Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 May 14.
Published in final edited form as: Brain Res. 2014 Oct 5;1607:47–74. doi: 10.1016/j.brainres.2014.09.064

A Fruitful Endeavor: Modeling ALS in the Fruit Fly

Ian Casci a,b, Udai Bhan Pandey a,*
PMCID: PMC4385417  NIHMSID: NIHMS633417  PMID: 25289585

Abstract

For over a century Drosophila melanogaster, commonly known as the fruit fly, has been instrumental in genetics research and disease modeling. In more recent years, it has been a powerful tool for modeling and studying neurodegenerative diseases, including the devastating and fatal amyotrophic lateral sclerosis (ALS). The success of this model organism in ALS research comes from the availability of tools to manipulate gene/protein expression in a number of desired cell-types, and the subsequent recapitulation of cellular and molecular phenotypic features of the disease. Several Drosophila models have now been developed for studying the roles of ALS-associated genes in disease pathogenesis that allowed us to understand the molecular pathways that lead to motor neuron degeneration in ALS patients. Our primary goal in this review is to highlight the lessons we have learned using Drosophila models pertaining to ALS research.

Keywords: Amyotrophic lateral sclerosis, ALS, Drosophila, Fly models, Animal models, Neurodegeneration, Motor neuron diseases, FTLD, SOD1, TDP-43, FUS, C9ORF72, VCP

1.0 Introduction

Amyotrophic lateral sclerosis (ALS), also called Lou Gehrig’s disease, is an adult onset, fatal, neurodegenerative disorder, for which there is currently no available cure. ALS was first described in the late 1860’s and early 1870’s by Jean-Martin Charcot (Charcot, 1874;Charcot and Joffroy, 1869), and a clear understanding of its pathogenesis at cellular and molecular levels is still eluding researchers today, nearly 150 years later. ALS is primarily characterized by the loss of upper and lower motor neurons, and as a result, death is usually due to respiratory failure (Gordon, 2013;Hardiman et al., 2011;Rowland, 1998). Clinically, ALS can be categorized into three primary types depending on the origin of onset (Hardiman et al., 2011). Approximately 65% of ALS patients are diagnosed with limb onset, 30% with bulbar and the remaining 5% with respiratory onset (Hardiman et al., 2011). In general, bulbar and respiratory have a worse prognosis than limb onset (Chio et al., 2009). Occurrence of the disease is highest between approximately 60 and 85 years of age (Alonso et al., 2009;Fang et al., 2009;Logroscino et al., 2010;O'Toole et al., 2008), with males typically having an overall higher rate of incidence than females (Alonso et al., 2009;Johnston et al., 2006). Disease progression is very rapid, and the average life expectancy post-diagnosis is 2-5 years (Byrne et al., 2012;Huisman et al., 2011;Lee et al., 2013;Logroscino et al., 2008;Mortara et al., 1984;Pugliatti et al., 2013).

Approximately 5% of all ALS cases show signs of inheritance and are referred to as familial ALS (FALS) (Byrne et al., 2011). The remaining 95% are not inherited, and are referred to as sporadic ALS (SALS). To date, several genes have been linked to both FALS and SALS, including SOD1 (Jones et al., 1994a;Jones et al., 1994b;Rosen et al., 1993), ALS2 (Hadano et al., 2001;Yang et al., 2001), VAPB (Nishimura et al., 2004a;Nishimura et al., 2004b), TARDBP (Arai et al., 2006;Mackenzie et al., 2007;Neumann et al., 2006), FUS (Belzil et al., 2009;Corrado et al., 2010;DeJesus-Hernandez et al., 2010;Kwiatkowski, Jr. et al., 2009;Vance et al., 2009), VCP (Johnson et al., 2010), OPTN (Del et al., 2011;Iida et al., 2012;Maruyama et al., 2010), SQSTM1 (Fecto et al., 2011), UBQLN2 (Deng et al., 2011), C9ORF72 (DeJesus-Hernandez et al., 2011;Gijselinck et al., 2012;Majounie et al., 2012;Renton et al., 2011), PFN1 (Wu et al., 2012a), hnRNPA1 and hnRNPA2B1 (Kim et al., 2013), and most recently MATR3 (Johnson et al., 2014), reviewed in (Leblond et al., 2014;Renton et al., 2014). In order to study these genes, researchers have developed animal models that recapitulate several of the cellular and molecular phenotypic features of ALS pathogenesis and neuronal cell death that are observed in human patients. Some of these features include the formation of cellular aggregates (Chou et al., 1996a;Chou et al., 1996b;Kato et al., 1997;Niwa et al., 2002;Shibata et al., 1996a;Shibata et al., 1994;Shibata et al., 1996b), mitochondrial abnormalities (Hirano et al., 1984a;Nakano et al., 1987;Swerdlow et al., 1998;Wiedemann et al., 2002), axonal abnormalities (Hirano et al., 1984a;Hirano et al., 1984b), dysfunctions in glutamate signaling and subsequent excitotoxicity (Kawahara et al., 2004;Rothstein et al., 1990;Takuma et al., 1999), and oxidative damage (Bowling et al., 1993).

2.0 History

It was more than 100 years ago that researchers began to realize the usefulness of Drosophila melanogaster in genetics research. Shortly after the turn of the 20th century, many key discoveries were made in Drosophila that helped shape our understanding of chromosome structure and heredity, reviewed in (Arias, 2008), (Bellen et al., 2010) and (Rubin and Lewis, 2000). From there it became instrumental for studying genetic mutations, and today researchers around the world use Drosophila as a model system for studying human diseases, including neurodegeneration. Many key discoveries that have helped in advancing the neurodegenerative disease field have been made using flies as a model system including, but not limited to, the identification of candidate genes as causes and therapies of disease (Fernandez-Funez et al., 2000;Steffan et al., 2001;Warrick et al., 1999), in vivo evidence for RNA-mediated neurodegeneration (Jin et al., 2003;Li et al., 2008), the identification of ER-golgi trafficking inhibition as a contributor to alpha-synuclein-associated toxicity in Parkinson’s disease (Cooper et al., 2006) and the identification of a molecular link between the ubiquitin-proteasome system and autophagy (Pandey et al., 2007).

Thomas Morgan, Alfred Sturtevant, Calvin Bridges and Hermann Muller were among the first pioneers of genetic research in Drosophila. They worked together in Morgan’s “Fly Room” at Columbia University (Sturtevant, 1965) where they studied heredity and clarified the relationships between genes, chromosomes and phenotypes, and ultimately won Morgan the 1933 Nobel Prize in Physiology or Medicine. In 1913, just a few short years after beginning his work with Drosophila, Sturtevant published the first ever genetic map (Sturtevant, 1913). A year later Bridges showed that genes and chromosomes were not separate entities, but that genes existed within chromosomes (a concept that was unclear at the time) (Bridges, 1914). Muller was instrumental in the development of balancer chromosomes that allowed researchers to stably and indefinitely maintain stocks of flies as heterozygotes for a desired mutation that is otherwise lethal to homozygotes (Muller, 1918). Several years later, Muller also discovered that gene mutations could be caused, as well as intentionally induced, by x-radiation (Muller, 1927). This finding won him the 1946 Nobel Prize in physiology or medicine.

With the work performed by Morgan and his colleagues, Drosophila was thrust to the forefront of genetics research. Researchers moved beyond using fruit flies to understand basic genetic concepts, and began identifying genes and understanding molecular pathways that are relevant to human development and functional processes. Though the importance of these discoveries cannot be overstated, some of the most important milestones in Drosophila genetics were those that led to its establishment as a model for studying human diseases. In 1982 Gerald M. Rubin and Allan C. Spradling were the first to use the P transposable element to insert exogenous DNA into the Drosophila genome (Spradling and Rubin, 1982). They used this method of creating transgenic flies to successfully integrate an exogenous gene and rescue a mutant phenotype (Rubin and Spradling, 1982). In 2000, the fully sequenced Drosophila genome was published (Adams et al., 2000;Myers et al., 2000). Since then, it has been estimated that as many as 77% of the human disease-associated genes reported in the Online Mendelian Inheritance in Man (OMIM) database have similar, likely orthologous, genes in Drosophila (Pandey and Nichols, 2011;Reiter et al., 2001). These discoveries laid the foundation upon which modern Drosophila models were created. Diseases are now being studied in an animal that is cheaper, faster, and easier to maintain than other models such as mice, rats and primates. The earliest Drosophila models of neurodegenerative diseases were developed in the 1990’s, and more are created every year (Bonini, 1999;Jackson et al., 1998;Lloyd and Taylor, 2010;Min and Benzer, 1997;Rincon-Limas et al., 2012;Warrick et al., 1998). In this review, we will highlight the Drosophila models that have been developed over the past five years for studying ALS.

3.0 Drosophila models of ALS

3.1 Superoxide dismutase (SOD1)

SOD1 is a protein that, under normal conditions, protects cells against oxidative stress by catalyzing the conversion of superoxide anions into oxygen and hydrogen peroxide (Fridovich, 1986). As it has been hypothesized that oxidative damage is linked to the aging process (HARMAN, 1956), SOD1 Drosophila models have been used to determine whether altering SOD1 expression levels can extend or shorten lifespans (Klichko et al., 1999;Parkes et al., 1998;Sun and Tower, 1999). Indeed, overexpression of SOD1 does appear to improve lifespans of Drosophila, although this effect appears to be tissue-specific, and does not occur in cells of the nervous system or in muscle cells (Martin et al., 2009). SOD1 may also play less of a role in the aging of organisms with naturally longer lifespans, as experiments in ants have shown lower SOD1 expression in queens than in male and worker ants, even though queens live much longer (Parker et al., 2004). In 1993, SOD1 became the first gene responsible for causing ALS (Rosen et al., 1993). We now know that SOD1 mutations account for approximately 1% of SALS cases and 12% of FALS cases (Renton et al., 2014). After the initial report in 1993, researchers were quick to utilize Drosophila to understand how SOD1 mutations lead to ALS (Table 1). By 1995, a fly model was developed to study the effects of mutant SOD1 protein subunits on wild type (WT) and mutant protein dimerization (Phillips et al., 1995). In more recent years a model was made by comparing transgenic flies overexpressing either Drosophila SOD1 (dSOD1), human SOD1 (hSOD1), or mutant hSOD1 (A4V or G85R) (Watson et al., 2008). Flies expressing WT hSOD1 or either mutant had a reduced climbing ability compared to dSOD1 flies, suggesting a motor neuron dysfunction due to expression of either WT or mutant versions of hSOD1. The differences in climbing did not become apparent until day 14, an observation that suggests a progressive loss of motor function in the WT and mutant hSOD1-expressing flies. In addition, these flies exhibited reduced synaptic transmission in dorsal longitudinal muscles of the giant fiber motor pathway (Watson et al., 2008). Consistent with SOD1-associated ALS pathogenesis in mouse models (Bruijn et al., 1997;Bruijn et al., 1998;Watanabe et al., 2001), the motor neurons developed aggregates of hSOD1 protein. Altogether, this model provided evidence that WT and ALS-causing mutant hSOD1 are toxic to motor neurons.

Table 1.

