Bidir Transcr Repeats

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Human Molecular Genetics, 2010, Vol.

19, Review Issue 1 R77–R82


doi:10.1093/hmg/ddq132
Advance Access published on April 4, 2010

Partners in crime: bidirectional transcription


in unstable microsatellite disease
Ranjan Batra, Konstantinos Charizanis and Maurice S. Swanson ∗

Department of Molecular Genetics and Microbiology and the Genetics Institute, University of Florida,
College of Medicine, Gainesville, FL, USA

Received March 11, 2010; Revised and Accepted April 2, 2010

Nearly two decades have passed since the discovery that the expansion of microsatellite trinucleotide
repeats is responsible for a prominent class of neurological disorders, including Huntington disease and
fragile X syndrome. These hereditary diseases are characterized by genetic anticipation or the intergenera-
tional increase in disease severity accompanied by a decrease in age-of-onset. The revelation that the vari-
able expansion of simple sequence repeats accounted for anticipation spawned a number of pathogenesis
models and a flurry of studies designed to reveal the molecular events affected by these expansions. This
work led to our current understanding that expansions in protein-coding regions result in extended homopo-
lymeric amino acid tracts, often polyglutamine or polyQ, and deleterious protein gain-of-function effects. In
contrast, expansions in noncoding regions cause RNA-mediated toxicity. However, the realization that the
transcriptome is considerably more complex than previously imagined, as well as the emerging regulatory
importance of antisense RNAs, has blurred this distinction. In this review, we summarize evidence for bidir-
ectional transcription of microsatellite disease genes and discuss recent suggestions that some repeat
expansions produce variable levels of both toxic RNAs and proteins that influence cell viability, disease
penetrance and pathological severity.

INTRODUCTION transmission (6 – 11). Alternatively, RNA gain-of-function


models have been proposed for diseases where microsatellite
Microsatellites are inherently polymorphic and variations in expansions are positioned in noncoding regions. In the case
the lengths of some of these repeats result in more than 30 of myotonic dystrophy type 1 (DM1), transcription of
neurological and neuromuscular diseases (1,2). For disorders CTGexp mutations in the 3′ -untranslated region (3′ -UTR) of
where the expansion mutation occurs in the coding region, the DMPK gene gives rise to stable RNA hairpin structures
such as Huntington disease (HD) and six spinocerebellar that serve as both scaffolds and triggers which alter the activi-
ataxias (SCA1, 2, 3, 6, 7 and 17), (CAG)n expansions ties of certain alternative splicing factors, including MBNL1
(CAGexp) result in the synthesis of polyglutamine (polyQ) and CUGBP1, respectively (12,13). Lately, this distinction
tracts in their respective proteins, which accumulate in intra- between coding and noncoding regions has become blurred
cellular, and often ubiquinated, inclusions (3,4). A similar by the discovery of pervasive transcription across the human
scenario has been described for coding region GCGexp genome and the functional relevance of natural antisense
mutations, polyalanine (polyA) expression and the presence RNAs to gene expression (14– 16).
of intranuclear filamentous inclusions in oculopharyngeal
muscular dystrophy (5). Toxicity is thought to ensue by
a protein gain-of-function mechanism. For example, polyQ- Antisense regulatory issues
containing proteins have been reported to disrupt various Recent evaluations of genome-wide transcription have demon-
cellular pathways, including transcriptional regulation, the strated that antisense transcripts can be detected for many
ubiquitin proteasome system, autophagy and synaptic mammalian genes, including the wild-type alleles associated

To whom correspondence should be addressed at: Department of Molecular Genetics and Microbiology, Cancer Genetics Research Complex,
University of Florida, 2033 Mowry Rd, Gainesville, FL, USA 32610-3610, USA. Tel: +1 3522738076; Fax: +1 3522738284; Email: mswanson@
ufl.edu

# The Author 2010. Published by Oxford University Press. All rights reserved.
For Permissions, please email: journals.permissions@oxfordjournals.org
R78 Human Molecular Genetics, 2010, Vol. 19, Review Issue 1

