamoto, A., Lucas, J.J. and Hen, R. (2000) Reversal of neuropathology disequilibrium between the myotonic dystrophy locus and a new
and motor dysfunction in a conditional model of Huntington’s disease. polymorphic DNA marker. Am. J. Hum. Genet., 49, 68–75.
Cell, 101, 57–66. 72. Imbert, G., Kretz, C., Johnson, K. and Mandel, J.L. (1993) Origin of the
59. Marsh, J.L., Walker, H., Theisen, H., Zhu, Y.-Z., Fielder, T., Purcell, J. expansion mutation in myotonic dystrophy. Nat. Genet., 4, 72–76.
and Thompson, L.M. (2000) Expanded polyglutamine peptides alone are 73. Richards, R.I., Holman, K., Friend, K., Kremer, E., Hillen, D., Staples, A.,
intrinsically cytotoxic and cause neurodegeneration in Drosophila. Brown, W.T., Goonewardena, P., Tarleton, J., Schwartz, C. et al. (1992)
Hum. Mol. Genet., 9, 13–25. Evidence of founder chromosomes in fragile X syndrome. Nat. Genet., 1,
60. Ordway, J.M., Tallaksen-Greene, S., Gutekunst, C.-A., Berstein, E.M., 257–260.
Cearley, J.A., Wiener, H.W., Dure, L.S., Lindsey, R., Hersch, S.M., 74. MacDonald, M.E., Novelletto, A., Lin, C., Tagle, D., Barnes, G.,
Jope, R.S. et al. (1997) Ectopically expressed CAG repeats cause
Bates, G., Taylor, S., Allitto, B., Altherr, M., Myers, R. et al. (1992) The
intranuclear inclusions and a progressive late onset neurological
Huntington’s disease candidate region exhibits many different haplotypes.
phenotype in the mouse. Cell, 91, 753–763.
Nat. Genet., 1, 99–103.
61. Fernandez-Funez, P., Nino-Rosales, M.L., de Gouyan, B., She, W.C.,
Luchak, J.M., Martinez, P., Turiegano, E., Benito, J., Capovilla, M., Skinner, 75. Chung, M.Y., Ranum, L.P., Duvick, L.A., Servadio, A., Zoghbi, H.Y. and
P.J. et al. (2000) Identification of genes that modify ataxin-1-induced Orr, H.T. (1993) Evidence for a mechanism predisposing to intergenerational
neurodegeneration. Nature, 408, 101–106. CAG repeat instability in spinocerebellar ataxia type I. Nat. Genet., 5,
62. Chan, H.Y.E., Warrick, J.M., Gray-Board, G.L., Paulson, H.L. and 254–258.
Bonini, N.M. (2000) Mechanisms of chaperone suppression of 76. Takiyama, Y., Igarashi, S., Rogaeva, E.A., Endo, K., Rogaev, E.I.,
polyglutamine disease: selectivity, synergy and modulation of protein Tanaka, H., Sherrington, R., Sanpei, K., Liang, Y. and Saito M. (1995)
solubility in Drosophila. Hum. Mol. Genet., 9, 2811–2820. Evidence for inter-generational instability in the CAG repeat in the MJD1
63. Penrose, L.S. (1948) The problem of anticipation in pedigrees of gene and for conserved haplotypes at flanking markers amongst Japanese
dystrophia myotonica. Ann. Eugenics, 14, 125–132. and Caucasian subjects with Machado-Joseph disease Hum. Mol. Genet.,
64. Tiscornia, G. and Mahadevan, M.S. (2000) Myotonic dystrophy: the role 4, 1137–1146.
of the CUG triplet repeats in splicing of a novel DMPK exon and altered 77. Tanaka, F., Doyu, M., Ito, Y., Matsumoto, M., Mitsuma, T., Abe, K.,
cytoplasmic DMPK mRNA isoform ratios. Mol. Cell, 5, 959–967. Aoki, M., Itoyama, Y., Fischbeck, K.H. and Sobue G. (1996) Founder
65. Miller, J.W., Urbanti, C.R., Teng-umnuay, P., Stenberg, M.G., effect in spinal and bulbar muscular atrophy (SBMA). Hum. Mol. Genet.,
Byrne, B.J., Thornton, C.A. and Swanson, M.S. (2000) Recruitment of 5, 1253–1257.
