ChapterPDF Available

Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular Mechanism and Development of Therapeutics

Authors:

Abstract and Figures

Spinal muscular atrophy (SMA) is an autosomal recessive neurodegenerative disease and one of the most common genetic causes of infant death. The loss or mutation of the SMN1 gene results in reduced SMN protein level leading to motor neuron death and progressive muscle atrophy. Although recent progress has been made in our understanding of the molecular mechanisms underlying the pathogenesis of the disease, there is currently no cure for SMA. In this review, we summarize the clinical manifestations, molecular pathogenesis, diagnostic strategy and development of therapeutic regimes for the better understanding and treatment of SMA.
Content may be subject to copyright.
Chapter 23
Spinal Muscular Atrophy: Classification,
Diagnosis, Background, Molecular Mechanism
and Development of Therapeutics
Faraz Tariq Farooq, Martin Holcik and
Alex MacKenzie
Additional information is available at the end of the chapter
http://dx.doi.org/10.5772/53800
1. Introduction
Spinal muscular atrophy (SMA) is an autosomal recessive neurodegenerative disease and one
of the most common genetic causes of infant death. The loss or mutation of the SMN1 gene
results in reduced SMN protein level leading to motor neuron death and progressive muscle
atrophy. Although recent progress has been made in our understanding of the molecular
mechanisms underlying the pathogenesis of the disease, there is currently no cure for SMA.
In this review, we summarize the clinical manifestations, molecular pathogenesis, diagnostic
strategy and development of therapeutic regimes for the better understanding and treatment
of SMA.
2. Epidemiology
Spinal muscular atrophy (SMA) is an autosomal recessive neuromuscular disorder character‐
ized by the loss of motor neurons from the anterior horn of the spinal cord which leads to
muscle weakness, hypotonia and ultimately muscle atrophy [1]. With a pan ethnic incidence
of 1:11,000 live births and a carrier frequency of 1:50, SMA is one of the leading genetic causes
of infant death globally [1-5].
© 2013 Farooq et al.; licensee InTech. This is an open access article distributed under the terms of the Creative
Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits unrestricted use,
distribution, and reproduction in any medium, provided the original work is properly cited.
3. Clinical classification
Due to the range of clinical severity, SMA is broadly classified into four major categories
characterized by the age of onset as well as severity of the disease [6-9]. SMA type I, which
was originally described by Werdnig and Hoffmann in the late 18th century is the most severe
and prevalent form of the disease and accounts for more than 50% of the known diagnosed
cases of SMA. Type I SMA presents within the first six months after birth and although
historically patients succumbed within the first 2 years of life, with better ventilatory and
nutritional support, the life expectancy of children with type I SMA can be increased beyond
the 5th birthday. Infants with type I SMA experience a rapid loss of skeletal muscle mass with
profound hypotonia and general muscle weakness characterized by poor head control,
difficulty with suckling, swallowing and an inability to sit without support. These children
develop problems with breathing over time due to impaired bulbar function and respiratory
muscle weakness leading to respiratory insufficiency. Respiratory failure due to aspiration
pneumonia is an important cause of SMA mortality [6, 10, 11]. The intermediate form of SMA,
known as type II, has an onset between 6 and 18 months of age. Patients with type II SMA can
sit unaided but still develop progressive muscle weakness and can never stand or walk on
their own. Other symptoms and physical signs include respiratory insufficiency due to
reduced bulbar function, poor weight gain, fine hand tremors and joint contractures [6]. SMA
type III has an onset between 18 months to 30 years of age. Patients are able to stand and walk
unaided, however they develop variable degree of muscle weakness which leads to a broad
spectrum of physical signs and symptoms. While most walk independently, some lose
ambulation during early adulthood and require wheelchair assistance. Others develop cramps
and joint overuse problems; some develop scoliosis [6, 12, 13]. Type IV SMA is the mildest
form of the condition and is characterized by adult onset with normal mobility. They have
mild muscle weakness in adulthood with normal longevity [6] (Table 1).
SMA Type Other Names Age of Onset Life
Span
Highest Motor Activity
Type I
years
(Severe)
Werdnig- Hoffmann
disease
0-6 months 2-5 Never sit
Type II
(Intermediate)
SMA, Dubowitz
type
7-18 months >2 years Sit, Never stand
Type III
(Mild)
Kugelberg-
Welander
disease
>18 months Adult Stand and walk
(may require assistance)
Type IV
(Adult)
------------- Adulthood Normal Walk during
adulthood-unassisted
(some muscle weakness)
Table 1. Classification of SMA disease
Neurodegenerative Diseases562
4. Diagnosis and treatment
The diagnosis of SMA is made by a thorough patient history and physical examination
followed by genetic testing. The survival of motor neuron (SMN) -1 genotyping has to a large
degree replaced electromyography (EMG) and muscle biopsies (Fig 1) [2, 14]. There is in 2012
no cure for SMA; current treatment is symptomatic and supportive. This includes clinical
management through family education and counselling along with attention to pulmonary,
gastrointestinal/nutrition and orthopedics/rehabilitation in an effort to managing symptoms
of the patients [9].
SMA with
clinical features
Genetic testing
(SMN1 gene test)
Homozygous
SMN1 deletion
Confirmation of
5q SMA diagnosis
No homozygous
deletion of SMN1
Repeat Clinical tests
EMG, NCS/RNS and CK
Demelinating or
axonal neuropathy,
NMJ disorder,
Myopathy &
Muscular atrophy.
Uncommon SMA
features but
neurogenic EMG,
normal CK
Diffuse
weakness,
normal EMG,
normal NCS
normal CK
Proximal>distal
weakness,
neurogenic EMG,
normal CK
Muscle or nerve
biopsy,
genetic tests for
muscular
dystrophies,
myopathies or
neuropathies.
Consider other
motor neuron
disorders such as
SMARD, Distal
SMA, X-SMA,
juvenile ALS
MRI brain and
spinal cord,
conduct
metabolic
screens
SMN1 gene copy
count
Two
SMN1
copy
One
SMN1
copy
Perform SMN1
gene sequencing
No mutation
Mutation found
Diagnosis of SMN-related
SMA remains unconfirmed
Confirmation of 5q SMA
diagnosis
Figure 1. SMA diagnosis
5. Genetics of the disease
The SMA disease causing SMN1 gene maps to a complex genomic region of chromosome
5q13.1. This region is characterized by an inverted duplication of the element with 4 genes
Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular mechanism…
http://dx.doi.org/10.5772/53800
563
(SMN, neuronal apoptosis inhibitor protein {NAIP}, SERF and GTFH2) present in telomeric and
centromeric copies (Fig 2a) [15, 16]. In 1995, it was reported that homozygous deletions of the
SMN1 gene were observed in and thus likely the cause of 95% of SMA patients [15]. All SMA
patients have one or more copies of a nearly identical gene, SMN2. These two genes are
distinguished by five nucleotide changes in exon 7 and 8. The critical nucleotide difference
which makes SMN2 only partially functional is a C to T transition at position 6 of exon 7. This
change leads to the exclusion of exon 7 in the majority of transcripts. This mRNA is subse‐
quently translated to form an unstable truncated non-oligomerizing isoform of SMN protein.
However, SMN2 gene still produces 5-10% functional full length SMN transcripts (Fig 1.2b)
[15, 17, 18]. The SMN2 gene is present in variable copy numbers in the population; all SMA
patients have one or more copy of the SMN2 gene which, due to its partial functionality, acts
as a positive disease modifier. There is thus an inverse correlation between the number of
SMN2 gene (which can produce between 10-50% of SMN protein depending on copy number)
and the severity of the disease [2]. Low levels of SMN protein allows embryonic development
but is not enough, in the long term, to allow motor neurons to survive in the spinal cord [19,
20]. Type I patients usually have 2 copies whereas Type II have 3 copies of SMN2. Type III and
IV have 3-4 copies of the SMN2 gene. Individuals with 5 or more copies of the SMN2 gene,
despite having no functional SMN1 gene are completely asymptomatic and are protected
against the disease manifestation.
a.
GTFH2 NAIP SMN2 SERF SERF SMN1 NAIP GTFH2
500 kb (Centromeric copy) 500 kb (Telomeric copy)
Chromosome 5q
SMN2 SMN1
T
*C
*
Chromosome 5q13
SMN∆7FL-SMN
~10 % ~90 %
Full-length Unstable, truncated
& degraded
Gene
mRNA
Protein
FL-SMN
~100 %
Full-length
b.
Figure 2.
Figure 2. (a) Human SMN locus and (b) genetics of SMA patients
Neurodegenerative Diseases
564
6. Pathology
The pathological hallmark of all forms of SMA is the loss of motor neurons from the lower
brainstem and the anterior horn of the spinal cord [21]. Anterograde axonal degeneration
results in denervation of the myocytes within the motor unit. This sometimes leads to rein‐
nervation of muscle, where adjacent uninjured motor neurons sprout leads to fiber type
grouping of myocytes. Histopathologic assessment of SMA muscle tissues reveals a large
number of rounded atrophic fibers resulting from denervation. The widely held notion had
been that SMA is primarily a neuronopathy (involving the cell body) with secondary degen‐
eration of the axons. However, more recent observations in the field have shifted the focus of
SMA pathology from the motor neuron cell body to the distal axon [22, 23] and the possibility
of a synaptopathic defect [20, 24]. Specifically it has been suggested that the presynaptic
transcriptome may be in some manner dysregulated; the direct inference is that SMN plays a
role in the peripheral transport of critical mRNA, among which is that species encoding beta-
actin. Regardless of the subcellular location of SMN mediated pathology, SMA is primarily
considered as a motor neuron disease and consequently treatment strategies focus on drugs
which can cross the blood brain barrier (BBB) to target the central nervous system (CNS).
However, motor neuron autonomy of SMA pathogenesis has recently been called into question
as multi-system involvement (including cardiovascular, peripheral necrosis and liver defects)
have been reported recently in both SMA patients and SMA mice models [25-33]. In addition,
one report has outlined the superiority of systemic SMN antisense oligonucleotide (ASO)
therapy compared with intrathecal delivery in severe murine SMA calling into question the
exclusive role of the motor neuron in disease causation [33].
