+ All Categories
Home > Documents > Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

Date post: 08-Dec-2016
Category:
Upload: lucia
View: 213 times
Download: 0 times
Share this document with a friend
11
Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy Roc ıo Ruiz and Luc ıa Tabares Department of Medical Physiology and Biophysics School of Medicine, University of Seville, Seville, Spain Abstract Spinal muscular atrophy is a genetic disease which severity depends on the amount of SMN protein, the product of the genes SMN1 and SMN2. Symptomatology goes from severe neuromuscular impairment leading to early death in infants to slow progressing motor deficits during adulthood. Much of the knowledge about the pathophysiology of SMA comes from studies using genetically engineered animal models of the disease. Here we investigated one of the milder models, the homozygous A2G SMA mice, in which the level of the protein is restored to almost normal levels by the addition of a mutated transgene to the severe SMN-deficient background. We examined neuromuscular function and found that calcium-dependent neurotransmitter release was significantly decreased. In addition, the amplitude of spontaneous endplate potentials was decreased, the morphology of NMJ altered, and slight changes in short-term synaptic plasticity were found. In spite of these defects, excitation contraction coupling was well preserved, possibly due to the safety factor of this synapse. These data further support that the quasi-normal restoration of SMN levels in severe cases preserves neuromuscular function, even when neurotransmitter release is significantly decreased at motor nerve terminals. Nevertheless, this deficit could represent a greater risk of motor impairment during aging or after injuries. Key words: motor neuron; neurodegeneration; neuromuscular junction; spinal muscular atrophy; synapse. Introduction SMN (survival motor neuron) is an essential and ubiquitous protein which best known functions are its participation in the assembly of small nuclear ribonucleoproteins and in mRNA splicing (Fischer et al. 1997; Liu et al. 1997; Meister et al. 2001; Pellizzoni et al. 2002). A deficiency in SMN expression produces spinal muscular atrophy (SMA), one of the most severe diseases of genetic origin in childhood (Crawford & Pardo, 1996). The disease manifests by deteri- oration in motor function due to the selective loss of sub- sets of a-motor neurons in the spinal cord and reduction in excitatory inputs onto motor neurons (Ling et al. 2010; Park et al. 2010; Mentis et al. 2011; Martinez et al. 2012) . There are several clinic forms of SMA which classification is based on the age of onset and the severity of the symp- toms (Munsat & Davies, 1992). In SMA type I, by far the most frequent form of SMA, the disease is diagnosed within 6 months of birth and is characterized by profound generalized muscular weakness that usually results in death before 2 years of age. In SMA type II, onset of the disease occurs between 6 and 18 months of age and infants are unable to walk. SMA type III and type IV are less severe and symptoms start after 18 months of age (type III), or during adulthood (type IV). Although patients of type III and IV are able to walk without aid, at least for a time, their motor function progressively deteriorates. The different clinic forms of SMA are directly related to the amount of SMN expressed in patients (McAndrew et al. 1997; Feldkotter et al. 2002; Harada et al. 2002). SMN is coded by two genes, SMN1 and SMN2 (Lefebvre et al. 1995), and whereas SMN1 produces full length SMN (SMN- FL), SMN2 produces only 10% of SMN-FL and 90% of a trun- cated form of the protein (SMND7) (McAndrew et al. 1997). As the number of copies of SMN2 varies in the population, the pathology in patients with deleted or mutated SMN1 will depend greatly on the number of SMN2 copies in each individual (Coovert et al. 1997; Lefebvre et al. 1997). SMA type I or type II patients have one or two copies of SMN2, whereas type III or type IV patients have three or more copies of SMN2. The generation of animal models of SMA has contributed enormously to the understanding of the pathophysiology Correspondence Luc ıa Tabares, Department of Medical Physiology and Biophysics, School of Medicine, University of Seville, Avenida S anchez Pizju an, 4, 41009 Seville, Spain. E: [email protected] Accepted for publication 20 February 2013 © 2013 Anatomical Society J. Anat. (2013) doi: 10.1111/joa.12038 Journal of Anatomy
Transcript
Page 1: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

Neurotransmitter release in motor nerve terminals of amouse model of mild spinal muscular atrophyRoc�ıo Ruiz and Luc�ıa Tabares

Department of Medical Physiology and Biophysics School of Medicine, University of Seville, Seville, Spain

Abstract

Spinal muscular atrophy is a genetic disease which severity depends on the amount of SMN protein, the

product of the genes SMN1 and SMN2. Symptomatology goes from severe neuromuscular impairment leading

to early death in infants to slow progressing motor deficits during adulthood. Much of the knowledge about

the pathophysiology of SMA comes from studies using genetically engineered animal models of the disease.

Here we investigated one of the milder models, the homozygous A2G SMA mice, in which the level of the

protein is restored to almost normal levels by the addition of a mutated transgene to the severe SMN-deficient

background. We examined neuromuscular function and found that calcium-dependent neurotransmitter

release was significantly decreased. In addition, the amplitude of spontaneous endplate potentials was

decreased, the morphology of NMJ altered, and slight changes in short-term synaptic plasticity were found. In

spite of these defects, excitation contraction coupling was well preserved, possibly due to the safety factor of

this synapse. These data further support that the quasi-normal restoration of SMN levels in severe cases

preserves neuromuscular function, even when neurotransmitter release is significantly decreased at motor nerve

terminals. Nevertheless, this deficit could represent a greater risk of motor impairment during aging or after

injuries.

Key words: motor neuron; neurodegeneration; neuromuscular junction; spinal muscular atrophy; synapse.

Introduction

SMN (survival motor neuron) is an essential and ubiquitous

protein which best known functions are its participation in

the assembly of small nuclear ribonucleoproteins and in

mRNA splicing (Fischer et al. 1997; Liu et al. 1997; Meister

et al. 2001; Pellizzoni et al. 2002). A deficiency in SMN

expression produces spinal muscular atrophy (SMA), one of

the most severe diseases of genetic origin in childhood

(Crawford & Pardo, 1996). The disease manifests by deteri-

oration in motor function due to the selective loss of sub-

sets of a-motor neurons in the spinal cord and reduction in

excitatory inputs onto motor neurons (Ling et al. 2010;

Park et al. 2010; Mentis et al. 2011; Martinez et al. 2012) .

There are several clinic forms of SMA which classification is

based on the age of onset and the severity of the symp-

toms (Munsat & Davies, 1992). In SMA type I, by far the

most frequent form of SMA, the disease is diagnosed

within 6 months of birth and is characterized by

profound generalized muscular weakness that usually

results in death before 2 years of age. In SMA type II,

onset of the disease occurs between 6 and 18 months of

age and infants are unable to walk. SMA type III and

type IV are less severe and symptoms start after

18 months of age (type III), or during adulthood (type

IV). Although patients of type III and IV are able to walk

without aid, at least for a time, their motor function

progressively deteriorates.

The different clinic forms of SMA are directly related to

the amount of SMN expressed in patients (McAndrew et al.