Table 1. Bidirectional transcription in microsatellite disease genes

Disease Gene(s) Repeat normal/expanded Current pathogenic Sense/antisense


mechanisma readsb

Dentorubral pallidoluysian atrophy (DRPLA) DRPLA (CAG)3 – 36/49 – 88 Protein GOF (polyQ) 741/27
Fragile X syndrome (FRAXA) FMR1 sense (CGG)6 – 52/230 to .2000 Protein LOF (FMRP) 227/1
Fragile X-associated tremor/ataxia syndrome (FXTAS) ASFMR1, FMR4 antisense (CGG)6 – 52/60 – 200 RNA GOF (CGGexp)
Friedreich ataxia (FRDA) FXN (GAA)6 – 32/.200 Protein LOF (frataxin) 102/3
Huntington disease (HD) HTT (CAG)6 – 35/36 – 121 Protein GOF (polyQ) 2554/43
Huntington disease-like 2 (HDL2) JPH3 (CAG)6 – 28/40 – 59 RNA GOF (CUGexp) 31/188
Myotonic dystrophy type 1 (DM1) DMPK (CTG)5 – 37/50 to .3500 RNA GOF (CUGexp) 483/27
Myotonic dystrophy type 2 (DM2) ZNF9 (CCTG)10 – 26/75 to 11 000 RNA GOF (CCUGexp) 2091/0
Oculopharyngeal muscular dystrophy (OPMD) PABPN1 (GCG)10/12 – 17 Protein GOF (polyA) 292/1
Spinal and bulbar muscular atrophy (SBMA) AR (CAG)9 – 36/40 – 55 Protein GOF (polyQ) 103/3
Spinocerebellar ataxia type 1 (SCA1) ATXN1 (CAG)6 – 39/39 – 81 Protein GOF (polyQ) 2541/59
Spinocerebellar ataxia type 2 (SCA2) ATXN2 (CAG)13 – 33/.34 Protein GOF (polyQ) 759/13
Spinocerebellar ataxia type 3 (SCA3) ATXN3 (CAG)13 – 44/.55 Protein GOF (polyQ) 177/20
Spinocerebellar ataxia type 6 (SCA6) CACNA1A (CAG)4 – 18/20 – 29 Protein GOF/LOF (?) 242/48
Spinocerebellar ataxia type 7 (SCA7) ATXN7 (CAG)4 – 35/37 – 306 Protein GOF (polyQ) 1313/34
Spinocerebellar ataxia type 8 (SCA8) ATXN8 sense (CTG),50/74 – 1300 Protein GOF (polyQ)sense 95/4
ATXN8OS antisense RNA GOF (CUGexp)antisense
Spinocerebellar ataxia type 10 (SCA10) ATXN10 (ATTCT)10 – 29/800 – 4500 Unknown 1560/9
Spinocerebellar ataxia type 12 (SCA12) PPP2R2B (CAG)4 – 32/51 – 78 Unknown 264/35
Spinocerebellar ataxia type 17 (SCA17) TBP (CAG)25 – 42/47 – 63 Protein GOF (polyQ) 66/2
Spinocerebellar ataxia type 31 (SCA31) TK2 (TGGAA)0/≥110 RNA GOF (UGGAAexp) 412/4

Twenty microsatellite expansion diseases are listed. For a complete list see Ref. (1).
a
Either protein or RNA gain-of-function (GOF) or protein loss-of-function (LOF) with the toxic (GOF) or affected (LOF) moiety indicated in parentheses (e.g.
polyQ). For SCA8 and FRAXA/FXTAS, the corresponding antisense and sense genes are indicated (superscript).
b
Cumulative number of sequencing reads from ASSAGE study of non-expanded alleles from normal peripheral blood mononuclear cells (PBMC) and four cell
lines (Jurkat, T cell leukemia; HCT116, colorectal cancer; MiaPaCa2, pancreatic cancer; MRC5, normal lung fibroblast) from Ref. (14). Since none of these cells
have a neuronal origin, it is important to note that antisense expression of these genes in the CNS may vary significantly from these values.

with microsatellite expansion diseases [Table 1 (14,17– 20)]. X-associated tremor/ataxia syndrome (FXTAS) are two dis-
In most cases, the major sense strand is protein coding and eases caused by variations in a CGG†CCG expansion at the
is generally present at a higher copy level than the correspond- FMR1/ASFMR1,FMR4 locus. While loss of FMR1 expression