7. Function of the SMN protein
SMN is a 294 amino acid long ubiquitously expressed protein with a molecular weight of 38
kilodaltons (kD). SMN is found in both the nucleus and cytoplasm. Within the nucleus, it is
localized both throughout the nucleoplasm as well as in nuclear structures called Gems and
Cajal bodies [34]. It is also found in abundance within the growth cones of the motor neurons
[35]. SMN has been implicated in ribonucleoprotein biogenesis (e.g. assembly, metabolism and
transport of various ribonucleoproteins), as well as playing a major role in the splicing
machinery. It is part of a multiprotein complex comprised of Gemins [2-8], spliceosomal U-
snRNPs, Sm proteins and profilins called the SMN complex. This complex is essential for the
biogenesis of snRNPs [36-45]. Given the variety of roles that SMN has been implicated in, not
surprisingly, the complete absence of SMN genes is embryonically lethal in virtually all
metazoan life forms tested, indeed even cell cultures cannot survive without SMN [19, 20, 46].
8. Molecular mechanism: splicing defect in SMA
Splicing is mediated by a complex called the spliceosome, the activity of which depends on a
number of factors. In particular, various cis- and trans-acting elements regulate the splicing of
Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular mechanism…
http://dx.doi.org/10.5772/53800
565
both SMN1 and SMN2. The C-T transition at position 6 of exon 7 in SMN2 gene disrupts the
function of an exonic splice enhancer (ESE; recognized by SF2/ASF to promote exon 7 inclusion)
and/or creates an exonic splice suppressor (ESS; recognized by hnRNP A1/A2) which results
in exon 7 skipping (Fig 3) [47-53].
Exon 6 Exon 8 Exon 7
C
Exon 6 Exon 8 Exon 7
U
SF2/
ASF
hnRNP
A1/A2
SF2/
ASF
SMN1 derived mRNA
SMN2 derived mRNA
100%
~10%
~90%
Full length
SMN mRNA
SMN ∆7
mRNA
ESE/ESS
ESE/ESS
ESE
ESE
Figure 3.
Figure 3. Splicing in SMA
9. Therapeutic strategies
Although there is no cure for SMA, the SMN2 gene locus serves as a target for SMA treatment.
The general treatment strategies for SMA are to compensate fully or in part for the absence of
SMN1 gene by increasing the levels of functional SMN protein levels though three distinct
approaches: i) to induce the expression of SMN2, ii) to modulate splicing of SMN2 transcript,
and iii) to stabilize the full length SMN mRNA and/or protein. In addition, gene and stem cell
therapies are also under development for the treatment of SMA. These and other strategies are
discussed below.
1. SMN dependent therapies: As outlined above, there is an inverse correlation between
the SMN2 gene copy number and disease severity [54, 55] which implies that directly
targeting the SMN2 gene in SMA patients through different pathways could be one key
for the development of a SMA drug treatment. Alternatively, SMN protein can also be
produced through gene replacement therapy.
Neurodegenerative Diseases
566
a. Activation of SMN2 promoter: Histone deacetylases (HDACs) repress transcription of
genes including SMN2 by chromatin condensation. Thus, HDAC inhibitors can increase
transcription of the SMN2 gene and can produce more full length SMN transcripts and
protein which may have a beneficial effect in patients. Various HDAC inhibitors have
been analyzed in cell culture, mouse models and in clinical trials as potential therapeutic
for SMA. Sodium butyrate, Valproic acid (VPA) and phenylbutyrate showed promise in
cell culture and mouse models and were also well tolerated by the patients [56-63].
However, no clinical improvement was observed in SMA patients with HDAC inhibitors
[61-63].
Recent studies with other HDAC inhibitors, LBH589, Trichostatin A (TSA) and Suberoy‐
lanilide hydroxamic acid (SAHA) showed SMN2 gene induction in culture as well as in
a number of animal models of neurodegeneration [62, 64-66]. In addition to these
compounds, we have shown that the lactation hormone prolactin (PRL) which can both
cross the blood brain barrier and, through binding to its receptor, activate the JAK2/STAT5
pathway also upregulates SMN2 gene transcription [68]. Interestingly the degree of
induction in SMN seen with the prolactin in the genetically engineered ∆7 SMA mouse
model (where SMN2 gene is the only source of SMN protein) is significantly greater than
that seen in cell culture and wild type mice. We have determined that this is because of
the difference between the promoter regions in SMN1 and SMN2 genes, the latter uniquely
having STAT5a transcription binding motifs. This might prove beneficial as all SMA
patients have SMN2 as the only source of SMN protein. Since PRL has been successfully
tested and proven safe in humans for the treatment of lactation deficient mothers [67], it
may bypass other compounds which are yet to be tested for clinical safety and join the
short list of drugs which may have immediate potential SMA therapeutic potential [68].
b. Correction of splicing: The suppression of exon 7 skipping to produce more full length
transcript from the SMN2 gene is another treatment strategy being explored for SMA.
HDAC inhibitors such as VPA, TSA and sodium butyrate appear to have a dual effect on
SMN mRNA expression; they not only open chromatin structure and therefore increase
the rate of transcription but also appear to affect the splicing process [56-58, 64]. The
antibiotic aclarubicin has been shown to increase full length SMN transcript by altering
the splicing process in vitro [69]. The most promising compounds which correct splicing
by preventing SMN2 exon 7 skipping are antisense oligos (ASOs). An ASO complemen‐
tary to SMN2 exon 7 pre-mRNA sequences has been shown to inhibit binding of negative
splicing factors and increase full length SMN transcript and protein production [30, 33,
70, 71]. The major hurdle in using ASOs for SMA therapeutics, however, is their inability
to cross the blood brain barrier. However, Hua et al. 2011 documented a marked im‐
provement in motor function along with an increase in survival in SMA mice with
systemic delivery of ASO which results into increase in SMN levels mostly in peripheral
tissues especially in liver. Interestingly, they documented only a slight increase in SMN
levels in CNS tissues [33]. However, there are a number of issues which need to be
addressed before clinical introduction of ASOs for SMA treatment (clinical safety, quantity
of ASO, cost, immune response etc) [72].
Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular mechanism…
http://dx.doi.org/10.5772/53800
567
c. Full length SMN transcript stabilization: In this relatively new approach by by Singh et
al., decapping enzyme DcpS, an integral part of the RNA degradation machinery, was
targeted by C5-substituted quinazolines which interact and open the enzyme into a
catalytically inactivated conformation. Full length SMN mRNA decay is in this fashion
blocked, ultimately increasing SMN protein in cell culture [73].
In a different approach, SMN mRNA has been shown to have a specific AU rich element
(ARE) region in its 3’ UTR which marks the mRNA for degradation. Our laboratory has
shown that activation of the p38 pathway results in the accumulation of RNA binding
protein HuR in the cytoplasm which then binds to the ARE in 3’UTR region of SMN mRNA
and stabilizes the transcript. Importantly, transcript stabilization is not associated with
any discernible inhibition of SMN protein translation. This study provided a novel
mechanism through which SMN mRNA could be stabilized using p38 activating com‐
pounds which can cross the blood brain barrier to develop new therapeutics for SMA
treatment [74].
d. Full length SMN protein stabilization: Aminoglycosides are class of antibiotics which
have been shown to mask premature stop codon mutations in some genes, allowing read
through translation to occur. This moderates translation termination through an alteration
in the conformation of the ribosomal reading site. Various aminoglycosides including
tobramycin and amikacin have been used successfully in patient fibroblasts to increase
SMN protein levels. However, their in vivo efficacy and safety has yet to be demonstrated
[75-77].
An alternative potential therapeutic approach involves targeting the ubiquitin-protea‐
some pathway which mediates intracellular protein turnover. Proteins are marked with
poly ubiquitin (Ub) molecules by the action of the enzymes E1 (Ub activating enzyme),
E2 (Ub conjugating enzyme) and E3 (Ub ligase). The polyubiquitin modification marks
the protein for destruction by the proteasome complex. SMN is one of the many proteins
degraded by the ubiquitin proteasome pathway. It has been shown that FDA approved
proteasome inhibitor bortezomib increases SMN both in vitro and in vivo by blocking
proteolysis of SMN protein. However, it should be noted that bortezomib cannot cross
the BBB; thus, it must be used in combination with other drugs which can cross the BBB
for the treatment of SMA [78].
e. Gene therapy: One of the most encouraging SMA therapeutic advances is the use of gene
therapy which shows significant promise. In the last three years several groups have used
self complementary adeno-associated virus (scAAV) 8 and 9 vectors carrying the SMN1
cDNA to treat mice models of SMA, resulting in the most dramatic extension in the life
span of mice yet observed combined with an overall amelioration of disease phenotype
[79-82]. However, early pre-symptomatic intervention is necessary for the success of this
therapy as is seen with other treatment strategies as well. Moreover, several challenges
must be addressed for this mode of SMA treatment before bringing it to clinical application
successfully. The most pressing issues are clinical safety, dealing with the cross-species
barriers, the cost of virus production along with the possibility of an immune response to
AAV which can neutralize its impact [83].
Neurodegenerative Diseases
568
2. SMN-independent strategies: There have been some recent advances in SMN-independ‐
ent strategies for the treatment of SMA. These include:
a. Stem cell therapy: Stem cell therapy has generated much attention as a treatment for
motor neuron diseases, including SMA, through replacement of the lost motor neurons
and, more realistically perhaps, supporting the existing neuron population. Primary
murine neuronal stem cells as well as embryonic stem cell-derived neural stem cells
injected into the spinal cord of animal models of SMA have been shown to ameliorate
disease phenotype and increase survival [84, 85]. It is unclear whether this is through
motor neuron and other cell replacement and/or through neuroprotection of host motor
neurons by the numerous factors released from the donor cells. Although induced
pluripotent stem (iPS) cells from an SMA patient have been differentiated into motor
neurons [86, 87], there are several obstacles which hinder their use as a therapeutic for
SMA treatment. These challenges include the production of the large number of stem cells
and their successful transplantation into the patients, which could populate and cover the
entire nervous system. Also, lentivirus vectors are used to deliver the cocktail of factors,
required to produce iPS cells in vitro; these would be unsuitable for use in patients as they
have the potential for insertional mutagenesis which could result into oncogenesis.