1997; Feldkotter et al. 2002; Harada et al. 2002). SMN is

coded by two genes, SMN1 and SMN2 (Lefebvre et al.

1995), and whereas SMN1 produces full length SMN (SMN-

FL), SMN2 produces only 10% of SMN-FL and 90% of a trun-

cated form of the protein (SMND7) (McAndrew et al. 1997).

As the number of copies of SMN2 varies in the population,

the pathology in patients with deleted or mutated SMN1

will depend greatly on the number of SMN2 copies in each

individual (Coovert et al. 1997; Lefebvre et al. 1997). SMA

type I or type II patients have one or two copies of SMN2,

whereas type III or type IV patients have three or more

copies of SMN2.

The generation of animal models of SMA has contributed

enormously to the understanding of the pathophysiology

Correspondence

Luc�ıa Tabares, Department of Medical Physiology and Biophysics,

School of Medicine, University of Seville, Avenida S�anchez Pizju�an, 4,

41009 Seville, Spain. E: [email protected]

Accepted for publication 20 February 2013

© 2013 Anatomical Society

J. Anat. (2013) doi: 10.1111/joa.12038

Journal of Anatomy

Page 2: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

of the disease. For example, it has been demonstrated that

the homozygous absence of Smn, the homologous SMN

gene in animals, is embryonically lethal in mice (Schrank et

al. 1997; Hsieh-Li et al. 2000). This early lethality can be pre-

vented by transgenic expression of the human gene SMN2

(Hsieh-Li et al. 2000; Monani et al. 2000b). Remarkably, mice

expressing one or two SMN2 copies have pathologic charac-

teristics similar to those of SMA type I patients (Hsieh-Li

et al. 2000; Monani et al. 2000b), whereas mice expressing

four copies of SMN2 have a normal life span and present

minor alterations such as peripheral necrosis in ears and tail.

On the other hand, expression of 8–16 SMN2 copies in null-

Smn mice results in almost a normal phenotype (Hsieh-Li

et al. 2000; Monani et al. 2000b). Therefore, as in humans,

in mice the disease severity is inversely correlated with the

SMN dosage.

Several SMA type I mouse models have been generated

so far. One of the most used is the so-called SMND7 (Le

et al. 2005). Against a severe background (Smn�/�, plus 1–

2 copies of SMN2), these mice express an additional gene

(SMN1 SMND7) which produces only truncated SMN pro-

tein. As a result, there is a slight increase in maximal survival

(from 5 to 6 days when SMN1 SMND7 is not incorporated

up to 2 weeks after the transgene is included). Importantly,

this provides a wider time window for studying the patho-

logical changes that low levels of SMN produces. For exam-

ple, it has been demonstrated that postnatal maturation

of the NMJ does not take place in most affected muscles

(Kariya et al. 2008; Murray et al. 2008; Kong et al. 2009; Ruiz

et al. 2010), as evidenced at the presynaptic side by the

anomalous distribution and significant decrease in synaptic

vesicles and active zones, reduction in mitochondria, imma-

ture organization of neurofilaments (NF) and microtubules

(Torres-Benito et al. 2011), and up to a ~ 50% reduction in

neurotransmitter release (Kong et al. 2009; Ling et al. 2010;

Ruiz et al. 2010; Torres-Benito et al. 2011, 2012). At the

postsynaptic side, small and immature end-plates are

observed (Kariya et al. 2008; Murray et al. 2008; Kong et al.

2009; Ruiz et al. 2010). In addition to the NMJ changes, loss

of proprioceptive sensory synaptic inputs on motor neurons

is also found (Ling et al. 2010; Mentis et al. 2011). Following

these changes, neurodegenerative signs such as NF accumu-

lation in motor nerve terminals, muscle fiber denervation,

and loss of spinal motor neurons are found (Michaud et al.;

Hsieh-Li et al. 2000; Jablonka et al. 2000; Monani et al.

2000a, 2003; Le et al. 2005; Avila et al. 2007; Kariya et al.

2008; Bowerman et al. 2009; Park et al. 2010; Riessland et

al. 2010; Torres-Benito et al. 2011; Ling et al. 2012).

Much less is known about the pathophysiology of the

milder forms of SMA. However, several mouse models

that resemble SMA type III have already been generated

and are starting to provide valuable information (Jablonka

et al. 2000; Monani et al. 2003; Kariya et al. 2008;

Gladman et al. 2010; Park et al. 2010; Simon et al. 2010;

Osborne et al. 2012). One of these models results in the

expression of the SMN1 A2G mutated transgene on the

severe Smn�/�;SMN2 background (Monani et al. 2003).

A2G is a common missense mutation in the SMN1 gene

in SMA patients (Parsons et al. 1998). Remarkably, the

hemizygous expression of the SMN1 A2G in mouse ame-

liorates the SMA phenotype (from type I towards type

III), significantly improving motor function and extend-

ing mean survival from 5 to 225 days. Even so, these

mice still suffer from motor neuron loss, muscle atrophy,

axonal NF accumulation and immature NMJs (Monani

et al. 2003). In the present study, we investigated a

milder SMA model which expresses the A2G transgene

in homozygosis on the severe Smn�/�;SMN2 back-

ground (Monani et al. 2003). We found that at 1 year

of age, homozygous A2G SMA mice, although smaller,

did not present alterations in the maturation of the

NMJ nor NF accumulation but did have a significant

reduction in neurotransmitter release at motor nerve

terminals of the transversus abdominis (TVA) muscle. In

addition, NMJ morphology appeared slightly altered.

However, assessment of motor function by electromyog-

raphy showed no apparent impairment of hind limbs.

These results allow a better understanding of the alter-

ations present in very mild SMA models in which motor

performance is not apparently affected initially but in

which reduction in the neuromuscular transmission

safety factor could limit motor function in case of

injury, disease or aging.

Methods

Animal model

The A2G mouse line was kindly provided by Dr. A. Burghes

(Department of Molecular Genetics, College of Biological

Sciences, The Ohio State University, Columbus, OH, USA).

Double transgenic Smn+/�;SMN2+/+;SMN A2G+/+ mice on a

FVB/N background were interbred. Genotyping was done

by PCR using tail DNA as previously described (Monani et al.

2003). Control mice were Smn+/+;SMN2+/+;SMN A2G+/+. Smn-

deficient mice were Smn�/�;SMN2+/+;SMN A2G+/+ (referred

to henceforth as homozygous A2G SMA mice). Only male

mice older 1 year or older were used for experiments and

controls were age-matched littermates of homozygous A2G

SMA mice. All experiments were performed according to

the guidelines of the European Council Directive for the

Care of Laboratory Animals.