ALS Drosophila Models Developed Since 2008

Disease Phenotype Reference
SOD1 Expression of WT or mutant hSOD1 caused climbing defect, reduced synaptic transmission and
hSOD1 aggregation in motor neurons.
Watson et al. 2008
Ubiquitous expression of zinc-deficient SOD1 results in locomotor dysfunction, altered mitochondrial
structure & reduced ATP levels.
Bahadorani et al. 2013
Alsin Age dependent reduction in adult locomotion in flies homozygous for mutant dALS2 compared to
WT controls.
Takayama et al. 2014
VAPB dVAP-null mutant flies exhibit morphological defects at the NMJ, quantal content and size and the
number of presynaptic active zones.
Chai et al. 2008
WT dVAP overexpression exhibit morphological changes at the NMJ. dVAP P58S mutant protein
expression causes morphological changes at the NMJ and induced the formation of cytoplasmic
aggregates.
Ratnaparkhi et al. 2008
Only N-terminal-WT dVAP that includes an MSP domain is cleaved at the cell membrane and
secreted as a ligand for Eph receptors.
Tsuda et al. 2008
dVAP T48I mutant protein expression causes external eye degeneration and elevated Hsp70
expression in brain.
Chen et al. 2010
dVAP-null mutant flies exhibit morphological defects of muscle mitochondria. Han et al. 2012
Endogenous dVAP protein interacts with Down syndrome cell adhesion molecule (Dscam). Yang et al. 2012
dVAP P85S mutant protein sequesters Sac1 in cytoplasmic granules, increasing PtdIns4P levels,
causing degenerative phenotypes.
Forrest et al. 2013
Oxysterol binding protein (Osbp) interacts with WT dVAP. Osbp localizes to cytoplasmic aggregates
in dVAP P85S mutant-expressing cells, and localizes to golgi in WT dVAP-expressing cells.
Moustaqim-Barrette et al. 2014
dVAP overexpression caused eye degeneration, wing posture, pupal eclosion rates, protein
aggregation and defects at the NMJ.
Sanhueza et al. 2014
TDP-43 Reduced lifespan, adult locomotion, axonal branching at neuromuscular synapses and reduction in
the number of synaptic boutons upon reduced expression of endogenous dTDP-43.
Feiguin et al. 2009
Reduction of dendritic branching of larval sensory neurons upon reduced expression of WT hTDP-43
or WT dTDP-43. Increase of dendritic branching upon overexpression of dTDP-43.
Lu et al. 2009
Reduced HDAC6 mRNA levels in TBPH knockout flies. Fiesel et al. 2010
Axon loss and neuronal death with hTDP-43 overexpression and dTDP-43 knockdown in mushroom
bodies. Overexpressing hTDP-43 in motor neurons causes cytoplasmic accumulation of TDP-43
aggregates, cell death, neuron swelling and NMJ morphological defects.
Li et al. 2010
hTDP-43 overexpression in flies caused age-dependent eye degeneration, reduced lifespan and
adult wing malformations.
Hanson et al. 2010
Mild degeneration caused by dVCP is enhanced by coexpression with dTDP-43, and suppressed
with knockdown of dTDP-43. Coexpression of either dVCP R152H with WT hTDP-43 or WT dVCP
with hTDP-43 M337V causes enhanced eye degeneration.
Ritson et al. 2010
Flies exhibit suppressed ALS-associated phenotypes when expressed TDP-43 mutants lack RNA-
binding ability.
Voigt et al. 2010
dAtx2 overexpression enhances eye degeneration and decreases adult lifespan when coexpressed
with WT hTDP-43.
Elden et al. 2010
Larvae expressing dTDP-43 in neurons as well as larvae carrying dTDP-43 null mutants exhibit
changes at the NMJ. Overexpression in mushroom bodies leads to a decrease in lobe size, as well
as the formation of TDP-43-positive cytoplasmic aggregates.
Lin et al. 2011
Hyperphosphorylation of hTDP-43 reduces TDP-43 cytoplasmic aggregation in neurons.
Hypophosphorylation increases aggregation.
Li et al. 2010
hTDP-43 NLS-deficient mutants cause severe degeneration in adult eyes, lethality when expressed
in larval neurons, and reduced adult lifespan when expressed in adult neurons.
Miguel et al. 2011
Expressing hTDP-43 A315T mutant causes eye degeneration. Expressing either WT or A315T
mutant TDP-43 causes mislocalization and aggregate formation in axons of eye imaginal disks, with
WT TDP-43-expressing flies exhibiting more severe aggregation.
Estes et al. 2011
Upregulation of endogenous chaperones reduces TDP-43 and TDP-25-associted phenotypes,
including cellular aggregation.
Gregory et al. 2012
Larvae overexpressing dTDP-43 produce a different expression profile in Drosophila CNS than larvae
carrying dTDP-43-null mutants.
Hazelett et al. 2012
Expression of Drosophila Itp-r83A mutants in motor neurons causes a mild increase of hTDP-43
cytoplasmic mislocalization, as well as a partial rescue hTDP-43-associated lifespan and climbing
defects in adult flies.
Kim et al. 2012
Overexpressing human TDP-43 in adult motor neurons causes a reduction in lifespan that is not rescued by expression of the human caspase inhibitor, P35, or the Drosophila homolog, dIAP. Zhan et al. 2013
Pan neuronal overexpression and reduced expression of dTDP-43 both result in a reduction of adult
lifespan, as well as reduced adult and larval locomotion. Overexpression of dTDP-43 in third instar
larvae eye imaginal disks does not result in cytoplasmic mislocalization of the protein.
Diaper et al. 2013b
Reducing dTDP-43 protein levels in glial and muscle cells causes movement disorders in adult flies.
Overexpressing dTDP-43 resulted in early lethality and formation of sarcoplasmic aggregates.
Altering dTDP-43 protein levels in glia and neurons also alters dEAAT1 and dEAAT2 mRNA levels.
Diaper et al. 2013a
In glial cells, WT hTDP-43 as well as A315T, D169G, G298S and N345K mutants mislocalize to the
cytoplasm, reduce larval motor function, reduce bouton number at the NMJ, and increase the number
of postsynaptic glutamate receptors.
Estes et al. 2013
WT hTDP-43 and G298S and M337V mutants cause retinal degeneration in adult eyes. WT and
mutant proteins lacking a NLS undergo cytoplasmic mislocalization that corresponds to severe retinal
degeneration.
Ihara et al. 2013
Altering levels of stress-granule associated proteins, ROX8, PEK and GADD34, causes changes in
TDP-43 toxicity in neurons of adult flies. Inhibiting stress granule formation reduces toxicity.
Kim et al. 2014
FUS Expression of WT hFUS and mutant hFUS (R524S and P525L) causes eye degeneration in adult
flies. Neurons for both WT and mutant are enlarged, bouton number is reduced, and larval
locomotion is decreased.
Chen et al. 2011
Conditional expression of WT hFUS and mutant hFUS (R518K, R521C and R521H) in adult neurons
reduces lifespan and locomotion. hFUS R521C mutants have a reduced number of presynaptic
active zones at the NMJ.
Lanson, Jr. et al. 2011
Shahidullah et al. 2013
dFUS-null mutant causes reduced adult lifespan and locomotion, and reduced pupae eclosion when
expressed pan-neuronally. Coexpression with dFUS or hFUS rescues all phenotypes.
Wang et al. 2011
Cytoplasmic accumulation of insoluble, non-aggregated hFUS occurs in neurons over time, and
corresponds to neurodegenerative phenotypes. Reduction of insoluble FUS with the chaperone
protein reduces phenotypes.
Miguel et al. 2012
RNAi knockdown of dFUS in neurons causes reduced adult climbing, bouton number and synaptic
branches.
Sasayama et al. 2012
Expression of WT dFUS causes motor neuron (MN) apoptosis. dFUS-null mutants do not cause MN
apoptosis.
Xia et al. 2012
Perturbing the RNA-binding ability of WT and mutant FUS strongly suppresses toxicity, including eye
degeneration, larval eclosion and mobility and cytoplasmic mislocalization into stress granules.
Daigle et al. 2013
TAF15 Expression of WT human TAF15 in Drosophila eyes and neurons causes eye degeneration and
locomotor dysfunctions respectively.
Couthouis et al. 2011
EWSR1 WT and mutant (G511A and P552L) EWSR1 expression induce degeneration in adult eyes, reduced
life span and climbing ability.
Couthouis et al. 2012
C9orf72 30-repeat expansion of riboGGGGCC causes eye degeneration, reduced adult locomotion, and
colocalization of the RNA-binding protein, Pur α, with ubiquitin in cellular inclusions.
Xu et al. 2013
Flies expressing arginine-rich dipeptide-repeats exhibit eye degeneration, reduced eclosion rate, and
reduced adult lifespan.
Mizielinska et al. 2014
hnRNPA2 hnRNPA2 D290V mutant expression causes degeneration of indirect flight muscles and mislocalizes
to cytoplasmic aggregates. Removal of prion-like domain function rescues these phenotypes.
Kim et al. 2013

ALS is thought to be a non-cell autonomous neurodegenerative disease (Boillee et al., 2006). Therefore, researchers have also studied non-neuronal cells in animal models of ALS. The Drosophila model developed in 2008 was utilized a few years later by another group to examine how different cell-types expressing mutant and WT hSOD1 are affected by ALS-linked environmental insults (Islam et al., 2012). One advantage to using Drosophila as a model organism is the ability to direct expression of transgenes to a wide range of specific tissue types using the UAS-gal4 system (Brand and Perrimon, 1993;McGuire et al., 2004). Flies expressing hSOD1 specifically in glial cells showed a decrease in lifespan when exposed to the neurotoxin β-N-methylamino-L-alanine (BMAA) compared to dSOD1-expressing controls. Interestingly, however, flies expressing either mutant or WT hSOD1 specifically in motor neurons survived longer than controls when exposed to BMAA. This is an unexpected and intriguing result, considering the role of motor neurons in ALS pathogenesis. Conversely, expression of the hSOD1-G85R mutant in either motor neurons or glia reduced Drosophila lifespan compared to controls when exposed to a different environmental insult, hydrogen peroxide (Kumimoto et al., 2013). In the absence of induced environmental factors, expression of hSOD1-G85R in either motor neuron or glial cells led to a decrease in lifespan, but expression in both cell types simultaneously increased lifespan compared to controls. However, another study showed that expression of the FALS SOD1 mutant, hSOD1-G41S, in Drosophila motor neurons increases lifespan when compared to controls (Elia et al., 1999). Although the differences caused by the G85R and G41S mutants may be due to their locations in different domains of the protein (Turner and Talbot, 2008), these observations highlight the complexities of studying ALS in animal models, as well as the variability that is observed among different fly models.

Mitochondrial dysfunction is a pathological hallmark of ALS in humans (Sasaki and Iwata, 1999;Wiedemann et al., 2002;Wiedemann et al., 1998), and may be caused by altered protein activity of SOD1 mutants (Ferri et al., 2006;Liu et al., 2004;Pasinelli et al., 2004). ALS-associated mutations can hinder binding of SOD1protein to zinc ions, resulting in cellular toxicity (Crow et al., 1997;Lyons et al., 1996). Recently, researchers using transgenic Drosophila expressing zinc-deficient hSOD1 obtained evidence supporting a link between mutant SOD1 and mitochondrial dysfunction (Bahadorani et al., 2013). These flies harbored a D83S mutation in a zinc-binding domain of exon 4 that inhibits binding of SOD1 protein to zinc, without affecting binding to its other cofactor, copper. They found that ubiquitous expression of zinc-deficient hSOD1 led to age-dependent locomotor dysfunction that was not present in control flies. However, these flies exhibited normal lifespans and showed no brain or eye degeneration. When they examined the effects of zinc deficiency on mitochondria, they found that mitochondrial structure was altered in flight muscles of mutant flies. ATP production was also reduced in the fly heads compared to controls, but not in fly bodies, possibly suggesting a reduction of ATP production specifically in neurons. In addition, mutant hSOD1 D83S expression resulted in an increased sensitivity to paraquat and zinc exposure based on declines in lifespan of these flies (Bahadorani et al., 2013). Taken together, this model provides evidence that zinc deficiency due to hSOD1 mutations may result in ALS pathogenesis through reduced mitochondrial function.

3.2 Alsin

In 2001, mutations of the gene, ALS2, were identified in individuals with ALS, specifically ALS2 (Hadano et al., 2001;Yang et al., 2001). Alsin (also called ALS2), the protein encoded by ALS2, is a guanine nucleotide exchange factor for Ras-related in brain 5 (Rab5) (Otomo et al., 2003). Since Rab5 is a GTPase involved in the endocytic pathway (Bucci et al., 1992), it has been proposed that loss of Alsin function due to ALS-causing mutations perturb endosomal dynamics, thereby implicating these processes in the pathogenesis of the disease (Otomo et al., 2003).

A Drosophila model for Alsin-associated ALS was recently developed using the Drosophila homolog of Alsin, dALS2 (Table 1) (Takayama et al., 2014). Consistent with the disease state in ALS2 patients, where mutations of Alsin result in premature cessation of translation and subsequent production of truncated protein products (Yamanaka et al., 2003), the mutations used in the Drosophila model were dALS2 constructs lacking approximately 30% of their coding sequence. Mutant flies did not show significant phenotypic changes at the NMJ compared to WT controls, apart from a slight increase in the bouton number of abdominal motor axons of adult mutant flies. They did, however, show a significant age-dependent reduction in locomotion when compared to WT flies. This reduction was rescued by ubiquitous WT dALS2 overexpression, suggesting dALS2 loss-of-function as the cause of the phenotype. Interestingly, however, this rescue was not as significantly recapitulated by overexpressing WT dALS2 in motor neurons (Takayama et al., 2014). It is possible that differences in transgene expression levels between larvae and adult flies, depending on whether expression is ubiquitous or directed to motor neurons, might explain these results. The locomotor reduction observed with ubiquitous expression, however, may prove to be a useful output in future experiments aimed at identifying modifiers for the dALS2-related phenotype with this model (Takayama et al., 2014).

3.3 VAMP-associated protein B (VAPB)

VAMP (vesicle associated membrane protein)-associated protein B (VAPB), encoded by the gene, VAPB, localizes to both the endoplasmic reticulum (ER) and mitochondrial-associated membranes, where it is involved in the ER-associated unfolded protein response (Kanekura et al., 2006) and ER-mitochondria calcium exchange (De Vos et al., 2012) processes, respectively. A missense mutation in the coding region for the major sperm protein (MSP) domain of VAPB was first identified in human ALS patients in 2004 (Nishimura et al., 2004a;Nishimura et al., 2004b). Since then, Drosophila models have been used by researchers to better understand WT VAPB function in various tissue-types, as well as how different ALS-linked mutations in VAPB lead to toxicity and subsequent pathogenesis of the disease.