Finally, even if motor neurons could develop in situ, the prospect that they would at a
meaningful level connect with the host CNS must be viewed as highly unlikely at this
time.
b. Modifying neuromuscular junctions through actin dynamics: The pharmacological
Rho-kinase inhibitor (downstream effector of RhoA-GTP which plays role in actin
dynamics) dramatically increases the life span of a mild SMA mouse model and improves
disease phenotype. This improvement in the disease phenotype is independent of SMN
increase, mainly through making neuromuscular junctions (NMJ) better, larger and more
mature [88]. This suggests that there are novel SMN independent avenues for the
development of therapeutics for SMA.
10. Future directions
Combination therapy: The impressive results seen so far with gene therapy and ASO's in the
field of SMA will be difficult to equal with a monotherapy approach. However, unless and
until gene therapy and ASO treatments are cleared for clinical safety as a therapeutic option
for SMA treatment, combinatorial approaches for SMA shall likely be necessary to target not
only CNS but also other tissues which are affected because of a lack of SMN. As outlined above,
SMA can be targeted through different approaches, we can in a safe combination use com‐
pounds which are already FDA approved and can increase SMN levels through SMN2 gene
activation (such as PRL) along with SMN2 transcript stabilizers (p38 pathway activators such
as celecoxib) and/or SMN protein stabilizer (proteasome inhibitor bortezomib) [18] and/or
neuroprotective compounds (Rho kinase inhibitor) [19], or a cocktail of the best suitable
combination of these compounds (Fig 5.1). I believe that this approach will speed up the
Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular mechanism…
http://dx.doi.org/10.5772/53800
569
process of finding the best possible and safest treatment of SMA. This approach is currently
being assayed in our laboratory and others, showing some positive and promising results in
the severe mouse model of the disease. More work is required to assess the potential drug
interactions and their side effects in the animal models of the disease before pushing this
approach for human clinical trials.
Designing clinical trials for SMA: In the last 5 years, a tremendous amount of promising
translational work has been done using animal models of the SMA which is progressing
rapidly towards the pre-clinical stage. However there are major challenges for designing a
perfect clinical trial for SMA which includes 1) Variability of the disease phenotype, 2) lack of
molecular biomarkers, 3) Accessibility of treatment centers and 4) lack of agreement for
standard of care and disease management. However these issues are likely to be resolved as
recently there has been a remarkable cooperation and collaboration between researchers,
clinicians, industry, government and volunteer organizations which is bringing everyone on
the same page to address these issues and reach a consensus for designing standard human
clinical trials for SMA internationally.
Early intervention: New born screening: We and others have seen, irrespective of the
modality, that early timing of the treatment is critical for maximum benefit in the mouse model
of the disease. Presymptomatic identification of infants with SMA through new-born screening
represents an important step in the effective treatment of SMA. In essence we shall need to
intervene before the damage is done; to do so we need to rapidly identify infants with SMA,
cases who will also serve as the best candidates to show the efficacy of promising therapeutic
Figure 4. Development of therapeutics for SMA
Neurodegenerative Diseases
570
treatments in the near future. Children in which the disease has already progressed may also
benefit with the use of best combinational approach, however the aim will be more towards
ameliorating the disease progress and preserving the function of remaining motor neurons
and other tissues rather than a complete reversal of the disease phenotype.
p38 activator
p38
SMN mRNA
SMN Protein
HuR
SMN mRNA
P
HuR
mRNA stabilization
Nucleus
STAT5
PRLR
PRL
SMN mRNA
P
P
SMN 1/2 gene
Transcriptional
upregulation
SMA disease
progression
Protein stabilization
Proteasome
Inhibitor
(Bortezomid)
Rho kinase
Inhibitor
Neuro protective
(SMN independent)
Figure 5. Proposed model of combination therapy for SMA treatment.
Author details
Faraz Tariq Farooq1,2*, Martin Holcik1,2 and Alex MacKenzie1,2
*Address all correspondence to: faraztfarooq@gmail.com
1 University of Ottawa, Ottawa, Canada
2 Apoptosis Research Center, CHEO Research Institute, CHEO, Ottawa, Canada
No conflicts of interest are reported.
Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular mechanism…
http://dx.doi.org/10.5772/53800
571
References
[1] Pearn, J. (1978). Incidence, prevalence, and gene frequency studies of chronic child‐
hood spinal muscular atrophy. J. Med. Genet. , 15, 409-413.
[2] Feldkotter, M, Schwarzer, V, Wirth, R, Wienker, T. F, & Wirth, B. (2002). Quantitative
analyses of SMN1 and SMN2 based on real-time lightCycler PCR: fast and highly re‐
liable carrier testing and prediction of severity of spinal muscular atrophy. Am. J.
Hum. Genet. , 70, 358-368.
[3] Ogino, S, Leonard, D. G, Rennert, H, Ewens, W. J, & Wilson, R. B. (2002). Genetic risk
assessment in carrier testing for spinal muscular atrophy. Am. J. Med. Genet. , 110,
301-307.
[4] Ogino, S, Wilson, R. B, & Gold, B. (2004). New insights on the evolution of the SMN1
and SMN2 region: simulation and meta-analysis for allele and haplotype frequency
calculations. Eur. J. Hum. Genet. , 12, 1015-1023.
[5] Sugarman, E. A., Nagan, N., Zhu, H., Akmaev, V. R., Zhou, Z., Rohlfs, E. M., Flynn,
K., Hendrickson, B. C., Scholl, T., Sirko-Osadsa, D. A., Allitto, B.A. 2012. Pan-ethnic
carrier screening and prenatal diagnosis for spinal muscular atrophy: clinical labora‐
tory analysis of >72,400 specimens. Eur. J. Hum. Genet. 20(1):27-32.
[6] Zerres, K, & Davies, K. E. (1999). th ENMC International Workshop: Spinal Muscular
Atrophies: recent progress and revised diagnostic criteria 17-19 April 1998, Soestdui‐
nen, The Netherlands. Neuromuscul. Disord. , 9, 272-278.
[7] Dubowitz, V. (1995). Chaos in the classification of SMA: a possible resolution. Neuro‐
muscul. Disord. , 5, 3-5.
[8] Iannaccone, S. T, Russman, B. S, Browne, R. H, Buncher, C. R, White, M, & Samaha,
F. J. (2000). Prospective analysis of strength in spinal muscular atrophy. DCN/Spinal
Muscular Atrophy Group. J. Child Neurol. , 15, 97-101.
[9] Wang, C. H, Finkel, R. S, Bertini, E. S, Schroth, M, Simonds, A, Wong, B, Aloysius, A,
Morrison, L, Main, M, Crawford, T. O, et al. (2007). Consensus statement for stand‐
ard of care in spinal muscular atrophy. J. Child Neurol. , 22, 1027-1049.
[10] Hagen, O, Glanzman, J. M, Mcdermott, A. M, Ryan, M. P, Flickinger, P. A, Quigley, J,
Riley, J, Sanborn, S, Irvine, E, Martens, C, et al. (2007). An expanded version of the
Hammersmith Functional Motor Scale for SMA II and III patients. Neuromuscul. Dis‐
ord. , 17, 693-697.
[11] Thomas, N. H, & Dubowitz, V. (1994). The natural history of type I (severe) spinal
muscular atrophy. Neuromuscul. Disord. , 4, 497-502.
[12] Zerres, K, Rudnik-schoneborn, S, Forrest, E, Lusakowska, A, Borkowska, J, & Haus‐
manowa-petrusewicz, I. (1997). A collaborative study on the natural history of child‐
Neurodegenerative Diseases
572
hood and juvenile onset proximal spinal muscular atrophy (type II and III SMA): 569
patients. J. Neurol. Sci. , 146, 67-72.
[13] Zerres, K, Wirth, B, & Rudnik-schoneborn, S. (1997). Spinal muscular atrophy--clini‐
cal and genetic correlations. Neuromuscul. Disord. , 7, 202-207.
[14] Prior, T. W, & Russman, B. S. (1993). Spinal Muscular Atrophy. In GeneReviews. R.A.
Pagon, T.D. Bird, C.R. Dolan, and K. Stephens, editors. University of Washington, Se‐
attle. Seattle (WA).
[15] Lefebvre, S, Burglen, L, Reboullet, S, Clermont, O, Burlet, P, Viollet, L, Benichou, B,
Cruaud, C, Millasseau, P, & Zeviani, M. (1995). Identification and characterization of
a spinal muscular atrophy-determining gene. Cell , 80, 155-165.
[16] Mahadevan, M. S, Korneluk, R. G, & Roy, N. MacKenzie, A., and Ikeda, J. (1995).
SMA genes: deleted and duplicated. Nat. Genet. , 9, 112-113.
[17] Lorson, C. L, Hahnen, E, Androphy, E. J, & Wirth, B. (1999). A single nucleotide in
the SMN gene regulates splicing and is responsible for spinal muscular atrophy.
Proc. Natl. Acad. Sci. U. S. A. , 96, 6307-6311.
[18] Monani, U. R, Lorson, C. L, Parsons, D. W, Prior, T. W, Androphy, E. J, Burghes, A.
H, & Mcpherson, J. D. (1999). A single nucleotide difference that alters splicing pat‐
terns distinguishes the SMA gene SMN1 from the copy gene SMN2. Hum. Mol. Gen‐
et. , 8, 1177-1183.
[19] Schrank, B, Gotz, R, Gunnersen, J. M, Ure, J. M, Toyka, K. V, Smith, A. G, & Sendtner,
M. (1997). Inactivation of the survival motor neuron gene, a candidate gene for hu‐
man spinal muscular atrophy, leads to massive cell death in early mouse embryos.
Proc. Natl. Acad. Sci. U. S. A. , 94, 9920-9925.
[20] Simic, G. (2008). Pathogenesis of proximal autosomal recessive spinal muscular atro‐
phy. Acta Neuropathol. , 116, 223-234.
[21] Murayama, S, Bouldin, T. W, & Suzuki, K. (1991). Immunocytochemical and ultra‐
structural studies of Werdnig-Hoffmann disease. Acta Neuropathol. , 81, 408-417.
[22] Pinter, M. J, Waldeck, R. F, Wallace, N, & Cork, L. C. (1995). Motor unit behavior in
canine motor neuron disease. J. Neurosci. , 15, 3447-3457.
[23] Cork, L. C, Struble, R. G, & Gold, B. G. DiCarlo, C., Fahnestock, K.E., Griffin, J.W.,
and Price, D.L. (1989). Changes in size of motor axons in hereditary canine spinal
muscular atrophy. Lab. Invest. , 61, 333-342.
