ArticlePDF Available

Antisense Oligonucleotide-Mediated Terminal Intron Retention of the SMN2 Transcript

Authors:

Abstract and Figures

The severe childhood disease spinal muscular atrophy (SMA) arises from the homozygous loss of the survival motor neuron 1 gene (SMN1). A homologous gene potentially encoding an identical protein, SMN2 can partially compensate for the loss of SMN1; however, the exclusion of a critical exon in the coding region during mRNA maturation results in insufficient levels of functional protein. The rate of transcription is known to influence the alternative splicing of gene transcripts, with a fast transcription rate correlating to an increase in alternative splicing. Conversely, a slower transcription rate is more likely to result in the inclusion of all exons in the transcript. Targeting SMN2 with antisense oligonucleotides to influence the processing of terminal exon 8 could be a way to slow transcription and induce the inclusion of exon 7. Interestingly, following oligomer treatment of SMA patient fibroblasts, we observed the inclusion of exon 7, as well as intron 7, in the transcript. Because the normal termination codon is located in exon 7, this exon/intron 7-SMN2 transcript should encode the normal protein and only carry a longer 3′ UTR. Further studies showed the extra 3′ UTR length contained a number of regulatory motifs that modify transcript and protein regulation, leading to translational repression of SMN. Although unlikely to provide therapeutic benefit for SMA patients, this novel technique for gene regulation could provide another avenue for the repression of undesirable gene expression in a variety of other diseases.
Content may be subject to copyright.
Original Article
Antisense Oligonucleotide-Mediated Terminal
Intron Retention of the SMN2 Transcript
Loren L. Flynn,
1,2,4
Chalermchai Mitrpant,
2,3,4
Ianthe L. Pitout,
1,2
Sue Fletcher,
1,2
and Steve D. Wilton
1,2
1
Centre for Comparative Genomics, Murdoch University, Perth, WA, Australia;
2
Perron Institute for Neurological and Translational Science, Perth, WA, Australia;
3
Department of Biochemistry, Mahidol University, Bangkok, Thailand
The severe childhood disease spinal muscular atrophy (SMA)
arises from the homozygous loss of the survival motor neuron 1
gene (SMN1). A homologous gene potentially encoding an
identical protein, SMN2 can partially compensate for the loss
of SMN1; however, the exclusion of a critical exon in the coding
region during mRNA maturation results in insufcient levels
of functional protein. The rate of transcription is known to in-
uence the alternative splicing of gene transcripts, with a fast
transcription rate correlating to an increase in alternative
splicing. Conversely, a slower transcription rate is more likely
to result in the inclusion of all exons in the transcript. Target-
ing SMN2 with antisense oligonucleotides to inuence the pro-
cessing of terminal exon 8 could be a way to slow transcription
and induce the inclusion of exon 7. Interestingly, following
oligomer treatment of SMA patient broblasts, we observed
the inclusion of exon 7, as well as intron 7, in the transcript.
Because the normal termination codon is located in exon 7,
this exon/intron 7-SMN2 transcript should encode the normal
protein and only carry a longer 30UTR. Further studies showed
the extra 30UTR length contained a number of regulatory
motifs that modify transcript and protein regulation, leading
to translational repression of SMN. Although unlikely to pro-
vide therapeutic benet for SMA patients, this novel technique
for gene regulation could provide another avenue for the
repression of undesirable gene expression in a variety of other
diseases.
INTRODUCTION
With a frequency of 1 in 10,000 live births,
1
the neurodegenerative
disease spinal muscular atrophy (SMA) is the leading genetic cause
of infant death.
2
SMA arises from inadequate levels of the survival
motor neuron (SMN) protein that ultimately results in the death of
motor neurons. While the survival motor neuron 1 (SMN1) gene is
missing in most SMA patients, copies of the homologous gene,
SMN2, potentially compensate for SMN production
3
; however, a
C > T base change in SMN2 exon 7 results in exclusion of the exon
from 90% of neuronal SMN2 transcripts.
3,4
To date, the main RNA
therapeutic focus for SMA has been the use of antisense oligonucleo-
tides (AOs) to enhance SMN2 exon 7 inclusion and increase
SMN levels (for review, see Porensky and Burghes
5
). In particular, a
20O-methoxyethyl (MOE) AO covering the ISS-N1 splicing domain
(Anti-ISS-N1) has shown promise in clinical trials
68
and has recently
received approval by the U.S. Food and Drug Administration.
9
How-
ever, the therapy is by no means denitive, with unknown conse-
quences of long-term AO exposure and further improvements in
AO efcacy needed before this therapy can be considered a qualied
success. While other studies have focused on targeting AOs to in-
tronic splice silencing motifs to enhance exon 7 inclusion,
8,1012
AO-mediated splice modication has broader potential.
The strategy described here was focused on targeting AOs to the last
exon in an attempt to slow transcription rates and concurrent pre-
mRNA processing to temporarily stall the spliceosome machinery.
Others have shown that a slow RNA polymerase II elongation rate
during transcription can increase the window of opportunityfor
upstream splicing events, with alternative exons more likely to be
included in the mature transcript.
13,14
To determine whether slower
transcription elongation could be induced by an AO, we targeted
AOs to SMN2 exon 8 in an attempt to increase the inclusion of
SMN2 exon 7 in the transcript.
Unexpectedly, AOs targeting SMN2 exon 8 induced the retention of
exon 7 and intron 7 in the mature transcript. Interestingly, an AO
covering the exon 8 acceptor site has been reported by others to induce
exon 7 and intron 7 retention, yet this work was not pursued further.
15
Because the normal termination codon is located within exon 7, this
induced transcript should therefore encode the normal full-length
protein; however, the size of the 30UTR is increased. It is well docu-
mented that the length of the 30UTR can affect transcript stability
and protein translation, with longer 30UTRs having more opportunity
for the binding of microRNAs and regulatory elements (for review, see
Barrett et al.
16
). However, the consequences of intron retention within
the mature transcript, and more specically within the 30UTRs, are a
more recently explored and less well understood area.
A study by Braunschweig and colleagues
17
reported that three-quar-
ters of mammalian multi-exon genes exhibit intron retention within
Received 20 December 2017; accepted 25 January 2018;
https://doi.org/10.1016/j.omtn.2018.01.011.
4
These authors contributed equally to this work.
Correspondence: Steve D. Wilton, Centre for Comparative Genomics, Health
Research Centre, Building 390, Murdoch University, 90 South Street, Perth, WA
6150, Australia.
E-mail: swilton@ccg.murdoch.edu.au
Molecular Therapy: Nucleic Acids Vol. 11 June 2018 ª2018 The Authors. 91
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
the mature transcript as a result of alternative splicing events. While
6%16% of 30UTRs are suggested to contain introns,
18
it is unclear at
this stage what percentage of these have the propensity to retain an
intron within the mature message. While transcripts containing in-
trons within the 30UTR were once believed to be non-functional
due to nonsense-mediated decay,
19,20
there is now evidence to show
that intron retention within the mature message is an important
mechanism for transcript and protein regulation (for review, see
Bicknell et al.
18
and Ge and Porse
20
). Tissue-specic transcript regu-
lation by intron retention is particularly common in neuronal cells
during differentiation and maturity,
21
and recent studies have re-
vealed a role for intron retention in hematopoietic cellular differenti-
ation.
22
Furthermore, intron retention within the 30UTR has been
shown to play a role in transcript autoregulation to maintain protein
homeostasis, a mechanism that is particularly common in proteins
involved in forming the spliceosome and in regulating pre-mRNA
processing.
23,24
A number of factors have been reported to regulate splicing events re-
sulting in intron retention, with a correlation observed between
intron retention and the presence of certain regulatory cis elements.
17
Of particular interest, intron retention has been suggested to be the
result of stalling of the RNA polymerase II elongation due to poor
splicing factor recruitment and weakened splicing in non-essential
transcripts.
17,25
Other factors inuencing this mechanism include
the position of the intron within the transcript, reduced intron length,
an increase in G/C content within the intron, and weak splice site
strength.
17
While factors that determine intron retention have been studied in
canonical splicing events, it is unknown what role they play in medi-
ating AO-induced intron retention and transcript expression. Conse-
quently, this study focused on gaining a further understanding of the
mechanisms inuencing AO-induced intron retention and, further-
more, investigating how it can impact transcript and protein expres-
sion as a potential strategy in treating genetic disease.
RESULTS
Targeting AOs to Exon 8 Results in Exon 7 and Intron 7 Retention
in SMN2 Transcripts
SMA type I broblasts (Coriell GM03813) were transfected with
20O-methyl AOs targeting SMN2 exon 8 (for binding coordinates
and AO sequences, see Table 3) at 300, 150, and 75 nM and incubated
(37C) for 48 hr. RT-PCR analysis (Figure 1A) of SMN2 showed an
increase in abundance of an approximately 850-bp product, which
was conrmed by sequencing (Figure 1B) to be the SMN2 transcript
retaining exon 7, as well as intron 7 (848 bp). This product is referred
to as exon/intron 7-SMN2 and is labeled ex/in7 in the gures. Because
the stop codon is located within exon 7, the addition of an extra
444-bp intronic sequence should encode the same protein as SMN1,
but increases the length of the 30UTR (Figure 1C). These results
were reproducible in two unrelated SMA patient primary cell strains
(data not shown), including an SMA type II patient (prepared in-
house) and an SMA type I patient with only one copy of SMN2 (Cor-
iell GM00232). Two additional bands were observed at approximately
100 bp above and 100 bp below the exon/intron 7-SMN2 transcript.
The larger band was deemed to be a PCR artifact because it was un-
able to be re-amplied and disappeared following increasing primer
annealing temperature. The lower band was conrmed by sequencing
to be the naturally occurring D5-SMN2 transcript containing intron 7
(data not shown).
The initial screening of AO sequences 118 is shown in Figure S1.
Following preliminary screening, additional AOs were designed by
microwalking around promising AO target sites, shifting up or down-
stream of the original sites (Table 3). Analysis of SMN2 transcripts
following transfection with rened AO sequences showed an
improvement in AO-induced exon/intron 7 retention (Figure 1A).
A clear dose response was observed in all AO-treated cells, with
AOs 10, 18, 24, and 25 consistently inducing the highest levels of in-
clusion across experiments (n = 6). These promising AOs were there-
fore selected for further evaluation, including protein analysis.
Splice Site Analysis Shows a Weak Exon 7 Donor Splice Site
To further investigate the exon/intron 7-SMN2 transcript induced by
AOs targeting exon 8, we analyzed splice site scores (Table 1) using
the online Human Splicing Finder 3.0 website.