3.3.1 Overexpression and Knockdown

Several Drosophila models have been developed to elucidate the function of WT human VAPB (hVAPB) and its Drosophila homolog DVAP33-A (referred to as dVAP in this review), encoded by the gene, DVAP33-A (Chai et al., 2008;Forrest et al., 2013;Han et al., 2012;Moustaqim-Barrette et al., 2014;Ratnaparkhi et al., 2008;Tsuda et al., 2008;Yang et al., 2012). hVAPB and dVAP proteins are structurally and functionally similar, sharing approximately 62% homology, and one protein can compensate for the loss of the other in maintaining normal cellular processes (Chai et al., 2008). In addition, creating equivalent human disease-causing mutations in dVAP leads to neurodegenerative phenotypes that mimic those seen when hVAP with ALS-causing mutations are expressed in Drosophila (Chai et al., 2008;Chen et al., 2010;Forrest et al., 2013;Oyanagi et al., 2008;Ratnaparkhi et al., 2008;Sanhueza et al., 2014;Tsuda et al., 2008).

Models utilizing either homolog have shown that altering their expression levels in neurons results in functional and morphological changes of the neuromuscular junction (NMJ). Loss of dVAP function in dVAP-null mutant flies exhibit an increase in bouton size, but a decrease of overall bouton number at the NMJ, compared to normal flies. Exogenous expression of hVAPB in dVAP-null flies rescues this loss of boutons, providing evidence to support functional overlap between the two proteins (Chai et al., 2008). When dVAP is overexpressed in neurons, bouton size decreases, with a concurrent decrease in the number of vesicles, but the overall bouton number increases (Chai et al., 2008;Pennetta et al., 2002;Ratnaparkhi et al., 2008). These changes are also observed in neurons overexpressing hVAPB (Chai et al., 2008). Synaptic transmission is also affected by hVAPB/dVAP expression, suggesting a possible role of these proteins in maintaining synaptic functions. Loss of dVAP function in dVAP-null mutants reduces the quantal content at the NMJ by approximately 40%, compared to normal control flies. Conversely, overexpressing dVAP increases the quantal content, consistent with an increase in the bouton number, but the overall quantal size is reduced (Chai et al., 2008). It was concluded that quantal changes are a response to the morphological changes occurring at the NMJ in these fly lines, in an effort to keep the synaptic transmission at normal levels (Chai et al., 2008). Loss of dVAP function increases the number of presynaptic active zones at the NMJ (Chai et al., 2008). Different groups looking at the effects of dVAP overexpression have reported differences in the number of active zones, although these differences may be due to the different techniques used to visualize the NMJ’s. One group, utilizing electron microscopy, has reported no change in the number of active zones upon overexpression of dVAP (Chai et al., 2008), whereas a second group, utilizing confocal microscopy with antibodies specific to the active zone protein, Bruchpilot, reported a decrease in active zones (Ratnaparkhi et al., 2008).

Mitochondrial dysfunction has become a hallmark of ALS pathogenesis in humans (Sasaki and Iwata, 1999;Wiedemann et al., 2002;Wiedemann et al., 1998). Similarly to a recently developed SOD1 model (Bahadorani et al., 2013), dVAP-null mutant flies have morphological defects of their muscle mitochondria (Han et al., 2012), suggesting that mitochondrial dysfunction may be a cause of toxicity in VAPB-associated ALS.

Pan-neuronal overexpression of dVAP P58S in Drosophila (equivalent to the ALS-causing mutant, P56S, in humans) appears to be toxic, causing a decrease in larval mobility and an increase in larval central neuron cell death (Chai et al., 2008;Nishimura et al., 2004a;Nishimura et al., 2004b) . One possible mechanism for dVAP P58S toxicity is through its ability to induce morphological changes at the NMJ. Similarly to dVAP-null flies, P58S mutant expression in neurons increases bouton size at the NMJ. Not surprisingly, therefore, although WT dVAP is capable of rescuing the bouton phenotype in the dVAP-null background, dVAP P58S is not (Ratnaparkhi et al., 2008). dVAP P58S also hinders microtubule organization at the NMJ, although the mutant protein does not appear to have an effect on presynaptic active zones or the number of vesicles (Ratnaparkhi et al., 2008).

It has been proposed that dVAP P58S might exert toxicity through a dominant negative effect of sequestering normal proteins into cytoplasmic granules (Ratnaparkhi et al., 2008). dVAP P58S mutant protein forms ubiquitin-positive, cytoplasmic aggregates in larval neurons and muscle cells (Ratnaparkhi et al., 2008;Tsuda et al., 2008). Interestingly, WT dVAP is mislocalized from the cell membrane into these aggregates, suggesting that mutant dVAP is capable of recruiting WT protein, thereby hindering its normal activity in the cell (Chai et al., 2008;Ratnaparkhi et al., 2008;Tsuda et al., 2008). Under normal conditions, WT dVAP localizes to the cell membrane, where the N-terminal portion that includes the major sperm protein (MSP) domain is secreted as a ligand for Eph receptors (Tsuda et al., 2008). Secretion of the MSP domain is decreased in cells that are expressing dVAP P58S and are positive for cytoplasmic aggregates (Tsuda et al., 2008). In addition, evidence for a dominant negative effect can be seen on an organismal level. dVAP-null mutant pupae have a reduced eclosion rate, compared to normal controls. Expression of WT dVAP in these pupae rescues the eclosion rate. However, co-expression of WT dVAP with P58S mutant protein in the dVAP-null background does not rescue the phenotype, suggesting that mutant dVAP can suppress WT function, possibly through dimerization of WT and mutant proteins (Ratnaparkhi et al., 2008).

Another VAPB mutation was identified in FALS patients in 2010, VAPB T46I (Chen et al., 2010). A model was subsequently developed to characterize the Drosophila homolog of the VAPB mutant, dVAP T48I (Chen et al., 2010). Expressing dVAP T48I in adult fly eyes causes structural degeneration not present in control flies, confirming the deleterious nature of the mutation. When WT dVAP is expressed pan-neuronally, neurons of larval brain and nerve fibers show WT protein localized to cell and ER membranes. In contrast, dVAP T48I mutant protein localizes to bright aggregates. Interestingly, mutant dVAP expression in the brain and in muscle cells causes structural changes of the ER, including fragmentation, similarly to what has been observed in ALS patients (Oyanagi et al., 2008). The Hsp70 chaperone protein is also found to be elevated in these cells. Taken together, this model has provided evidence that expression of the dVAP T48I mutant leads to toxicity, and has thus provided insight into the nature of the VAPB T46I mutant in humans. Specifically, these results have identified stress response and morphological changes of the ER as possible modes of mutant protein toxicity in VAPB-related ALS (Chen et al., 2010).

A Drosophila model was recently developed to characterize a third VAPB mutation identified in 2012 in a single FALS patient with an additional C9ORF72 repeat expansion (Sanhueza et al., 2014;van et al., 2012). The human VAPB mutant protein, VAPB V234I, has an equivalent mutation in Drosophila, dVAP V260I. The experiments performed in this model focus on comparing overexpression of mutant dVAP with WT dVAP. dVAP V260I and WT dVAP overexpression exhibited similar phenotypic results, with the mutant dVAP producing more severe phenotypes (Sanhueza et al., 2014). Expression of both WT and dVAP V260I cause morphological changes at the NMJ including decreased bouton size with increased bouton number, as well as the formation of satellite boutons. The formation of these satellite boutons appears to be the result of alterations in microtubule architecture (Sanhueza et al., 2014). On a cellular level, both WT and mutant protein expression form aggregates in muscle cells. These cells also exhibit altered nuclear morphology including elongation, an increase in overall volume and clustering of nuclei with neighboring cells. Interestingly, the dVAP V260I mutant protein localizes to the muscle nuclei (Sanhueza et al., 2014). Altered nuclear morphology is also observed in neurons of larval brains, and dVAP V260I protein forms large cytoplasmic, not nuclear, inclusions in these cells. Similarly to dVAP T48I overexpression (Chen et al., 2010), WT and dVAP V260I expression in muscle cells induces a heat shock stress response, identified by an increase in Hsp70 expression and localization to nuclei (Sanhueza et al., 2014). Experiments focused on changes on the organismal level revealed reduced larval locomotion, pupal eclosion rates and altered adult wing posture when WT and dVAP V260I were neuronally expressed. Altered wing posture was also observed upon expression in muscle cells. When targeted to adult eyes, both WT and mutant protein caused structural degeneration (Sanhueza et al., 2014). It has been observed that both WT and mutant protein expression induced the altered phenotypes, with the dVAP V260I mutant producing more severe phenotypes, suggesting a possible gain of mutant protein function. The formation of aggregates in muscle cells and possible protein sequestration, however, may suggest a loss-of-function, so it was suggested that dVAP V260I toxicity may be due to both a loss and a gain-of-function (Sanhueza et al., 2014).

3.3.2 Modifiers of VAPB-Associated Phenotypes

Recently, proteins that interact with endogenous dVAP have been identified, including Down syndrome cell adhesion molecule (Dscam) (Yang et al., 2012), Suppressor of Actin 1 (Sac1) (Forrest et al., 2013) and Oxysterol binding protein (Osbp). Pan-neuronal, RNAi-mediated knockdown of the endogenous phosphoinositide phosphatase, Sac1, in Drosophila causes morphological changes at the NMJ, as well as structural degeneration of the adult eyes (Forrest et al., 2013). Interestingly, Sac1 and WT dVAP colocalize in larval brains, eye imaginal disks and salivary glands. When the ALS-associated dVAP P85S mutant is overexpressed in neuronal cells of these tissue-types, Sac1 is sequestered into mutant dVAP-induced cytoplasmic aggregates. Furthermore, RNAi-mediated knockdown of dVAP and Sac1in these tissues, as well as overexpression of dVAP P85S mutant protein, elevated expression of phosphatidylinositol 4-phosphate (PtdIns4P). Increased PtdIns4P expression results in degenerative phenotypes at the NMJ (Forrest et al., 2013). It has therefore been proposed that dVAP P85S mutant toxicity occurs through Sac1 interaction and sequestration into cellular aggregates, resulting in increased expression of PtdIns4P and subsequent morphological changes at the NMJ (Forrest et al., 2013). Taken together, these results identify Sac1 and PtdIns4P proteins as candidates for involvement in ALS pathogenesis, as well as possible therapeutic targets for developing effective therapies for ALS in the future.

The lipid binding protein, Osbp, interacts with WT dVAP, but not the dVAP P85S mutant (Moustaqim-Barrette et al., 2014). In normal cells, Osbp colocalizes with dVAP at the cell and ER membranes. In dVAP P85S mutant-expressing cells, however, Osbp mislocalizes to cytoplasmic aggregates, similarly to dVAP (Moustaqim-Barrette et al., 2014). Interestingly, in dVAP-null mutants, Osbp mislocalizes to the golgi. Since overexpression of human Osbp (OsbpL8) in ER is sufficient to rescue loss-of-function-associated phenotypes of dVAP-null flies, it has been proposed that the P85S mutation of dVAP incurs a loss-of-function to the protein in disease pathogenesis (Moustaqim-Barrette et al., 2014).

3.4 TAR DNA-binding protein 43 (TDP-43)

TDP-43 is a 43 kDa protein encoded by the gene TARDBP. It was first identified in 1995 as a transcriptional regulator that bound to pyrimidine-rich motifs of the human immunodeficiency virus 1 (HIV-1) regulatory element, TAR (Ou et al., 1995). TDP-43 is a primarily nuclear protein, and was first linked to ALS in 2006 when it was identified as a component of ubiquitinated, cytoplasmic granules in the neurons of patients with ALS and FTLD (frontotemporal lobar degeneration) (Neumann et al., 2006). Shortly thereafter, researchers began identifying several TARDBP mutations in FALS and SALS patients (Kabashi et al., 2008;Kuhnlein et al., 2008;Rutherford et al., 2008;Sreedharan et al., 2008;Van Deerlin et al., 2008). To date, a total of 47 missense mutations and 1 nonsense mutation in TARDBP have been identified (Lattante et al., 2013), some of which have been directly linked to cytoplasmic mislocalization of the TDP-43 protein (Barmada et al., 2010;Winton et al., 2008). We now know that TDP-43 is involved in regulating multiple aspects of RNA metabolism, including alternative splicing (Buratti et al., 2001), mRNA stability (Volkening et al., 2009) and likely mRNA transport (Alami et al., 2014;Wang et al., 2008a), thus implicating aberrations of these processes in disease pathogenesis.

Several Drosophila models for TDP-43-associated ALS have been generated and extensively characterized (Table 1). One of the major goals of researchers has been to determine if TDP-43-associated ALS pathogenesis is due to a gain or a loss of protein function. Drosophila models have provided evidence to support both hypotheses, and the truth may actually be a combination of the two.

3.4.1 Overexpression and Knockdown

TDP-43 is an RNA-binding protein that is highly conserved across organisms, including Drosophila. Researchers, therefore, commonly use Drosophila models to observe the effects of overexpression or knockdown of either human TDP-43 (hTDP-43) or the drosophila ortholog, dTDP-43, which is encoded by the gene TBPH (TAR DNA-binding protein-43 homolog). These gain-of function/loss-of-function experiments have been helpful for elucidating TDP-43 function as well as validating new models through recapitulation of key phenotypic features of ALS.