ing antisense transcript(s) which is noncoding and only par- clearly underlies FRAXA, potential roles for the antisense
tially overlaps the sense transcriptional unit, often in the ASFMR1 and FMR4 transcripts in FXTAS will be reviewed.
promoter and 3′ terminal regions. Antisense transcription has Finally, we evaluate the curious molecular etiology of
been suggested to regulate gene expression at multiple HD-like 2 (HDL2), where the current evidence indicates that
levels, including (i) transcriptional interference of the sense this disease is associated with a CTG expansion in the JPH3
strand due to potential collisions between RNA polymerase gene and CUGexp RNA toxicity. Given the clinical similarity
II holoenzyme complexes transcribing opposite strands; (ii) of HDL2 to HD and the overwhelming evidence for a
modulation of gene expression by recruitment of various chro- primary role of polyQ toxicity in the latter disease, it is sur-
matin remodeling complexes which mediate histone modifi- prising that HDL2 CAGexp antisense transcripts have not
cations and DNA methylation; (iii) masking splice and been detected in the brain.
polyadenylation sites of pre-mRNAs transcribed from sense Prior to beginning our discussion, it is important to note that
strands leading to alternative splice, and poly(A), site selec- any potential pathological impacts of antisense transcription
tion; (iv) duplex formation between sense and antisense tran- likely depend on several parameters, including the genomic
scripts that induces RNA editing and modulates RNA organization of the disease locus, spatial and temporal
interference, stability and translation (21– 23). The expansion expression patterns of the affected genes, repeat sequence
of microsatellite repeats could disrupt these regulatory steps and repeat length together with location in the gene(s). In
by altering the relative levels of sense versus antisense tran- addition, the expression of neighboring genes may also be
scripts or by affecting transcriptional start or termination sites. altered which could be particularly relevant to disease loci
In the following sections, we focus on three neurological where gene density is high, such as the DMPK gene in DM1
disorders where bidirectional transcription has been implicated (24 –27).
in disease pathogenesis and progression although antisense
RNAs have been reported at other microsatellite expansion
loci (2,13). First, we review progress on spinocerebellar SCA8: a toxic cocktail of polyQ protein and CUGexp RNA?
ataxia type 8 (SCA8) where both a polyQ protein produced For the majority of unstable microsatellite disease genes, it is
from the sense strand and CUGexp antisense RNAs have unknown if antisense RNAs are transcribed through the
been suggested to play significant roles in pathogenesis. In expanded repeat regions (14,20). An exception is SCA8, a
contrast to SCA8, fragile X syndrome (FRAXA) and fragile dominant SCA with reduced penetrance, where there is evi-
Human Molecular Genetics, 2010, Vol. 19, Review Issue 1 R79