26
SMN2 exon 7 was
predicted to have a very strong acceptor site with a score of 98.2
out of a possible 100, while the donor splice site was weaker, scoring
82.81 out of 100. The exon 8 acceptor splice site had a predicted score
of 91.9 out of 100. While these splice site scores are only a predicted
measurement of the likelihood of the site being recognized by the
splicing machinery, the comparatively weaker exon 7 donor splice
site could lead to reduced splicing at the exon/intron 7 junction
when the intron 7/exon 8 junction is further compromised following
AO treatment.
PMO Delivery by Electroporation Improves Exon/Intron 7
Inclusion, Inducing a Decrease in SMN Protein
Previously identied optimal 20O-methyl AO sequences 10, 18, 24,
and 25 were resynthesized as phosphorodiamidate morpholino olig-
omers (PMOs) by Genetools (Philomath, OR, USA), and are now
cited as PMOs 10, 18, 24, and 25. PMOs were administered to cells
using nucleofection for optimal delivery at 1 and 0.5 mM for SMN
transcript and protein analysis by RT-PCR and western blot,
respectively. Nucleofection of PMOs showed increased levels of
exon/intron 7 retention in the mature transcript compared with the
same sequences tested as 20O-methyl AOs, with a clear reduction in
the levels of FL-SMN and D7-SMN transcripts. In particular, PMO-
10 induced almost 100% exon/intron 7 inclusion as determined by
RT-PCR (Figure 2A).
Interestingly, western blot analysis of SMN protein levels revealed a
signicant decrease in the amount of SMN detected in samples trans-
fected with exon-8-targeting PMOs (Figures 2B and 2C). PMO-10
and PMO-24 were the most effective compounds inducing a respec-
tive 50% (p = 0.022) and 33% (p = 0.027) decrease in SMN protein
when compared with the level observed in untreated broblasts
Molecular Therapy: Nucleic Acids
92 Molecular Therapy: Nucleic Acids Vol. 11 June 2018
(n = 4). The Anti-ISS-N1 PMO sequence was transfected as a positive
control and was shown to increase SMN levels by up to 80%
compared with that in untreated SMA patient broblasts (p = 0.032).
PMO-Induced SMN Knockdown Is Reproducible in Unaffected
Fibroblasts
PMO-24 and PMO-25 were evaluated in non-SMA broblasts to
determine the effects of intron 7 retention on SMN protein levels in
cells with a higher baseline of SMN. PMOs were transfected by nucle-
ofection at 1 and 0.5 mM, and incubated for 3 days prior to western
blot analysis. RT-PCR analysis of the total SMN transcripts conrmed
that exon-8-targeting AOs induce almost 100% exon/intron 7 reten-
tion, and hence this must represent both the SMN1 and the SMN2
transcripts (Figure 3A). Consistent with the ndings in SMA patient
broblasts, western blot analysis (Figures 3B and 3C) demonstrated
that PMO-24 and PMO-25 effectively decreased the levels of SMN
protein in non-SMA broblasts by 55% (p = 0.041) and 38%
(p = 0.072), respectively (n = 3). Anti-ISS-N1 was transfected as a
positive control and was shown to increase the levels of SMN protein
by up to 35% as seen in non-SMA cells transfected with the low AO
dose; however, this was not statistically signicant. Furthermore,
an AO designed to induce exon 7 skipping was transfected into
non-SMA broblasts as a positive control for downregulating SMN
levels. Fibroblasts transfected with this PMO show a 76% decrease
Figure 1. SMN Transcript Analysis following 20OMethyl AO Transfection
Analysis of SMA fibroblasts following transfection with exon 8 targeting 20O-methyl AOs, showing (A) RT-PCR analysis of SMN2 transcripts from transfected fibroblasts (300,
150, 75 nM). Anti-ISS-N1 was used as a positive control for transfection efficiency, and a sham control AO was used as a transfection negative control. A 100-bp marker was
used for comparison, and an RT-PCR no-template negative control was loaded in the final lane, (B) sequencing chromatogram across SMN2 exon 7 to intron 7 and intron 7 to
exon 8, confirming the presence of intron 7 within the mature transcript, and (C) schematic showing SMN2 transcripts identified within fibroblasts transfected with exon-8-
targeting AOs, including full-length SMN (FL-SMN), the D7-SMN transcript missing exon 7, and the exon/intron 7 retained transcript (ex/in7-SMN) with the extended 30UTR.
www.moleculartherapy.org
Molecular Therapy: Nucleic Acids Vol. 11 June 2018 93
in SMN levels compared with sham control and untreated broblasts
(p = 0.030).
Functional SMN protein forms aggregates with the gemin proteins
that uoresce as bright sparkling foci, reminiscent of gems after anti-
body staining. Therefore, the presence of gems within the cell indi-
cates functional localization of SMN protein. The percentages of
broblast nuclei staining positive for gems were counted (Figure 4A).
PMOs 10, 18, 24, and 25 were all transfected into non-SMA cells by
nucleofection at 1 mM and incubated for 3 days prior to xation
and immunouorescent staining. Interestingly, the sham control
PMO induced an increase in nuclei containing gems from 16.3% in
untreated broblasts to 20.6% following control AO transfection.
However, PMOs 10, 18, 24, and 25 all decreased the number of nuclei
containing gems, with as low as 7.3% of broblasts containing gems
following transfection with PMO-10. In comparison, 25.9% of bro-
blasts transfected with the Anti-ISS-N1 PMO sequence express gems,
while only 3.3% of those transfected with the exon skipping control
PMO express gems. Representative images of broblasts transfected
with each PMO are shown in Figure 4B.
PMOs targeting exon 8 induce more efcient retention of exon and
intron 7 in both the SMN1 and the SMN2 transcripts compared
with the 20O-methyl AOs of the same sequence, and analysis of the
SMN protein by western blot and immunouorescence shows a
further decrease in SMN expression following transfection. While
the exon/intron 7-SMN transcript occurs naturally at low levels in un-
transfected cells, it appears that the extended 30UTR introduces a
number of new regulatory mechanisms into the transcript that nega-
tively impact on protein expression.
Intron Retention Introduces Negative Regulatory Elements to
the 30UTR
In silico analysis of the extended 30UTR was carried out using the
online tools UTRscan,
27
miRBase,
28
Polyadq,
29
and DNA Functional
Site Miner (DNA FS Miner).
30
Table 2 lists the potential regulatory
elements identied within intron 7 from each of these databases,
Figure 2. SMN Transcript and Protein Analysis in
SMA Fibroblasts following PMO Nucleofection
SMN transcript and protein levels in SMA fibroblasts
transfected with PMOs by nucleofection at 1 and 0.5 mM,
showing (A) RT-PCR analysis of SMN2 products con-
firming exon/intron retention, (B) western blots showing
SMN protein levels compared with b-tubulin levels, and (C)
densitometric analysis showing changes in SMN protein
levels normalized against b-tubulin. SMN levels in trans-
fected fibroblasts are shown as an n-fold change
compared with those in samples from untreated cells.
Error bars represent the SEM.
Table 1. Exons 7 and 8 Splice Site Predictions Using Human Splicing Finder
3.0
Exon Splice Site Type Splice Site Motif Consensus Value (0100)
7 acceptor ttttccttacagGG 98.2
7 donor GGAgtaagt 82.81
8 acceptor tctcatttgcagGA 91.9
Molecular Therapy: Nucleic Acids
94 Molecular Therapy: Nucleic Acids Vol. 11 June 2018
and Figure 5A illustrates the location of these elements within
intron 7. In silico analysis of the extended 30UTR by the online
tool UTRscan drew attention to two possible regulatory motifs known
as the bearded (BRD) box and the K box. Each of these motifs has
been shown by others to disrupt translation of neuronal gene tran-
scripts during Drosophila development by recruiting microRNAs.
31,32
The BRD box consensus sequence is AGCUUUA and for K box is
UGUGAU. A search for microRNA recognition sites using miRBase
revealed three potential microRNA binding sites within intron 7,
with E values below 10, that suggests that these sites are active. The
microRNAs hsa-miR-3118, hsa-miR-3976, and hsa-miR-5580-3p all
have complementary bases to the SMN intron 7 sequence within
the seed region. It is therefore possible that these microRNAs or the
BRD and K box motifs could disrupt SMN translation.
The exon/intron 7-SMN transcript was further analyzed for polyade-
nylation [poly(A)] signals using two online tools, Polyadq
29
and DNA
FS Miner.
30
Each tool identied two potential poly(A) sites with cor-
responding CA cleavage sites within intron 7. A potential poly(A)-1
(ATTAAA) signal was identied at 132 bases into intron 7, and a po-
tential poly(A)-2 (AATAAA) signal was identied at 238 bases into
intron 7. To determine whether early polyadenylation could destabi-
Figure 3. SMN Transcript and Protein Analysis in
Unaffected Fibroblasts following PMO
Nucleofection
SMN transcript and protein levels in unaffected fibroblasts
transfected with PMOs by nucleofection at 1 and 0.5 mM,
showing (A) RT-PCR analysis of SMN products confirming
exon/intron retention, (B) western blots showing SMN
protein levels compared with b-tubulin levels, and (C)
densitometric analysis showing changes in SMN protein
levels normalized against b-tubulin. SMN levels in trans-
fected fibroblasts are shown as an n-fold change
compared with those in samples from untreated cells.
Error bars represent the SEM.
lize the extended SMN transcript following
PMO treatment, we designed specic poly(A)
primers to target the predicted cleavage sites
and downstream sequences to amplify polyade-
nylated products (Figure 5B). Following nucleo-
fection of PMOs into unaffected broblasts,
RNA was collected at multiple time points
including 12, 24, 48, and 72 hr. Samples were
DNase treated and RNA was amplied using
the exon/intron 7 forward primer with the spe-
cic poly(A)-R1 and R2 primers (Figure 5C).
No differences were observed within each treat-
ment group over the 72-hr duration of the time
course.
Specic poly(A)-R1 primer binding to the rst
ATTAAA poly(A) site amplied a faint product
in some samples, suggesting this site could
initiate polyadenylation. RT-PCR using the specic poly(A)-R2
primer directed to the second AATAAA site resulted in two products,
a stronger amplicon amplied by the second cleavage site, as well as a
fainter non-specic amplication of the rst cleavage site. The stron-
ger amplicon was sequenced and conrmed to have a poly(A) tail
extending past the primer annealing site (Figure 5D). This result
suggests that early polyadenylation is occurring at this second
AATAAA site within intron 7, and as a result could destabilize the
exon/intron 7-SMN transcript and therefore result in decreased pro-
tein levels.