It has been demonstrated, with the use of both exogenous and endogenous WT TDP-43 in Drosophila, that maintaining normal TDP-43 protein levels in various tissue-types is critical for proper physiological functions of the organism. Pan-neuronal reduction of endogenous Drosophila TBPH levels, either through deletion mutants or through dTDP-43 RNAi, reduces adult fly lifespan and locomotion (typically walking and climbing) (Diaper et al., 2013b;Feiguin et al., 2009). Interestingly, co-expressing human TDP-43 rescues these phenotypes, suggesting a possible natural feedback mechanism for maintaining expression levels, that is conserved in these orthologs (Feiguin et al., 2009). Pan-neuronal overexpression of dTDP-43 has also shown a reduction in adult lifespan, as well as a reduction in both adult fly and larval locomotion (Diaper et al., 2013b). In sensory neurons of Drosophila larvae, lowered expression of dTDP-43 reduces dendritic branching, and overexpression increases branching (Lu et al., 2009). In mushroom bodies of the Drosophila brain, both overexpression of human TDP-43, as well as reduced Drosophila TDP-43 expression, causes loss of axons and neuronal death (Li et al., 2010). When dTDP-43 is overexpressed in this tissue-type, there is a reduction of lobe size, as well as the formation of TDP-43-positive cytoplasmic aggregates (Lin et al., 2011).

Since ALS affects motor neurons in humans, several experiments in Drosophila models are carried out on larval and adult fly motor neurons. When dTDP-43 levels are decreased specifically in larval motor neurons, dysmorphologies of the NMJ are observed, including reductions in axonal branching (Feiguin et al., 2009). There are, however, differences in reported bouton numbers at the NMJ between different groups. Some have reported a decrease in bouton number when endogenous dTDP-43 levels are reduced via deletion mutants in a W1118 fly strain (Feiguin et al., 2009). Others have reported an increase in bouton number when using a y1 w67c23 fly strain carrying a different deletion of dTDP-43 (Lin et al., 2011). It has been suggested that the conflicting observations may be due to the differences in genetic backgrounds of the flies used in the experiments, as well as the differences in the dTDP-43 deletion constructs (Lin et al., 2011). Expression of human TDP-43 in larval motor neurons resulted in reduced axonal branching and fewer synaptic boutons, in addition to axon swelling, motor neuron cell death and the formation of hTDP-43-positive cytoplasmic aggregates (Estes et al., 2011;Li et al., 2010). Some groups, however, have reported only small amounts of cytoplasmic hTDP-43, as well as little to no aggregation (Hanson et al., 2010;Voigt et al., 2010). Therefore, it is possible that a small amount of cytoplasmic TDP-43 is enough to cause toxicity in motor neurons, even without the presence of large aggregates. TDP-43-positive cytoplasmic aggregates, a hallmark of TDP-43-associated ALS in humans (Arai et al., 2006;Mackenzie et al., 2007;Neumann et al., 2006;Tan et al., 2007), have also been reported in adult Drosophila motor neurons overexpressing endogenous TDP-43 (Lin et al., 2011).

It has previously been shown that overexpressing human TDP-43 in adult fly motor neurons causes a significant reduction in lifespan (Hanson et al., 2010) compared to controls. Interestingly, overexpression of hTDP-43 with either the human caspase inhibitor, P35, or the Drosophila homolog, dIAP (Drosophila inhibitor of apoptosis), in adult motor neurons does not improve lifespan compared to flies expressing hTDP-43 alone (Zhan et al., 2013). This result suggests that human WT TDP-43 toxicity may not be due to neuronal apoptosis through programmed cell death mechanisms. However, when overexpressing endogenous Drosophila TDP-43 in pupae, caspase inhibition rescues CCAP/bursicon neuron apoptosis, suggesting that in these cells, at this stage of development and using endogenous protein, loss of neurons does occur through programmed cell death (Vanden Broeck et al., 2013).

Adult eyes are often used for transgene expression in Drosophila models of neurodegenerative diseases to produce easily observable eye phenotypes (Cushman-Nick et al., 2013;Hindle et al., 2013;Mohan et al., 2014;Yadav and Tapadia, 2013). Death of neurons in the eyes produces degenerative phenotypes that can be seen both inside and on the surface of the eye. These phenotypes can range from structural changes to loss of pigmentation. Expressing human TDP-43 in adult fly eyes causes age and dose-dependent structural degeneration (Elden et al., 2010;Estes et al., 2011;Hanson et al., 2010;Li et al., 2010;Ritson et al., 2010). This degeneration has also been directly correlated with mislocalization of hTDP-43 to the cytoplasm, and the mislocalization alone is sufficient to cause the degeneration (Ritson et al., 2010). In contrast, however, overexpression of endogenous Drosophila TDP-43 in eye imaginal disks of third instar larvae does not appear to cause cytoplasmic mislocalization, despite the presence of observed neurodegenerative phenotypes in adult flies (Diaper et al., 2013b).

Muscle atrophy and excitotoxicity from aberrant glutamate signaling are hallmarks of ALS pathogenesis in humans (Kawahara et al., 2004;Rothstein et al., 1990;Takuma et al., 1999). Recently, endogenous Drosophila TDP-43 expression has been identified in muscle and glial cells (Diaper et al., 2013a). Specifically knocking down dTDP-43 in adult muscle cells causes an overall reduction in motor function, based on an observed reduction in physical activity, compared to controls expressing normal protein levels. Overexpressing dTDP-43 in larval muscle cells results in protein mislocalization from nuclei into sarcoplasmic aggregates. These larvae also exhibit reduced locomotion and early lethality compared to controls. Targeted overexpression in glial cells is lethal to larvae. However, unlike in muscle cells, knockdown of protein levels in adult glial cells causes motor dysfunction that is only apparent in aged flies (Diaper et al., 2013a).

It has also been shown that overexpression and loss of function of dTDP-43 in glial cells and neurons alters the mRNA levels of dEAAT1 and dEAAT2, the Drosophila homologs of the glutamate transporters EAAT1 and EAAT2(Diaper et al., 2013a). EAAT2 mRNA has previously been identified as a possible target of TDP-43 (Sephton et al., 2011;Tollervey et al., 2011). This transporter is important for the clearance of glutamate released at the NMJ, and its dysfunction or loss has been linked to ALS through the identification of increased glutamate levels in the plasma and cerebral spinal fluid of ALS patients, potentially resulting in neuron excitotoxicity (Couratier et al., 1993;Plaitakis and Caroscio, 1987;Rothstein et al., 1992;Rothstein et al., 1995;Shaw et al., 1995;Spreux-Varoquaux et al., 2002;Trotti et al., 2001;Trotti et al., 1999). Consistent with the disease state in humans, dTDP-43 null mutant flies express lower overall levels of both dEAAT1 and dEAAT2 mRNA compared to controls. WT dTDP-43 decreases mRNA levels when overexpressed in glial cells, but increases mRNA levels when overexpressed in neurons (Diaper et al., 2013a).

Recent experiments using Drosophila models to study the effects of expressing human TDP-43 in glial cells have provided evidence to support its toxicity in this cell-type. The results have shown that WT hTDP-43, as well as A315T, D169G, G298S and N345K mutants, mislocalize to the cytoplasm when overexpressed (Estes et al., 2013). WT and mutant hTDP-43 overexpression in both glial and motor neuron cells also reduces bouton number at larval NMJ’s and hinders larval motor function, specifically their ability to turn over after being flipped upside down. When expressed in glial cells, WT and hTDP-43 mutants cause an increase in the number of postsynaptic glutamate receptors, but have no effect on the number of presynaptic active zones. In contrast, when expressed in motor neurons, WT and mutant proteins increase the number of active zones, but only WT hTDP-43 concurrently increases the number of glutamate receptors. The effect of these changes in both cell types is an imbalance in the ratio of active zones to glutamate receptors at the NMJ. This imbalance coincides with the motor dysfunction observed in the larvae, suggesting that this may be the mechanism that leads to toxicity of the overexpressed proteins. It was hypothesized that the increase in both active zone and glutamate receptor number, caused by overexpression of WT hTDP-43 in motor neurons, suggests a different mechanism of toxicity from the mutants in this cell type (Estes et al., 2013). However, the fact that WT and mutant TDP-43 exert similar effects in glial cells makes it difficult to identify the consequences of the ALS-linked mutations. These results suggest that the mutations themselves are not necessary for toxicity in glial cells. Taken together, however, these experiments provide further evidence linking glial and muscle cells to ALS, and support a role of glutamate transport dysfunction in disease pathogenesis.

In addition to studying overexpression and knockdown of WT TDP-43, Drosophila models have been used to study ALS-associated TDP-43 mutants. Interestingly, in many of these experiments, mutant and WT toxicity is either not compared, or cannot be distinguished from each other due to having comparable effects. As a result, it is difficult to determine what ALS-associated mutations actually do to normal protein function, and how large of a role they play, if any, in the actual pathogenesis of the disease. Overexpression of human TDP-43 M337V in Drosophila eyes causes mild degeneration and shows both full-length TDP-43, as well as a 25kD fragment, similar to what has been identified in FTLD patient samples (Neumann et al., 2006;Ritson et al., 2010). Dose and age-dependent adult eye degeneration is also seen in flies expressing D169G, G298S and N345K mutants, when compared to non-hTDP-43 expressing control flies (Estes et al., 2013). This phenotype is accompanied by axonal aggregation of these mutant proteins in eye neuroepithelium, providing further evidence for a link between the aberrant mislocalization and aggregation of hTDP-43, and neurodegeneration (Estes et al., 2013). The Drosophila TDP-43 Q367X mutant reduces miR-9a expression in larvae compared to WT TDP-43-expressing flies, and subsequently results in improper specification of sensory organ precursor cells(Li et al., 2013). This suggests that dTDP-43 controls specification through miR-9a, and that a loss of function through an ALS-associated mutation causes dysregulation of the process (Li et al., 2013).

Pan-neuronal expression of A315T, G348C, A382T, G287S, or N390D reduces adult lifespan compared to control flies expressing driver alone, the extent of which varies between different mutants (Voigt et al., 2010). However, these results do not directly implicate the mutations themselves in toxicity, as WT hTDP-43 overexpression caused the largest reduction in lifespan compared to the mutant and control flies. Interestingly, RNA-binding-deficient hTDP-43 only mildly reduced adult lifespan, essentially rescuing the WT hTDP-43 phenotype (Voigt et al., 2010). These results were supported by later experiments performed in adult eyes (Ihara et al., 2013). Overexpression of WT hTDP-43 caused retinal degeneration that was rescued by inhibiting RNA-binding function through both removal and mutation of the RNA recognition motif 1 (RRM1). Degeneration caused by G298S and M337V mutants was also rescued by inhibiting the RNA-binding ability of these proteins. Removal of the nuclear localization signal (NLS) in WT hTDP-43 and G298S and M337V mutants resulted in “forced” cytoplasmic mislocalization when expressed in adult eyes that corresponded to increased degeneration compared to WT hTDP-43 controls with intact NLS’s. When the RNA-binding ability of these proteins was removed, cytoplasmic mislocalization remained, but the degeneration was rescued (Ihara et al., 2013). Taken together, the experiments performed by these groups provide evidence that the RNA-binding function of TDP-43 may be necessary for toxicity, as mislocalization alone does not appear to be sufficient for retinal degeneration.

Another group has also used human TDP-43 lacking a functional NLS to further study the effects of cytoplasmic mislocalization in disease pathogenesis (Miguel et al., 2011). The NLS-mutant hTDP-43 in these experiments causes severe degeneration in adult eyes, as well as reduced adult lifespan when expressed in neurons, and lethality when expressed in larval neurons (Miguel et al., 2011). Interestingly, increased death of adult flies begins prior to the formation of cytoplasmic aggregates, lending support to the conclusions of others that aggregation of TDP-43 is not necessary for toxicity (Hanson et al., 2010;Miguel et al., 2011). In contrast, however, studies of FTLD show a link between cellular toxicity and aggregates that are positive for either full-length TDP-43 or the disease-linked, 25kD, C-terminal fragment, TDP-25 (Igaz et al., 2009;Igaz et al., 2008;Neumann et al., 2006;Zhang et al., 2009). In fact, it was recently shown in a Drosophila model that upregulating endogenous chaperone proteins in various tissue types can reduce TDP-43 and TDP-25 aggregation, a result that coincides with rescue of their ALS-associated phenotypes in the flies (Gregory et al., 2012). Furthermore, altering expression levels of the stress granule-associated PEK, ROX8 and GADD34 proteins in adult neurons affects TDP-43-associated toxicity (Kim et al., 2014). Specifically, knocking down PEK and ROX8, two proteins involved in stress granule formation and the Drosophila homologs of mammalian PERK and TIA1, respectively, (Brand and Bourbon, 1993;Farny et al., 2009) inhibits stress granule formation and subsequently suppresses TDP-43 toxicity. Knocking down GADD34, a phosphatase that prevents the formation of stress granules by dephosphorylating the critical stress granule component, eIF2α, (Novoa et al., 2001) effectively promotes stress granule formation and subsequently enhances TDP-43 toxicity (Kim et al., 2014). Therefore, due to the discrepancies in the data obtained by these different groups, it remains unclear as to whether or not cytoplasmic mislocalization and aggregation of TDP-43, a phenotype linked to ALS in humans, is a primary or secondary cause of toxicity.