Electromyography

Compound muscular action potentials (CMAPs) were

recorded as described (Ruiz et al. 2005; Simon et al. 2010) in

anesthetized mice (tribromoethanol 2%, 0.15 mL per 10 g

body weight, i.p.). Stimulating needle electrodes were

placed at the sciatic notch and the head of the fibula. The

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares2

Page 3: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

active recording electrode (a circumferential surface elec-

trode) was placed at midthigh. The reference electrode was

inserted at the base of the fifth foot phalanx. A ground

electrode was placed at the base of the tail. Supramaximal

responses were first recorded, followed by responses to

incremental currents, in very small steps, from subthreshold

levels until the progressive recruitment of 10–15 responses.

Each incremental current pulse was applied three times,

and responses were considered stable and therefore accept-

able if they were identical. The mean single motor unit

action potential (SMUAP) was calculated averaging the size

increments of the first 10 responses.

Muscle preparation

Mice were anesthetized with tribromoethanol (2%,

0.15 mL per 10 g body weight, i.p.) and killed by exsangui-

nation. The TVA muscles were dissected with their nerve

branches intact and pinned to the bottom of a 2-mL cham-

ber, over a bed of cured silicone rubber (Sylgard, Dow Corn-

ing). Preparations were continuously perfused with a

solution of the following composition (in mM): 125 NaCl, 5

KCl, 2 CaCl2, 1 MgCl2, 25 NaHCO3 and 15 glucose. The solu-

tion was continuously gassed with 95% O2 and 5% CO2,

which maintained the pH at 7.35. The muscles were used

either for intracellular electrical recordings or were fixed

for performing immunohistochemistry.

Intracellular electrical recording and analysis

The nerve was stimulated by means of a suction elec-

trode. The stimulation consisted of square-wave pulses of

0.5 ms duration and 2–40 V amplitude, at variable fre-

quencies (0.5–100 Hz). A glass microelectrode (10–20 MO)filled with 3 M KCl was connected to an intracellular

recording amplifier (Neuro Data IR283; Cygnus Technol-

ogy) and used to impale single muscle fibers near the

motor nerve endings. Evoked endplate potentials (EPPs)

and miniature EPPs (mEPPs) were recorded from different

NMJs within the muscle as described previously (Ruiz et

al. 2010). Muscular contraction was prevented by includ-

ing in the bath 3–4 lM l-conotoxin GIIIB (Alomone Labo-

ratories), a specific blocker of muscular voltage gated

sodium channels. Recording was performed at room tem-

perature (22–23 °C). The mean amplitudes of the EPP and

mEPPs recorded at each NMJ were linearly normalized to

�70 mV resting membrane potential. EPP amplitudes

were corrected for nonlinear summation (Martin, 1955).

The kinetics of EPP and mEPP were characterized by their

rise time (10–90%) and decay time constant (calculated

from the exponential fit of the decay phase). Quantal

content (QC) was estimated by the direct method, which

consists of recording mEPPs and EPPs (nerve stimulation

0.5 Hz) simultaneously and then calculating the ratio:

QC = average peak EPP/average peak mEPP.

Immunohistochemistry

Dissected TVA muscles from control and A2G SMA mice

were fixed with 4% paraformaldehyde (PFA) for 2 h. Mus-

cles were bathed with 0.1 M glycine in phosphate-bufferd

saline (PBS) for 20 min, then permeabilized with 1% (v/v)

Triton X-100 in PBS for 30 min, and incubated then in a

blocking solution containing 5% (w/v) bovine serum albu-

min (BSA), 1% Triton X-100 in PBS for 2 h. Samples were

incubated overnight at 4 °C with a mouse primary antibody

against 160 kDa neurofilament (NF; 1 : 750, Millipore). Next

day, muscles were rinsed for 1 h in PBS containing 1%

Triton X-100, incubated for 1 h both with 4 lg mL–1 Alexa

488-conjugated goat anti-mouse secondary antibody (Invi-

trogen) and 10 ng mL–1 rhodamine-bungarotoxin (Sigma),

and rinsed again with PBS for 90 min. Finally, muscles were

mounted in glycerol containing DABCO and imaged with

an upright Olympus FV1000 confocal microscope. Z-stack

projections were made from serial scanning every 0.5 lm to

reconstruct the NMJ.

Statistics

Results are reported as means � SEM, with n being the

number of muscle fibers per group and N the number of

mice per group. Student’s t-test was performed as unpaired,

two-tailed sets when the distribution was normal and the

Mann–Whitney rank sum test was used when the distribu-

tion was not normal. Results were considered statistically

different when P < 0.05.

Results

Homozygous expression of SMN1 A2G on the SMA severe

mouse background extended survival for more than 1 year,

as previously described (Monani et al. 2003). Body weight

was, however, reduced (~ 20%) in comparison with their

control littermates (Fig. 1A) and mice presented shorter

tails. No other external alterations, such as ear or tail necro-

sis (Hsieh-Li et al. 2000; Osborne et al. 2012), were apparent.

As motor performance was also normal and no signs of

limb muscle wasting or strength loss were observed in

homozygous A2G SMA mice (data not shown), we explored

the morphological characteristics of the NMJ in TVA, one of

the most affected muscles in the severe SMA mouse model

SMND7 (Murray et al. 2008; Ruiz et al. 2010; Torres-Benito

et al. 2011). In the severe model, important defects in post-

natal maturation of the pre- and postsynaptic motor termi-

nals, together with axonal and terminal NF accumulation

and denervation have been reported. In addition, in

8-month-old hemizygous A2G SMA mice poor maturation

of end-plates has been also described at the oblique

abdominal muscle (Kariya et al. 2008).

We stained postsynaptic receptors with bungarotoxin-

rhodamine, and axons and intraterminal axonal branches

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares 3

Page 4: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

with an antibody anti-NF, revealed by an Alexa488 second-

ary antibody. Figure 1B shows representative examples of

NMJs from a 1-year-old control (left panel) and a littermate

homozygous A2G SMA mouse. Contrary to what it was

found in most severe mouse models, NMJs from homozy-

gous SMA mice typically presented mature characteristics

[i.e. acetylcholine receptors (AChRs) grouped in bands, and

perforated end-plates], and no signs of denervation. How-

ever, ~ 84% of NMJs showed a fragmented appearance and

separated preterminal axonal branches (Fig. 1B, right),

contrary to the pretzel-like form found in ~ 63% of control

littermates (Fig. 1B, left). In contrast, no difference in NF

accumulation was observed between homozygous A2G

SMA and control mice.

Synaptic transmission reduction in motor nerve

terminals

To investigate whether synaptic transmission was affected

in motor nerve terminals from the TVA in homozygous

A2G SMA mice, we studied the properties of miniature

(mEPP) and evoked (EPP) end-plate potentials in 1-year-

old mice and compared them with those of their control

littermates.

Smaller miniature end-plate potentials in A2G SMA mice

Spontaneous neurotransmitter release was investigated by

recording the miniature end-plate potentials (mEPPs). In

A2G SMA mice, the mEPPs were smaller than in controls, as

can be observed directly from the recording traces (Fig. 2A)

and from their different mEPP size distributions (Fig. 2B).