To compare the stability and cleavage of the exon/intron 7-SMN and
FL-SMN transcripts, we designed primers to target downstream of the
poly(A) sites, and we tested them with a forward primer targeting the
exon/intron 7 boundary (Figure 5C). Interestingly, both primer sets
produce a strong amplicon extending beyond the poly(A) signal
and do not show diminished expression of the transcript following
AO treatment. Taken together, these results show that while early
polyadenylation appears to occur at the second poly(A) signal, the
transcript level remains stable, suggesting that transcription may be
occur at a faster rate than polyadenylation and cleavage, or that the
early polyadenylation is not destabilizing the transcript.
www.moleculartherapy.org
Molecular Therapy: Nucleic Acids Vol. 11 June 2018 95
DISCUSSION
The original intent of this study was to inuence the rate of SMN2
transcription by targeting AOs to the terminal exon in an attempt
to increase exon 7 inclusion. However, another splice-switching
mechanism for manipulating expression was revealed. Selected AOs
targeting SMN2 exon 8 promoted exon 7 and intron 7 inclusion in
the mature SMN message, revealing a novel AO application: inducing
terminal intron retention. In silico analysis of the SMN2 exon 7 splice
sites suggests that this action may be the result of a strong acceptor
splice site (scoring 98 out of 100) and a weaker donor splice site
Figure 4. Immunofluorescence Staining and Analysis of SMN Protein following PMO Nucleofection
Immunofluorescence analysis of SMN shown as gems in PMO nucleofected unaffected fibroblasts (1 mM), compared with untreated fibroblasts. (A) Graph displaying the
percentages of cell nuclei containing gems, as indicated above each bar, and (B) representative images showing anti-SMN- (green) and Hoechst (blue)-stained SMA fi-
broblasts following PMO nucleofection. Typical gems stained within the cell nucleus are indicated with white arrows. Images were taken at 20objectiv e. Scale bar, 25 mm.
Molecular Therapy: Nucleic Acids
96 Molecular Therapy: Nucleic Acids Vol. 11 June 2018
(scoring 83 out of 100). The weaker donor site might be subject to
poor recognition by the splicing machinery, abrogating splicing be-
tween exon 7 and the following exon, especially when the splicing
of exon 8 is compromised by AO binding.
The correlation in splice-switching efcacy between 20O-methyl and
PMO compounds of the same sequence is well established for induced
exon skipping in dystrophin transcripts as a therapy for Duchenne
muscular dystrophy,
33
as well as for the efcacy of the Anti-ISS-N1
sequence.
34
Furthermore, the PMO chemistry has been shown to be
safe, with no signicant adverse events reported in a 5-year study of
Duchenne muscular dystrophy patients receiving weekly treatment
with Exondys51, a PMO drug granted accelerated approval by the
U.S. Food and Drug Administration (FDA) in 2016 for amenable dys-
trophin mutations.
35,36
Consistent with the results of exon 8 targeting 20O-methyl AOs,
PMOs of the same sequence were effective at inducing exon/
intron 7 retention. Interestingly, while PMOs targeting exon 8
increased the levels of the exon/intron 7-SMN transcript, these
PMOs also induced a 50% decrease in SMN protein as assessed by
western blot. This result was reproducible between SMA patient -
broblasts and unaffected broblasts following AO transfection. Simi-
larly, immunouorescence staining of the transfected broblasts
showed fewer nuclei containing functional SMN in the form of
gemswhen compared with sham-control PMO transfected cells.
It is likely that the observed decrease in SMN protein could be due
to the longer than normal 30UTR within the exon/intron 7-SMN
transcript, and therefore many regulatory factors could contribute
to protein downregulation.
Endogenous intron retention has been shown by others to act as
a form of gene repression, often through the introduction of a
premature termination codon, rendering the transcript susceptible
to nonsense-mediated decay.
17,37
In the study presented here,
nonsense-mediated decay is unlikely to be the cause of SMN downre-
gulation due to the retained intron occurring after the normal termi-
nation codon. However, as a result of the extended 30UTR, inuences
of downstream sequences could lead to this transcript not being ef-
ciently translated.
The length of the 30UTR can be a critical factor in regulating tran-
script stability and protein expression. A longer 30UTR can increase
the opportunity for sequence-specic recognition motifs to recruit
regulatory factors, including microRNAs.
16
Furthermore, the possi-
bility of altered mRNA secondary structures can inuence the avail-
ability of such sequences to these factors.
38
A number of online
databases can identify potential microRNA and regulatory factor
binding sites in NCBI documented transcripts. However, there is a
lack of appropriate resources whereby an altered 30UTR sequence
can be analyzed, limiting the search possibilities for this study. It is
probable that there are many factors inuencing the translational
knockdown observed for the exon/intron 7-SMN transcript.
The microRNA prediction database miRBase allows the user to input
an mRNA sequence for analysis, and analysis of the SMN intron 7
sequence revealed three potential microRNA binding sites. Of these,
miR-3976 is a validated microRNA and is reportedly overexpressed in
pancreatic cancer.
39
However, how miR-3976 impacts translation,
and whether it is expressed in the dermal broblast used in this study,
is yet to be determined. Further in silico analysis of the sequence using
the online UTRscan database
27
drew attention to a number of motifs,
including the BRD and K box motifs, as well as two alternative poly-
adenylation signals. The BRD box and K box motifs have been shown
to recruit microRNAs that act as translational inhibitors of certain
proteins during Drosophila development.
31,32,40
The functionality of
these motifs in human sequences is unknown and, therefore, any in-
uence on SMN translation is only speculative.
Others have tested the use of AOs to prevent microRNAs from
binding to a transcript, by either binding of the microRNA as an
antagomir, or to act as a decoy, binding directly to the transcript of
interest.
41
To further investigate the mechanism of translational
knockdown presented here, future studies could use AOs targeting
the microRNAs themselves, their binding sites, as well as the BRD
and K box motifs. Examining SMN expression in such studies could
indicate whether these regulatory elements play a role in inhibiting
SMN translation in the exon/intron 7-SMN transcript. However, it
is unlikely that this will be of clinical benet to SMA patients.
Further in silico analysis of the SMN2 intron 7 sequence identied two
potential polyadenylation signals with corresponding CA cleavage
sites that could prematurely cleave the mRNA, potentially resulting
in transcript and protein destabilization. Premature polyadenylation
and mRNA cleavage have been shown by others to cause less efcient
Table 2. In Silico Analysis of Potential Regulatory Elements within SMN
Intron 7
The prediction value or score indicates the strength of the regulatory site within the
SMN sequence. For poly(A) signals, scores >0.5 are true predictions for Polyadq and
scores >0.6 are true predictions for DNA FS Miner. For miRBase, E values <10 may
have binding potential.
NA, not applicable.
www.moleculartherapy.org
Molecular Therapy: Nucleic Acids Vol. 11 June 2018 97
processing of transcripts when compared with those cleaved at the
distal 30or canonical poly(A) site.
42
In this study, primers were
designed to anneal to the sites identied within intron 7, and should
amplify a product only if polyadenylation has occurred. The primer
designed to anneal to the ATTAAA site failed to generate a product,
while the primer annealing to the AATAAA site amplied a clean,
consistent product whose levels increased with an increase in intron
7 retention.
It appears that the early AATAAA site within intron 7 initiates poly-
adenylation prior to use of the canonical poly(A) site within exon 8,
and as such negatively affects SMN translation. However, primers
amplifying downstream of the early poly(A) site produced a strong
and consistent product that is not diminished following AO treat-
ment at any time point, indicating that mRNA cleavage does not
Figure 5. Analysis of Predicted SMN Intron
Regulatory Elements
Analysis of predicted SMN intron 7 regulatory elements,
showing (A) a schematic of the in silico analysis of regu-
latory elements within intron 7, (B) a schematic of the
location of potential poly(A) signal and cleavage sites and
the primer binding sites, (C) RT-PCR across SMN and
Cyclin-D (housekeeping gene) transcripts in fibroblasts
nucleofected with 1 mM (1) PMO-24, (2) PMO-25, (3)
control PMO, and (4) an untreated sample, using indicated
primers, and (D) sequencing chromatogram of the RT-
PCR product amplified using the exon 6/7 forward and
poly(A)R2 reverse primers.
occur downstream of either of the poly(A) sig-
nals. Taken together, these results show that
while early polyadenylation appears to occur at
the second poly(A) signal, the transcript level re-
mains stable, suggesting that transcription in
this case may occur at a faster rate than polyade-
nylation and cleavage.
Interestingly, the product amplied by the spe-
cic poly(A)-R2 primer was also observed in un-
treated samples, suggesting polyadenylation
may be a natural mechanism for controlling
SMN levels. Furthermore, it has been reported
that the canonical poly(A) signal and cleavage
site in SMN exon 8 is inefcient at recruiting
cleavage factors, and consequently SMN polya-
denylation is subjected to additional regulation
by U1A, a component of the U1 snRNP.
43
The
study showed that overexpression of U1A can
inhibit SMN polyadenylation and cleavage,
decreasing the levels of SMN protein.
43
Given
the inefciency of the canonical poly(A) site, it
is probable that the intron 7 poly(A) signal could
be more favorable for initiating polyadenylation
and cleavage. However, the presence of U1A
may still inhibit cleavage at this site, and therefore explain the lack
of mRNA cleavage observed in the RT-PCR experiments within the
current study.
The presence of polyadenylation signals in terminal introns could
contribute to the process of protein regulation by intron retention.
To examine this theory further, future studies should assess the
occurrence of polyadenylation in alternatively spliced 30UTR-intron
retention transcripts. It would be interesting to investigate whether
this effect of premature polyadenylationoccurs in other gene
transcripts.
AOs targeting the terminal exon to induce intron retention could be
useful in those diseases where protein repression is essential for treat-
ment, including many types of cancer. We speculate that in the study
Molecular Therapy: Nucleic Acids
98 Molecular Therapy: Nucleic Acids Vol. 11 June 2018
presented here, the extended SMN transcript is regulated by a number
of motifs within intron 7 that are involved in translational repression
due to the retention of this sequence in the 30UTR. However,
this mechanism may only apply to a select number of genes.