Drosophila and human TDP-43 overexpression and knockdown flies were utilized to analyze global gene expression changes in various tissue types and stages of Drosophila development (Hazelett et al., 2012;Vanden Broeck et al., 2013;Zhan et al., 2013). These studies have specifically led to the identification of genes, including Hey, Nup50 (Zhan et al., 2013) and Map205 (Vanden Broeck et al., 2013), as well as pathways such as Notch (Zhan et al., 2013), Wnt and BMP (Hazelett et al., 2012), that may be linked to TDP-43-associated ALS pathogenesis. In a more general sense, they have shed light on the role and importance of TDP-43 in development, and the need for an organism to maintain normal TDP-43 protein levels for physiological functions. Changes in Drosophila TDP-43 expression levels in the larval central nervous system (CNS) result in largely non-overlapping gene expression patterns. Specifically, dTDP-43 overexpression leads to downregulation of one set of genes, whereas larvae carrying loss-of-function mutations in dTDP-43 exhibit an upregulation of a different set of genes (Hazelett et al., 2012). However, when dTDP-43 levels are altered ubiquitously at the late pupal stage, there is a significant overlap between upregulated and downregulated genes (Vanden Broeck et al., 2013). Taken together, the results of these studies provide evidence that TDP-43 plays a large role in gene expression that may change throughout Drosophila development. Interestingly, a ubiquitous reduction of dTDP-43 in the late pupal stage causes increased expression of the endogenous microtubule associated protein, Map205. Aberrant expression of this protein, in turn, appears to induce cytoplasmic mislocalization of the ecdysteroid receptor (EcR), which may result in a failure to transition from pupal to adult gene expression patterns and subsequent neuronal death (Vanden Broeck et al., 2013). Overall, these Drosophila models have identified candidate genes for future study, and suggest that the pathogenesis of TDP-43-associated ALS is a very complex process, involving many genes and biological processes.

It is still an enigma if WT or mutant TDP-43 toxicity is due to a gain or a loss of function. Observing WT toxicity can be helpful to identify how perturbations of normal TDP-43 levels affect disease pathogenesis. Unfortunately, it is challenging to assess the effects of ALS-associated TDP-43 mutants when WT protein is itself toxic. WT toxicity in and of itself suggests a gain of function. However, loss of function models have also been developed (Hazelett et al., 2012). As mentioned previously, flies overexpressing WT Drosophila TDP-43 produce different expression profiles in the CNS during larval development than flies carrying dTDP-43-null mutations (Hazelett et al., 2012). Therefore, it is possible that both a gain and a loss of TDP-43 function are toxic, but for different reasons, and through different mechanisms. Taken together, however, Drosophila models have shown that WT and ALS-associated mutant TDP-43 expression level changes, cellular mislocalization, and protein functions such as RNA-binding are all factors that may play a role in ALS pathogenesis.

3.4.2 Modifiers of TDP-43-Associated Phenotypes

Drosophila has been a useful model organism for performing genetic screens to identify genes and pathways that modify phenotypes associated with neurodegenerative diseases, including spinocerebellar ataxia (Fernandez-Funez et al., 2000;Ren et al., 2011;Shieh and Bonini, 2011), Wolfram syndrome (Jones et al., 2014), Alzheimer’s disease (Moran et al., 2013;Shulman and Feany, 2003) and Huntington’s disease (Kazemi-Esfarjani and Benzer, 2000). Screens can be targeted, testing the effects of an array of known genes on a particular phenotype (Jones et al., 2014;Ren et al., 2011), or they can be unbiased, looking for any genes that modify a phenotype. Unbiased screens are typically performed using either deficiency lines that carry heterozygous deletions of mapped regions of the chromosome (Li et al., 2014;Simonsen et al., 2007), or lines with random enhancer-promoter (EP)-transposable element insertions (Fernandez-Funez et al., 2000;Shulman and Feany, 2003). Drosophila at any stage of development can be used, depending on the phenotype of interest (Batlevi et al., 2010;Rorth et al., 1998). Eye degeneration is a commonly used phenotypic output (Kazemi-Esfarjani and Benzer, 2000;Moran et al., 2013), although genetic screens have also been performed using survival (Li et al., 2014), wing posture (Shieh and Bonini, 2011), climbing (Shieh and Bonini, 2011) defects and axonal and synaptic degeneration (Wishart et al., 2012). Here we highlight the genes that have been identified as modifiers of TDP-43-associated neurodegeneration in Drosophila models of ALS.

Histone deacetylase 6 (HDAC6) is a human class II histone deacetylase that has previously been linked to neurodegenerative diseases (Grozinger et al., 1999;Pandey et al., 2007). Unlike other histone deacetylases, HDAC6 functions primarily in the cytoplasm where it plays a critical role in the clearance of cytoplasmic, ubiquitinated, misfolded proteins (Hubbert et al., 2002;Kawaguchi et al., 2003;Verdel et al., 2000). Homozygous knockout of TBPH in 1st-instar larval brains, as well as heterozygous knockout in adults, shows a reduction in HDAC6 mRNA levels (Fiesel et al., 2010). These results suggest that altering TDP-43 levels in neurons also affects HDAC6 levels, and may therefore result in toxicity by inhibiting the removal of accumulated cytoplasmic proteins in patients with ALS (Fiesel et al., 2010). Another possible mechanism of toxicity is through aberrant synaptic transmission. Bruchpilot (BRP), a protein that is part of the presynaptic dense body and regulates synaptic transmission by gathering synaptic vesicles together for release, was recently identified as a target of HDAC6 deacetylation in the neurons of adult Drosophila (Miskiewicz et al., 2014). Increasing HDAC6 expression reduces BRP acetylation, thereby increasing its binding to vesicles and creating a larger vesicle pool for release. Consistent with the identification of HDAC6 as a target of TDP-43 (Fiesel et al., 2010), this phenotype is recapitulated when human TDP-43 is overexpressed, with a larger effect observed in flies harboring ALS-associated TDP-43-mutations (A315T and A382T) (Miskiewicz et al., 2014). Conversely, TBPH-null mutant flies show an increase in BRP acetylation and subsequently smaller vesicle pools. Following these observations, it was found that reducing HDAC6 levels in the A315T mutant flies rescues the acetylation levels. In addition, reducing HDAC6 expression also rescues motor dysfunction observed in A315T mutant adult flies (Miskiewicz et al., 2014). Taken together, these results suggest that TDP-43-associated ALS pathogenesis may be due to aberrant synaptic transmission, resulting from increased HDAC6 expression and subsequently reduced BRP acetylation.

Ubiquilin is a cytosolic protein capable of binding both poly-ubiquitinated chains and 19s proteosome subunits. It is therefore believed to play a role in the degradation of misfolded proteins (Ko et al., 2004). A Drosophila model was developed in 2010 to show the effect of ubiquilin overexpression on TDP-43 associated neurodegeneration. Consistent with other models, overexpression of hTDP-43 in motor neurons reduces the lifespans of adult flies, reduces pupal eclosion rates and causes wing malformations. Co-expression with ubiquilin enhances these phenotypes. Interestingly, however, human TDP-43 in cells co-expressing ubiquilin is almost exclusively nuclear, and is not found in cytoplasmic aggregates, despite the overall enhanced degeneration in these flies. In addition to linking ubiquilin to TDP-associated pathology in a model of ALS, these results provide more evidence that aggregation of mislocalized hTDP-43 is not necessary for progression of the disease (Hanson et al., 2010).

Valosin-containing protein (VCP) is an ATPase that has been linked to the pathogenesis of ALS (Johnson et al., 2010), as well as other neurodegenerative diseases including inclusion body myopathy associated with Paget disease of bone and frontotemporal dementia (IBMPFD) (Higashiyama et al., 2002;Watts et al., 2004). The Drosophila homolog of VCP is encoded by the gene, ter94. One mutation of VCP that causes IBMPFD, R155H, has an equivalent mutation in Drosophila, dVCP R152H (Ritson et al., 2010). A drosophila model for TDP-43-associated ALS has revealed that coexpression of dVCP R152H and WT human TDP-43 in fly eyes enhances the structural degeneration and cytoplasmic mislocalization that is observed when WT hTDP-43 is overexpressed alone. This enhancement is also produced when WT Drosophila VCP is expressed with the ALS-associated TDP-43 mutant, hTDP-43 M337V. Expression of dVCP alone causes mild degeneration that is suppressed when Drosophila TDP-43 is knocked down, and enhanced when dTDP-43 is overexpressed (Ritson et al., 2010). Taken together, these results identify VCP is a modifier of TDP-43 pathogenesis, and provide in vivo evidence linking a single mutation to the pathogenesis of both ALS and IBMPFD.

CAG repeat expansion within the coding region of ataxin-2, ATXN2, causes spinocerebellar ataxia type 2 (SCA2) (Imbert et al., 1996;Lorenzetti et al., 1997;Pulst et al., 1996;Sanpei et al., 1996). The Drosophila homolog of ataxin-2, dAtx2, enhances eye degeneration and decreases adult lifespan when coexpressed with WT human TDP-43 in eyes and motor neurons, respectively (Elden et al., 2010). When endogenous dAtx2 levels are reduced, eye degeneration and lifespan improve in flies expressing hTDP-43. It is important to note that altering dAtx2 in these experiments does not alter hTDP-43 levels. Parallel work in cell culture has shown that ataxin-2 and TDP-43 physically interact in an RNA-dependent manner (Elden et al., 2010). Thus, ataxin-2 appears to be a modifier of ALS-related TDP-43 toxicity. As with IBMPFD, these experiments have identified a possible link between ALS and SCA2.

Inositol-1,4,5-trisphosphate receptor type 1 (ITPR1), encoded by the gene ITPR1, is an IP3 receptor bound to the endoplasmic reticulum, and is involved in controlling the release of intracellular calcium ions (Nucifora, Jr. et al., 1995). A recent high content RNAi screen in HeLa cells identified ITPR1 as a modifier of TDP-43 cytoplasmic mislocalization (Kim et al., 2012). Specifically, knockdown of ITPR1 increased TDP-43 cytoplasmic mislocalization. These results were confirmed in Drosophila using the ortholog of ITPR1, Itp-r83A. Expression of Itp-r83A mutants in motor neurons resulted in a very mild increase in cytoplasmic mislocalization of human TDP-43. These mutants also partially rescued adult climbing and lifespan phenotypes associated with hTDP-43 overexpression toxicity. It was thus proposed that TDP-43 overexpression toxicity is due to high nuclear concentrations. The toxic phenotypes are rescued when nuclear concentrations are reduced through the reduction of calcium signaling and subsequent cytoplasmic localization and degradation of TDP-43 (Kim et al., 2012).

Cacophony, a voltage-gated calcium channel encoded by the gene, cac, was recently identified as a potential modifier of Drosophila TDP-43 toxicity (Hazelett et al., 2012). Cacophony calcium channels have been identified at presynaptic active zones of the NMJ (Kawasaki et al., 2004) of Drosophila, and play an important role in regulating function, development and plasticity of neurons through calcium signaling (Peng and Wu, 2007). More recent analyses in a Drosophila loss of function model of TDP-43-associated ALS have shown that flies homozygous for Drosophila TDP-43-null mutations exhibit lower CAC expression levels than control flies (Chang et al., 2013). Both pan-neuronal and motor-neuron-specific overexpression of CAC in these mutant flies rescues the crawling defects that have been previously reported (Chang et al., 2013;Hazelett et al., 2012). In addition, loss of dTDP-43 reduces CAC levels at the NMJ. Furthermore, in normal flies, WT dTDP-43 associates with CAC mRNA transcripts, and loss of dTDP-43 function in dTDP-43-null flies results in mild alternative splicing of CAC mRNA (Chang et al., 2013). Considering the role of CAC at the NMJ, and previous reports linking abnormalities of the NMJ to ALS (Estes et al., 2013;Feiguin et al., 2009), these results provide evidence that CAC dysfunction may be involved in ALS pathogenesis and is a potential therapeutic target.

Recently it has been shown that expression of hTDP-43 can suppress the neurodegenerative phenotypes of a Drosophila model of Fragile X-associated tremor/ataxia syndrome (FXTAS) in an RNA-dependent manner. Interestingly, it was found that ALS-causing mutations of TDP-43 enhanced the phenotypes associated with FXTAS. In addition, phenotypic suppression by TDP-43 appears to be through an interaction with Heterogeneous nuclear ribonucleoprotein A2B1 (hnRNPA2B1), suggesting a link between these two RNA-binding proteins and FXTAS (He et al., 2014).

3.4.3 Posttranslational Modifications

Hyperphosphorylated TDP-43 in cytoplasmic aggregates was reported in neurons of patients with TDP-43-associated ALS (Neumann et al., 2006). It has been proposed that aberrant phosphorylation of TDP-43 is associated with progression of the disease (Arai et al., 2006). Phosphomimetic TDP-43 shows a diffuse pattern in neurons of Drosophila. In contrast, phospho-deficient mutants form aggregates in perikarya and dendrites (Li et al., 2011). These results suggest that hyperphosphorylation prevents aggregation of TDP-43, which does not appear to coincide with what is observed in ALS patients. Recently Doubletime (DBT), the Drosophila homolog of casein kinase I〈, was identified as a specific TDP-43 kinase that phosphorylates Ser409/410 of WT human TDP-43 as well as Q331K and M337V mutants (Choksi et al., 2014). Coexpression of DBT with TDP-43 Q331K in adult fly eyes produces severe retinal degeneration that does not exist when DBT is coexpressed with either WT TDP-43 or the M337V mutant. In addition, only the Q331K mutants form aggregates in neurons when phosphorylated by DBT (Choksi et al., 2014). It is unclear as to why toxicity and aggregate formation, following phosphorylation by DBT, is specific to TDP-43 Q331K mutant proteins in these experiments. However, in contrast to the earlier phosphorylation study (Li et al., 2011), the effects that DBT phosphorylation has on the Q331K mutant more accurately reflect the aggregation and toxicity observed in ALS patients (Hasegawa et al., 2008;Inukai et al., 2008).