This was further confirmed in the cumulative frequency dis-

tribution plot, which showed a significant shift to the left

of the mutant curve with respect to that of controls

(Fig. 2C), with median values of 0.39 mV in mutants and

0.57 mV in controls (P < 0.001; Mann–Whitney test). The

origin of this difference was not clear but among the possi-

ble explanations are a decrease in the amount of neuro-

transmitter content (quantal size), a decrease in the

number of postsynaptic acetylcholine receptors, and a

decrease in the sensitivity of the postsynaptic receptors.

On the other hand, when the mean frequency of sponta-

neous neurotransmitter release per fiber was compared, no

statistical difference between controls (1.11 � 0.19 events/s;

n = 18, N = 3) and A2G SMA mice (0.97 � 0.21 events/s;

n = 18,, N = 3; P = 0.56) was found (Fig. 2D).

Decrease in calcium-dependent neurotransmission in

A2G SMA mice

Next, we investigated the postsynaptic electrical response

(end-plate potential, EPP) following presynaptic single

action potential elicited by the electrical stimulation of the

nerve (Fig. 3A). The mean size of the EPP in each fiber was

obtained from recordings lasting 180–200 s during which

the nerve was stimulated repetitively at 0.5 Hz. In A2G

SMA mice, the mean amplitude of the EPP was signifi-

cantly less (P = 0.004; two tailed t-test) than in controls

(23.28 mV � 0.39; n = 18, N = 3, vs. 46.70 mV � 0.48;

n = 18, N = 3, respectively; Fig. 3B). We also compared

quantal content (the amount of synaptic vesicles that fuse

with the plasma membrane in response to an action

potential) and found that it was reduced ~ 21% in A2G

SMA mice (P = 0.03) in comparison with controls

(54.16 � 4.58; n = 18, N = 3, vs. 68.89 � 6.21; n = 18,

N = 3, respectively; Fig. 3C).

The analysis of the kinetics of the EPP, quantified by mea-

suring the rise time between the 10 and 90% of the rising

phase, and the decay time constant of the falling phase,

showed an small, but significant, slower rising phase in A2G

SMA mice than in controls (1.09 ms � 0.05; n = 18, N = 3,

vs. 0.96 ms � 0.05; n = 18, N = 3, respectively (P = 0.02;

Fig. 3A, left graph), suggesting a slight impairment

between pre- and postsynaptic coupling. No significant dif-

ference was found in the decay time constant (A2G SMA:

2.88 ms � 0.24; n = 18, N = 3; controls: 2.58 ms � 0.18; n

= 18, N = 3; P = 0.22; Fig. 3A, right graph).

Small alterations in short-term plasticity

To explore the ability of A2G SMA motor terminals to

change their synaptic strength, trains of repetitive stimuli of

05

10152025303540

0 100 200 300 400

Bod

y w

eigh

t (g)

Age (days)

ControlA2G SMA

A2G SMAControl

NF

BTX-Rho

Merge

A

B

Fig. 1 Body weight and NMJ in homozygous A2G SMA mice. (A)

Body weight in control (open symbols) and homozygous A2G SMA

(closed symbols) mice. (B) Representative NMJ confocal images from

control and homozygous A2G SMA mice. Neurofilaments have been

labeled by specific antibodies (green) and AChR have been labeled

with BTX-Rho (red). Scale bar: 15 lm.

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares4

Page 5: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

different frequencies and durations were applied to the

nerve and the amplitude of the EPP measured in control

and A2G SMA fibers. Figure 4A shows representative exam-

ples of postsynaptic responses and appears to illustrate no

large differences between genotypes. This was confirmed

by statistical analysis in a number of terminals as shown in

the graph in Fig. 4B, representing the mean size of the EPPs

along the train at 50 Hz in control and A2G SMA mice

fibers, normalized to the size of the first response. At this

frequency, the degree of facilitation and depression along

the train was not different between genotypes.

We also measured the paired-pulse facilitation (EPP2/

EPP1), which gives an indication of the initial vesicular

release probability. No differences between controls and

A2G SMA mice were found with 1-s stimulations at differ-

ent frequencies (10–100 Hz; Fig. 4C). Finally, when the train

index was calculated (see figure legend), which is an indica-

tion of the amount of short-term depression, a trend

toward less depression in mutant in comparison with con-

trol terminals was observed, although this only reached sta-

tistical significance at 20 Hz (0.97 � 0.02; n = 18, N = 3, vs.

0.91 � 0.02; n = 23, N = 3; P = 0.04; Fig. 4D). These results

show that short-term plasticity in A2G SMAmouse terminals

was, in general, well preserved in spite of the decrease in

quantal content.

No failures

In contrast to frequent failures in neurotransmission

reported in 9-month-old hemizygous A2G SMA mice (Kariya

et al. 2008), we found no signs of failures in 1-year-old

homozygous mice at any of the stimulation frequencies

explored (0.5–100 Hz; data not shown).

Electromyography

Although no signs of limb muscle weakness were appar-

ent in homozygous A2G SMA mice, EMG recordings were

performed to test whether a reduction in compound mus-

cular action potential (CMAP) amplitude or alteration in

the size of the motor units was detectable, as occurs in

SMA patients (Swoboda et al. 2005) and in other mild

SMA mouse models (Kariya et al. 2008; Simon et al.

2010).

Recordings were done by means of a surface ring elec-

trode placed at the middle of the thigh in 1-year-old male

mice homozygous for A2G and wild-type littermates. Maxi-

mum compound muscle action potential (CMAP) and motor

unit (MU) recruitment were recorded after sciatic nerve

stimulation.

As previously reported (Monani et al. 2003), no spontane-

ous activity (fibrillation or fasciculation potentials) in resting

0

20

40

60

80

100

0 0.5 1 1.5 2 2.5 3

Control (946 events; 18.3)A2G SMA (910 events; 18.3) 0.2

0.6

1

1.4

0.5 mV

mEPPs amplitude (mV)

Cum

. fre

quen

cy %

mE

PP

s pe

r sec

.(18.3) (18.3)

Control A2G SMA

100

80

60

40

20

02.01.51.00.50.0

mEPP size (mV)

200

150

100

50

02.01.51.00.50.0

mEPP size (mV)

# of

obs

erva

tions

# of

obs

erva

tions

500 ms

ControlA2G SMA

A

B

C DFig. 2 Spontaneous end-plate potentials in

homozygous A2G SMA mice.

(A) Representative recordings from the TVA

muscle in control (black traces) and in

homozygous A2G SMA mice (gray traces).

(B) Frequency histograms of mEPP sizes in

representative examples in both genotypes.

(C) Cumulative mEPP size curves in both

genotypes. (D) Frequency of mEPP

occurrence. The numbers between

parentheses are the number of fibers and the

number of mice studied, respectively.