Alternatively, if the stop codon were to be in the nal exon, intron
retention could disrupt the reading frame or introduce a premature
termination codon. To identify genes where intron retention could
be applied, it will be necessary to look at a number of factors within
the gene, most importantly the splice site scores for the anking
exons. If the donor site is strong, then intron retention may not be
possible.
Aside from weakened splice sites of retained introns, additional cis
and trans factors suggested to inuence natural intron retention
include the position of the intron in the transcript, an increase in
G/C content, and reduced intron length.
17
Interestingly, while the
G/C content of SMN intron 7 is only 33.8%, being positioned adjacent
to an alternatively spliced exon increases the probability of the intron
being retained. Furthermore, at 444 nt long, intron 7 is relatively short
compared with the median human intron length of 1,334 nt in the
coding region and 1,303 nt within the 30UTR.
44
It will be interesting
in future studies to compare the relevance of these factors across tran-
scripts to identify markers that could predict the likelihood of effec-
tive AO-induced intron retention.
In this study we present a novel application for splice-switching
PMOs in initiating terminal intron retention. It is unfortunate that
this model is unlikely to provide therapeutic benet to SMA
patients, yet further work could see intron retention being applied
to a number of other diseases. This study is reective of the ever-ex-
panding complexity of gene regulation and undoubtedly sheds new
light on splicing and AO mechanisms that may offer new avenues
of therapy.
MATERIALS AND METHODS
AO Design and Synthesis
AOs were designed to target the exon 8 acceptor splice site and exon
splice enhancers (ESEs) as predicted by the online SpliceAid predic-
tion tool,
45
available at http://www.introni.it/splicing.html.AO
nomenclature was based on that described by Mann et al.
46
All
20O-methyl PS-AOs were synthesized in-house on an Expedite
8909 nucleic acid synthesizer with a phosphorothioate backbone.
Following identication of optimal 20O-methyl AO sequences, these
AOs were prepared as PMOs, purchased through Genetools (Philo-
math, OR, USA). Table 3 lists the details of all AOs used in this study.
20O-Methyl AO Transfection
SMA type I patient broblasts (GM03183; Coriell Cell Repositories,
Camden, NJ, USA) and normal human dermal broblasts prepared
in-house (Murdoch University Human Research Ethics Committee
Approval #2013/156) were proliferated and seeded in 10% fetal
bovine serum (FBS) DMEM and incubated at 37C for 24 hr prior
to transfection. All 20O-methyl PS-AOs were transfected using
Lipofectin (Life Technologies, Melbourne, Australia) at a 2:1 ratio
of lipofectin to total AO, according to manufacturers protocols,
and incubated for 48 hr.
Nucleofection of PMOs
PMO delivery by nucleofection was performed using a Nucleofection
X unit with the Nucleofection P2 kit, using the CA-137 program
(Lonza, Melbourne, Australia). PMOs were transfected at 1 and
0.5 mM, as determined by the nal transfection volume, supplemented
with 5% FBS DMEM and incubated for 72 hr.
RNA Extraction and PCR
RNA was extracted using the MagMAX-96 Total RNA Isolation Kit,
including a DNase treatment (Life Technologies), according to the
manufacturers instructions. RT-PCRs were performed using the
One-Step SuperScript III RT-PCR kit with Platinum Taq Polymerase
(Life Technologies) according to manufacturers instructions. All
primer sequences used in this study are detailed in Supplemental
Materials and Methods. Products were amplied with the tempera-
ture prole, 55C for 30 min, 94C for 2 min, followed by 2530 cycles
of 94C for 30 s, 56C for 30 s, and 68C for 1 min. Amplicon se-
quences were identied by Sanger sequencing at the Australian
Genome Research Facility (AGRF, Perth, Australia).
Western Blot Analysis
Cell lysates were prepared in 125 mM Tris/HCl (pH 6.8), 15% SDS,
10% glycerol (v/v), 1.25 mM PMSF (Sigma-Aldrich), protease inhibi-
tor cocktail (3 mL/100 mL; P8340; Sigma-Aldrich), 0.004% bromophe-
nol blue, and 2.5 mM dithiothreitol. Pellets were sonicated six times
for 1-s pulses and samples denatured at 94C for 5 min.
Approximately 10 mg of total protein (as determined by BCA assay)
was loaded per sample on a NuPAGE Novex 4%12% BIS/Tris gel
(Life Technologies). Proteins were transferred onto a Pall Fluorotrans
polyvinylidene uoride (PVDF) membrane at 350 mA for 1 hr
in western transfer buffer. MANSMA7 (1:1,000; Developmental
Studies Hybridoma Bank) and b-tubulin (1:20,000; DSHB) mono-
clonal primary antibodies were incubated overnight at 4C prior
to detection using a Western Breeze Chemiluminescent Immunode-
tection System (Life Technologies), according to the manufacturers
instructions. Western blot images were captured on a Vilber
Lourmat Fusion FX system using Fusion software, and Bio-1D soft-
ware was used for densitometry analysis. All p values were calculated
using a paired two-tailed t test, and SE bars were used to represent
the SEM.
Immunofluorescence
Cells on coverslips were xed using ice-cold acetone-methanol (1:1),
then blocked in 10% ltered goat serum in PBS containing 0.2%
Triton-X (PBT). SMN was detected with MANSMA1 (1:100;
DSHB) antibody, incubated overnight at 4C in PBT. Cells
were stained with Hoechst (Sigma-Aldrich) for nuclei detection
(1 mg/mL diluted 1:125), and the MANSMA1 primary antibody
was detected using Alexa Fluor 488 (1:400; Thermo Fisher Scientic).
Photos were overlaid, and the number of gems per nuclei was counted
www.moleculartherapy.org
Molecular Therapy: Nucleic Acids Vol. 11 June 2018 99
and recorded as a percentage of total nuclei, with at least 300 cells
counted per slide.
In Silico Analysis
A number of online databases were used to analyze the extended
30UTR sequence to identify potential regulatory elements. Splice
site scores were analyzed by Human Splice Finder version 3.0 avail-
able at http://www.umd.be/HSF3/.
26
Regulatory element binding
was predicted using UTR Scan available at http://itbtools.ba.itb.
cnr.it/.
27
Polyadenylation signals were analyzed using Polyadq avail-
able at http://rulai.cshl.edu/tools/polyadq/polyadq_form.html
29
and
DNA FS Miner available at http://dnafsminer.bic.nus.edu.sg/.
30
SUPPLEMENTAL INFORMATION
Supplemental Information includes Supplemental Materials and
Methods and one gure and can be found with this article online at
https://doi.org/10.1016/j.omtn.2018.01.011.
AUTHOR CONTRIBUTIONS
Conceived and designed experiments: L.L.F., C.M., S.F., and S.D.W.
Performed experiments: L.L.F., C.M., and I.L.P. Wrote and edited
manuscript: L.L.F., C.M., I.L.P., S.F., and S.D.W.
ACKNOWLEDGMENTS
The authors would like to acknowledge Project funding from
the Parry Foundation, Spinal Muscular Atrophy Association of
Australia, and the NHMRC (project grant 1086311). L.L.F. and
I.L.P. received scholarships from Team Spencer and Muscular Dys-
trophy WA.
REFERENCES
1. McAndrew, P.E., Parsons, D.W., Simard, L.R., Rochette, C., Ray, P.N., Mendell, J.R.,
Prior, T.W., and Burghes, A.H. (1997). Identication of proximal spinal muscular at-
rophy carriers and patients by analysis of SMNT and SMNC gene copy number. Am.
J. Hum. Genet. 60, 14111422.
Table 3. AO Sequences and SMN1 and SMN2 Binding Coordinates
Synthesis Round AO Number Annealing Coordinates Sequence 50/30Length (bp)
1: Initial screening
1 hSMN8A(18+7) GCA UUU CCU GCA AAU GAG AAA UUA G 25
2 hSMN8A(12+13) AUG CCA GCA UUU CCU GCA AAU GAG A 25
3 hSMN8A(7+18) GCU CUA UGC CAG CAU UUC CUG CAA A 25
4 hSMN8A(2+23) GUG CUG CUC UAU GCC AGC AUU UCC U 25
5 hSMN8A(+4+28) AUU UAG UGC UGC UCU AUG CCA GCA U 25
6 hSMN8A(+9+33) GUG UCA UUU AGU GCU GCU CUA UGC C 25
7 hSMN8A(+14+38) UAG UGG UGU CAU UUA GUG CUG CUC U 25
8 hSMN8A(+24+48) AUC GUU UCU UUA GUG GUG UCA UUU A 25
9 hSMN8A(+34+58) GAU CUG UCU GAU CGU UUC UUU AGU G 25
10 hSMN8A(+59+83) AUC UUC UAU AAC GCU UCA CAU UCC A 25
11 hSMN8A(+84+108) AUA UUU UGA AGA AAU GAG GCC AGU U 25
12 hSMN8A(+152+178) CAU AAC UUU UAA UCA AGA AGA GUU ACC 27
13 hSMN8A(+24+43) UUC UUU AGU GGU GUC AUU UA 20
14 hSMN8A(+44+63) UUC CAG AUC UGU CUG AUC GU 20
15 hSMN8A(+64+83) AUC UUC UAU AAC GCU UCA CA 20
16 hSMN8A(+29+48) AUC GUU UCU UUA GUG GUG UC 20
17 hSMN8A(+35+53) GUC UGA UCG UUU CUU UAG UG 20
18 hSMN8(+39+58) GAU CUG UCU GAU CGU UUC UU 20
2: Microwalking
19 hSMN8A(12+18) GCU CUA UGC CAG CAU UUC CUG CAA AUG AGA 30
20 hSMN8A(10+15) CUA UGC CAG CAU UUC CUG CAA AUG 24
21 hSMN8A(+39+63) UUC CAG AUC UGU CUG AUC GUU UCU U 25
22 hSMN8A(+44+68) UCA CAU UCC AGA UCU GUC UGA UCG U 25
23 hSMN8A(+55+79) UCU AUA ACG CUU CAC AUU CCA GAU C 25
24 hSMN8A(+57+81) CUU CUA UAA CGC UUC ACA UUC CAG A 25
25 hSMN8A(+59+78) CUA UAA CGC UUC ACA UUC CA 20
Control AOs
ISS-N1 hSMN2.7D(10-29) AUU CAC UUU CAU AAU GCU GG 20
Skipping mSmn7A(+7+36) UGA GCA CUU UCC UUC UUU UUU AUU UUG UCU 30
Sham scrambled sequence GCT ATT ACC TTA ACC CAG 18
Molecular Therapy: Nucleic Acids
100 Molecular Therapy: Nucleic Acids Vol. 11 June 2018
2. Shababi, M., Lorson, C.L., and Rudnik-Schöneborn, S.S. (2014). Spinal muscular at-
rophy: a motor neuron disorder or a multi-organ disease? J. Anat. 224,1528.