dence for bidirectional transcription that results in the pro-


duction of toxic noncoding RNAs from the antisense transcrip-
tional unit (ATXN8OS) and a nearly pure polyQ protein, ataxin
8, from ATXN8 (18). Unlike other forms of spinocerebellar
ataxia that contain CAGexp in the coding region of the sense
strand, SCA8 was initially thought to be an exclusively RNA-
mediated disease, similar to DM1, because the only detectable
RNAs were noncoding ATXN8OS transcripts containing a
CUGexp near the 3′ end (28). As mentioned previously,
RNA toxicity in the neuromuscular disease DM1 is caused
by transcription of a noncoding CTGexp in the DMPK gene.
Transcription results in CUGexp RNAs that effectively trap
the MBNL1 splicing factors and alter alternative exon use
during RNA processing to favor the production of fetal iso-
forms (12,13). These CUGexp RNA – MBNL1 complexes
accumulate in nuclear RNA foci that may be the more toxic
entity or simply a protective response (29).
The possibility that SCA8 is also caused by this type of
CUGexp toxicity mechanism is supported by several studies.
Transgenic expression of either the wild type or an expanded
human SCA8 gene in Drosophila causes progressive retinal Figure 1. Potential roles for bidirectional transcription in SCA8 pathogenesis.
neurodegeneration which is enhanced by muscleblind (the fly Transcription of the sense (ATXN8) strand (blue ribbon) results in the synthesis
homologue of MBNL1) loss-of-function mutations (30). of the ataxin8 polyQ (PolyQ) protein from the CAGexp transcript (blue line).
Moreover, CUGexp foci are detectable in the cerebellar While a direct role for ataxin8 has not been demonstrated, polyQ (green)
expression in other microsatellite expansion disorders, such as Huntington
cortex (Purkinje cells, Bergmann glia, molecular layer inter- disease, disrupts multiple regulatory pathways (transcription, ubiquitin protea-
neurons) of both SCA8 patients and transgenic mice that somal system, autophagy, synaptic transmission). In contrast, ATXN8OS tran-
express a human AXTN8/ATXN8OS gene with a (CTG)116 scription produces a toxic CUGexp RNA (red line, RNA stem-loop with U –U
expansion mutation (SCA8 BAC-exp) (31). These mutant mismatches shown as bulged triangles) that reverts RNA splicing patterns to a
mice develop a progressive neurological phenotype, striking fetal pattern due to MBNL1 sequestration and CELF1 hyperphosphorylation.
Also shown is transcriptional interference induced by divergently transcribing
gait disturbances and a decrease in cerebellar-cortical inhi- RNA polymerase II complexes (Pol II).
bition (18,31). While both low copy SCA8 BAC-exp and het-
erozygous Mbnl1 DE3/+ knockout mice are overtly normal,
crosses between these lines result in progeny with enhanced protein-mediated (polyQ intranuclear inclusions) events have
rotarod deficits. If the toxic RNA-mediated mechanism is been documented (Fig. 1). Nevertheless, the pathogenic picture
similar to DM1, then the reduction in cerebellar inhibition is incomplete. For sense transcription, it is not clear whether
could reflect altered splicing and subsequent mis-expression the levels of the ATXN8 transcription and polyQ protein, which
of a protein involved in cerebellar inhibitory pathways. cannot be detected by protein blot analysis, are sufficient to
Indeed, this appears to be the case since the levels of the cause cellular dysfunction although pure polyQ proteins appear
GABA transporter GAT4/GABT4 are elevated in both to be highly toxic, so susceptible cell populations may rapidly
human SCA8 and mouse SCA8 BAC-exp brains and this has be eliminated (32). For antisense transcription, expression of
been linked to MBNL1-regulated alternative splicing of CUGexp RNAs leads to RNA gain-of-function effects on RNA
GAT4 pre-mRNA. These data are consistent with a model in splicing and loss of cerebellar GABAergic inhibition but the
which mutant ATXN8OS CUGexp transcripts trap MBNL1 basis for the cerebellar atrophy observed in SCA8 remains an
proteins which in turn causes mis-splicing of MBNL target enigma since these neurodegenerative changes are not recapitu-
RNAs that encode proteins important for GABAergic lated in the SCA8 BAC-exp transgenic model. An additional
inhibition. complication is that another gene, KLHL1, is also transcribed
On the sense strand, ATXN8 transcription was initially missed in an antisense direction to the ATXN8OS gene and the 5′ ends
in human brain samples perhaps due to CAGexp amplification of these genes overlap in a head-to-head configuration (33).
problems (18). Fortunately, ATXN8 RNAs are detectable by Because antisense transcription could theoretically down-
strand-specific RT–PCR using primers immediately down- regulate expression of the opposite strand, ATXN8OS CTGexp
stream of the expansion in the mouse SCA8 BAC-exp model. transcripts might decrease KLHL1 mRNA levels in SCA8. In
Surprisingly, and unlike other CAGexp expansion diseases, this agreement with this possibility, Klhl1 knockout mice show
RNA is predicted to encode a nearly pure polyQ protein (18). mild atrophy of the molecular layer in the cerebellum, a charac-
Cerebellar polyQ inclusions, detectable by monoclonal antibody teristic feature of SCA8 patients that is missing in the transgenic
(mAb) 1C2 that recognizes expanded polyQ tracts, are observed SCA8 BAC-exp model. Confirmation of this antisense
in Purkinje cells in human SCA8 and SCA8 BAC-exp mouse loss-of-function model, which requires that KLHL1 gene
brains and brain stems, indicating that antisense transcription expression is reduced in SCA8 brains, has not been reported.
occurs through these microsatellite repeats. Thus, SCA8 is the One of the puzzling features of SCA8 is the low penetrance
first candidate for a microsatellite disease in which both associated with this disease and reports that some unaffected
RNA-mediated (RNA foci formation and mis-splicing) and individuals harbor ATXN8/ATXN8OS alleles with as many as
R80 Human Molecular Genetics, 2010, Vol. 19, Review Issue 1