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares 5

Page 6: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

(18.3) (18.3)EP

Ps

ampl

itude

(mV

)

0

10

20

30

40

50

**

(18.3) (18.3)

Qua

ntal

Con

tent

0

20

40

60

80

*

*

(18.3) (18.3) (18.3) (18.3)

Ris

e tim

e (m

s)

Dec

ay τ

(m

s)

5 ms 0

0.2

0.4

0.6

0.8

1

1.2

1

2

3

ControlA2G SMAA

B C

Fig. 3 Decreased calcium-dependent neurotransmitter release in motor nerve terminals of homozygous A2G SMA mice. (A) EPP traces from

control (black trace) and homozygous A2G SMA TVA muscles (gray trace). Recordings were normalized to the same peak amplitude for compari-

son. Rising (left graph) and decay (right graph) times of evoked EPPs in control and homozygous A2G mice. (B,C) Mean EPP size (B) and quantal

content (C) in control and homozygous A2G mice. *P < 0.05; ** P < 0.005.

0.5 mV

Control

0

0.2

0.4

0.6

Trai

n in

dex

10 20 50 100Frequency (Hz)

*

PP

F

10 20 50 100Frequency (Hz)

1 mV

50 ms

0.8

1

1.2

0 2 4 6 8 10EP

Ps

ampl

itude

(N)

Stimulus number

Control (17.3)

A2G SMA (13.3)

50 Hz

50 ms

A2G SMA

0.8

1

00.20.40.60.8

11.2

A

B C D

Fig. 4 Short-term plasticity in homozygous A2G SMA mice. (A) Representative EPP recordings from the TVA muscle in control (black trace) and in

homozygous A2G SMA mice (gray trace) during a train of stimuli at 50 Hz. (B) For each fiber, we constructed the mean curve observed for three

sequential stimulus trains normalized to the first EPP, separated by 20-s intervals. (C) PPF quantification (EPP2/EPP1) at different frequencies

(10–100 Hz) in control (white bars) and in homozygous A2G SMA mice (gray bars). (D) Normalized depression measured as train index,

[(EPP9 + EPP10)/2]/EPP1, at different frequencies (10–100 Hz) in control (white bars) and in homozygous A2G SMA mice (gray bars) terminals.

* P < 0.05.

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares6

Page 7: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

muscles was observed in homozygous A2G SMA mice

(n = 5), suggesting no denervation or anomalous nerve

excitability.

Recording of the maximal CMAP after supramaximal

nerve stimulation in both mouse groups showed no signifi-

cant differences in CMAP first peak amplitude between

control and A2G SMA mice (37.8 � 11.6 mV; n = 3; vs.

45.7 � 3.4 mV; n = 5; respectively, P = 0.58), suggesting no

differences in the number of functional motor units

between genotypes (Fig. 5A).

To test whether mutant muscles were more predisposed

to fatigue, the nerve was supramaximally stimulated at

20 Hz for 1 s while the successive CMAP were recorded

(Fig. 5B). The amount of fatigue was calculated as the per-

cent of decrease in the CMAP peak amplitude at the end of

the stimulation train relative to the first response. In control

mice the signal depression was 17 � 2% (N = 3) and in

homozygous A2G SMA mice it was 16 � 4% (N = 2); a non-

statistical difference (P = 0.67).

Finally, to estimate the average single motor unit action

potential (SMUAP) we applied 10 ‘successful’ successive

stimuli of increasing strength and recorded the increment

in the responses which represent the successive recruitment

of single motor units. Examples of the sizes of the

increments are shown in Fig. 5C, in control and A2G SMA

mice. In neither of the genotypes were large steps

observed, which is an indication that no giant motor units

were recruited. Moreover, quantification of the average

SMUAP amplitude in control and A2G SMA mice showed

no statistical differences (1.4 � 0.33 mV; N = 3 vs.

1 � 0.14 mV; N = 5, respectively; P = 0.34), further support-

ing that there was no axonal sprouting, similar to what it

has been reported in another mild SMA model (Gladman

et al. 2010; but see Monani et al. 2003). Therefore, these

data demonstrate that the average size of measured motor

units was not different between genotypes and support

that no significant loss of motor units occurs in A2G SMA

mice.

Discussion

SMA is not only a neurodegenerative disease in children

and adults but also a neurodevelopment disease when SMN

levels are low before birth. At early stages of development,

SMN is implicated in the control of neurite length, in the

correct development of axonal growth cone, in motor axon

guidance, and in the maturation and maintenance of

synaptic contacts in motor circuits. Most of this information

has been obtained from in vivo and in vitro SMA models in

which SMN is much decreased. However, less is known

about the pathological changes in mild animal models of

the disease in which SMN is not greatly reduced. The aim of

this work has been to investigate possible alterations in a

very mild SMA mouse model, the homozygous A2G SMA

mouse (Monani et al. 2003), with the aim of better

understanding which properties are preserved and which

are not in the milder forms of the disease.

In mouse models of slowly progressing motor neuron dis-

eases, loss of motor neurons is partially compensated by

axonal sprouting and reinnervation, a process in which

Control

30 mV

2 ms

0

0.4

0.8

1.2

0

0.4

0.8

1.2

0 2 4 6 8 100 2 4 6 8 10

200 ms

30 mV

ΔV (m

V)

Stimulus Stimulus

0

0.8

1.6

SM

UA

P (m

V)

ΔV (m

V)

0

20

40

60

CM

AP

(mV

)

0

0.4

0.8

A20

/A1

ControlA2G SMAA2G SMA

(2) (3)

(3) (5)

(3) (5)

A

B

C

Fig. 5 EMG characteristics of hind limb

muscles in homozygous A2G SMA mice.

(A) Representative recording of CMAP in both

genotypes. Graph represents the

quantification of the absolute first peak

amplitude of CMAP in response to a

supramaximal stimulation in control (white

bars) and in homozygous A2G SMA mice

(gray bars) terminals. (B) Representative

recordings during a train of stimuli at 100 Hz

in control (left trace) and in homozygous A2G

SMA mice (right trace). The bar graph shows

the depression of CMAP amplitudes (A20/A1)

in both genotypes. (C) Amplitudes of

SMUAPs in control and homozygous A2G

mice in response to stimuli of increasing

amplitude. Each number in the x-axis

represents a stimulus that elicited an

increment in the amplitude of the response.

SMUAP mean size is showed in the graph in

both studied groups. The numbers between

parentheses are the number of fibers and the

number of mice studied, respectively.

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares 7

Page 8: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

CNTF (ciliary neurotrophic factor) participates (Simon et al.