3. Lorson, C.L., Hahnen, E., Androphy, E.J., and Wirth, B. (1999). A single nucleotide in
the SMN gene regulates splicing and is responsible for spinal muscular atrophy. Proc .
Natl. Acad. Sci. USA 96, 63076311.
4. Monani, U.R., Lorson, C.L., Parsons, D.W., Prior, T.W., Androphy, E.J., Burghes,
A.H., and McPherson, J.D. (1999). A single nucleotide difference that alters splicing
patterns distinguishes the SMA gene SMN1 from the copy gene SMN2. Hum. Mol.
Genet. 8, 11771183.
5. Porensky, P.N., and Burghes, A.H. (2013). Antisense oligonucleotides for the treat-
ment of spinal muscular atrophy. Hum. Gene Ther. 24, 489498.
6. Chiriboga, C.A., Swoboda, K.J., Darras, B.T., Iannaccone, S.T., Montes, J., De Vivo,
D.C., Norris, D.A., Bennett, C.F., and Bishop, K.M. (2016). Results from a phase 1
study of nusinersen (ISIS-SMN(Rx)) in children with spinal muscular atrophy.
Neurology 86, 890897.
7. Finkel, R.S., Chiriboga, C.A., Vajsar, J., Day, J.W., Montes, J., De Vivo, D.C.,
Yamashita, M., Rigo, F., Hung, G., Schneider, E., et al. (2016). Treatment of infan-
tile-onset spinal muscular atrophy with nusinersen: a phase 2, open-label, dose-esca-
lation study. Lancet 388, 30173026.
8. Singh, N.K., Singh, N.N., Androphy, E.J., and Singh, R.N. (2006). Splicing of a critical
exon of human Survival Motor Neuron is regulated by a unique silencer element
located in the last intron. Mol. Cell. Biol. 26, 13331346.
9. U.S. Food and Drug Administration. (2016). FDA approves rst drug for
spinal muscular atrophy. December 23, 2016. https://www.fda.gov/NewsEvents/
Newsroom/PressAnnouncements/ucm534611.htm.
10. Osman, E.Y., Miller, M.R., Robbins, K.L., Lombardi, A.M., Atkinson, A.K., Brehm,
A.J., and Lorson, C.L. (2014). Morpholino antisense oligonucleotides targeting in-
tronic repressor Element1 improve phenotype in SMA mouse models. Hum. Mol.
Genet. 23, 48324845.
11. Singh, N.N., Lawler, M.N., Ottesen, E.W., Upreti, D., Kaczynski, J.R., and Singh, R.N.
(2013). An intronic structure enabled by a long-distance interaction serves as a novel
target for splicing correction in spinal muscular atrophy. Nucleic Aci ds Res. 41,8144
8165.
12. Hua, Y., Vickers, T.A., Okunola, H.L., Bennett, C.F., and Krainer, A.R. (2008).
Antisense masking of an hnRNP A1/A2 intronic splicing silencer corrects SMN2
splicing in transgenic mice. Am. J. Hum. Genet. 82, 834848.
13. Fong, N., Kim, H., Zhou, Y., Ji, X., Qiu, J., Saldi, T., Diener, K., Jones, K., Fu, X.D., and
Bentley, D.L. (2014). Pre-mRNA splicing is facilitated by an optimal RNA polymerase
II elongation rate. Genes Dev. 28, 26632676.
14. Dujardin, G., Lafaille, C., de la Mata, M., Marasco, L.E., Muñoz, M.J., Le Jossic-
Corcos, C., Corcos, L., and Kornblihtt, A.R. (2014). How slow RNA polymerase II
elongation favors alternative exon skipping. Mol. Cell 54, 683690.
15. Lim, S.R., and Hertel, K.J. (2001). Modulation of survival motor neuron pre-mRNA
splicing by inhibition of alternative 30splice site pairing. J. Biol. Chem. 276, 45476
45483.
16. Barrett, L.W., Fletcher, S., and Wilton, S.D. (2012). Regulation of eukaryotic gene
expression by the untranslated gene regions and other non-coding elements. Cell.
Mol. Life Sci. 69, 36133634.
17. Braunschweig, U., Barbosa-Morais, N.L., Pan, Q., Nachman, E.N., Alipanahi, B.,
Gonatopoulos-Pournatzis, T., Frey, B., Irimia, M., and Blencowe, B.J. (2014).
Widespread intron retention in mammals functionally tunes transcriptomes.
Genome Res. 24, 17741786.
18. Bicknell, A.A., Cenik, C., Chua, H.N., Roth, F.P., and Moore, M.J. (2012). Introns in
UTRs: why we should stop ignoring them. BioEssays 34, 10251034.
19. Jaillon, O., Bouhouche, K., Gout, J.F., Aury, J.M., Noel, B., Saudemont, B., Nowacki,
M., Serrano, V., Porcel, B.M., Ségurens, B., et al. (2008). Translational control of
intron splicing in eukaryotes. Nature 451, 359362.
20. Ge, Y., and Porse, B.T. (2014). The functional consequences of intron retention: alter-
native splicing coupled to NMD as a regulator of gene expression. BioEssays 36,
236243.
21. Yap, K., Lim, Z.Q., Khandelia, P., Friedman, B., and Makeyev, E.V. (2012).
Coordinated regulation of neuronal mRNA steady-state levels through developmen-
tally controlled intron retention. Genes Dev. 26, 12091223.
22. Wong, J.J., Ritchie, W., Ebner, O.A., Selbach, M., Wong, J.W., Huang, Y., Gao, D.,
Pinello, N., Gonzalez, M., Baidya, K., et al. (2013). Orchestrated intron retention reg-
ulates normal granulocyte differentiation. Cell 154, 583595.
23. Ni, J.Z., Grate, L., Donohue, J.P., Preston, C., Nobida, N., OBrien, G., Shiue, L., Clark,
T.A., Blume, J.E., and Ares, M., Jr. (2007). Ultraconserved elements are associated
with homeostatic control of splicing regulators by alternative splicing and
nonsense-mediated decay. Genes Dev. 21, 708718.
24. Bergeron, D., Pal, G., Beaulieu, Y.B., Chabot, B., and Bachand, F. (2015). Regulated
intron retention and nuclear pre-mRNA decay contribute to PABPN1 autoregula-
tion. Mol. Cell. Biol. 35, 25032517.
25. Wong, J.J., Gao, D., Nguyen, T.V., Kwok, C.T., van Geldermalsen, M., Middleton, R.,
Pinello, N., Thoeng, A., Nagarajah, R., Holst, J., et al. (2017). Intron retention is regu-
lated by altered MeCP2-mediated splicing factor recruitment. Nat. Commun. 8,
15134.
26. Desmet, F.O., Hamroun, D., Lalande, M., Collod-Béroud, G., Claustres, M., and
Béroud, C. (2009). Human Splicing Finder: an online bioinformatics tool to predict
splicing signals. Nucleic Acids Res. 37, e67.
27. Grillo, G., Turi, A., Licciulli, F., Mignone, F., Liuni, S., Ban, S., Gennarino, V.A.,
Horner, D.S., Pavesi, G., Picardi, E., and Pesole, G. (2010). UTRdb and UTRsite
(RELEASE 2010): a collection of sequences and regulatory motifs of the untranslated
regions of eukaryotic mRNAs. Nucleic Acids Res. 38, D75D80.
28. Kozomara, A., and Grifths-Jones, S. (2014). miRBase: annotating high condence
microRNAs using deep sequencing data. Nucleic Acids Res. 42, D68D73.
29. Tabaska, J.E., and Zhang, M.Q. (1999). Detection of polyadenylation signals in hu-
man DNA sequences. Gene 231,7786.
30. Liu, H., Han, H., Li, J., and Wong, L. (2005). DNAFSMiner: a web-based software
toolbox to recognize two types of functional sites in DNA sequences.
Bioinformatics 21,671673.
31. Lai, E.C., and Posakony, J.W. (1997). The Bearded box, a novel 30UTR sequence
motif, mediates negative post-transcriptional regulation of Bearded and Enhancer
of split Complex gene expression. Development 124, 48474856.
32. Lai, E.C., Burks, C., and Posakony, J.W. (1998). The K box, a conserved 30UTR
sequence motif, negatively regulates accumulation of enhancer of split complex tran-
scripts. Development 125, 40774088.
33. Adams, A.M., Harding, P.L., Iversen, P.L., Coleman, C., Fletcher, S., and Wilton, S.D.
(2007). Antisense oligonucleotide induced exon skipping and the dystrophin gene
transcript: cocktails and chemistries. BMC Mol. Biol. 8,57.
34. Mitrpant, C., Porensky, P., Zhou, H., Price, L., Muntoni, F., Fletcher, S., Wilton, S.D.,
and Burghes, A.H. (2013). Improved antisense oligonucleotide design to suppress
aberrant SMN2 gene transcript processing: towards a treatment for spinal muscular
atrophy. PLoS ONE 8, e62114.
35. Mendell, J.R., Goemans, N., Lowes, L.P., Alfano, L.N., Berry, K., Shao, J., Kaye, E.M.,
and Mercuri, E.; Eteplirsen Study Group and Telethon Foundation DMD Italian
Network (2016). Longitudinal effect of eteplirsen vs. historical control on ambulation
in DMD. Ann. Neurol. 79, 257271.
36. U.S. Food and Drug Administration. (2016). FDA News Release: FDA grants acceler-
ated approval to rst drug for Duchenne muscular dystrophy. September 19, 2016.
http://www.fda.gov/NewsEvents/Newsroom/PressAnnouncements/ucm521263.htm.
37. Aartsma-Rus, A., Bremmer-Bout, M., Janson, A.A., den Dunnen, J.T., van Ommen,
G.J., and van Deutekom, J.C. (2002). Targeted exon skipping as a potential gene
correction therapy for Duchenne muscular dystrophy. Neuromuscul. Disord. 12
(Suppl 1 ), S71S77.
38. Chen, J.M., Férec, C., and Cooper, D.N. (2006). A systematic analysis of disease-asso-
ciated variants in the 30regulatory regions of human protein-coding genes II: the
importance of mRNA secondary structure in assessing the functionality of 30UTR
variants. Hum. Genet. 120, 301333.