1000 CTG repeats (34 –36). Perhaps bidirectional transcrip- Antisense transcription may also play a role in FXTAS
tion through repeats leads to silencing of the genes at this pathogenesis (19). A promoter located in the intron 2 of
locus in some individuals significantly lowering the levels of FMR1 drives expression of the antisense ASFMR1 gene
toxic RNAs and proteins. However, there are no reports of het- through the CCG repeat region. ASFMR1 transcript levels
erochromatin formation at this locus. Are ATXN8 and increase in blood cells from FXTAS patients and the
ATXN8OS coexpressed in the same cells and, if they are, ASFMR1 transcript is spliced, polyadenylated, transported to
what are their relative expression levels? Laser microdissec- cytoplasm and may be translated into a polyproline protein
tion of specific cell types and strand-specific RT –PCR could using an initiation codon upstream of the CUGexp although
be used to address this important question. Moreover, there is no evidence that this occurs (19). Multiple alterna-
additional cell and animal model systems need to be devel- tively spliced forms of ASFMR1 RNA exist and the CGGexp
oped to determine the relative toxicity thresholds of mutation alters the splicing pattern. Another noncoding
rCUGexp RNA and polyQ. gene, FMR4, initiates upstream of FMR1 start site (48).
These two genes appear to share a bidirectional promoter
since both transcriptional units are silenced in fragile X, but
Bidirectional transcription in FXTAS upregulated in FXTAS, patient leukocytes. Of course, these
In DM1, the most severe form of the disease is present at birth antisense repeats could also fold into RNA hairpins that
and is caused by very large CTGexp (usually .1000 repeats) sequester rCCG-binding proteins.
mutations while shorter expansions are associated with the
adult-onset degenerative disease. A similar scenario occurs
for CGGexp mutations in the 5′ -UTR of the FMR1 gene (37). HDL2: one frame is not enough
FMR1 expansions cause two distinct diseases depending Huntington disease-like 2 (HDL2) is another autosomal domi-
upon repeat length (38). In the normal population, FMR1 nant and progressive neurodegenerative disease caused by a
alleles contain 6 – 52 CGG repeats and larger expansions CTG†CAG expansion mutation which, like HD, results in
(.200 repeats) cause FMR1 transcriptional silencing and the prominent loss of striatal neurons (49– 51). In contrast to
neurodevelopmental disorder fragile X syndrome. Although HD, the HDL2 sense transcript, encoded by the Junctophilin-3
longer repeats (55 – 200) were originally considered to be pre- (JPH3) gene, contains a CTGexp that is located in a 3′ -UTR or
mutations, these intermediate expansions are now known to in a coding region to produce either a polyalanine or polyleu-
cause a late adult-onset disorder, fragile X tremor/ataxia syn- cine protein depending on the splicing of JPH3 exon 2A that
drome (FXTAS) (39). Clinical manifestations of FXTAS contains three alternative 3′ splice sites upstream of the CTG
include cerebellar degeneration with loss of Purkinje cells, repeat. HDL2 expansions are relatively short (40 – 59 repeats)
spongiosis of the deep cerebellar white matter and axonal which probably reflects the location of the repeats in both
swelling. The primary neuropathological hallmark is eosino- coding and noncoding regions. While a combination of
philic ubiquitin-positive intranuclear inclusions disseminated protein- and RNA gain-of-function mechanisms could
through the brain and spinal column. explain HDL2 pathogenesis, the location of the CTGexp in a
Several lines of evidence suggest that RNA-mediated protein noncoding region suggests an RNA-mediated mechanism.
sequestration events similar to those observed in DM1 and Indeed, CUGexp RNA foci are detectable in cortical neurons
SCA8 may be important for FXTAS pathogenesis. FMR1 of HDL2 postmortem brain and these foci colocalize with
mRNA levels increase 2–10-fold in FXTAS and these RNAs MBNL1 (50). Mis-splicing of several MBNL1 targets,
can be detected in the ubiquitin-positive intranuclear inclusions MAPT exon 2 and APP exon 7, in the HDL2 brain has also
that have been reported to also contain an array of proteins, been reported.
including lamin A, Pura, hnRNPA2/B1, MBNL1, HSP40 and Because of the similarities between HD and HDL2, protein
the 20S proteasome complex (40,41). In support of the idea gain-of-function mechanisms have also been evaluated. One
that some of these proteins are effectively trapped by obvious candidate would be a polyQ protein expressed from
rCGGexp inclusions, overexpression of hnRNP A2/B1 or Pura JPH3 antisense CAGexp transcripts. In support of this possi-
suppresses neurodegeneration in a Drosophila (CGG)90-EGFP bility, the 1– 2 mm ubiquitin-positive peri-nucleolar inclusions
transgenic model and Pura knockout mice develop tremors observed in HDL2 brains are also recognized by the mAb 1C2,
and seizures by 4 weeks of age, suggesting an essential role which was originally elicited against the TATA-box-binding
of Pura in brain development and function (42–44). protein (TBP) that contains a polyQ tract (51). An argument
FXTAS-associated CGGexp repeats are also pathogenic inde- against a role for JPH3 antisense transcription is that it has
pendent of gene context. Mouse Fmr1 (CGG)98 and not been possible to detect CAGexp transcripts from HDL2
(CGG)120 knockin models have been reported which repro- brains using RT – PCR strategies and mAb 1C2 has been
duce several features of FXTAS (45,46). Recently, transgenic reported to cross-react with polyalanine and polyleucine,
mice have been generated in which expression of a (CGG)90 which are encoded on the sense transcript (50,51). However,
repeat upstream of either Fmr1 or EGFP is driven by the Pur- another study failed to confirm that 1C2 recognizes either
kinje cell-specific L7 promoter (47). In contrast to control polyalanine or polyleucine (18) and anti-JPH3 antibodies do
lines expressing the corresponding transgenes minus CGGexp, not detect expanded polyalanine or polyleucine proteins in
both Fmr1- and EGFP-driven (CGG)90 lines develop ubiqui- the HDL2 brain (51). Another confounding factor is that
nated intranuclear inclusions, Purkinje cell axonal swelling TBP is detectable in HDL2 intranuclear inclusions (51).
and increased neuronal loss. Interestingly, hnRNP A2/B1 and Similar to SCA8, these 1C2-positive inclusions do not gener-
Pura do not colocalize with these CGGexp inclusions. ally colocalize with nuclear RNA foci in the HDL2 brain so
Human Molecular Genetics, 2010, Vol. 19, Review Issue 1 R81