2010; Selvaraj et al. 2012). This remodeling gives rise to

anomalously large motor units that, in many instances, can

be shown by electromyography (EMG). In addition, other

anomalous features in motor neuron diseases, such as spon-

taneous activity and alterations in the normal size of

CMAPs, are good indications of the degree of affectation

and of the compensation capability of the system. For

example, in contrast to hemizygous A2G SMA mice and to

heterozygous (Smn+/�) mice, which present abnormal EMG

activity (Monani et al. 2003; Simon et al. 2010), homozygous

A2G SMA mice exhibited an apparent absence of EMG

abnormalities in hind limb muscles. This result is not unex-

pected given the higher SMN expression level in A2G SMA

mice in comparison with other type III SMA mouse models

(Monani et al. 2003). The absence of spontaneous muscle

electrical activity at rest in homozygous A2G SMA mice sug-

gests that end-plates were not denervated and that motor

axons were not anomalously excited. In patients, no sponta-

neous activity in the EMG has been reported in very mild

type III SMA forms of the disease (Hausmanowa-Petrusewicz

& Jozwik, 1986). In the same way, the size of the CMAP was

not different between control and homozygous A2G SMA

mice, which is indicative of either no alteration or the pres-

ence of compensatory large motor units resulting from

sprouting and reinnervation, as occurs in Smn+/� mice

(Simon et al. 2010). We favor the first possibility, as the

mean size of motor units was not different between control

and A2G SMA mice, contrary to what has been found in

Smn+/� mice (Simon et al. 2010) and in other mouse models

of motor neuron diseases, such in that of spinal muscular

atrophy with respiratory distress type I (Ruiz et al. 2005) or

in a model of ALS (Hegedus et al. 2009). Nevertheless, the

presence of high threshold anomalously large motor units

in this SMA model cannot be completely excluded, as the

technique used for this estimation is based on the recruit-

ment of 10–12 motor units with the lower thresholds. In

fact, in a mouse model of ALS it has been shown that fast-

twitch and fast-fatigable (FF) motor neurons, which usually

have a higher threshold for activation, are lost first, fol-

lowed by fast-twitch and fatigue-resistant (FR) motor neu-

rons, whereas slow-twitch (S) are less affected (Pun et al.

2006). For the most severe form of SMA, a similar suscepti-

bility may also exist (Kanning et al. 2010). In severe SMA

mouse models it has been shown that large motor neurons,

which usually have a high threshold for activation and

belong to motor units of the FF type (Hashizume et al.

1988), are more vulnerable than smaller ones (Baumer et al.

2009). Accordingly, in severe SMA patients, type II muscle

fibers, which normally are innervated by fast motor neu-

rons, are atrophic, whereas slow type I muscle fibers are

better preserved and even hypertrophic (Dubowitz, 1978).

However, this sole hypothesis is not sufficient to explain

why in severe SMA mouse models, muscles with predomi-

nant oxidative slow-twitch fibers, such as the TVA, are more

affected than glycolytic fast-twitch muscles, such the LAL

(Murray et al. 2008; Ruiz et al. 2010). Other proposed fac-

tors for motor neuron vulnerability have been the location

of motor neurons at the spinal cord, the synapsing pheno-

type or the metabolic demand (Monani et al. 2000b; Kariya

et al. 2008; Murray et al. 2008; Baumer et al. 2009; Ruiz

et al. 2010; Ling et al. 2012). Finally, it is of interest to note

that in SMA and ALS patients, extraocular muscles are very

well preserved during the course of the disease (Kubota

et al. 2000; Mitsumoto et al. 2006). Although these muscles

have fast-twitch muscle fibers, they also have extremely

small motor units. However, it is not clear whether this is

indeed a protective factor.

In homozygous A2G SMA mice, the absence of pathologi-

cal electromyography signs in hind limb muscles contrasts

with the alterations in neuromuscular transmission

observed in the TVA, namely smaller mEPPs and ~ 21%

reduction in the number of vesicles fused with the plasma

membrane per action potential (quantal content). The

reduction in mEPP size was unexpected considering the

smaller size of homozygous mice in comparison with con-

trol littermates. Smaller muscle fibers normally have a

higher input resistance, which implies a larger voltage

change for the same amount of current flow. Nevertheless,

our results are in accordance with a previously reported

reduction in miniature postsynaptic currents in a severe

SMA model (Martinez et al. 2012). Among the reasons that

could explain this reduction are a decrease in synaptic vesi-

cle ACh content, a reduction in the density of ACh postsyn-

aptic receptors, or a combination of both. On the other

hand, the reduction in quantal content in SMA mouse mod-

els is a constant in most muscles in different SMA mouse

models (Torres-Benito et al. 2011; Ling et al. 2012). In the

TVA, it is around 50% at 2 weeks of age in the SMND7

model (Ruiz et al. 2010; Torres-Benito et al. 2011) but

reduced only 21% in 1-year homozygous mice (Fig. 3C).

Given the high safety factor in neuromuscular transmission,

this deficit probably does not compromise function. Fur-

thermore, repetitive stimulation of the nerve terminals did

not result in larger synaptic depression than in control litter-

mates. The lack of alteration in short-term synaptic plastic-

ity, in spite of the decrease in quantum content, is not

surprising given the capability of the neuromuscular system

to adapt to chronic changes (Ferraiuolo et al. 2009). Also, at

the observational level, no postural abnormalities or respi-

ratory distress was observed in mutant mice as would be

expected if the TVA function were deficient. It cannot be

discarded, however, that some functional deficit could

appear at mice older than 1 year, when age-related motor

neuron loss occurs.

Paradoxically, in hemizygous A2G SMA mice at 8–9-

month-old mice, an increase in quantal content was

detected in the semitendinosus muscle in relation to litter-

mate controls, in addition to frequent neurotransmission

failures in a percentage of fibers (Kariya et al. 2008). These

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares8

Page 9: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

changes were accompanied by immature end-plates and NF

accumulation in axonal motor branches of the gastrocne-

mius muscle (Kariya et al. 2008). We did not observe these

alterations in the TVA muscle of the homozygous mice,

probably because the homozygous model is a form of the

disease that is even milder than the heterozygous Smn+/�(Jablonka et al. 2000; Simon et al. 2010). In accordance with

the apparent lack of muscle weakness, our morphological

and functional results showed no signs of denervation, con-

trary to what it is found in other mouse SMA models that

present different degrees of impairment, from large neuro-

muscular denervation in severe SMND7 mice (Ling et al.

2012) to reduced remodeling potential of NMJs in Smn2B/�

mice, apparently in relation to the loss of terminal Schwann

cells (Murray et al. 2012). On the other hand, we did not

find any signs of sprouting and reinnervation, contrary to

what it has been reported in SMA intermediate/mild models

such as hemizygous A2G (Monani et al. 2003;) and in Smn+/

� mice (Simon et al. 2010).

The ability of SMN1 A2G to rescue the SMA phenotype is

of great interest. It has been shown previously that A2G

does not efficiently self-associate, but it can form complexes

when some amount of SMN (full-length) exists (Monani

et al. 2003). Therefore, it is critical that SMN reaches a cer-

tain minimal threshold, as the requirement for this is higher

during the embryonic and early postnatal period than in

the adulthood, as demonstrated in the different therapeutic

assays in SMA animal models (Lorson et al. 2010; Sendtner,

2010; Hua et al. 2011; Porensky et al. 2012). If the SMN

level is below its critical threshold at birth and is not rap-

idly restored, the postnatal maturation of motor nerve ter-

minals stops (Kariya et al. 2008; Torres-Benito et al. 2011).