39. Madhavan, B., Yue, S., Galli, U., Rana, S., Gross, W., Müller, M., Giese, N.A., Kalthoff,
H., Becker, T., Büchler, M.W., and Zöller, M. (2015). Combined evaluation of a panel
www.moleculartherapy.org
Molecular Therapy: Nucleic Acids Vol. 11 June 2018 101
of protein and miRNA serum-exosome biomarkers for pancreatic cancer diagnosis
increases sensitivity and specicity. Int. J. Cancer 136, 26162627.
40. Lai, E.C. (2002). Micro RNAs are complementary to 30UTR sequence motifs that
mediate negative post-transcriptional regulation. Nat. Genet. 30, 363364.
41. Beg, M.S., Brenner, A.J., Sachdev, J., Borad, M., Kang, Y.K., Stoudemire, J., Smith, S.,
Bader, A.G., Kim, S., and Hong, D.S. (2017). Phase I study of MRX34, a liposomal
miR-34a mimic, administered twice weekly in patients with advanced solid tumors.
Invest New Drugs 35, 180188.
42. Beaudoing, E., Freier, S., Wyatt, J.R., Claverie, J.M., and Gautheret, D. (2000).
Patterns of variant polyadenylation signal usage in human genes. Genome Res. 10,
10011010.
43. Workman, E., Veith, A., and Battle, D.J. (2014). U1A regulates 30processing of the
survival motor neuron mRNA. J. Biol. Chem. 289, 37033712.
44. Hong, X., Scoeld, D.G., and Lynch, M. (2006). Intron size, abundance, and distribu-
tion within untranslated regions of genes. Mol. Biol. Evol. 23, 23922404.
45. Piva, F., Giulietti, M., Burini, A.B., and Principato, G. (2012). SpliceAid 2: a database
of human splicing factors expression data and RNA target motifs. Hum. Mutat. 33,
8185.
46. Mann, C.J., Honeyman, K., McClorey, G., Fletcher, S., and Wilton, S.D. (2002).
Improved antisense oligonucleotide induced exon skipping in the mdx mouse model
of muscular dystrophy. J. Gene Med. 4, 644654.
Molecular Therapy: Nucleic Acids
102 Molecular Therapy: Nucleic Acids Vol. 11 June 2018
... It is important to note that, based on the mechanism of SMN2 exon 7 splicing, "correction of the splicing defect of SMN2 exon 7" and "modulation of SMN2 exon 7 splicing" have been explored as strategies for treating SMA [7,8]. Many candidate therapeutic approaches using antisense oligonucleotides (ASOs) have been explored in pursuit of these strategies [9][10][11][12][13][14][15][16][17]. ...
... In this study, three ASOs were examined: ASO targeting ISS-N1 (ASO-NUS, named after nusinersen), ASO targeting the sequence within exon 8 (ASO-EX8, named based on the target exon), and ASO targeting the intron 7-exon 8 splice site junction (ASO-SSJ, named based on the 3 splice-site location) ( Figure 1A). The sequences of ASOs were as follows: ASO-NUS, 5 -UCA CUU UCA UAA UGC UGG-3 [16]; ASO-SSJ, 5 -CUA GUA UUU CCU GCA AAU GAG-3 [9]; and ASO-EX8, 5 -AUC UUC UAU AAC GCU UCA CAU UCC A-3 [17]. ...
... Many ASOs for treating SMA have been studied, such as ASOs inhibiting alternative 3 splice site pairing of SMN2 exon 8 [9], ASOs targeting an intronic splicing suppressor site in SMN2 intron 6 [10], peptide nucleic acid (PNA) with an arginine-serine (RS) domain that is a site for binding to exon 7 (known as the ESSENCE method) [11], ASOs containing a sequence complementary to exon 7 and a sequence non-complementary to some ESE motifs (known as the TOES method) [12,13], trans-splicing RNA carrying the exon 7 sequence [14], ASOs targeting the ISS-N1 site [15,16], and ASOs targeting the 5 region of exon 8 [17]. ...
Article
Full-text available
Spinal muscular atrophy (SMA) is caused by survival motor neuron 1 SMN1 deletion. The survival motor neuron 2 (SMN2) encodes the same protein as SMN1 does, but it has a splicing defect of exon 7. Some antisense oligonucleotides (ASOs) have been proven to correct this defect. One of these, nusinersen, is effective in SMA-affected infants, but not as much so in advanced-stage patients. Furthermore, the current regimen may exhibit a ceiling effect. To overcome these problems, high-dose ASOs or combined ASOs have been explored. Here, using SMA fibroblasts, we examined the effects of high-concentration ASOs and of combining two ASOs. Three ASOs were examined: one targeting intronic splicing suppressor site N1 (ISS-N1) in intron 7, and two others targeting the 3′ splice site and 5′ region of exon 8. In our experiments on all ASO types, a low or intermediate concentration (50 or 100 nM) showed better splicing efficiency than a high concentration (200 nM). In addition, a high concentration of each ASO created a cryptic exon in exon 6. When a mixture of two different ASOs (100 nM each) was added to the cells, the cryptic exon was included in the mRNA. In conclusion, ASOs at a high concentration or used in combination may show less splicing correction and cryptic exon creation.
... Antisense oligonucleotides (AOs) are synthetic nucleic acid analogues, specifically designed to regulate target gene transcripts and protein expression. Common mechanistic strategies include: RNase H-mediated transcript degradation [1]; preventing the initiation of protein translation [2]; and modulating pre-mRNA splicing to either produce alternative functional protein isoforms, to overcome mutations or errors in splicing, or to produce a non-functional transcript [3][4][5][6]. The application of splice-switching antisense therapies to treat a broad variety of genetic disorders has greatly expanded over the past decade, with antisense drugs to treat Duchenne muscular dystrophy and spinal muscular atrophy now available in the clinic. ...
... ACU UUC CUU CUU UUU UAU UUU GUC UGA AAC UCC UUC UUU UUU AUU UUG UCU G ACU UUC CUU CUU UUU UAU UUU GUC U UGA GCA CUU UCC UUC UUU UUU AUU UUG UCU CAC UUU CCU UCU UUU UUA UU AAU GAC AGA CUU ACU UCU UAA UUU G UUU AAA AUG ACA GAC UUA CUU CUU AAU UUG AUG ACA GAC UUA CUU CUU AAU UUG UAU GUG CGU UGA AUU UUA GAC CUG GCU AUA A AGG CGG CGG CGG CGG GCC GUU GAA U AAG GGG GGA GGG GGU AGU GGA GGC G AAC GGG GGC AUC CAG CAC GGC AG UAC UUA CUG GUG GUC CUG AAG GGA A Some sequence overlap with 2 -O-methoxyethyl AOs described in[25].2 Previously used as a control[4]. ...
Article
Full-text available
The literature surrounding the use of antisense oligonucleotides continues to grow, with new disease and mechanistic applications constantly evolving. Furthermore, the discovery and advancement of novel chemistries continues to improve antisense delivery, stability and effectiveness. For each new application, a rational sequence design is recommended for each oligomer, as is chemistry and delivery optimization. To confirm oligomer delivery and antisense activity, a positive control AO sequence with well characterized target-specific effects is recommended. Here, we describe splice-switching antisense oligomer sequences targeting the ubiquitously expressed human and mouse SMN and Smn genes for use as control AOs for this purpose. We report two AO sequences that induce targeted skipping of SMN1/SMN2 exon 7 and two sequences targeting the Smn gene, that induce skipping of exon 5 and exon 7. These antisense sequences proved effective in inducing alternative splicing in both in vitro and in vivo models and are therefore broadly applicable as controls. Not surprisingly, we discovered a number of differences in efficiency of exon removal between the two species, further highlighting the differences in splice regulation between species.
... The notion of ASO-mediated rescue of FMRP production by targeting the gene responsible for the disorder will further advance the treatment options for patients with FXS. In fact, ASOs have been successfully used therapeutically to mediate exonskipping events in conditions such as Duchene muscular dystrophy, rheumatoid arthritis, cancer, and SMA (22,23). Similarly, the pioneering work by Shah et al. paves the way to restore functional FMRP in a subset of FXS patients. ...
... The RNA solution was then stored at −20 • C until use. Primers used to assess expression of hnRNPA1, SMN and other transcripts are listed in the supplementary Table S2 [36][37][38][39][40][41][42][43]. Approximately 200 ng of total RNA from each extraction was used as templates for single-step RT-PCR (Invitrogen, Waltham, MA, USA). ...
Article
Full-text available
Spinal muscular atrophy (SMA) is a severe, debilitating neuromuscular condition characterised by loss of motor neurons and progressive muscle wasting. SMA is caused by a loss of expression of SMN1 that encodes the survival motor neuron (SMN) protein necessary for the survival of motor neurons. Restoration of SMN expression through increased inclusion of SMN2 exon 7 is known to ameliorate symptoms in SMA patients. As a consequence, regulation of pre-mRNA splicing of SMN2 could provide a potential molecular therapy for SMA. In this study, we explored if splice switching antisense oligonucleotides could redirect the splicing repressor hnRNPA1 to the hnRNPA1b isoform and restore SMN expression in fibroblasts from a type I SMA patient. Antisense oligonucleotides (AOs) were designed to promote exon 7b retention in the mature mRNA and induce the hnRNPA1b isoform. RT-PCR and western blot analysis were used to assess and monitor the efficiency of different AO combinations. A combination of AOs targeting multiple silencing motifs in hnRNPA1 pre-mRNA led to robust hnRNPA1b induction, which, in turn, significantly increased expression of full-length SMN (FL-SMN) protein. A combination of PMOs targeting the same motifs also strongly induced hnRNPA1b isoform, but surprisingly SMN2 exon 5 skipping was detected, and the PMO cocktail did not lead to a significant increase in expression of FL-SMN protein. We further performed RNA sequencing to assess the genome-wide effects of hnRNPA1b induction. Some 3244 genes were differentially expressed between the hnRNPA1b-induced and untreated SMA fibroblasts, which are functionally enriched in cell cycle and chromosome segregation processes. RT-PCR analysis demonstrated that expression of the master regulator of these enrichment pathways, MYBL2 and FOXM1B, were reduced in response to PMO treatment. These findings suggested that induction of hnRNPA1b can promote SMN protein expression, but not at sufficient levels to be clinically relevant.