sense or antisense transcripts may not be expressed at similar et al. (2010) Early increase in extrasynaptic NMDA receptor signaling and
levels in the same cell populations or the combination of toxic expression contributes to phenotype onset in Huntington’s disease mice.
Neuron, 65, 178 –190.
RNA and protein leads to under-representation due to rapid 11. Okamoto, S., Pouladi, M.A., Talantova, M., Yao, D., Xia, P., Ehrnhoefer,
cell death. Nevertheless, the issue of JPH3 antisense transcrip- D.E., Zaidi, R., Clemente, A., Kaul, M., Graham, R.K. et al. (2009)
tion in the brain should be re-investigated using RNA-seq Balance between synaptic versus extrasynaptic NMDA receptor activity
technologies particularly in light of a recent asymmetric influences inclusions and neurotoxicity of mutant huntingtin. Nat. Med.,
15, 1407–1413.
strand-specific analysis of gene expression (ASSAGE) 12. Cooper, T.A., Wan, L. and Dreyfuss, G. (2009) RNA and disease. Cell,
report. Using normal blood mononuclear cells as well as 136, 777–793.
several cell lines, this study demonstrated that JPH3 antisense 13. O’Rourke, J.R. and Swanson, M.S. (2009) Mechanisms of RNA-mediated
transcripts exist and indeed are more abundantly expressed disease. J. Biol. Chem., 284, 7419– 7423.
(6-fold more sequencing reads) than sense RNAs [Table 1 14. He, Y., Vogelstein, B., Velculescu, V.E., Papadopoulos, N. and Kinzler,
K.W. (2008) The antisense transcriptomes of human cells. Science, 322,
(14,20)]. 1855–1857.
15. Faghihi, M.A. and Wahlestedt, C. (2009) Regulatory roles of natural
antisense transcripts. Nat. Rev. Mol. Cell. Biol., 10, 637 –643.
Perspective 16. Taft, R.J., Pang, K.C., Mercer, T.R., Dinger, M. and Mattick, J.S. (2010)
Non-coding RNAs: regulators of disease. J. Pathol., 220, 126– 139.
While contemporary studies on SCA8, FXTAS and HDL2 17. Cho, D.H., Thienes, C.P., Mahoney, S.E., Analau, E., Filippova, G.N. and
have revealed the potential for bidirectional transcription to Tapscott, S.J. (2005) Antisense transcription and heterochromatin at the
play a formidable role in the pathogenesis of these diseases, DM1 CTG repeats are constrained by CTCF. Mol. Cell, 20, 483– 489.
many fundamental questions remain. Does repeat length influ- 18. Moseley, M.L., Zu, T., Ikeda, Y., Gao, W., Mosemiller, A.K., Daughters,
ence bidirectional expression of the affected locus? Is it poss- R.S., Chen, G., Weatherspoon, M.R., Clark, H.B., Ebner, T.J. et al. (2006)
Bidirectional expression of CUG and CAG expansion transcripts and
ible that toxic RNAs and proteins are coordinately produced intranuclear polyglutamine inclusions in spinocerebellar ataxia type 8.
from the same expansion mutation or do some cell types pre- Nat. Genet., 38, 758 –769.
ferentially synthesize one over the other? Are cells that 19. Ladd, P.D., Smith, L.E., Rabaia, N.A., Moore, J.M., Georges, S.A.,
produce both toxic RNAs and proteins particularly susceptible Hansen, R.S., Hagerman, R.J., Tassone, F., Tapscott, S.J. and Filippova,
G.N. (2007) An antisense transcript spanning the CGG repeat region of
to expansion-mediated cell death? Answers to these and many FMR1 is upregulated in premutation carriers but silenced in full mutation
related questions require the development of a new generation individuals. Hum. Mol. Genet., 16, 3174– 3187.
of cell- and animal-based models that are designed to specifi- 20. Dion, V. and Wilson, J.H. (2009) Instability and chromatin structure of
cally evaluate the potential for combinatorial protein and RNA expanded trinucleotide repeats. Trends Genet., 25, 288– 297.
toxicity induced by bidirectional transcription. 21. Lavorgna, G., Dahary, D., Lehner, B., Sorek, R., Sanderson, C.M. and
Casari, G. (2004) In search of antisense. Trends Biochem. Sci., 29, 88–94.
22. Lapidot, M. and Pilpel, Y. (2006) Genome-wide natural antisense