human muscleblind proteins to (CUG)n expansions associated with 78. Yanagisawa, H., Fujii, K., Nagafuchi, S., Nakahori, Y., Nakagome, Y.,
myotonic dystrophy EMBO J., 19, 4439–4448. Akane, A., Nakamura, M., Sano, A., Komure, O., Kondo, I. et al. (1996)
66. Mankodi, A., Logigian, E., Callahan, L., McClain, C., White, R., A unique origin and multistep process for the generation of expanded
Henderson, D., Krym, M. and Thornton, C.A. (2000) Myotonic dystrophy DRPLA triplet repeats. Hum. Mol. Genet., 5, 373–379.
in transgenic mice expressing an expanded CUG repeat. Science, 289, 79. Cossee, M., Schmitt, M., Campuzano, V., Reutenauer, L., Moutou, C.,
1769–1772. Mandel, J.L. and Koenig, M. (1997) Evolution of the Friedreich’s ataxia
67. Kirby, R.J., Hamilton, G.M., Finnegan, D.J., Johnson, K.J. and trinucleotide repeat expansion: founder effect and premutations.
Jarman, A.P. (2001) Drosophila homolog of the myotonic dystrophy- Proc. Natl Acad. Sci. USA, 94, 7452–7457.
associated gene, SIX5, is required for muscle and gonad development.
80. Didierjean, O., Cancel, G., Stevanin, G., Durr, A., Burk, K., Benomar, A.,
Curr. Biol., 11, 1044–1049.
Lezin, A., Belal, S., Abada-Bendid, M., Klockgether, T. and Brice, A.
68. Knight, S.J.L., Flannery, A.V., Hirst, M.C., Campbell, L., Christodoulou, Z.,
(1999) Linkage disequilibrium at the SCA2 locus. J. Med. Genet., 36,
Phelps, S.R., Pointon, J., Middleton-Price, H.R., Barnicoat, A.,
415–417.
Pembrey, M.E. et al. (1993) Trinucleotide repeat amplification and
hypermethylation of a CpG island in FRAXE mental retardation. Cell, 74, 81. Jonasson, J., Juvonen, V., Sistonen, P., Ignatius, J., Johansson, D.,
127–134. Bjorck, E.J., Wahlstrom, J., Melberg, A., Holmgren, G., Forsgren, L. and
69. Gecz, J., Oostra, B.A., Hockey, A., Carbonell, P., Turner, G., Haan, E.A., Holmberg, M. (2000) Evidence for a common Spinocerebellar ataxia type 7
Sutherland, G.R. and Mulley, J.C. (1997) FMR2 expression in families (SCA7) founder mutation in Scandinavia. Eur. J. Hum. Genet., 8, 918–922.
with FRAXE mental retardation. Hum. Mol. Genet., 6, 435–441. 82. Takashima, M., Ishikawa, K., Nagaoka, U., Shoji, S. and Mizusawa, H.
70. Gécz, J. (2000) FMR3 is a novel gene associated with FRAXE CpG island (2001) A linkage disequilibrium at the candidate gene locus for 16q-linked
and transcriptionally silent in FRAXE full mutations. J. Med. Genet., 37, autosomal dominant cerebellar ataxia type III in Japan. J. Hum. Genet.,
782–784. 46, 167–171.
71. Harley, H.G., Brook, J.D., Floyd, J., Rundle, S.A., Crow, S., Walsh, K.V., 83. Mori, M., Adachi, Y., Kusumi, M. and Nakashima, K. (2001) Spinocerebellar
Thibault, M.C., Harper, P.S. and Shaw, D.J. (1991) Detection of linkage ataxia type 6: founder effect in Western Japan. J. Neurol. Sci., 185, 43–47.

You might also like