3.5 Fused in sarcoma/translocated in liposarcoma (FUS/TLS)

FUS/TLS, often referred to as FUS, is a DNA/RNA binding protein (Baechtold et al., 1999;Zinszner et al., 1997). FUS regulates gene transcription (Tan and Manley, 2010;Uranishi et al., 2001;Wang et al., 2008b) and, similarly to TDP-43, plays a role in regulating different aspects of RNA metabolism, including nucleocytoplasmic RNA transport (Zinszner et al., 1997) and alternative splicing (Lagier-Tourenne et al., 2012;Rogelj et al., 2012;Yang et al., 1998). As its name suggests, FUS was originally linked to cancer (Crozat et al., 1993;Rabbitts et al., 1993). In more recent years, however, mutations of FUS have been identified in patients with neurodegenerative diseases including FALS (Kwiatkowski, Jr. et al., 2009;Vance et al., 2009), SALS (Belzil et al., 2009;Corrado et al., 2010;DeJesus-Hernandez et al., 2010) and FTLD (Neumann et al., 2009). The first Drosophila models for FUS-associated ALS were developed in 2011 (Chen et al., 2011;Lanson, Jr. et al., 2011;Wang et al., 2011) (Table 1). Similarly to TDP-43 and SOD1, these models, and those that followed (Table 1), have shown that expression of WT or mutant human FUS (hFUS) in various Drosophila tissue-types and stages of development recapitulates key pathological features of ALS.

3.5.1 Overexpression and Knockdown

Given their functional similarities and mutual link to ALS, it is logical to suspect that FUS and TDP-43 have overlapping mechanisms underlying their toxicity in neurons. FUS overexpression and knockdown in various tissue types has shown similar effects to TDP-43 experiments, and has provided some interesting new insights that may be applicable to TDP-43 pathogenesis as well. Similarly to TDP-43, expression of WT human FUS and ALS-associated mutants (R518K, R521C, R521H, R524S and P525L) in Drosophila eyes leads to degeneration that increases with age and expression level (Chen et al., 2011;Lanson, Jr. et al., 2011;Xia et al., 2012). We and others have also shown evidence that the degeneration is greater in hFUS mutant-expressing flies than in WT-expressing flies (Chen et al., 2011;Lanson, Jr. et al., 2011). Removal of the nuclear export signal (NES) in mutant constructs, thereby containing their localization to nuclei, rescues eye degeneration phenotypes (Lanson, Jr. et al., 2011). These results suggest that FUS mutations may confer either a toxic gain of cytoplasmic function, or a loss of nuclear function to the protein. In contrast, when targeted to mushroom bodies, both WT and mutant hFUS (R524S and P525L) cause decreased lobe sizes (Chen et al., 2011). Pan-neuronal expression of WT hFUS and R518K, R521C and R521H mutants in adult flies leads to reduced locomotion compared to controls, with mutant locomotion lower than WT flies. In addition, WT and R521H-expressing flies have a shorter lifespan than controls, with mutant flies having shorter lifespans than WT flies (Lanson, Jr. et al., 2011).

As ALS is primarily a motor neuron disease, researchers in the last few years have focused heavily on studying this cell type in Drosophila. It was recently shown that overexpression of hFUS WT and mutant protein causes presynaptic structural changes at the NMJ of Drosophila larvae (Shahidullah et al., 2013). Consistent with previous reports (Lanson, Jr. et al., 2011), bouton number is unaffected. However, contrary to mutant hTDP-43 expression in motor neurons (Estes et al., 2013), hFUS R521C-expressing larvae do have a reduced number of presynaptic active zones, as well as dysmorphology of the structures that form the active zones (Shahidullah et al., 2013). These structural abnormalities are accompanied by reduced synaptic transmission in the NMJ’s, a phenotype that is also observed in Drosophila FUS-null mutants. Despite these abnormalities, excitability of the neurons and propagation of the action potentials are unaffected in these mutant flies. Therefore, presynaptic abnormalities at the NMJ of motor neurons may in fact be involved in the initial stages of ALS pathogenesis (Shahidullah et al., 2013).

When expressed in larval motor neurons, WT hFUS cellular distribution is primarily nuclear, whereas ALS-associated mutant hFUS mislocalizes to the cytoplasm (Chen et al., 2011;Lanson, Jr. et al., 2011). These results are consistent with the disease state in ALS-patients (Kwiatkowski, Jr. et al., 2009;Vance et al., 2009). There is also a reduction of larval mobility, and an enlargement of the neurons themselves in both WT and mutant hFUS-expressing larvae (Chen et al., 2011;Lanson, Jr. et al., 2011;Xia et al., 2012). Interestingly, however, it appears that not all ALS-associated hFUS mutants produce the same morphological changes at the NMJ. R521G, R524S and P525L mutants cause a reduction in bouton number, but R518K, R521C and R521H mutants do not (Chen et al., 2011;Lanson, Jr. et al., 2011;Xia et al., 2012).

Over time, there is an accumulation of insoluble, non-aggregated WT hFUS when expressed in adult neurons that coincides with neurodegenerative phenotypes (Miguel et al., 2012). Reducing insoluble hFUS levels with coexpression of the chaperone protein, HSPA1L, reduces eye degeneration and improves lifespan. WT hFUS is primarily localized in the nucleus of these cells. These results suggest that the toxicity observed from WT hFUS overexpression is the result of nuclear, not cytoplasmic function (Miguel et al., 2012), consistent with the recently proposed cause of TDP-43-associated toxicity (Kim et al., 2012).

The Drosophila homolog of FUS is cabeza, encoded by the gene caz. The two proteins share approximately 53% identity (Stolow and Haynes, 1995). dFUS-null mutant flies show reduced pupae eclosion, adult lifespan, and adult locomotion compared to controls (Wang et al., 2011). Interestingly, expressing WT dFUS, WT human FUS, or ALS-associated human FUS mutants (R522G or P525L) rescues the eclosion phenotype. Adult lifespan and locomotion is rescued by WT dFUS and hFUS, but not by the human FUS mutants (Wang et al., 2011). Pan-neuronal expression of dFUS RNAi does not affect adult lifespan, but does inhibit climbing. RNAi knockdown in motor neurons also reduces bouton number and length of synaptic branches at the NMJ (Sasayama et al., 2012). Additionally, knockdown in fly eyes causes external degeneration (Azuma et al., 2014;Shimamura et al., 2014). As with TDP-43, it is still unknown whether FUS toxicity is due to a gain of function or a loss of function. As mentioned previously, some early experiments with human FUS mutants suggest a possible toxic gain of function (Chen et al., 2011;Lanson, Jr. et al., 2011). However, these studies using dFUS show that a loss of function recapitulates ALS-associated phenotypes similarly to overexpression models. Interestingly, both WT dFUS overexpression and null-mutant loss of function are capable of causing similar phenotypes in Drosophila, with the exception of motor neuron apoptosis (Xia et al., 2012). Overexpression of WT dFUS causes apoptosis of motor neurons, but dFUS-null mutants do not. The conclusion of these results is that although perturbations of dFUS homeostasis are toxic to flies, the mechanisms of toxicity may not always be the same (Xia et al., 2012). Taken together, it is possible that, similarly to TDP-43, FUS toxicity is due to a combination of both a loss of nuclear function and gain of cytoplasmic function.

We have recently shown evidence that the RNA-binding ability of FUS is necessary for ALS-associated toxicity in both Drosophila and yeast model systems (Daigle et al., 2013). We expressed WT hFUS and hFUS mutant proteins lacking RNA-binding capabilities in Drosophila and found that ALS-associated phenotypes were reduced compared to normal WT and mutant hFUS-expressing flies. Removing RNA-binding function resulted in reduced eye degeneration in adults, improved larval eclosion and mobility, reduced brain atrophy in larvae, reduced cytoplasmic mislocalization, and reduced incorporation into cytoplasmic stress granules following treatment with sodium arsenite (Daigle et al., 2013). These results are consistent with the previously discussed results showing that RNA-binding may also be necessary for TDP-43-associated toxicity in ALS pathogenesis (Ihara et al., 2013;Voigt et al., 2010).

3.5.2 Modifiers of FUS-Associated Phenotypes

In addition to similar pathological phenotypes in ALS, there is evidence that FUS and TDP-43 have similar and potentially overlapping mechanisms of toxicity (Kim et al., 2010;Ling et al., 2010). Our laboratory has previously shown that WT and mutant hTDP-43 modifies the hFUS-associated ALS phenotypes observed in Drosophila models. Both WT and ALS-associated mutant hTDP-43 (M337V) enhance WT and hFUS R521H mutant degeneration in fly eyes (Lanson, Jr. et al., 2011). Others have shown that Drosophila TDP-43 modifies the dFUS phenotype, and the results suggest that TBPH and caz function in a similar pathway (Wang et al., 2011).

In addition to modifying TDP-43-associated ALS phenotypes in Drosophila, VCP has recently been shown to modify FUS-associated ALS phenotypes as well (Azuma et al., 2014). RNAi mediated knockdown of dFUS in adult fly eyes causes external degeneration that can be partially rescued by coexpression with WT dVCP. Conversely, overexpressing dFUS in dVCP-null mutant flies enhances the degeneration. dVCP also affects dFUS levels in neurons. Reduction of nuclear levels of dFUS via pan-neuronal, RNAi-mediated knockdown in larvae is enhanced in dVCP-null mutant flies compared to knockdown in normal flies. Coexpression with WT dVCP, however, increases nuclear dFUS levels compared to dFUS knockdown alone. Adult flies expressing dFUS RNAi in neurons exhibit reduced climbing capabilities that gets worse with age. This climbing dysfunction is enhanced in dVCP-null mutants, but is mildly rescued by overexpressing WT dVCP. Consistent with these results, dVCP also modifies dFUS toxicity at NMJ’s (Azuma et al., 2014). Larvae expressing dFUS RNAi in neurons show a reduction in presynaptic terminal branching and number of boutons. Expressing dFUS RNAi in a dVCP-null mutant fly enhances these reductions. Both phenotypes are rescued by coexpression with WT dVCP, with bouton numbers even surpassing control larvae (Azuma et al., 2014). Not only do these results identify dVCP as a modifier of ALS-associated dFUS phenotypes, they provide further evidence, in addition to TDP-43 studies (Ritson et al., 2010), that VCP may be involved in ALS pathogenesis and may be a viable target for therapy.

Recent evidence suggests that dFUS is involved in the epidermal growth factor receptor (EGFR) signaling pathway through the EGFR-associated genes rhomboid-1, rhomboid-3 and mirror (Shimamura et al., 2014). Reduced expression of any one of these genes in adult eyes is sufficient to suppress the eye degeneration observed in dFUS knockdown flies. These results are consistent with previous reports linking perturbations of the mitogen-activated protein kinase (MAPK) signaling pathway to ALS (Ackerley et al., 2004;Bendotti et al., 2004;Bendotti et al., 2005;Dewil et al., 2007;Holasek et al., 2005;Kim and Choi, 2010;Shimamura et al., 2014;Tortarolo et al., 2003).

3.5.3 Posttranslational Modifications

Protein arginine methyltransferases (PRMT’s) catalyze the methylation of arginine residues of proteins. This is important for regulation and maintenance of protein function, reviewed in (Bedford and Richard, 2005;Fackelmayer, 2005). PRMT1 methylates FUS, regulating its function and cellular distribution (Dormann et al., 2012;Du et al., 2011;Tradewell et al., 2012;Yamaguchi and Kitajo, 2012). Using a Drosophila model previously developed in our laboratory (Lanson, Jr. et al., 2011) in combination with mammalian cell cultures, it has been recently shown that PRMT1 interacts with WT hFUS, as well as the ALS-associated hFUS mutant, R521H (Scaramuzzino et al., 2013). Targeted coexpression of PRMT1 RNAi with either WT hFUS or hFUS R521H in Drosophila eyes enhances the degeneration that occurs with WT or mutant hFUS alone. The results suggest that reduced arginine methylation of FUS by PRMT1 may be associated with ALS pathogenesis (Scaramuzzino et al., 2013).

3.6 TAF15 and EWSR1

The structural and functional similarities between TDP-43 and FUS, including their shared RNA-binding abilities, have led to the hypothesis that RNA metabolism plays a critical role in the pathogenesis of ALS (Lagier-Tourenne and Cleveland, 2009). For this reason, a yeast functional screen was recently performed on 133 human RNA-binding proteins, in an effort to identify ALS-causing candidates that exhibit cytoplasmic aggregation and toxicity similarly to TDP-43 and FUS (Couthouis et al., 2011). Of the proteins examined, TAF15 (RNA polymerase II, TATA box binding protein (TBP)-associated factor, 68kDa) and EWSR1 (EWS RNA-binding protein 1) not only exhibited aggregation and toxicity, but are also in the same protein family as FUS (Bertolotti et al., 1996;Couthouis et al., 2012;Couthouis et al., 2011). Therefore, TAF15 and EWSR1 were selected as candidates for further study.