[24] Kariya, S, Park, G. H, Maeno-hikichi, Y, Leykekhman, O, Lutz, C, Arkovitz, M. S,
Landmesser, L. T, & Monani, U. R. (2008). Reduced SMN protein impairs maturation
of the neuromuscular junctions in mouse models of spinal muscular atrophy. Hum.
Mol. Genet. , 17, 2552-2569.
[25] Menke, L. A, Poll-the, B. T, Clur, S. A, Bilardo, C. M, Van Der Wal, A. C, Lemmink,
H. H, & Cobben, J. M. (2008). Congenital heart defects in spinal muscular atrophy
Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular mechanism…
http://dx.doi.org/10.5772/53800
573
type I: a clinical report of two siblings and a review of the literature. Am. J. Med. Gen‐
et. A. 146A:, 740-744.
[26] Rudnik-schoneborn, S, Heller, R, Berg, C, Betzler, C, Grimm, T, Eggermann, T, Eg‐
germann, K, Wirth, R, Wirth, B, & Zerres, K. (2008). Congenital heart disease is a fea‐
ture of severe infantile spinal muscular atrophy. J. Med. Genet. , 45, 635-638.
[27] Bevan, A. K, Hutchinson, K. R, Foust, K. D, Braun, L, Mcgovern, V. L, Schmelzer, L,
Ward, J. G, Petruska, J. C, Lucchesi, P. A, Burghes, A. H, et al. (2010). Early heart fail‐
ure in the SMN∆7 model of spinal muscular atrophy and correction by postnatal
scAAV9-SMN delivery. Hum. Mol. Genet. , 19, 3895-3905.
[28] Heier, C. R, Satta, R, & Lutz, C. and DiDonato, C.J. (2010). Arrhythmia and cardiac
defects are a feature of spinal muscular atrophy model mice. Hum. Mol. Genet. , 19,
3906-3918.
[29] Shababi, M, Habibi, J, Yang, H. T, Vale, S. M, Sewell, W. A, & Lorson, C. L. (2010).
Cardiac defects contribute to the pathology of spinal muscular atrophy models. Hum.
Mol. Genet. , 19, 4059-4071.
[30] Williams, J. H, Schray, R. C, Patterson, C. A, Ayitey, S. O, Tallent, M. K, & Lutz, G. J.
(2009). Oligonucleotide-mediated survival of motor neuron protein expression in
CNS improves phenotype in a mouse model of spinal muscular atrophy. J. Neurosci. ,
29, 7633-7638.
[31] Rudnik-schoneborn, S, Vogelgesang, S, Armbrust, S, Graul-neumann, L, Fusch, C, &
Zerres, K. (2010). Digital necroses and vascular thrombosis in severe spinal muscular
atrophy. Muscle Nerve , 42, 144-147.
[32] Araujo AdeQ., Araujo, M., and Swoboda, K.J. (2009). Vascular perfusion abnormali‐
ties in infants with spinal muscular atrophy. J. Pediatr. , 155, 292-294.
[33] Hua, Y, Sahashi, K, Rigo, F, Hung, G, Horev, G, Bennett, C. F, & Krainer, A. R. (2011).
Peripheral SMN restoration is essential for long-term rescue of a severe spinal mus‐
cular atrophy mouse model. Nature , 478, 123-126.
[34] Liu, Q, & Dreyfuss, G. (1996). A novel nuclear structure containing the survival of
motor neurons protein. EMBO J. , 15, 3555-3565.
[35] Zhang, H, Xing, L, Rossoll, W, Wichterle, H, Singer, R. H, & Bassell, G. J. (2006). Mul‐
tiprotein complexes of the survival of motor neuron protein SMN with Gemins traf‐
fic to neuronal processes and growth cones of motor neurons. J. Neurosci. , 26,
8622-8632.
[36] Friesen, W. J, & Dreyfuss, G. (2000). Specific sequences of the Sm and Sm-like (Lsm)
proteins mediate their interaction with the spinal muscular atrophy disease gene
product (SMN). J. Biol. Chem. , 275, 26370-26375.
Neurodegenerative Diseases
574
[37] Buhler, D, Raker, V, Luhrmann, R, & Fischer, U. (1999). Essential role for the tudor
domain of SMN in spliceosomal U snRNP assembly: implications for spinal muscu‐
lar atrophy. Hum. Mol. Genet. , 8, 2351-2357.
[38] Jones, K. W, Gorzynski, K, Hales, C. M, Fischer, U, Badbanchi, F, Terns, R. M, &
Terns, M. P. (2001). Direct interaction of the spinal muscular atrophy disease protein
SMN with the small nucleolar RNA-associated protein fibrillarin. J. Biol. Chem. , 276,
38645-38651.
[39] Meister, G, Buhler, D, Laggerbauer, B, Zobawa, M, Lottspeich, F, & Fischer, U. (2000).
Characterization of a nuclear 20S complex containing the survival of motor neurons
(SMN) protein and a specific subset of spliceosomal Sm proteins. Hum. Mol. Genet. ,
9, 1977-1986.
[40] Mourelatos, Z, Abel, L, Yong, J, Kataoka, N, & Dreyfuss, G. (2001). SMN interacts
with a novel family of hnRNP and spliceosomal proteins. EMBO J. , 20, 5443-5452.
[41] Narayanan, U, Achsel, T, Luhrmann, R, & Matera, A. G. (2004). Coupled in vitro im‐
port of U snRNPs and SMN, the spinal muscular atrophy protein. Mol. Cell , 16,
223-234.
[42] Pellizzoni, L, Baccon, J, Charroux, B, & Dreyfuss, G. (2001). The survival of motor
neurons (SMN) protein interacts with the snoRNP proteins fibrillarin and GAR1.
Curr. Biol. , 11, 1079-1088.
[43] Pellizzoni, L, Charroux, B, & Dreyfuss, G. (1999). SMN mutants of spinal muscular
atrophy patients are defective in binding to snRNP proteins. Proc. Natl. Acad. Sci. U.
S. A. , 96, 11167-11172.
[44] Pellizzoni, L, Charroux, B, Rappsilber, J, Mann, M, & Dreyfuss, G. (2001). A function‐
al interaction between the survival motor neuron complex and RNA polymerase II. J.
Cell Biol. , 152, 75-85.
[45] Fischer, U, Liu, Q, & Dreyfuss, G. (1997). The SMN-SIP1 complex has an essential
role in spliceosomal snRNP biogenesis. Cell , 90, 1023-1029.
[46] Kariya, S, Mauricio, R, Dai, Y, & Monani, U. R. (2009). The neuroprotective factor
Wld(s) fails to mitigate distal axonal and neuromuscular junction (NMJ) defects in
mouse models of spinal muscular atrophy. Neurosci. Lett. , 449, 246-251.
[47] Hofmann, Y, Lorson, C. L, Stamm, S, Androphy, E. J, & Wirth, B. (2000). Htra2-beta 1
stimulates an exonic splicing enhancer and can restore full-length SMN expression to
survival motor neuron 2 (SMN2). Proc. Natl. Acad. Sci. U. S. A. , 97, 9618-9623.
[48] Chen, H. H, Chang, J. G, Lu, R. M, Peng, T. Y, & Tarn, W. Y. (2008). The RNA binding
protein hnRNP Q modulates the utilization of exon 7 in the survival motor neuron 2
(SMN2) gene. Mol. Cell. Biol. , 28, 6929-6938.
Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular mechanism…
http://dx.doi.org/10.5772/53800
575
[49] Hofmann, Y, & Wirth, B. (2002). hnRNP-G promotes exon 7 inclusion of survival mo‐
tor neuron (SMN) via direct interaction with Htra2-beta1. Hum. Mol. Genet. , 11,
2037-2049.
[50] Young, P. J. DiDonato, C.J., Hu, D., Kothary, R., Androphy, E.J., and Lorson, C.L.
(2002). SRp30c-dependent stimulation of survival motor neuron (SMN) exon 7 inclu‐
sion is facilitated by a direct interaction with hTra2 beta 1. Hum. Mol. Genet. , 11,
577-587.
[51] Cartegni, L, & Krainer, A. R. (2002). Disruption of an SF2/ASF-dependent exonic
splicing enhancer in SMN2 causes spinal muscular atrophy in the absence of SMN1.
Nat. Genet. , 30, 377-384.
[52] Cartegni, L, & Krainer, A. R. (2003). Correction of disease-associated exon skipping
by synthetic exon-specific activators. Nat. Struct. Biol. , 10, 120-125.
[53] Kashima, T, & Manley, J. L. (2003). A negative element in SMN2 exon 7 inhibits splic‐
ing in spinal muscular atrophy. Nat. Genet. , 34, 460-463.
[54] Mcandrew, P. E, Parsons, D. W, Simard, L. R, Rochette, C, Ray, P. N, Mendell, J. R,
Prior, T. W, & Burghes, A. H. (1997). Identification of proximal spinal muscular atro‐
phy carriers and patients by analysis of SMNT and SMNC gene copy number. Am. J.
Hum. Genet. , 60, 1411-1422.
[55] Campbell, L, Potter, A, Ignatius, J, Dubowitz, V, & Davies, K. (1997). Genomic varia‐
tion and gene conversion in spinal muscular atrophy: implications for disease proc‐
ess and clinical phenotype. Am. J. Hum. Genet. , 61, 40-50.
[56] Chang, J. G, Hsieh-li, H. M, Jong, Y. J, Wang, N. M, Tsai, C. H, & Li, H. (2001). Treat‐
ment of spinal muscular atrophy by sodium butyrate. Proc. Natl. Acad. Sci. U. S. A. ,
98, 9808-9813.
[57] Sumner, C. J, Huynh, T. N, Markowitz, J. A, Perhac, J. S, Hill, B, Coovert, D. D,
Schussler, K, Chen, X, Jarecki, J, Burghes, A. H, et al. (2003). Valproic acid increases
SMN levels in spinal muscular atrophy patient cells. Ann. Neurol. , 54, 647-654.
[58] Brichta, L, Hofmann, Y, Hahnen, E, Siebzehnrubl, F. A, Raschke, H, Blumcke, I, Eyu‐
poglu, I. Y, & Wirth, B. (2003). Valproic acid increases the SMN2 protein level: a well-
known drug as a potential therapy for spinal muscular atrophy. Hum. Mol. Genet. ,
12, 2481-2489.