transcription: coupling its regulation to its different regulatory
Conflict of Interest statement. None declared. mechanisms. EMBO Rep., 7, 1216– 1222.
23. Morris, K.V., Santoso, S., Turner, A.M., Pastori, C. and Hawkins, P.G.
(2008) Bidirectional transcription directs both transcriptional gene
FUNDING activation and suppression in human cells. PLoS Genet., 4, e1000258.
24. Wang, Y.H., Amirhaeri, S., Kang, S., Wells, R.D. and Griffith, J.D. (1994)
Our research is supported by the NIH (AR046799, NS048843, Preferential nucleosome assembly at DNA triplet repeats from the
NS058901) and the MDA. myotonic dystrophy gene. Science, 265, 669– 671.
25. Otten, A.D. and Tapscott, S.J. (1995) Triplet repeat expansion in
myotonic dystrophy alters the adjacent chromatin structure. Proc. Natl
Acad. Sci. USA, 92, 5465– 5469.
REFERENCES 26. Klesert, T.R., Otten, A.D., Bird, T.D. and Tapscott, S.J. (1997)
1. Mirkin, S.M. (2007) Expandable DNA repeats and human disease. Nature, Trinucleotide repeat expansion at the myotonic dystrophy locus reduces
447, 932–940. expression of DMAHP. Nat. Genet., 16, 402–406.
2. Lopez Castel, A., Cleary, J.D. and Pearson, C.E. (2010) Repeat instability 27. Thornton, C.A., Wymer, J.P., Simmons, Z., McClain, C. and Moxley,
as the basis for human diseases and as a potential target for therapy. Nat. R.T. 3rd (1997) Expansion of the myotonic dystrophy CTG repeat reduces
Rev. Mol. Cell. Biol., 11, 165–170. expression of the flanking DMAHP gene. Nat. Genet., 16, 407–409.
3. Orr, H.T. and Zoghbi, H.Y. (2007) Trinucleotide repeat disorders. Ann. 28. Koob, M.D., Moseley, M.L., Schut, L.J., Benzow, K.A., Bird, T.D., Day,
Rev. Neurosci., 30, 575– 621. J.W. and Ranum, L.P. (1999) An untranslated CTG expansion causes a
4. Brouwer, J.R., Willemsen, R. and Oostra, B.A. (2009) Microsatellite novel form of spinocerebellar ataxia (SCA8). Nat. Genet., 21, 379–384.
repeat instability and neurological disease. Bioessays, 31, 71–83. 29. Nykamp, K.R. and Swanson, M.S. (2004) Toxic RNA in the nucleus:
5. Messaed, C. and Rouleau, G.A. (2009) Molecular mechanisms underlying unstable microsatellite expression in neuromuscular disease. Prog. Mol.
polyalanine diseases. Neurobiol. Dis., 34, 397–405. Subcell. Biol., 35, 57–77.
6. Bennett, E.J., Shaler, T.A., Woodman, B., Ryu, K.Y., Zaitseva, T.S., 30. Mutsuddi, M., Marshall, C.M., Benzow, K.A., Koob, M.D. and Rebay, I.
Becker, C.H., Bates, G.P., Schulman, H. and Kopito, R.R. (2007) Global (2004) The spinocerebellar ataxia 8 noncoding RNA causes
changes to the ubiquitin system in Huntington’s disease. Nature, 448, neurodegeneration and associates with staufen in Drosophila. Curr. Biol.,
704–708. 14, 302– 308.
7. Gil, J.M. and Rego, A.C. (2008) Mechanisms of neurodegeneration in 31. Daughters, R.S., Tuttle, D.L., Gao, W., Ikeda, Y., Moseley, M.L., Ebner,
Huntington’s disease. Eur. J. Neurosci., 27, 2803–2820. T.J., Swanson, M.S. and Ranum, L.P. (2009) RNA gain-of-function in
8. Kumari, D. and Usdin, K. (2009) Chromatin remodeling in the noncoding spinocerebellar ataxia type 8. PLoS Genet., 5, e1000600.
repeat expansion diseases. J. Biol. Chem., 284, 7413– 7417. 32. Marsh, J.L., Walker, H., Theisen, H., Zhu, Y.Z., Fielder, T., Purcell, J. and
9. Zoghbi, H.Y. and Orr, H.T. (2009) Pathogenic mechanisms of a Thompson, L.M. (2000) Expanded polyglutamine peptides alone are
polyglutamine-mediated neurodegenerative disease, spinocerebellar intrinsically cytotoxic and cause neurodegeneration in Drosophila. Hum.
ataxia type 1. J. Biol. Chem., 284, 7425– 7429. Mol. Genet., 9, 13– 25.
10. Milnerwood, A.J., Gladding, C.M., Pouladi, M.A., Kaufman, A.M., Hines, 33. Nemes, J.P., Benzow, K.A., Moseley, M.L., Ranum, L.P. and Koob, M.D.
R.M., Boyd, J.D., Ko, R.W., Vasuta, O.C., Graham, R.K., Hayden, M.R. (2000) The SCA8 transcript is an antisense RNA to a brain-specific
R82 Human Molecular Genetics, 2010, Vol. 19, Review Issue 1