TAF15 and EWSR1 are RNA-binding proteins that function in transcriptional regulation (Bertolotti et al., 1996;Lee et al., 2004;May et al., 1993) and alternative splicing (Ibrahim et al., 2013;Paronetto et al., 2011). Traditionally, TAF15 and EWSR1 have been associated with cancer through the formation of fusion genes that result in aberrant transcription (Law et al., 2006). Both genes are linked to extraskeletal mixoid chondrosarcoma (Attwooll et al., 1999;Labelle et al., 1995) and acute leukemia (Martini et al., 2002). Following their identification in the recent yeast functional screen, TAF15 and EWSR1 were found to harbor mutations in patients with SALS (Couthouis et al., 2012;Couthouis et al., 2011). Drosophila models using human WT and mutant forms of each protein were subsequently developed (Table 1) (Couthouis et al., 2012;Couthouis et al., 2011). In the TAF15 model, overexpression of WT protein caused eye degeneration and an age-dependent decrease in climbing ability, when expression was targeted to the eyes and motor neurons, respectively, compared to normal control flies. Overexpressing WT TAF15, as well as two different ALS-linked mutants, R408C and G391E, in motor neurons reduces lifespan compared to normal controls, with the mutants exhibiting shorter lifespans than the WT-expressing flies (Couthouis et al., 2011).

In the EWSR1 model pan-neuronal expression of WT protein reduces lifespan of the flies, and causes an age-dependent reduction of climbing ability compared to normal controls (Couthouis et al., 2012). Directed expression of human WT EWSR1 to the adult eyes caused dose-dependent structural degeneration. Interestingly, two different ALS-linked mutants, G511A and P552L, produced relatively equivalent levels of degeneration. This group suggests that observing similar effects of both mutant and WT protein expression in the Drosophila eyes may represent a limitation of the model, or may be evidence that mutant EWSR1 is not the cause of pathogenicity in ALS patients (Couthouis et al., 2012). Further experimentation with this model will be necessary to determine the reason for this observation. Taken together, these models have provided in vivo data confirming the function of both TAF15 and EWSR1, and further implicate RNA metabolism in disease pathogenesis.

3.7 C9ORF72

One of the more recently discovered genes to be linked to ALS is C9orf72. It was first identified in ALS patients in 2011, and its toxicity is due to a GGGGCC hexanucleotide repeat expansion within the first intron of the gene (DeJesus-Hernandez et al., 2011;Gijselinck et al., 2012;Renton et al., 2011). A recent study has shown this hexanucleotide repeat expansion of C9orf72 present in ALS patients in populations all over the world. The combined data showed an overall frequency of approximately 6.3% of SALS cases and 37.6% of FALS cases, making it the most common known genetic cause of SALS worldwide (Majounie et al., 2012). The first Drosophila model for C9orf72-associated ALS was developed in 2013 (Table 1) (Xu et al., 2013). This model shows that expression of 30 riboGGGGCC-repeat constructs recapitulates ALS-associated phenotypes when compared to control flies expressing 3 riboGGGGCC repeats. Expression of 30-repeat constructs in Drosophila eyes causes severe degeneration that increases with age and is not seen in control flies. When targeted to motor neurons, the GGGGCC expansion also reduces adult locomotion over time compared to controls that show no effect at all. The RNA-binding protein, Pur α, was concurrently found to bind to the expanded riboGGGGCC repeats in mammalian cell culture experiments performed by this group (Xu et al., 2013). Co-expression of Pur α with expanded riboGGGGCC in Drosophila eyes suppressed the degeneration phenotype. In addition, Pur α colocalizes with ubiquitin in cellular inclusions in the flies expressing the expanded riboGGGGCC repeats, but not in controls. Taken together, these results not only add further in vivo evidence to support the role of GGGGCC repeats in the pathogenesis of C9orf72-associated ALS, but they also provide evidence to support the hypothesis that sequestration of RNA-binding proteins is a mechanism for disease development and progression (Xu et al., 2013).

There is a growing body of evidence that GGGGCC repeat toxicity may also be due to the production of dipeptide-repeat (DPR) proteins through repeat-associated non-ATG (RAN) translation (Ash et al., 2013;Baborie et al., 2014;Kwon et al., 2014;Mizielinska et al., 2014;Mori et al., 2013;Zhang et al., 2014;Zu et al., 2011;Zu et al., 2013). Through this process, both the sense and antisence GGGGCC repeat sequences can be translated into five different DPR proteins (proline:arginine, glycine:arginine, proline:alanine, glycine:alanine and glycine:proline/proline:glycine) that vary based on the reading frame of the transcripts (Kwon et al., 2014). A Drosophila model was recently developed in which adult flies expressing proline:arginine and glycine:arginine DPR proteins in the eyes exhibited eye degeneration that was not present in proline:alanine or glycine:alanine-expressing flies (Table 1) (Mizielinska et al., 2014). Survival from egg to adult stages was also decreased in the flies expressing the arginine-rich DPR proteins. Adult lifespans are reduced when the arginine-rich DPR proteins are expressed pan-neuronally, compared to the proline:alanine and glycine:alanine-expressing flies (Mizielinska et al., 2014). Consistent with the previous Drosophila model (Xu et al., 2013), toxicity is also observed in flies expressing expanded GGGGCC-repeat RNA, compared to controls (Mizielinska et al., 2014). These results suggest that both DPR proteins and expanded GGGGCC-repeat RNA are sufficient for cellular toxicity.

3.8 Heterogeneous nuclear ribonucleoprotein A1 and A2B1 (hnRNPA1 and hnRNPA2B1)

hnRNPA1 and hnRNPA2B1 were first linked to ALS in 2013 (Kim et al., 2013). hnRNP’s are nuclear, RNA-binding proteins involved in mRNA processing (Hutchison et al., 2002). hnRNPA1 and hnRNPA2B1 have been shown previously to bind directly to TDP-43 for mRNA splicing (Buratti et al., 2005). Mutations within the prion-like domains of hnRNPA1 and hnRNPA2B1 were found in familial cases of multisystem proteinopathy and ALS (Kim et al., 2013). A newly developed model (Table 1) has shown that one of the identified hnRNPA2 mutants, D290V, causes severe degeneration when expressed in indirect flight muscle cells of Drosophila (Kim et al., 2013). This degeneration is rescued by removing a portion of the prion-like domain. In addition, D290V mutants form cytoplasmic aggregates, whereas WT hnRNPA2, as well as mutant proteins lacking prion-like domain function, are localized to nuclei. These results, in combination with yeast and cell culture data, add hnRNPA1 and hnRNPA2B1 to the list of ALS-causing genes, and suggest that toxicity is specifically due to the activity of the prion-like domains of the proteins (Kim et al., 2013). These results also imply that other proteins with prion-like domains may be involved in ALS pathogenesis.

4.0 Translation from Flies to Mammalian Models

Drosophila melanogaster is a very useful organism for modeling human disease. As an animal model, Drosophila has an advantage over cell culture models of disease, in that the data obtained is more relevant to a living organism. Drosophila are less expensive to maintain than other animal models such as mice, rats and non-human primates. They also have shorter lifespans than these larger organisms, allowing for faster matings and production of experimental progeny, as well as shorter run-times on lifespan experiments. For these reasons, fruit flies are a fast, inexpensive tool for the initial studies of a disease. However, on a physiological level, Drosophila models may not be the best model of choice per se, as the mammalian nervous system is much more complicated than the fly nervous system. Therefore, it is critical that the studies performed in flies are translated and verified in mammals.

Most of the mammalian models of SOD1-associated ALS are mice and rats, although dog models have also been developed (Joyce et al., 2011). ALS-linked SOD1 mutations have been more thoroughly modeled in rodents than in Drosophila. More than 10 of these mutant genes are expressed in mouse models compared to two in Drosophila, reviewed in (Joyce et al., 2011). In both flies and mice, expressing human SOD1 A4V and G85R mutant proteins results in the development of neurodegenerative phenotypes that are reminiscent of those seen in ALS patients, including the formation of SOD1-positive inclusions (Bruijn et al., 1997;Deng et al., 2006;Watson et al., 2008). Interestingly, however, SOD1 A4V protein-positive inclusions are only formed in mice when WT human SOD1 is also expressed, an observation not seen in flies (Deng et al., 2006;Watson et al., 2008). SOD1 G85R-expressing mice also show an acceleration of SOD1-positive inclusion formation when WT human SOD1 is concurrently expressed (Wang et al., 2009). Although both Drosophila and mammalian models will continue to be useful in studying SOD1-associated ALS, the more extensive set of mouse models available will likely make them the most useful to researchers in the near future.

Several mammalian model systems have been generated for TDP-43-associated ALS since TDP-43 has been linked to neurodegenerative diseases. Interestingly, these systems reflect the same gain of function/loss of function dilemma that exists in Drosophila models (Ihara et al., 2013). Overexpression of WT and ALS-associated mutant human TDP-43 causes loss of neurons in mice (Igaz et al., 2011;Wegorzewska et al., 2009;Wils et al., 2010), suggesting a gain of protein function as the cause of the disease. However, loss of endogenous TDP-43 expression in mouse spinal motor neurons causes accumulation of ubiquitinated, cytoplasmic proteins in these cells, as well as cell death, suggesting a loss of protein function as the cause of disease pathogenesis (Wu et al., 2012b). Furthermore, ALS-like neurodegenerative phenotypes are observed in rats overexpressing ALS-linked mutant human TDP-43, but not WT protein (Zhou et al., 2010). Although seemingly contradictory, these mammalian models, as with the Drosophila models of TDP-43 associated ALS, provide evidence that both a loss and gain of protein function are toxic, although possibly through different mechanisms.

The observation that WT and ALS-linked mutant human FUS are toxic in Drosophila (Chen et al., 2011;Daigle et al., 2013;Lanson, Jr. et al., 2011) has been confirmed in rats (Huang et al., 2012;Huang et al., 2011) and in mice (Qiu et al., 2014). The toxicity also appears to be greater in the mutant FUS-expressing rats than in the WT-expressing animals (Huang et al., 2011). These results are also consistent with work done in ALS-patient skin fibroblasts harboring mutations in the 3’ untranslated region of FUS, in which these mutations caused overexpression of WT hFUS and are likely the cause of ALS in these patients (Sabatelli et al., 2013).

GGGGCC hexanucleotide repeat expansion of C9ORF72 was found to be toxic in Drosophila and mammalian cell models, simultaneously (Xu et al., 2013). Expression of C9ORF72 with as few as 30 repeats was sufficient to cause ALS-associated phenotypes in the flies, and cell death of neuro-2a cells, a mouse-derived neuronal cell line (Xu et al., 2013), and suggested that whatever we learned from a fly model of C9ORF72 could be translated to a mammalian system.

5.0 Summary and Future Directions

Drosophila melanogaster has been established as a useful model organism for studying ALS through its ability to recapitulate phenotypic features of the disease. This, in addition to a fully sequenced genome, has made Drosophila models well-suited for unbiased genetic screens that identify modifiers of ALS-associated phenotypes. These screens helped in discovering candidate genes that may be involved in mediating the disease process as well as in modifying neurodegenerative phenotypes. The ability to manipulate the expression of genes with and without ALS-causing mutations in specific tissue types, and at different stages of development, has also been useful for understanding the roles of those genes and tissues in disease progression. In spite of its usefulness in ALS research, however, Drosophila melanogaster is a non-mammalian organism and therefore has its limitations as a physiological model system of human disease. It is critical that the knowledge gained from Drosophila continues to be translated into mammalian models, followed by validation in human patient samples. The overexpression/knockdown paradigm is overused in Drosophila, and it is possible that expressing high levels of human protein in flies may not be completely relevant to the disease process in humans. Translating the knowledge gained from flies into mammalian and human model systems will turn these insights into a usable understanding of ALS that will lead to effective treatments, and eventually cures.

Highlights.

  • Over a century Drosophila melanogaster has been a powerful model organism for genetics research

  • Drosophila models have been used in ALS research for understanding the molecular basis of disease

  • Identifying modifiers of ALS by genetic and drug screening would help in understanding the disease

Acknowledgments

This work was supported by the Robert Packard Center for ALS at Johns Hopkins, the Amyotrophic Lateral Sclerosis Association and the National Institute of Health Grant #R01-NS081303 to UBP.