[59] Tsai, L. K, Tsai, M. S, Ting, C. H, & Li, H. (2008). Multiple therapeutic effects of val‐
proic acid in spinal muscular atrophy model mice. J. Mol. Med. (Berl) , 86, 1243-1254.
[60] Andreassi, C, Angelozzi, C, Tiziano, F. D, Vitali, T, De Vincenzi, E, Boninsegna, A,
Villanova, M, Bertini, E, Pini, A, Neri, G, et al. (2004). Phenylbutyrate increases SMN
expression in vitro: relevance for treatment of spinal muscular atrophy. Eur. J. Hum.
Genet. , 12, 59-65.
Neurodegenerative Diseases
576
[61] Swoboda, K. J, Scott, C. B, Reyna, S. P, & Prior, T. W. LaSalle, B., Sorenson, S.L.,
Wood, J., Acsadi, G., Crawford, T.O., Kissel, J.T. et al. (2009). Phase II open label
study of valproic acid in spinal muscular atrophy. PLoS One 4:e5268.
[62] Garbes, L, Riessland, M, Holker, I, Heller, R, Hauke, J, Trankle, C, Coras, R, Blumcke,
I, Hahnen, E, & Wirth, B. (2009). LBH589 induces up to 10-fold SMN protein levels by
several independent mechanisms and is effective even in cells from SMA patients
non-responsive to valproate. Hum. Mol. Genet. , 18, 3645-3658.
[63] Mercuri, E, Bertini, E, Messina, S, Solari, A, Amico, D, Angelozzi, A, Battini, C, Berar‐
dinelli, R, Boffi, A, Bruno, P, et al. (2007). Randomized, double-blind, placebo-con‐
trolled trial of phenylbutyrate in spinal muscular atrophy. Neurology , 68, 51-55.
[64] Avila, A. M, Burnett, B. G, Taye, A. A, Gabanella, F, Knight, M. A, Hartenstein, P, &
Cizman, Z. Di Prospero, N.A., Pellizzoni, L., Fischbeck, K.H. et al. (2007). Trichosta‐
tin A increases SMN expression and survival in a mouse model of spinal muscular
atrophy. J. Clin. Invest. , 117, 659-671.
[65] Narver, H. L, Kong, L, Burnett, B. G, Choe, D. W, Bosch-marce, M, Taye, A. A, Eck‐
haus, M. A, & Sumner, C. J. (2008). Sustained improvement of spinal muscular atro‐
phy mice treated with trichostatin A plus nutrition. Ann. Neurol. , 64, 465-470.
[66] Riessland, M, Ackermann, B, Forster, A, Jakubik, M, Hauke, J, Garbes, L, Fritzsche, I,
Mende, Y, Blumcke, I, Hahnen, E, et al. (2010). SAHA ameliorates the SMA pheno‐
type in two mouse models for spinal muscular atrophy. Hum. Mol. Genet. , 19,
1492-1506.
[67] Powe, C. E, Allen, M, Puopolo, K. M, Merewood, A, Worden, S, Johnson, L. C,
Fleischman, A, & Welt, C. K. (2010). Recombinant human prolactin for the treatment
of lactation insufficiency. Clin. Endocrinol. (Oxf) , 73, 645-653.
[68] Farooq, F, Molina, F. A, & Hadwen, J. MacKenzie, D., Witherspoon, L., Osmond, M.,
Holcik, M., and MacKenzie, A. (2011). Prolactin increases SMN expression and sur‐
vival in a mouse model of severe spinal muscular atrophy via the STAT5 pathway. J.
Clin. Invest. , 121, 3042-3050.
[69] Andreassi, C, Jarecki, J, Zhou, J, Coovert, D. D, Monani, U. R, Chen, X, Whitney, M,
Pollok, B, Zhang, M, Androphy, E, et al. (2001). Aclarubicin treatment restores SMN
levels to cells derived from type I spinal muscular atrophy patients. Hum. Mol. Gen‐
et. , 10, 2841-2849.
[70] Singh, N. N, Singh, R. N, & Androphy, E. J. (2007). Modulating role of RNA structure
in alternative splicing of a critical exon in the spinal muscular atrophy genes. Nucleic
Acids Res. , 35, 371-389.
[71] Singh, N. N, Shishimorova, M, Cao, L. C, Gangwani, L, & Singh, R. N. (2009). A short
antisense oligonucleotide masking a unique intronic motif prevents skipping of a
critical exon in spinal muscular atrophy. RNA Biol. , 6, 341-350.
Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular mechanism…
http://dx.doi.org/10.5772/53800
577
[72] MacKenzieA. (2012). Sense in antisense therapy for spinal muscular atrophy. N. Engl.
J. Med. , 366, 761-763.
[73] Singh, J, Salcius, M, Liu, S. W, Staker, B. L, Mishra, R, Thurmond, J, Michaud, G, Mat‐
toon, D. R, Printen, J, Christensen, J, et al. (2008). DcpS as a therapeutic target for spi‐
nal muscular atrophy. ACS Chem. Biol. , 3, 711-722.
[74] Farooq, F, Balabanian, S, Liu, X, & Holcik, M. and MacKenzie, A. (2009). Mitogen-ac‐
tivated protein kinase stabilizes SMN mRNA through RNA binding protein HuR.
Hum. Mol. Genet. 18:4035-4045., 38.
[75] Heier, C. R. and DiDonato, C.J. (2009). Translational readthrough by the aminoglyco‐
side geneticin (G418) modulates SMN stability in vitro and improves motor function
in SMA mice in vivo. Hum. Mol. Genet. , 18, 1310-1322.
[76] Mattis, V. B, Rai, R, Wang, J, Chang, C. W, Coady, T, & Lorson, C. L. (2006). Novel
aminoglycosides increase SMN levels in spinal muscular atrophy fibroblasts. Hum.
Genet. , 120, 589-601.
[77] Wolstencroft, E. C, Mattis, V, Bajer, A. A, Young, P. J, & Lorson, C. L. specific re‐
quirement for SMN protein activity: the role of aminoglycosides in inducing elevated
SMN protein levels. Hum. Mol. Genet. , 14, 1199-1210.
[78] Kwon, D. Y, Motley, W. W, Fischbeck, K. H, & Burnett, B. G. (2011). Increasing ex‐
pression and decreasing degradation of SMN ameliorate the spinal muscular atrophy
phenotype in mice. Hum. Mol. Genet. , 20, 3667-3677.
[79] Foust, K. D, Wang, X, Mcgovern, V. L, Braun, L, Bevan, A. K, Haidet, A. M, Le, T. T,
Morales, P. R, Rich, M. M, Burghes, A. H, et al. (2010). Rescue of the spinal muscular
atrophy phenotype in a mouse model by early postnatal delivery of SMN. Nat. Bio‐
technol. , 28, 271-274.
[80] Passini, M. A, Bu, J, Roskelley, E. M, Richards, A. M, Sardi, S. P, Riordan, O, Klinger,
C. R, Shihabuddin, K. W, Cheng, L. S, & Cns-targeted, S. H. gene therapy improves
survival and motor function in a mouse model of spinal muscular atrophy. J. Clin.
Invest. , 120, 1253-1264.
[81] Valori, C. F, Ning, K, Wyles, M, Mead, R. J, Grierson, A. J, Shaw, P. J, & Azzouz, M.
(2010). Systemic delivery of scAAV9 expressing SMN prolongs survival in a model of
spinal muscular atrophy. Sci. Transl. Med. 2:35ra42.
[82] Dominguez, E, Marais, T, Chatauret, N, Benkhelifa-ziyyat, S, Duque, S, Ravassard, P,
Carcenac, R, Astord, S, De Moura, A. P, Voit, T, et al. (2011). Intravenous scAAV9 de‐
livery of a codon-optimized SMN1 sequence rescues SMA mice. Hum. Mol. Genet. ,
20, 681-693.
[83] MacKenzieA. (2010). Genetic therapy for spinal muscular atrophy. Nat. Biotechnol. ,
28, 235-237.
Neurodegenerative Diseases
578
[84] Corti, S, Nizzardo, M, Nardini, M, Donadoni, C, & Salani, S. Del Bo, R., Papadimi‐
triou, D., Locatelli, F., Mezzina, N., Gianni, F. et al. (2009). Motoneuron transplanta‐
tion rescues the phenotype of SMARD1 (spinal muscular atrophy with respiratory
distress type 1). J. Neurosci. , 29, 11761-11771.
[85] Corti, S, Nizzardo, M, Nardini, M, Donadoni, C, Salani, S, Ronchi, D, Simone, C, Fal‐
cone, M, Papadimitriou, D, Locatelli, F, et al. (2010). Embryonic stem cell-derived
neural stem cells improve spinal muscular atrophy phenotype in mice. Brain , 133,
465-481.
[86] Dimos, J. T, Rodolfa, K. T, Niakan, K. K, Weisenthal, L. M, Mitsumoto, H, Chung, W,
Croft, G. F, Saphier, G, Leibel, R, Goland, R, et al. (2008). Induced pluripotent stem
cells generated from patients with ALS can be differentiated into motor neurons. Sci‐
ence , 321, 1218-1221.
[87] Ebert, A. D, Yu, J, & Rose, F. F. Jr, Mattis, V.B., Lorson, C.L., Thomson, J.A., and
Svendsen, C.N. (2009). Induced pluripotent stem cells from a spinal muscular atro‐
phy patient. Nature , 457, 277-280.
[88] Bowerman, M, Beauvais, A, Anderson, C. L, & Kothary, R. (2010). Rho-kinase inacti‐
vation prolongs survival of an intermediate SMA mouse model. Hum. Mol. Genet. ,
19, 1468-1478.
Spinal Muscular Atrophy: Classification, Diagnosis, Background, Molecular mechanism…
http://dx.doi.org/10.5772/53800
579
... The SMN1 is highly conserved and presents as a single copy in the genome of all eukaryotic organisms [35,36]. A normal individual has two forms of the SMN gene, which are telomeric SMN1 and its paralog, centromeric SMN2 [11,37] (Figure 2). Both genes are nearly identical, with only a difference in five base pairs. ...