... However, our laboratory first recognised the applicability of these oligomers to splice switching in the mid-2000s [167][168][169], and subsequently, a collaboration was established to explore PMOs as agents to treat Duchenne muscular dystrophy (DMD) in association with AVI Pty Ltd, now known as Sarepta Therapeutics, Cambridge, MA. Our laboratory has since evaluated PMOs for translation blockade, exon skipping and exon inclusion, and selection of transcription start sites [170][171][172][173]. The first of our splice-switching applications to gain clinical approval is the PMO Exondys 51, targeting exon 51 of the DMD transcript. ...
Article
Full-text available
Polyglutamine (polyQ) ataxias are a heterogenous group of neurological disorders all caused by an expanded CAG trinucleotide repeat located in the coding region of each unique causative gene. To date, polyQ ataxias encompass six disorders: spinocerebellar ataxia types 1, 2, 3, 6, 7, and 17 and account for a larger group of disorders simply known as polyglutamine disorders, which also includes Huntington’s disease. These diseases are typically characterised by progressive ataxia, speech and swallowing difficulties, lack of coordination and gait, and are unfortunately fatal in nature, with the exception of SCA6. All the polyQ spinocerebellar ataxias have a hallmark feature of neuronal aggregations and share many common pathogenic mechanisms, such as mitochondrial dysfunction, impaired proteasomal function, and autophagy impairment. Currently, therapeutic options are limited, with no available treatments that slow or halt disease progression. Here, we discuss the common molecular and clinical presentations of polyQ spinocerebellar ataxias. We will also discuss the promising antisense oligonucleotide therapeutics being developed as treatments for these devastating diseases. With recent advancements and therapeutic approvals of various antisense therapies, it is envisioned that some of the studies reviewed may progress into clinical trials and beyond.
Article
Full-text available
Here we report a Survival Motor Neuron 2 (SMN2) super minigene, SMN2Sup, encompassing its own promoter, all exons, their flanking intronic sequences and the entire 3′-untranslated region. We confirm that the pre-mRNA generated from SMN2Sup undergoes splicing to produce a translation-competent mRNA. We demonstrate that mRNA generated from SMN2Sup produces more SMN than an identical mRNA generated from a cDNA clone. We uncover that overexpression of SMN triggers skipping of exon 3 of SMN1/SMN2. We define the minimal promoter and regulatory elements associated with the initiation and elongation of transcription of SMN2. The shortened introns within SMN2Sup preserved the ability of camptothecin, a transcription elongation inhibitor, to induce skipping of exons 3 and 7 of SMN2. We show that intron 1-retained transcripts undergo nonsense-mediated decay. We demonstrate that splicing factor SRSF3 and DNA/RNA helicase DHX9 regulate splicing of multiple exons in the context of both SMN2Sup and endogenous SMN1/SMN2. Prevention of SMN2 exon 7 skipping has implications for the treatment of spinal muscular atrophy (SMA). We validate the utility of the super minigene in monitoring SMN levels upon splicing correction. Finally, we demonstrate how the super minigene could be employed to capture the cell type-specific effects of a pathogenic SMN1 mutation.
Article
Manipulation of RNA splicing machinery has emerged as a drug modality. Here, we illustrate the potential of this novel paradigm to correct aberrant splicing events focused on the recent therapeutic advances in spinal muscular atrophy (SMA). SMA is an incurable neuromuscular disorder and at present the primary genetic cause of early infant death. This Review summarizes the exciting journey from the first reported SMA cases to the currently approved splicing-switching treatments, i.e., antisense oligonucleotides and small-molecule modifiers. We emphasize both chemical structures and molecular bases for recognition. We briefly discuss the advantages and disadvantages of these treatments and include the remaining challenges and future directions. Finally, we also predict that these success stories will contribute to further therapies for human diseases by RNA-splicing control.
Article
Monocytes play a crucial role in maintaining homeostasis and mediating a successful innate immune response. They also act as central players in diverse pathological conditions, thus making them an attractive therapeutic target. Within the bone marrow, monocytes arise from a committed precursor termed cMoP (Common Monocyte Progenitor). However, molecular mechanisms that regulate the differentiation of cMoP to various monocytic subsets remain unclear. Herein, we purified murine myeloid precursors for deep poly‐A enriched RNA sequencing to understand the role of alternative splicing in the development and differentiation of monocytes under homeostasis. Our analyses revealed intron retention to be the major alternative splicing mechanism involved in the monocyte differentiation cascade, especially in the differentiation of Ly6Chi monocytes to Ly6Clo monocytes. Furthermore, we found that the key genes regulated by intron retention in the differentiation of murine Ly6Chi to Ly6Clo monocytes were also conserved in humans. Our data highlight the unique role of intron retention in the regulation of the monocytic differentiation pathway.
Article
Full-text available
It has been estimated that 80% of the pre-mRNA undergoes alternative splicing, which exponentially increases the flow of biological information in cellular processes and can be an attractive therapeutic target. It is a crucial mechanism to increase genetic diversity. Disturbed alternative splicing is observed in many disorders, including neuromuscular diseases and carcinomas. Spinal Muscular Atrophy (SMA) is an autosomal recessive neurodegenerative disease. Homozygous deletion in 5q13 (the region coding for the motor neuron survival gene (SMN1)) is responsible for 95% of SMA cases. The nearly identical SMN2 gene does not compensate for SMN loss caused by SMN1 gene mutation due to different splicing of exon 7. A pathologically low level of survival motor neuron protein (SMN) causes degeneration of the anterior horn cells in the spinal cord with associated destruction of α-motor cells and manifested by muscle weakness and loss. Understanding the regulation of the SMN2 pre-mRNA splicing process has allowed for innovative treatment and the introduction of new medicines for SMA. After describing the concept of splicing modulation, this review will cover the progress achieved in this field, by highlighting the breakthrough accomplished recently for the treatment of SMA using the mechanism of alternative splicing.
Article
Full-text available
While intron retention (IR) is considered a widely conserved and distinct mechanism of gene expression control, its regulation is poorly understood. Here we show that DNA methylation directly regulates IR. We also find reduced occupancy of MeCP2 near the splice junctions of retained introns, mirroring the reduced DNA methylation at these sites. Accordingly, MeCP2 depletion in tissues and cells enhances IR. By analysing the MeCP2 interactome using mass spectrometry and RNA co-precipitation, we demonstrate that decreased MeCP2 binding near splice junctions facilitates IR via reduced recruitment of splicing factors, including Tra2b, and increased RNA polymerase II stalling. These results suggest an association between IR and a slower rate of transcription elongation, which reflects inefficient splicing factor recruitment. In summary, our results reinforce the interdependency between alternative splicing involving IR and epigenetic controls of gene expression.
Article
Full-text available
Purpose Naturally occurring tumor suppressor microRNA-34a (miR-34a) downregulates the expression of >30 oncogenes across multiple oncogenic pathways, as well as genes involved in tumor immune evasion, but is lost or under-expressed in many malignancies. This first-in-human, phase I study assessed the maximum tolerated dose (MTD), safety, pharmacokinetics, and clinical activity of MRX34, a liposomal miR-34a mimic, in patients with advanced solid tumors. Patients and Methods Adult patients with solid tumors refractory to standard treatment were enrolled in a standard 3 + 3 dose escalation trial. MRX34 was given intravenously twice weekly (BIW) for three weeks in 4-week cycles. Results Forty-seven patients with various solid tumors, including hepatocellular carcinoma (HCC; n = 14), were enrolled. Median age was 60 years, median prior therapies was 4 (range, 1–12), and most were Caucasian (68%) and male (57%). Most common adverse events (AEs) included fever (all grade %/G3%: 64/2), fatigue (57/13), back pain (57/11), nausea (49/2), diarrhea (40/11), anorexia (36/4), and vomiting (34/4). Laboratory abnormalities included lymphopenia (G3%/G4%: 23/9), neutropenia (13/11), thrombocytopenia (17/0), increased AST (19/4), hyperglycemia (13/2), and hyponatremia (19/2). Dexamethasone premedication was required to manage infusion-related AEs. The MTD for non-HCC patients was 110 mg/m2, with two patients experiencing dose-limiting toxicities of G3 hypoxia and enteritis at 124 mg/m2. The half-life was >24 h, and Cmax and AUC increased with increasing dose. One patient with HCC achieved a prolonged confirmed PR lasting 48 weeks, and four patients experienced SD lasting ≥4 cycles. Conclusion MRX34 treatment with dexamethasone premedication was associated with acceptable safety and showed evidence of antitumor activity in a subset of patients with refractory advanced solid tumors. The MTD for the BIW schedule was 110 mg/m2 for non-HCC and 93 mg/m2 for HCC patients. Additional dose schedules of MRX34 have been explored to improve tolerability.
Article
Full-text available
Objective: To examine safety, tolerability, pharmacokinetics, and preliminary clinical efficacy of intrathecal nusinersen (previously ISIS-SMNRx), an antisense oligonucleotide designed to alter splicing of SMN2 mRNA, in patients with childhood spinal muscular atrophy (SMA). Methods: Nusinersen was delivered by intrathecal injection to medically stable patients with type 2 and type 3 SMA aged 2-14 years in an open-label phase 1 study and its long-term extension. Four ascending single-dose levels (1, 3, 6, and 9 mg) were examined in cohorts of 6-10 participants. Participants were monitored for safety and tolerability, and CSF and plasma pharmacokinetics were measured. Exploratory efficacy endpoints included the Hammersmith Functional Motor Scale Expanded (HFMSE) and Pediatric Quality of Life Inventory. Results: A total of 28 participants enrolled in the study (n = 6 in first 3 dose cohorts; n = 10 in the 9-mg cohort). Intrathecal nusinersen was well-tolerated with no safety/tolerability concerns identified. Plasma and CSF drug levels were dose-dependent, consistent with preclinical data. Extended pharmacokinetics indicated a prolonged CSF drug half-life of 4-6 months after initial clearance. A significant increase in HFMSE scores was observed at the 9-mg dose at 3 months postdose (3.1 points; p = 0.016), which was further increased 9-14 months postdose (5.8 points; p = 0.008) during the extension study. Conclusions: Results from this study support continued development of nusinersen for treatment of SMA. Classification of evidence: This study provides Class IV evidence that in children with SMA, intrathecal nusinersen is not associated with safety or tolerability concerns.