transcript encoding a novel actin-binding protein (KLHL1). Hum. Mol. coupled cellular proliferation as revealed by genetic inactivation in the
Genet., 9, 1543– 1551. mouse. Mol. Cell. Biol., 23, 6857–6875.
34. Day, J.W., Schut, L.J., Moseley, M.L., Durand, A.C. and Ranum, L.P. 43. Jin, P., Duan, R., Qurashi, A., Qin, Y., Tian, D., Rosser, T.C., Liu, H.,
(2000) Spinocerebellar ataxia type 8: clinical features in a large family. Feng, Y. and Warren, S.T. (2007) Pur alpha binds to rCGG repeats and
Neurology, 55, 649– 657. modulates repeat-mediated neurodegeneration in a Drosophila model of
35. Worth, P.F., Houlden, H., Giunti, P., Davis, M.B. and Wood, N.W. (2000) fragile X tremor/ataxia syndrome. Neuron, 55, 556– 564.
Large, expanded repeats in SCA8 are not confined to patients with 44. Sofola, O.A., Jin, P., Qin, Y., Duan, R., Liu, H., de Haro, M., Nelson, D.L.
cerebellar ataxia. Nat. Genet., 24, 214– 215. and Botas, J. (2007) RNA-binding proteins hnRNP A2/B1 and CUGBP1
36. Ikeda, Y., Daughters, R.S. and Ranum, L.P. (2008) Bidirectional suppress fragile X CGG premutation repeat-induced neurodegeneration in
expression of the SCA8 expansion mutation: one mutation, two genes. a Drosophila model of FXTAS. Neuron, 55, 565–571.
Cerebellum, 7, 150–158. 45. Oostra, B.A. and Willemsen, R. (2009) FMR1: a gene with three faces.
37. Verkerk, A.J., Pieretti, M., Sutcliffe, J.S., Fu, Y.H., Kuhl, D.P., Pizzuti, Biochim. Biophys. Acta, 1790, 467–477.
A., Reiner, O., Richards, S., Victoria, M.F., Zhang, F.P. et al. (1991) 46. Wenzel, H.J., Hunsaker, M.R., Greco, C.M., Willemsen, R. and Berman,
Identification of a gene (FMR-1) containing a CGG repeat coincident with R.F. Ubiquitin-positive intranuclear inclusions in neuronal and glial cells
a breakpoint cluster region exhibiting length variation in fragile X
in a mouse model of the fragile X premutation. Brain Res., 1318,
syndrome. Cell, 65, 905–914.
155–166.
38. Oostra, B.A. and Willemsen, R. (2003) A fragile balance: FMR1
47. Hashem, V., Galloway, J.N., Mori, M., Willemsen, R., Oostra, B.A.,
expression levels. Hum. Mol. Genet., 12 (Spec No 2), R249–R257.
Paylor, R. and Nelson, D.L. (2009) Ectopic expression of CGG containing
39. Jacquemont, S., Hagerman, R.J., Leehey, M., Grigsby, J., Zhang, L.,
Brunberg, J.A., Greco, C., Des Portes, V., Jardini, T., Levine, R. et al. mRNA is neurotoxic in mammals. Hum. Mol. Genet., 18, 2443– 2451.
(2003) Fragile X premutation tremor/ataxia syndrome: molecular, clinical, 48. Khalil, A.M., Faghihi, M.A., Modarresi, F., Brothers, S.P. and Wahlestedt,
and neuroimaging correlates. Am. J. Hum. Genet., 72, 869– 878. C. (2008) A novel RNA transcript with antiapoptotic function is silenced
40. Iwahashi, C.K., Yasui, D.H., An, H.J., Greco, C.M., Tassone, F., Nannen, in fragile X syndrome. PLoS One, 3, e1486.
K., Babineau, B., Lebrilla, C.B., Hagerman, R.J. and Hagerman, P.J. 49. Greenstein, P.E., Vonsattel, J.P., Margolis, R.L. and Joseph, J.T. (2007)
(2006) Protein composition of the intranuclear inclusions of FXTAS. Huntington’s disease like-2 neuropathology. Mov. Disord., 22,
Brain, 129, 256– 271. 1416– 1423.
41. Tassone, F., Beilina, A., Carosi, C., Albertosi, S., Bagni, C., Li, L., 50. Rudnicki, D.D., Holmes, S.E., Lin, M.W., Thornton, C.A., Ross, C.A. and
Glover, K., Bentley, D. and Hagerman, P.J. (2007) Elevated FMR1 Margolis, R.L. (2007) Huntington’s disease-like 2 is associated with CUG
mRNA in premutation carriers is due to increased transcription. RNA, 13, repeat-containing RNA foci. Ann. Neurol., 61, 272–282.
555– 562. 51. Rudnicki, D.D., Pletnikova, O., Vonsattel, J.P., Ross, C.A. and Margolis,
42. Khalili, K., Del Valle, L., Muralidharan, V., Gault, W.J., Darbinian, N., R.L. (2008) A comparison of Huntington disease and Huntington-
Otte, J., Meier, E., Johnson, E.M., Daniel, D.C., Kinoshita, Y. et al. (2003) disease-like 2 neuropathology. J. Neuropathol. Exp. Neurol., 67,
Puralpha is essential for postnatal brain development and developmentally 366–374.

You might also like