... The SMN1 gene produces full-length, functional SMN (FL-SMN) protein. A synonymous C-to-T base substitution (c.840C > T) at the position 6 of SMN2 exon 7 disrupts the proper splicing and leads to a majority (~90%) of exon 7-skipped transcript (∆7-transcript) [37][38][39][40]. Subsequent translation of such transcript results in a truncated and unstable SMN protein [17,41]. ...
... Most of the patients, typically present with two or three copies of the SMN2 gene [56], have generalized muscle weakness, including an inability of controlling the movements of the head and inability to sit unaided, among others. Due to weakness of the respiratory muscles, they have breathing distress and increased the risk of aspiration [37,[57][58][59]. Babies have difficulty swallowing and sucking, leading to difficulty with feeding and a failure to thrive [2]. ...
Article
Full-text available
Spinal muscular atrophy (SMA), one of the leading inherited causes of child mortality, is a rare neuromuscular disease arising from loss-of-function mutations of the survival motor neuron 1 (SMN1) gene, which encodes the SMN protein. When lacking the SMN protein in neurons, patients suffer from muscle weakness and atrophy, and in the severe cases, respiratory failure and death. Several therapeutic approaches show promise with human testing and three medications have been approved by the U.S. Food and Drug Administration (FDA) to date. Despite the shown promise of these approved therapies, there are some crucial limitations, one of the most important being the cost. The FDA-approved drugs are high-priced and are shortlisted among the most expensive treatments in the world. The price is still far beyond affordable and may serve as a burden for patients. The blooming of the biomedical data and advancement of computational approaches have opened new possibilities for SMA therapeutic development. This article highlights the present status of computationally aided approaches, including in silico drug repurposing, network driven drug discovery as well as artificial intelligence (AI)-assisted drug discovery, and discusses the future prospects.
... They generally have an almost normal life span. Type 4 (late SMA, OMIM #271150): It is a very rare type that usually starts in young adulthood resulting in mild motor impairment [6]. Some classifications tend to categorize patients with significant muscle weakness and respiratory distress at birth into a distinct type designated as SMA type 0/prenatal SMA. ...
... In more than 95% of cases, the disease results from the loss of SMN1 gene. However, intragenic gene mutations, including missense, nonsense, frameshift and splice-site variations, account for the remaining 5% of cases [6,8]. The SMN1 is located within a genomic segment of inverted duplication. ...
Article
Full-text available
Background Spinal muscular atrophy (SMA) is an autosomal recessive neuromuscular disease representing the most prevalent monogenic cause of infant mortality. It results from the loss of SMN1 gene, but retention of its paralog SMN2 whose copy number can modulate the disease severity and guide the therapeutic regimen. Methods For SMA molecular analysis, 236 unrelated Egyptian patients were enrolled at our institution. The Multiplex ligation-dependent probe amplification analysis (MLPA) was applied to investigate the main genetic defect in the enrolled patients ( SMN1 loss) and to determine a possible genotype–phenotype correlation between the copy number of other genes in the SMN locus (5q13.2) and disease severity in Egyptian patients with SMA. A small cohort of healthy subjects ( n = 57) was also included to investigate the possible differences in the distributions of SMN2 and NAIP genes between patients and healthy individuals. Results Disease diagnosis was confirmed in only 148 patients (62.7%) highlighting the clinical overlapping of the disease and emphasizing the importance of molecular diagnosis. In patients with homozygous SMN1 loss, the disease was mediated by gene deletion and conversion in 135 (91.2%) and 13 (8.8%) patients, respectively. In the study cohort, SMN2 and NAIP copy numbers were inversely correlated with disease severity. However, no significant association was detected between GTF2H2A and SERF1B copy numbers and patient phenotype. Significant differences were demonstrated in the copy numbers of SMN2 and NAIP between SMA patients and healthy subjects. Conclusion Molecular analysis of SMA is essential for disease diagnosis. Consistent with previous studies on other populations, there is a close relationship between SMN2 and NAIP copy numbers and clinical phenotype. Additionally, potential differences in these two genes distributions are existing between patients and healthy subjects. National program for carrier screening should be established as a preventive disease strategy. On the other hand, neonatal testing would provide accurate estimation for disease incidence.
... (Farooq & Machenzie, 2013) Para possibilitar o funcionamento adequado de 100% do gene da proteína SMN, é necessário um gene que forneça informações, e este é o gene SMN1, que em 1995 foi realizada a identificação da deleção homozigótica (desequilíbrio do cromossomo, por perda de um segmento cromossômico) do mesmo, o que levou os pesquisadores a conclusão que 95% dos pacientes com AME é ocasionado por esse erro cromossômico. (Tizzano, 2007;Chrun et al., 2017;Farooq & Mackenzie, 2013) A proteína SMN pode ser também produzida por um gene quase idêntico ao SMN1, o SMN2. A diferença entre ambos ocorre no nucleotídeo (responsável pela formação dos ácidos nucleicos do DNA e RNA) no exon 7 e 8, isso porque no gene SMN1 o transcrito C é gerado de forma completa, enquanto no gene SMN2 esse transcrito C se transforma em T na posição 6 do exon 7, o que vai resultar na exclusão do exon 7. Dando origem a proteína SMN de forma diferente e incompleta de acordo com sua função fisiológica. ...
... Sendo que, pessoas que apresentem 5 ou mais cópias do gene SMN2, apesar da ausência do gene SMN1 são considerados assintomáticos e protegidos contra a manifestação da doença. (Tizzano, 2007;Farooq & Mackenzie, 2013;Zanoteli et al., 2020;Baione & Ambiel, 2010) ...
Article
Full-text available
A Amiotrofia Muscular Espinhal (AME) é uma patologia de origem genética de herança autossômica recessiva, neuromuscular degenerativa. Acomete principalmente os neurônios motores do corno anterior da medula espinhal. De acordo com a idade de inicío das manifestações clinicas, e pelo comprometimento motor, a AME é classificada em quatro tipos (I a IV). O referido estudo de revisão literária tem como objetivo expor a etiologia, fisiopatologia, diagnóstico, tratamento e a importância da intervenção fisioterapêutica em pacientes portadores da Amiotrofia Muscular Espinhal, em especifico na AME tipo I. Na presente pesquisa foram considerados para criterios de inclusão artigos que envolvesse a tematica no que se refere á intervenção fisioterapêutica na Amiotrofia Muscular Espinhal tipo 1. A publicação dos dados coletados ocorreu entre o período de 2007 a 2021, em lingua portuguesa, inglesa e espanhola. O levantamento bibibliográfico foi realizado nas bases de dados: Scielo, Pubmed, PEDro, Lilacs e Google acadêmico, no período de março a outubro. No decorrer da pesquisa foi observado á intervenção fisioterapêutica na Amiotrofia Muscular Espinhal em especifico na AME tipo 1. Concluindo que o modo de atuação da fisioterapia preconiza minimizar/retardar as manifestações clínicas que afetam o sistema musculoesquelético e as complicações no sistema respiratório, contribuindo para promover qualidade e prolongar a vida, e consequentemente evitando o óbito desses pacientes.
... The advantages include its affordable operational costs as well as its ability to predict the results of in vitro analysis and reveal molecular mechanisms ( [19][20][21][22]). Information regarding these molecular mechanisms can be used to develop diagnostic and therapeutic methods [23][24][25]. In this study, in silico examination was carried out using molecular docking, MD simulation, and analysis of α-glucosidase enzymes and their ligands. ...
Article
Full-text available
Diabetes is a chronic disease with a high mortality rate worldwide and can cause other diseases such as kidney damage, narrowing of blood vessels, and heart disease. The concomitant use of drugs such as metformin, sulfonylurea, miglitol, and acarbose may cause side effects with long-term administration. Therefore, natural ingredients are the best choice, considering that their long-term side effects are not significant. One of the compounds that can be used as a candidate antidiabetic is mangostin; however, information on the molecular mechanism needs to be further analyzed through molecular docking, simulating molecular dynamics, and testing the in silico antidiabetic potential. This study focused on modeling the protein structure, molecular docking, and molecular dynamics simulations and analyses. This process produces RMSD values, free energies, and intermolecular hydrogen bonding. Based on the analysis results, all molecular dynamics simulations can occur under physiological conditions, and γ-mangostin is the best among the test compounds.
... All of the patients in the presented cases were diagnosed with SMA type III because their onsets were between 18 months and 30 years old. Patients with SMA type 3 usually develop more problems with back pain and can include disorders of the backbone such as scoliosis [10][11][12]. Before we could reach the diagnosis of SMA type III, some considerations of other neuromuscular disorders were also taken into account. ...
Article
Full-text available
Background Spinal muscular atrophy is a recessively inherited autosomal neuromuscular disorder, with characteristic progressive muscle weakness. Most spinal muscular atrophy cases clinically manifest during infancy or childhood, although it may first manifest in adulthood. Although spinal muscular atrophy has come to the era of newborn screening and promising treatments, genetically confirmed spinal muscular atrophy patients are still rare in third world countries, including Indonesia. Case presentations We presented three Indonesian patients with spinal muscular atrophy genetically confirmed during adulthood. The first case was a 40-year-old male who presented with weakness in his lower limbs that started when he was 9 years old. At the age of 16 years, he could no longer walk and started using a wheelchair. He first came to our clinic at the age of 38 years, and was diagnosed with spinal muscular atrophy 2 years later. The second patient was a 58-year-old male who presented with lower limb weakness since he was 12 years old. Owing to the geographical distance and financial problems, he was referred to our clinic at the age of 56 years, when he already used a walker to walk. Lastly, the third patient was a 28-year-old woman, who was in the first semester of her second pregnancy, and who presented with slowly progressing lower limb weakness. Her limb weakness began at the age of 8 years, and slowly progressed until she became dependent on her wheelchair 8 years later until now. She had successfully given birth to a healthy daughter 3 years before her first visit to our clinic. All three patients were diagnosed with neuromuscular disorder diseases, with the differential diagnoses of Duchenne muscular dystrophy, spinal muscular atrophy, and Becker muscular dystrophy. These patients were finally confirmed to have spinal muscular atrophy due to SMN1 deletion by polymerase chain reaction and restriction fragment length polymorphism. Conclusions Many genetic diseases are often neglected in developing countries owing to the difficulty in diagnosis and unavailable treatment. Our case series focused on the disease courses, diagnosis difficulties, and clinical presentations of three patients that finally lead to diagnoses of spinal muscular atrophy.