Besides this effect, neurodegeneration changes at the NMJ

and other synapses takes place. Furthermore, when the

SMN level is well below its normal threshold, the degener-

ative changes are so fast that there is no time for

compensatory mechanisms such as sprouting and reinner-

vation. The correlation, if any, between synapse matura-

tion defects and neurodegeneration is not clear as yet. It is

evident that alteration of the NMJ can exist in apparently

mature terminals, as in the SMA model in the present

study. However, it could be also that an inappropriate

maturation process of the NMJ could accelerate the neu-

rodegeneration process.

In summary, we found that homozygous A2G SMA mice,

in spite of having no apparent motor functional deficits

and having mature NMJs, present a significant reduction in

neurotransmitter release and morphological NMJ altera-

tions (fragmentation). However, evaluation of neuromuscu-

lar coupling by electromyography showed no alterations in

hind limb muscles. These results suggest that although in

mild cases of SMA, motor symptoms could be initially

absent or very mild, the safety factor at the NMJ is

decreased by the reduction in neurotransmitter release.

Therefore, complementary strategies for avoiding age-

related or disease-related impairment of motor function

might be required in patients with the milder form of the

disease. For example, in future studies it might be of inter-

est to test whether caloric restriction or programmed exer-

cise could prevent NMJ alterations.

Acknowledgements

The authors thank A. Burghes for the generous gift of this mouse

model. The financial support of the following research grants is

gratefully acknowledged: Spanish Ministry of Science and Innova-

tion BFU2010-21648, and Junta de Andaluc�ıa P09-CVI-4862.

References

Avila AM, Burnett BG, Taye AA, et al. (2007) Trichostatin A

increases SMN expression and survival in a mouse model of

spinal muscular atrophy. J Clin Invest 117, 659–671.

Baumer D, Lee S, Nicholson G, et al. (2009) Alternative splicing

events are a late feature of pathology in a mouse model of

spinal muscular atrophy. PLoS Genet 5, e1000773.

Bowerman M, Anderson CL, Beauvais A, et al. (2009) SMN,

profilin IIa and plastin 3: a link between the deregulation

of actin dynamics and SMA pathogenesis. Mol Cell Neurosci

42, 66–74.

Coovert DD, Le TT, McAndrew PE, et al. (1997) The survival

motor neuron protein in spinal muscular atrophy. Hum Mol

Genet 6, 1205–1214.

Crawford TO, Pardo CA (1996) The neurobiology of childhood

spinal muscular atrophy. Neurobiol Dis 3, 97–110.

Dubowitz V (1978) Muscle disorders in childhood. Major Probl

Clin Pediatr 16, iii-xiii, 1–282.

Feldkotter M, Schwarzer V, Wirth R, et al. (2002) Quantitative

analyses of SMN1 and SMN2 based on real-time lightCycler PCR:

fast and highly reliable carrier testing and prediction of severity

of spinal muscular atrophy. Am J Hum Genet 70, 358–368.

Ferraiuolo L, De Bono JP, Heath PR, et al. (2009) Transcriptional

response of the neuromuscular system to exercise training and

potential implications for ALS. J Neurochem 109, 1714–1724.

Fischer U, Liu Q, Dreyfuss G (1997) The SMN-SIP1 complex has

an essential role in spliceosomal snRNP biogenesis. Cell 90,

1023–1029.

Gladman JT, Bebee TW, Edwards C, et al. (2010) A humanized

Smn gene containing the SMN2 nucleotide alteration in exon

7 mimics SMN2 splicing and the SMA disease phenotype. Hum

Mol Genet 19, 4239–4252.

Harada Y, Sutomo R, Sadewa AH, et al. (2002) Correlation

between SMN2 copy number and clinical phenotype of

spinal muscular atrophy: three SMN2 copies fail to rescue

some patients from the disease severity. J Neurol 249,

1211–1219.

Hashizume K, Kanda K, Burke RE (1988) Medial gastrocnemius

motor nucleus in the rat: age-related changes in the number

and size of motoneurons. J Comp Neurol 269, 425–430.

Hausmanowa-Petrusewicz I, Jozwik A (1986) The application of

the nearest neighbor decision rule in the evaluation of elec-

tromyogram in spinal muscular atrophy (SMA) of childhood.

Electromyogr Clin Neurophysiol 26, 689–703.

Hegedus J, Putman CT, Gordon T (2009) Progressive motor unit

loss in the G93A mouse model of amyotrophic lateral sclerosis

is unaffected by gender. Muscle Nerve 39, 318–327.

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares 9

Page 10: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

Hsieh-Li HM, Chang JG, Jong YJ, et al. (2000) A mouse model

for spinal muscular atrophy. Nat Genet 24, 66–70.

Hua Y, Sahashi K, Rigo F, et al. (2011) Peripheral SMN restora-

tion is essential for long-term rescue of a severe spinal muscu-

lar atrophy mouse model. Nature 478, 123–126.

Jablonka S, Schrank B, Kralewski M, et al. (2000) Reduced sur-

vival motor neuron (Smn) gene dose in mice leads to motor

neuron degeneration: an animal model for spinal muscular

atrophy type III. Hum Mol Genet 9, 341–346.

Kanning KC, Kaplan A, Henderson CE (2010) Motor neuron

diversity in development and disease. Annu Rev Neurosci 33,

409–440.

Kariya S, Park GH, Maeno-Hikichi Y, et al. (2008) Reduced SMN

protein impairs maturation of the neuromuscular junctions in

mouse models of spinal muscular atrophy. Hum Mol Genet 17,

2552–2569.

Kong L, Wang X, Choe DW, et al. (2009) Impaired synaptic vesi-

cle release and immaturity of neuromuscular junctions in

spinal muscular atrophy mice. J Neurosci 29, 842–851.

Kubota M, Sakakihara Y, Uchiyama Y, et al. (2000) New ocular

movement detector system as a communication tool in venti-

lator-assisted Werdnig-Hoffmann disease. Dev Med Child

Neurol 42, 61–64.

Le TT, Pham LT, Butchbach ME, et al. (2005) SMNDelta7, the

major product of the centromeric survival motor neuron

(SMN2) gene, extends survival in mice with spinal muscular

atrophy and associates with full-length SMN. Hum Mol Genet

14, 845–857.

Lefebvre S, Burglen L, Reboullet S, et al. (1995) Identification

and characterization of a spinal muscular atrophy-determining

gene. Cell 80, 155–165.

Lefebvre S, Burlet P, Liu Q, et al. (1997) Correlation between

severity and SMN protein level in spinal muscular atrophy. Nat

Genet 16, 265–269.

Ling KK, Lin MY, Zingg B, et al. (2010) Synaptic defects in the

spinal and neuromuscular circuitry in a mouse model of spinal

muscular atrophy. PLoS ONE 5, e15457.