Article
Full-text available
Objective: To continue evaluation of the long-term efficacy and safety of eteplirsen, a phosphorodiamidate morpholino oligomer designed to skip DMD exon 51 in patients with Duchenne muscular dystrophy (DMD). Three-year progression of eteplirsen-treated patients was compared to matched historical controls (HC). Methods: Ambulatory DMD patients who were ≥7 years old and amenable to exon 51 skipping were randomized to eteplirsen (30/50mg/kg) or placebo for 24 weeks. Thereafter, all received eteplirsen on an open-label basis. The primary functional assessment in this study was the 6-Minute Walk Test (6MWT). Respiratory muscle function was assessed by pulmonary function testing (PFT). Longitudinal natural history data were used for comparative analysis of 6MWT performance at baseline and months 12, 24, and 36. Patients were matched to the eteplirsen group based on age, corticosteroid use, and genotype. Results: At 36 months, eteplirsen-treated patients (n = 12) demonstrated a statistically significant advantage of 151m (p < 0.01) on 6MWT and experienced a lower incidence of loss of ambulation in comparison to matched HC (n = 13) amenable to exon 51 skipping. PFT results remained relatively stable in eteplirsen-treated patients. Eteplirsen was well tolerated. Analysis of HC confirmed the previously observed change in disease trajectory at age 7 years, and more severe progression was observed in patients with mutations amenable to exon skipping than in those not amenable. The subset of patients amenable to exon 51 skipping showed a more severe disease course than those amenable to any exon skipping. Interpretation: Over 3 years of follow-up, eteplirsen-treated patients showed a slower rate of decline in ambulation assessed by 6MWT compared to untreated matched HC.
Article
Full-text available
The poly(A)-Binding Protein Nuclear 1 is encoded by the PABPN1 gene whose mutations result in oculopharyngeal muscular dystrophy, a late onset disorder for which the molecular basis remains unknown. Despite recent studies investigating the functional roles of PABPN1, little is known about its regulation. Here we show that PABPN1 negatively controls its own expression to maintain homeostatic levels in human cells. Transcription from the PABPN1 gene results in the accumulation of two major isoforms: an unspliced nuclear transcript that retains the 3' -terminal intron and a fully spliced cytoplasmic mRNA. Increased dosage of PABPN1 protein causes a significant decrease in the spliced:unspliced ratio, reducing the levels of endogenous PABPN1 protein. We also show that PABPN1 autoregulation requires inefficient splicing of its 3' -terminal intron. Our data suggest that autoregulation occurs via the binding of PABPN1 to an adenosine (A)-rich region in its 3' UTR, which promotes retention of the 3' -terminal intron and clearance of intron-retained pre-mRNAs by the nuclear exosome. Our findings unveil a mechanism of regulated intron retention coupled to nuclear pre-mRNA decay that functions in the homeostatic control of PABPN1 expression. Copyright © 2015, American Society for Microbiology. All Rights Reserved.
Article
Full-text available
Alternative splicing modulates expression of most human genes. The kinetic model of cotranscriptional splicing suggests that slow elongation expands and that fast elongation compresses the "window of opportunity" for recognition of upstream splice sites, thereby increasing or decreasing inclusion of alternative exons. We tested the model using RNA polymerase II mutants that change average elongation rates genome-wide. Slow and fast elongation affected constitutive and alternative splicing, frequently altering exon inclusion and intron retention in ways not predicted by the model. Cassette exons included by slow and excluded by fast elongation (type I) have weaker splice sites, shorter flanking introns, and distinct sequence motifs relative to "slow-excluded" and "fast-included" exons (type II). Many rate-sensitive exons are misspliced in tumors. Unexpectedly, slow and fast elongation often both increased or both decreased inclusion of a particular exon or retained intron. These results suggest that an optimal rate of transcriptional elongation is required for normal cotranscriptional pre-mRNA splicing. © 2014 Fong et al.; Published by Cold Spring Harbor Laboratory Press.
Article
Full-text available
Alternative splicing (AS) of precursor RNAs is responsible for greatly expanding the regulatory and functional capacity of eukaryotic genomes. Of the different classes of AS, intron retention (IR) is the least well understood. In plants and unicellular eukaryotes, IR is the most common form of AS, whereas in animals, it is thought to represent the least prevalent form. Using high-coverage poly(A)(+) RNA-seq data, we observe that IR is surprisingly frequent in mammals, affecting transcripts from as many as three-quarters of multiexonic genes. A highly correlated set of cis features comprising an "IR code" reliably discriminates retained from constitutively spliced introns. We show that IR acts widely to reduce the levels of transcripts that are less or not required for the physiology of the cell or tissue type in which they are detected. This "transcriptome tuning" function of IR acts through both nonsense-mediated mRNA decay and nuclear sequestration and turnover of IR transcripts. We further show that IR is linked to a cross-talk mechanism involving localized stalling of RNA polymerase II (Pol II) and reduced availability of spliceosomal components. Collectively, the results implicate a global checkpoint-type mechanism whereby reduced recruitment of splicing components coupled to Pol II pausing underlies widespread IR-mediated suppression of inappropriately expressed transcripts.
Article
Full-text available
Splicing is functionally coupled to transcription, linking the rate of RNA polymerase II (Pol II) elongation and the ability of splicing factors to recognize splice sites (ss) of various strengths. In most cases, slow Pol II elongation allows weak splice sites to be recognized, leading to higher inclusion of alternative exons. Using CFTR alternative exon 9 (E9) as a model, we show here that slowing down elongation can also cause exon skipping by promoting the recruitment of the negative factor ETR-3 onto the UG-repeat at E9 3' splice site, which displaces the constitutive splicing factor U2AF65 from the overlapping polypyrimidine tract. Weakening of E9 5' ss increases ETR-3 binding at the 3' ss and subsequent E9 skipping, whereas strengthening of the 5' ss usage has the opposite effect. This indicates that a delay in the cotranscriptional emergence of the 5' ss promotes ETR-3 recruitment and subsequent inhibition of E9 inclusion.
Article
Background: Nusinersen is a 2'-O-methoxyethyl phosphorothioate-modified antisense drug being developed to treat spinal muscular atrophy. Nusinersen is specifically designed to alter splicing of SMN2 pre-mRNA and thus increase the amount of functional survival motor neuron (SMN) protein that is deficient in patients with spinal muscular atrophy. Methods: This open-label, phase 2, escalating dose clinical study assessed the safety and tolerability, pharmacokinetics, and clinical efficacy of multiple intrathecal doses of nusinersen (6 mg and 12 mg dose equivalents) in patients with infantile-onset spinal muscular atrophy. Eligible participants were of either gender aged between 3 weeks and 7 months old with onset of spinal muscular atrophy symptoms between 3 weeks and 6 months, who had SMN1 homozygous gene deletion or mutation. Safety assessments included adverse events, physical and neurological examinations, vital signs, clinical laboratory tests, cerebrospinal fluid laboratory tests, and electrocardiographs. Clinical efficacy assessments included event free survival, and change from baseline of two assessments of motor function: the motor milestones portion of the Hammersmith Infant Neurological Exam-Part 2 (HINE-2) and the Children's Hospital of Philadelphia Infant Test of Neuromuscular Disorders (CHOP-INTEND) motor function test, and compound motor action potentials. Autopsy tissue was analysed for target engagement, drug concentrations, and pharmacological activity. HINE-2, CHOP-INTEND, and compound motor action potential were compared between baseline and last visit using the Wilcoxon signed-rank test. Age at death or permanent ventilation was compared with natural history using the log-rank test. The study is registered at ClinicalTrials.gov, number NCT01839656. Findings: 20 participants were enrolled between May 3, 2013, and July 9, 2014, and assessed through to an interim analysis done on Jan 26, 2016. All participants experienced adverse events, with 77 serious adverse events reported in 16 participants, all considered by study investigators not related or unlikely related to the study drug. In the 12 mg dose group, incremental achievements of motor milestones (p<0·0001), improvements in CHOP-INTEND motor function scores (p=0·0013), and increased compound muscle action potential amplitude of the ulnar nerve (p=0·0103) and peroneal nerve (p<0·0001), compared with baseline, were observed. Median age at death or permanent ventilation was not reached and the Kaplan-Meier survival curve diverged from a published natural history case series (p=0·0014). Analysis of autopsy tissue from patients exposed to nusinersen showed drug uptake into motor neurons throughout the spinal cord and neurons and other cell types in the brainstem and other brain regions, exposure at therapeutic concentrations, and increased SMN2 mRNA exon 7 inclusion and SMN protein concentrations in the spinal cord. Interpretation: Administration of multiple intrathecal doses of nusinersen showed acceptable safety and tolerability, pharmacology consistent with its intended mechanism of action, and encouraging clinical efficacy. Results informed the design of an ongoing, sham-controlled, phase 3 clinical study of nusinersen in infantile-onset spinal muscular atrophy. Funding: Ionis Pharmaceuticals, Inc and Biogen.
Article
Late diagnosis contributes to pancreatic cancer (PaCa) dismal prognosis, urging for reliable, early detection. Serum-exosome protein and/or miRNA markers might be suitable candidates, which we controlled for patients with PaCa. Protein markers were selected according to expression in exosomes of PaCa cell line culture supernatants, but not healthy donors' serum-exosomes. miRNA was selected according to abundant recovery in microarrays of PaCa patients', but not healthy donors' serum-exosomes and exosome-depleted serum. According to these preselections, serum-exosomes were tested by flow cytometry for the PaCa-initiating cell (PaCIC) markers CD44v6, Tspan8, EpCAM, MET and CD104. Serum-exosomes and exosome-depleted serum was tested for miR-1246, miR-4644, miR-3976, and miR-4306 recovery by qRT-PCR. The majority (95%) of PaCa-patients' (131) and patients with nonPa-malignancies reacted with a panel of anti-CD44v6, -Tspan8, -EpCAM and -CD104. Serum-exosomes of healthy donors' and patients with non-malignant diseases were not reactive. Recovery was tumor grading and staging independent including early stages. The selected miR-1246, miR-4644, miR-3976 and miR-4306 were significantly upregulated in 83% of PaCa serum-exosomes, but rarely in control groups. These miRNA were also elevated in exosome-depleted serum of PaCa patients, but at a low level. Concomitant evaluation of PaCIC and miRNA serum-exosome marker panels significantly improved sensitivity (1.00, CI 0.95-1) with a specificity of 0.80 (CI: 0.67-0.90) for PaCa versus all others groups and of 0.93 (CI: 0.81-0.98) excluding nonPa-malignancies. Thus, the concomitant evaluation of PaCIC and PaCa-related miRNA marker panels awaits retrospective analyses of larger cohorts, as it should allow for a highly sensitive, minimally-invasive PaCa diagnostics. © 2014 Wiley Periodicals, Inc.