... [1,2]. Classification of SMA is based on the age of onset and maximum motor function milestone achieved: SMA type I (Werdnig-Hoffmann disease) has onset in the first month of life, with a disability to sit unsupported and needing breathing support, with a survival chance of only up to 2-years old; SMA type II has onset between 6 and 18-months old, with the best motor ability to sit alone without support; SMA type III (Kugelberg-Welander disease) has onset between 18 months and 30-years old, with the ability to stand and walk without support; and SMA type IV has adulthood onset and tends to have normal mobility with mild muscle weakness [3,4]. The estimated incidence of SMA is 1 in 10,000 newborns and the prevalence is 1-2 people per 100,000, with an overall carrier frequency of 1 in 54 in the general population. ...
Article
Full-text available
Background Spinal muscular atrophy is a genetic disorder characterized by degeneration of lower motor neurons, leading to progressive muscular atrophy and even paralysis. Spinal muscular atrophy usually associated with a defect of the survival motor neuron 1 (SMN-1) gene. Classification of spinal muscular atrophy is based on the age of onset and maximum motor function milestone achieved. Although spinal muscular atrophy can be screened for in newborns, and even confirmed earlier genetically, this remains difficult in Third World countries such as Indonesia. Case presentation A 28-year-old Asian woman in the first trimester of her second pregnancy, was referred to the neurology department from the obstetric department. Her milestone history showed she was developmentally delayed and the ability to walk independently was reached at 26 months old. At 8 years old, she started to stumble and lose balance while walking. At this age, spinal muscular atrophy was suspected because of her clinical presentations, without any molecular genetic testing. She was married at the age of 25 years and was soon pregnant with her first child. At the gestational age of 32 weeks, her first pregnancy was ended by an emergency caesarean section because of premature rupture of the membranes. In this second pregnancy, she was referred early to the general hospital from the district hospital to receive multidisciplinary care. She and her first daughter underwent genetic testing for spinal muscular atrophy, which has been readily available in our institution since 2018, to confirm the diagnosis and prepare for genetic counseling. Conclusions Managing pregnancy in a patient with spinal muscular atrophy should be performed collaboratively. In this case, genetic testing of spinal muscular atrophy and the collaborative management of this patient allowed the clinical decision making and genetic counseling throughout her pregnancy and delivery.
... homozygous gene in positively suspected patients then clinical testing are repeated with some invasive procedures such as muscle biopsy of samples can also be performed. Neurogenic EMG, MRI and CPK values can also be used for diagnosis, but molecular tools are considered as most superior, non-invasive and accurate for identification of disease ( Figure 2) [91]. With advancement of technology, screening of SMA was done from dried blood spot of upto 1-8 year old DBS filter paper as compare to DNA isolated from fresh blood by Kato et al. [92]. ...
Article
Spinal Muscular Atrophy (SMA) is a pan-ethnic disorder and generally characterized as prevalent lethal genetic disease of infants. It is an autosomal recessive neuromuscular disease caused by degeneration of alpha motor neurons in the spinal cord, resulting in progressive proximal muscle weakness and paralysis. Due to the high carrier frequency (1:50), the burden of this genetic disorder is very heavy in developing countries. Till date no absolute cure or effective treatment of the disease is available in clinical practice, whereas minor enhancement of SMN protein levels can be beneficial. It can be achieved by augmenting SMN2 transcription, stimulating exon 7 splicing and protein stabilization. Due to its low prevalence among population, costly screening and diagnosis, the disease is still lacking proper management. SMN is expressed almost in all tissues of body, still the reason why only lower motor neurons are affected in SMA is unknown. Research is still going on and with advancement of innovative therapies and gene modification, improved outcome may come in near future. Presently, supportive care including respiratory, nutritional, psychiatric and orthopaedic management can ameliorate clinical symptoms and improve survival rates if SMA is diagnosed early in life. Routine prenatal and new-born screening can help with potential benefits and timely management. In this review, the concept of newer methodological system and recent advances for molecular diagnosis of SMA with the variability in the clinical features is stressed. The public health community should remain alert to the rapidly changing developments in early detection and treatment of SMA.
... People with SMA type 3 learn to walk but might lose the ability to walk over time. Seventy percent of people with SMA types 2 and 3 are alive at age 25 [4][5][6]. ...
Article
Full-text available
Background: Spinal muscular atrophy (SMA) is an autosomal-recessive neuromuscular disorder resulting in progressive muscle weakness. In December 2016, the U.S. Food and Drug Administration approved the first treatment for SMA, a drug named nusinersen (Spinraza) that is administered intrathecally. However many children with SMA have neuromuscular scoliosis or spinal instrumentation resulting in challenging intrathecal access. Therefore alternative routes must be considered in these complex patients. Objective: To investigate routes of drug access, we reviewed our institutional experience of administering intrathecal nusinersen in all children with spinal muscular atrophy regardless of spinal anatomy or instrumentation. Materials and methods: We reviewed children with SMA who were referred for intrathecal nusinersen injections from March to December 2017 at our institution. In select children with spinal hardware, spinal imaging was requested to facilitate pre-procedure planning. Standard equipment for intrathecal injections was utilized. All children were followed up by their referring neurologist. Results: A total of 104 intrathecal nusinersen injections were performed in 26 children with 100% technical success. Sixty procedures were performed without pre-procedural imaging and via standard interspinous technique. The remaining 44 procedures were performed in 11 complex (i.e. neuromuscular scoliosis or spinal instrumentation) patients requiring pre-procedural imaging for planning purposes. Nineteen of the 44 complex procedures were performed via standard interspinous technique from L2 to S1. Twenty-two of the 44 complex procedures were performed using a neural-foraminal approach from L3 to L5. Three of the 44 complex procedures were performed via cervical puncture technique. There were no immediate or long-term complications but there was one child with short-term complications of meningismus and back pain at the injection site. Conclusion: Although we achieved 100% technical success in intrathecal nusinersen administration, our practices evolved during the course of this study. As a result of our early experience we developed an algorithm to assist in promoting safe and effective nusinersen administration in children with spinal muscular atrophy regardless of SMA type, abnormal spinal anatomy and complex spinal instrumentation.
Article
Further understanding of the molecular mechanisms is necessary since it is important for designing new drugs. This study aimed to understand the molecular mechanisms involved in the design of drugs that are inhibitors of the α-glucosidase enzyme. This research aims to gain further understanding of the molecular mechanisms underlying antidiabetic drug design. The molecular docking process yielded 4 compounds with the best affinity energy, including γ-Mangostin, 1,6-dimethyl-ester-3-isomangostin, 1,3,6-trimethyl-ester-α-mangostin, and 3,6,7-trimethyl-ester-γ-mangostin. Free energy calculation with molecular mechanics with generalized born and surface area solvation indicated that the 3,6,7-trimethyl-γ-mangostin had a better free energy value compared to acarbose and simulated maltose together with 3,6,7-trimethyl-γ-mangostin compound. Based on the analysis of electrostatic, van der Waals, and intermolecular hydrogen interactions, 3,6,7-trimethyl-γ-mangostin adopts a noncompetitive inhibition mechanism, whereas acarbose adopts a competitive inhibition mechanism. Consequently, 3,6,7-trimethyl-ester-γ-mangostin, which is a derivative of γ-mangostin, can provide better activity in silico with molecular docking approaches and molecular dynamics simulations. This research commenced with retrieving protein structures from the RCSB database, generating the formation of ligands using the ChemDraw Professional software, conducting molecular docking with the Autodock Vina software, and performing molecular dynamics simulations using the Amber software, along with the evaluation of RMSD values and intermolecular hydrogen bonds. Free energy, electrostatic interactions, and Van der Waals interaction were calculated using MM/GBSA. Acarbose, used as a positive control, and maltose are simulated together with test compound that has the best free energy. The forcefields used for molecular dynamics simulations are ff19SB, gaff2, and tip3p.
Article
Full-text available
Objective This study evaluated the feasibility of a matching-pair test using eye-tracking technology to assess nusinersen effectiveness in patients with advanced spinal muscular atrophy (SMA) type I. Methods This prospective, observational study enrolled patients with 5q-SMA type I who had lost gross motor function. Three different levels of matching-pair tests were conducted using the eye-gaze system (My Tobii; TobiiDynavox Inc.) at baseline, and after 9 and 24 weeks of nusinersen treatment. The primary endpoint was the change from baseline in matching-pair test scores and response times (i.e., the time to answer matching-pair test) at 24 weeks from baseline. Children's Hospital of Philadelphia Infant Test of Neuromuscular Disorders (CHOP-INTEND), Pediatric Quality of Life inventory for patients with Neuromuscular Disease (PedsQL-NM) and Numerical Rating Scale (NRS) scores were also assessed as secondary endpoints. Analysis of ocular fixation was performed as an additional analysis. This study was registered at https://www.umin.ac.jp/ctr/ (UMIN000033935). Results Seven patients (one male, six female) aged 5–21 years (median 11 years) were enrolled; all patients were bedridden and six patients were ventilated. All seven patients were able to conduct level 1 matching-pair tests at each assessment; five patients were also able to conduct levels 2 and 3. Two patients (those with the highest CHOP-INTEND scores) were able to complete all tests correctly within 60 s. There was a non-significant trend toward improvement in CHOP-INTEND, PedsQL-NM, and NRS scores over the 6-month period. There were no significant differences in the number of actions, errors, correct answers, or response times between baseline and Week 9 or 24 at any level. However, the result of an additional analysis suggests that detection of eye movement would be useful to evaluate for advanced SMA. Conclusions Eye-tracking systems are possibly feasible for the assessment of treatment efficacy in patients with advanced SMA type I.
Article
Full-text available
Spinal muscular atrophy (SMA) is the most common genetic disease leading to infant mortality. This neuromuscular disorder is caused by the loss or mutation of the telomeric copy of the 'survival of motor neuron' (Smn) gene, termed SMN1. Loss of SMN1 leads to reduced SMN protein levels, inducing degeneration of motor neurons (MN) and progressive muscle weakness and atrophy. To date, SMA remains incurable due to the lack of a method to deliver therapeutically active molecules to the spinal cord. Gene therapy, consisting of reintroducing SMN1 in MNs, is an attractive approach for SMA. Here we used postnatal day 1 systemic injection of self-complementary adeno-associated virus (scAAV9) vectors carrying a codon-optimized SMN1 sequence and a chimeric intron placed downstream of the strong phosphoglycerate kinase (PGK) promoter (SMNopti) to overexpress the human SMN protein in a mouse model of severe SMA. Survival analysis showed that this treatment rescued 100% of the mice, increasing life expectancy from 27 to over 340 days (median survival of 199 days) in mice that normally survive about 13 days. The systemic scAAV9 therapy mediated complete correction of motor function, prevented MN death and rescued the weight loss phenotype close to normal. This study reports the most efficient rescue of SMA mice to date after a single intravenous injection of an optimized SMN-encoding scAAV9, highlighting the considerable potential of this method for the treatment of human SMA.