Ling KK, Gibbs RM, Feng Z, et al. (2012) Severe neuromuscular

denervation of clinically relevant muscles in a mouse model of

spinal muscular atrophy. Hum Mol Genet 21, 185–195.

Liu Q, Fischer U, Wang F, et al. (1997) The spinal muscular atro-

phy disease gene product, SMN, and its associated protein

SIP1 are in a complex with spliceosomal snRNP proteins. Cell

90, 1013–1021.

Lorson CL, Rindt H, Shababi M (2010) Spinal muscular atrophy:

mechanisms and therapeutic strategies. Hum Mol Genet 19,

R111–118.

Martin AR (1955) A further study of the statistical composition

on the end-plate potential. J Physiol 130, 114–122.

Martinez TL, Kong L, Wang X, et al. (2012) Survival motor neu-

ron protein in motor neurons determines synaptic integrity in

spinal muscular atrophy. J Neurosci 32, 8703–8715.

McAndrew PE, Parsons DW, Simard LR, et al. (1997) Identifica-

tion of proximal spinal muscular atrophy carriers and patients

by analysis of SMNT and SMNC gene copy number. Am J Hum

Genet 60, 1411–1422.

Meister G, Buhler D, Pillai R, et al. (2001) A multiprotein com-

plex mediates the ATP-dependent assembly of spliceosomal U

snRNPs. Nat Cell Biol 3, 945–949.

Mentis GZ, Blivis D, Liu W, et al. (2011) Early functional impair-

ment of sensory-motor connectivity in a mouse model of

spinal muscular atrophy. Neuron 69, 453–467.

Michaud M, Arnoux T, Bielli S, et al. (2010) Neuromuscular

defects and breathing disorders in a new mouse model of

spinal muscular atrophy. Neurobiol Dis 38, 125–135.

Mitsumoto H, Floyd A, Tang MX, et al. (2006) Transcranial mag-

netic stimulation for upper motor neuron involvement in

amyotrophic lateral sclerosis (ALS). Suppl Clin Neurophysiol

59, 327–332.

Monani UR, Coovert DD, Burghes AH (2000a) Animal models of

spinal muscular atrophy. Hum Mol Genet 9, 2451–2457.

Monani UR, Sendtner M, Coovert DD, et al. (2000b) The

human centromeric survival motor neuron gene (SMN2) res-

cues embryonic lethality in Smn(�/�) mice and results in a

mouse with spinal muscular atrophy. Hum Mol Genet 9,

333–339.

Monani UR, Pastore MT, Gavrilina TO, et al. (2003) A transgene

carrying an A2G missense mutation in the SMN gene modu-

lates phenotypic severity in mice with severe (type I) spinal

muscular atrophy. J Cell Biol 160, 41–52.

Munsat TL, Davies KE (1992) International SMA consortium

meeting. 26–28 June 1992, Bonn, Germany. Neuromuscul

Disord 2, 423–428.

Murray LM, Comley LH, Thomson D, et al. (2008) Selective vul-

nerability of motor neurons and dissociation of pre- and post-

synaptic pathology at the neuromuscular junction in mouse

models of spinal muscular atrophy. Hum Mol Genet 17,

949–962.

Murray LM, Beauvais A, Bhanot K, et al. (2012) Defects in neu-

romuscular junction remodelling in the Smn(2B/–) mouse

model of spinal muscular atrophy. Neurobiol Dis 49C, 57–67.

Osborne M, Gomez D, Feng Z, et al. (2012) Characterization of

behavioral and neuromuscular junction phenotypes in a novel

allelic series of SMA mouse models. Hum Mol Genet 21,

4431–4447.

Park GH, Maeno-Hikichi Y, Awano T, et al. (2010) Reduced sur-

vival of motor neuron (SMN) protein in motor neuronal pro-

genitors functions cell autonomously to cause spinal muscular

atrophy in model mice expressing the human centromeric

(SMN2) gene. J Neurosci 30, 12005–12019.

Parsons DW, McAndrew PE, Iannaccone ST, et al. (1998) Intra-

genic telSMN mutations: frequency, distribution, evidence of

a founder effect, and modification of the spinal muscular

atrophy phenotype by cenSMN copy number. Am J Hum

Genet 63, 1712–1723.

Pellizzoni L, Yong J, Dreyfuss G (2002) Essential role for the

SMN complex in the specificity of snRNP assembly. Science

298, 1775–1779.

Porensky PN, Mitrpant C, McGovern VL, et al. (2012) A single

administration of morpholino antisense oligomer rescues

spinal muscular atrophy in mouse. Hum Mol Genet 21,

1625–1638.

Pun S, Santos AF, Saxena S, et al. (2006) Selective vulnerability

and pruning of phasic motoneuron axons in motoneuron

disease alleviated by CNTF. Nat Neurosci 9, 408–419.

Riessland M, Ackermann B, Forster A, et al. (2010) SAHA amelio-

rates the SMA phenotype in two mouse models for spinal

muscular atrophy. Hum Mol Genet 19, 1492–1506.

Ruiz R, Lin J, Forgie A, et al. (2005) Treatment with trkC agonist

antibodies delays disease progression in neuromuscular

degeneration (nmd) mice. Hum Mol Genet 14, 1825–1837.

Ruiz R, Casanas JJ, Torres-Benito L, et al. (2010) Altered intracel-

lular Ca2+ homeostasis in nerve terminals of severe spinal mus-

cular atrophy mice. J Neurosci 30, 849–857.

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares10

Page 11: Neurotransmitter release in motor nerve terminals of a mouse model of mild spinal muscular atrophy

Schrank B, Gotz R, Gunnersen JM, et al. (1997) Inactivation of

the survival motor neuron gene, a candidate gene for human

spinal muscular atrophy, leads to massive cell death in early

mouse embryos. Proc Natl Acad Sci U S A 94, 9920–9925.

Selvaraj BT, Frank N, Bender FL, et al. (2012) Local axonal func-

tion of STAT3 rescues axon degeneration in the pmn model of

motoneuron disease. J Cell Biol 199, 437–451.

Sendtner M (2010) Therapy development in spinal muscular

atrophy. Nat Neurosci 13, 795–799.

Simon CM, Jablonka S, Ruiz R, et al. (2010) Ciliary neurotrophic

factor-induced sprouting preserves motor function in a mouse

model of mild spinal muscular atrophy. Hum Mol Genet 19,

973–986.

Swoboda KJ, Prior TW, Scott CB, et al. (2005) Natural history of

denervation in SMA: relation to age, SMN2 copy number, and

function. Ann Neurol 57, 704–712.

Torres-Benito L, Neher MF, Cano R, et al. (2011) SMN require-

ment for synaptic vesicle, active zone and microtubule postna-

tal organization in motor nerve terminals. PLoS ONE 6, e26164.

Torres-Benito L, Ruiz R, Tabares L (2012) Synaptic defects in

SMA animal models. Dev Neurobiol 72, 126–133.

11

© 2013 Anatomical Society

Neurotransmission in mild SMA, R. Ruiz and L. Tabares


Recommended