Article
Full-text available
Spinal muscular atrophy (SMA), a common motor neuron disease in humans, results from loss of functional survival motor neuron (SMN1) alleles. A nearly identical copy of the gene, SMN2, fails to provide protection from SMA because of a single translationally silent nucleotide difference in exon 7. This likely disrupts an exonic splicing enhancer and causes exon 7 skipping, leading to abundant production of a shorter isoform, SMN2Delta7. The truncated transcript encodes a less stable protein with reduced self-oligomerization activity that fails to compensate for the loss of SMN1. This report describes the identification of an in vivo regulator of SMN mRNA processing. Htra2-beta1, an SR-like splicing factor and ortholog of Drosophila melanogaster transformer-2, promoted the inclusion of SMN exon 7, which would stimulate full-length SMN2 expression. Htra2-beta1 specifically functioned through and bound an AG-rich exonic splicing enhancer in SMN exon 7. This effect is not species-specific as expression of Htra2-beta1 in human or mouse cells carrying an SMN2 minigene dramatically increased production of full-length SMN2. This demonstrates that SMN2 mRNA processing can be modulated in vivo. Because all SMA patients retain at least one SMN2 copy, these results show that an in vivo modulation of SMN RNA processing could serve as a therapeutic strategy to prevent SMA.
Article
Full-text available
Spinal muscular atrophy (SMA) is a common neurodegenerative disease caused by deletion or loss-of-function mutations of the survival of motor neurons (SMN) protein. SMN is in a complex with several proteins, including Gemin2, Gemin3 and Gemin4, and it plays important roles in small nuclear ribonucleoprotein (snRNP) biogenesis and in pre-mRNA splicing. Here, we characterize three new hnRNP proteins, collectively referred to as hnRNP Qs, which are derived from alternative splicing of a single gene. The hnRNP Q proteins interact with SMN, and the most common SMN mutant found in SMA patients is defective in its interactions with them. We further demonstrate that hnRNP Qs are required for efficient pre-mRNA splicing in vitro. The hnRNP Q proteins may provide a molecular link between the SMN complex and splicing.
Article
Full-text available
Spinal muscular atrophy (SMA) is a motor neuron disease and the leading genetic cause of infant mortality; it results from loss-of-function mutations in the survival motor neuron 1 (SMN1) gene. Humans have a paralogue, SMN2, whose exon 7 is predominantly skipped, but the limited amount of functional, full-length SMN protein expressed from SMN2 cannot fully compensate for a lack of SMN1. SMN is important for the biogenesis of spliceosomal small nuclear ribonucleoprotein particles, but downstream splicing targets involved in pathogenesis remain elusive. There is no effective SMA treatment, but SMN restoration in spinal cord motor neurons is thought to be necessary and sufficient. Non-central nervous system (CNS) pathologies, including cardiovascular defects, were recently reported in severe SMA mouse models and patients, reflecting autonomic dysfunction or direct effects in cardiac tissues. Here we compared systemic versus CNS restoration of SMN in a severe mouse model. We used an antisense oligonucleotide (ASO), ASO-10-27, that effectively corrects SMN2 splicing and restores SMN expression in motor neurons after intracerebroventricular injection. Systemic administration of ASO-10-27 to neonates robustly rescued severe SMA mice, much more effectively than intracerebroventricular administration; subcutaneous injections extended the median lifespan by 25 fold. Furthermore, neonatal SMA mice had decreased hepatic Igfals expression, leading to a pronounced reduction in circulating insulin-like growth factor 1 (IGF1), and ASO-10-27 treatment restored IGF1 to normal levels. These results suggest that the liver is important in SMA pathogenesis, underscoring the importance of SMN in peripheral tissues, and demonstrate the efficacy of a promising drug candidate.
Article
Full-text available
Spinal muscular atrophy (SMA) is a leading inherited cause of infant death with a reported incidence of ~1 in 10,000 live births and is second to cystic fibrosis as a common, life-shortening autosomal recessive disorder. The American College of Medical Genetics has recommended population carrier screening for SMA, regardless of race or ethnicity, to facilitate informed reproductive options, although other organizations have cited the need for additional large-scale studies before widespread implementation. We report our data from carrier testing (n = 72,453) and prenatal diagnosis (n = 121) for this condition. Our analysis of large-scale population carrier screening data (n = 68,471) demonstrates the technical feasibility of high throughput testing and provides mutation carrier and allele frequencies at a level of accuracy afforded by large data sets. In our United States pan-ethnic population, the calculated a priori carrier frequency of SMA is 1/54 with a detection rate of 91.2%, and the pan-ethnic disease incidence is calculated to be 1/11,000. Carrier frequency and detection rates provided for six major ethnic groups in the United States range from 1/47 and 94.8% in the Caucasian population to 1/72 and 70.5% in the African American population, respectively. This collective experience can be utilized to facilitate accurate pre- and post-test counseling in the settings of carrier screening and prenatal diagnosis for SMA.
Article
Infantile spinal muscular atrophy (SMA) caused by homozygous SMN1 gene deletions/mutations is characterized by neuronal loss and axonopathy of motor neurons. We report two unrelated patients with severe SMA type I who had only one SMN2 copy and developed ulcerations and necroses of the fingers and toes. Sural nerve biopsy was normal in patient 1, whose affected skin displayed necroses and thrombotic occlusions of small vessels. Corresponding to a mouse model and other patients with similar findings, we believe that severe survival motor neuron (SMN) deficiency may present as vasculopathy. Muscle Nerve 42: 144–147, 2010
Article
A newborn girl presented with asphyxia, joint contractures and diminished spontaneous movements. Echocardiography showed hypoplastic left heart. Spinal muscular atrophy type I (SMA I) was diagnosed by detecting a homozygous deletion in the survival motor neuron 1 gene (SMN1). In the first trimester of a subsequent pregnancy, SMA I, hypoplastic left heart, and contractures were identified again. Congenital heart defects (CHD) have now been reported in 20 patients with SMA I, including three previously reported siblings and our two siblings, leading us to hypothesize that SMA I/CHD represents a unique phenotype of SMA I rather than a coincidental association. The homozygous SMN1 deletion may play a role in the development of CHD when it occurs in the presence of mutations or polymorphisms in other genes important for cardiac development. © 2008 Wiley-Liss, Inc.
Chapter
Spinal muscular atrophy (SMA) is characterized by progressive muscle weakness resulting from degeneration and loss of the anterior horn cells (i.e., lower motor neurons) in the spinal cord and the brain stem nuclei. Onset ranges from before birth to adolescence or young adulthood. Poor weight gain, sleep difficulties, pneumonia, scoliosis, and joint contractures are common complications. Before the genetic basis of SMA was understood, it was classified into clinical subtypes; however, it is now apparent that the phenotype of SMA associated with disease-causing mutations of the SMN1 gene spans a continuum without clear delineation of subtypes. Nonetheless, classification by age of onset and maximum function achieved is useful for prognosis and management; subtypes include: SMA 0 (proposed), with prenatal onset and severe joint contractures, facial diplegia, and respiratory failure; SMA I, with onset before age six months; SMA II, with onset between age six and 12 months; SMA III, with onset in childhood after age 12 months; and SMA IV, with adult onset. The diagnosis of SMA is based on molecular genetic testing. The two genes associated with SMA are SMN1 and SMN2. SMN1 (survival motor neuron 1) is the primary disease-causing gene. About 95%-98% of individuals with SMA are homozygous for a deletion or truncation of SMN1 and about 2%-5% are compound heterozygotes for an SMN1 deletion or truncation and an SMN1 intragenic mutation. SMN1 deletion or truncation is typically detected by demonstrating the absence of exon 7 because it can be differentiated from exon 7 of SMN2. SMA carrier testing, a PCR-based dosage assay, allows determination of the number of SMN1 copies by measuring the number of exon 7-containing SMN1 copies. This test can be difficult to interpret because instead of having the normal two copies of SMN1, one on each chromosome, some carriers have the two SMN1 copies on one chromosome and some carriers have an SMN1 intragenic mutation that is not detected by this test. Furthermore, 2% of individuals with SMA have one de novo mutation, meaning that only one parent is a carrier. Because of these difficulties in SMA carrier test interpretation, SMA carrier testing should be provided in the context of formal genetic counseling. Treatment of manifestations: When nutrition is a concern in SMA, placement of a gastrostomy tube is appropriate. As respiratory function deteriorates, tracheotomy or non-invasive respiratory support is offered. Sleep disorder breathing can be treated with nighttime use of continuous positive airway pressure. Surgery for scoliosis in individuals with SMA II and SMA III can be carried out safely if the forced vital capacity is greater than 40%. A power chair and other equipment may improve quality of life. Surveillance: evaluation every six months or more frequently for children who are weak to assess nutritional state, respiratory function, and orthopedic status (spine, hips, and joint range of motion). SMA is inherited in an autosomal recessive manner. Each pregnancy of a couple who have had a child with SMA has an approximately 25% chance of producing an affected child, an approximately 50% chance of producing an asymptomatic carrier, and an approximately 25% chance of producing an unaffected child who is not a carrier. These recurrence risks deviate slightly from the norm for autosomal recessive inheritance because, in about 2% of cases, the affected individual has a de novo SMN1 mutation on one allele; in these instances only one parent is a carrier of an SMN1 mutation, and thus the sibs are not at increased risk for SMA. Carrier testing for at-risk relatives and prenatal testing for pregnancies at increased risk are possible if the disease-causing mutations in a family have been identified.
Article
A signal feature of spinal muscular atrophy (SMA) is the dying back of motor neurons. This is caused by a loss-of-function mutation in SMN1. Administration of an oligonucleotide that modifies the processing of SMN2, a homologous gene, extends the life span in a mouse model of SMA.