Abbreviations

ALS

Amyotrophic Lateral Sclerosis

OMIM

Online Mendelian Inheritance in Man

SOD1

Superoxide dismutase

FALS

Familial ALS

SALS

Sporadic ALS

WT

Wild type

dSOD1

Drosophila SOD1

hSOD1

Human SOD1

BMAA

β-N-methylamino-L-alanine

Rab5

Ras-related in brain 5

VAPB

VAMP (vesicle-associated membrane protein)-associated protein B

ER

Endoplasmic reticulum

MSP

Major sperm protein

hVAPB

Human VAPB

dVAP

Drosophila VAPB (DVAP33-A)

NMJ

Neuromuscular junction

Dscam

Down syndrome cell adhesion molecule

PtdIns4P

Phosphatidylinositol 4-phosphate

Osbp

Oxysterol binding protein

TDP-43

TAR DNA-binding protein 43

dTDP-43

Drosophila TDP-43

hTDP-43

Human TDP-43

HIV-1

Human immunodeficiency virus 1

FTLD

Frontotemporal lobar degeneration

TBPH

TAR DNA-binding protein-43 homolog

dIAP

Drosophila inhibitor of apoptosis

EP

Enhancer promoter

HDAC6

Histone deacetylase 6

BRP

Bruchpilot

VCP

Valosin-containing protein

IBMPFD

Inclusion body myopathy associated with Paget disese of bone and frontotemporal dementia

SCA2

Spinocerebellar ataxia type 2

RRM

RNA recognition motif

NLS

Nuclear localization signal

FUS/TLS

Fused in sarcoma/translocated in liposarcoma

NES

Nuclear export signal

EGFR

Epidermal growth factor receptor

MAPK

Mitogen-activated protein kinase

PRMT

Protein arginine methyltransferase

CNS

Central nervous system

ITPR1

inositol-1,4,5-trisphosphate receptor type 1

FXTAS

Fragile X-associated tremor/ataxia syndrome

TAF15

RNA polymerase II, TATA box binding protein (TBP)-associated factor, 68kDa

EWSR1

EWS RNA-binding protein 1

RAN

Repeat-associated non-ATG

DPR

Dipeptide-repeat protein

hnRNP

Heterogeneous nuclear ribonucleoproteins

EcR

Ecdysteroid receptor

RRM1

RNA recognition motif 1

CAC

Cacophony

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Ackerley S, Grierson AJ, Banner S, Perkinton MS, Brownlees J, Byers HL, Ward M, Thornhill P, Hussain K, Waby JS, Anderton BH, Cooper JD, Dingwall C, Leigh PN, Shaw CE, Miller CC. P38alpha Stress-Activated Protein Kinase Phosphorylates Neurofilaments and Is Associated With Neurofilament Pathology in Amyotrophic Lateral Sclerosis. Mol Cell Neurosci. 2004;26:354–364. doi: 10.1016/j.mcn.2004.02.009. [DOI] [PubMed] [Google Scholar]
  2. Adams MD, Celniker SE, Holt RA, Evans CA, Gocayne JD, Amanatides PG, Scherer SE, Li PW, Hoskins RA, Galle RF, George RA, Lewis SE, Richards S, Ashburner M, Henderson SN, Sutton GG, Wortman JR, Yandell MD, Zhang Q, Chen LX, Brandon RC, Rogers YH, Blazej RG, Champe M, Pfeiffer BD, Wan KH, Doyle C, Baxter EG, Helt G, Nelson CR, Gabor GL, Abril JF, Agbayani A, An HJ, Andrews-Pfannkoch C, Baldwin D, Ballew RM, Basu A, Baxendale J, Bayraktaroglu L, Beasley EM, Beeson KY, Benos PV, Berman BP, Bhandari D, Bolshakov S, Borkova D, Botchan MR, Bouck J, Brokstein P, Brottier P, Burtis KC, Busam DA, Butler H, Cadieu E, Center A, Chandra I, Cherry JM, Cawley S, Dahlke C, Davenport LB, Davies P, de PB, Delcher A, Deng Z, Mays AD, Dew I, Dietz SM, Dodson K, Doup LE, Downes M, Dugan-Rocha S, Dunkov BC, Dunn P, Durbin KJ, Evangelista CC, Ferraz C, Ferriera S, Fleischmann W, Fosler C, Gabrielian AE, Garg NS, Gelbart WM, Glasser K, Glodek A, Gong F, Gorrell JH, Gu Z, Guan P, Harris M, Harris NL, Harvey D, Heiman TJ, Hernandez JR, Houck J, Hostin D, Houston KA, Howland TJ, Wei MH, Ibegwam C, Jalali M, Kalush F, Karpen GH, Ke Z, Kennison JA, Ketchum KA, Kimmel BE, Kodira CD, Kraft C, Kravitz S, Kulp D, Lai Z, Lasko P, Lei Y, Levitsky AA, Li J, Li Z, Liang Y, Lin X, Liu X, Mattei B, McIntosh TC, McLeod MP, McPherson D, Merkulov G, Milshina NV, Mobarry C, Morris J, Moshrefi A, Mount SM, Moy M, Murphy B, Murphy L, Muzny DM, Nelson DL, Nelson DR, Nelson KA, Nixon K, Nusskern DR, Pacleb JM, Palazzolo M, Pittman GS, Pan S, Pollard J, Puri V, Reese MG, Reinert K, Remington K, Saunders RD, Scheeler F, Shen H, Shue BC, Siden-Kiamos I, Simpson M, Skupski MP, Smith T, Spier E, Spradling AC, Stapleton M, Strong R, Sun E, Svirskas R, Tector C, Turner R, Venter E, Wang AH, Wang X, Wang ZY, Wassarman DA, Weinstock GM, Weissenbach J, Williams SM, Woodage T, Worley KC, Wu D, Yang S, Yao QA, Ye J, Yeh RF, Zaveri JS, Zhan M, Zhang G, Zhao Q, Zheng L, Zheng XH, Zhong FN, Zhong W, Zhou X, Zhu S, Zhu X, Smith HO, Gibbs RA, Myers EW, Rubin GM, Venter JC. The Genome Sequence of Drosophila Melanogaster. Science. 2000;287:2185–2195. doi: 10.1126/science.287.5461.2185. [DOI] [PubMed] [Google Scholar]
  3. Alami NH, Smith RB, Carrasco MA, Williams LA, Winborn CS, Han SS, Kiskinis E, Winborn B, Freibaum BD, Kanagaraj A, Clare AJ, Badders NM, Bilican B, Chaum E, Chandran S, Shaw CE, Eggan KC, Maniatis T, Taylor JP. Axonal Transport of TDP-43 MRNA Granules Is Impaired by ALS-Causing Mutations. Neuron. 2014;81:536–543. doi: 10.1016/j.neuron.2013.12.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Alonso A, Logroscino G, Jick SS, Hernan MA. Incidence and Lifetime Risk of Motor Neuron Disease in the United Kingdom: a Population-Based Study. Eur J Neurol. 2009;16:745–751. doi: 10.1111/j.1468-1331.2009.02586.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Arai T, Hasegawa M, Akiyama H, Ikeda K, Nonaka T, Mori H, Mann D, Tsuchiya K, Yoshida M, Hashizume Y, Oda T. TDP-43 Is a Component of Ubiquitin-Positive Tau-Negative Inclusions in Frontotemporal Lobar Degeneration and Amyotrophic Lateral Sclerosis. Biochem Biophys Res Commun. 2006;351:602–611. doi: 10.1016/j.bbrc.2006.10.093. [DOI] [PubMed] [Google Scholar]
  6. Arias AM. Drosophila Melanogaster and the Development of Biology in the 20th Century. Methods Mol Biol. 2008;420:1–25. doi: 10.1007/978-1-59745-583-1_1. [DOI] [PubMed] [Google Scholar]
  7. Ash PE, Bieniek KF, Gendron TF, Caulfield T, Lin WL, DeJesus-Hernandez M, van Blitterswijk MM, Jansen-West K, Paul JW, III, Rademakers R, Boylan KB, Dickson DW, Petrucelli L. Unconventional Translation of C9ORF72 GGGGCC Expansion Generates Insoluble Polypeptides Specific to C9FTD/ALS. Neuron. 2013;77:639–646. doi: 10.1016/j.neuron.2013.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Attwooll C, Tariq M, Harris M, Coyne JD, Telford N, Varley JM. Identification of a Novel Fusion Gene Involving HTAFII68 and CHN From a T(9;17)(Q22;Q11.2) Translocation in an Extraskeletal Myxoid Chondrosarcoma. Oncogene. 1999;18:7599–7601. doi: 10.1038/sj.onc.1203156. [DOI] [PubMed] [Google Scholar]
  9. Azuma Y, Tokuda T, Shimamura M, Kyotani A, Sasayama H, Yoshida T, Mizuta I, Mizuno T, Nakagawa M, Fujikake N, Ueyama M, Nagai Y, Yamaguchi M. Identification of Ter94, Drosophila VCP, As a Strong Modulator of Motor Neuron Degeneration Induced by Knockdown of Caz, Drosophila FUS. Hum Mol Genet. 2014 doi: 10.1093/hmg/ddu055. [DOI] [PubMed] [Google Scholar]
  10. Baborie A, Griffiths TD, Jaros E, Perry R, McKeith IG, Burn DJ, Masuda-Suzukake M, Hasegawa M, Rollinson S, Pickering-Brown S, Robinson AC, Davidson YS, Mann DM. Accumulation of Dipeptide Repeat Proteins Predates That of TDP-43 in Frontotemporal Lobar Degeneration Associated With Hexanucleotide Repeat Expansions in C9ORF72 Gene. Neuropathol Appl Neurobiol. 2014 doi: 10.1111/nan.12178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Baechtold H, Kuroda M, Sok J, Ron D, Lopez BS, Akhmedov AT. Human 75-KDa DNA-Pairing Protein Is Identical to the Pro-Oncoprotein TLS/FUS and Is Able to Promote D-Loop Formation. J Biol Chem. 1999;274:34337–34342. doi: 10.1074/jbc.274.48.34337. [DOI] [PubMed] [Google Scholar]
  12. Bahadorani S, Mukai ST, Rabie J, Beckman JS, Phillips JP, Hilliker AJ. Expression of Zinc-Deficient Human Superoxide Dismutase in Drosophila Neurons Produces a Locomotor Defect Linked to Mitochondrial Dysfunction. Neurobiol Aging. 2013;34:2322–2330. doi: 10.1016/j.neurobiolaging.2013.03.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Barmada SJ, Skibinski G, Korb E, Rao EJ, Wu JY, Finkbeiner S. Cytoplasmic Mislocalization of TDP-43 Is Toxic to Neurons and Enhanced by a Mutation Associated With Familial Amyotrophic Lateral Sclerosis. J Neurosci. 2010;30:639–649. doi: 10.1523/JNEUROSCI.4988-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Batlevi Y, Martin DN, Pandey UB, Simon CR, Powers CM, Taylor JP, Baehrecke EH. Dynein Light Chain 1 Is Required for Autophagy, Protein Clearance, and Cell Death in Drosophila. Proc Natl Acad Sci U S A. 2010;107:742–747. doi: 10.1073/pnas.0907967107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bedford MT, Richard S. Arginine Methylation an Emerging Regulator of Protein Function. Mol Cell. 2005;18:263–272. doi: 10.1016/j.molcel.2005.04.003. [DOI] [PubMed] [Google Scholar]
  16. Bellen HJ, Tong C, Tsuda H. 100 Years of Drosophila Research and Its Impact on Vertebrate Neuroscience: a History Lesson for the Future. Nat Rev Neurosci. 2010;11:514–522. doi: 10.1038/nrn2839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Belzil VV, Valdmanis PN, Dion PA, Daoud H, Kabashi E, Noreau A, Gauthier J, Hince P, Desjarlais A, Bouchard JP, Lacomblez L, Salachas F, Pradat PF, Camu W, Meininger V, Dupre N, Rouleau GA. Mutations in FUS Cause FALS and SALS in French and French Canadian Populations. Neurology. 2009;73:1176–1179. doi: 10.1212/WNL.0b013e3181bbfeef. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bendotti C, Atzori C, Piva R, Tortarolo M, Strong MJ, DeBiasi S, Migheli A. Activated P38MAPK Is a Novel Component of the Intracellular Inclusions Found in Human Amyotrophic Lateral Sclerosis and Mutant SOD1 Transgenic Mice. J Neuropathol Exp Neurol. 2004;63:113–119. doi: 10.1093/jnen/63.2.113. [DOI] [PubMed] [Google Scholar]
  19. Bendotti C, Bao CM, Cheroni C, Grignaschi G, Lo CD, Peviani M, Tortarolo M, Veglianese P, Zennaro E. Inter- and Intracellular Signaling in Amyotrophic Lateral Sclerosis: Role of P38 Mitogen-Activated Protein Kinase. Neurodegener Dis. 2005;2:128–134. doi: 10.1159/000089617. [DOI] [PubMed] [Google Scholar]
  20. Bertolotti A, Lutz Y, Heard DJ, Chambon P, Tora L. HTAF(II)68, a Novel RNA/SsDNA-Binding Protein With Homology to the Pro-Oncoproteins TLS/FUS and EWS Is Associated With Both TFIID and RNA Polymerase II. EMBO J. 1996;15:5022–5031. [PMC free article] [PubMed] [Google Scholar]
  21. Boillee S, Vande VC, Cleveland DW. ALS: a Disease of Motor Neurons and Their Nonneuronal Neighbors. Neuron. 2006;52:39–59. doi: 10.1016/j.neuron.2006.09.018. [DOI] [PubMed] [Google Scholar]
  22. Bonini NM. A Genetic Model for Human Polyglutamine-Repeat Disease in Drosophila Melanogaster. Philos Trans R Soc Lond B Biol Sci. 1999;354:1057–1060. doi: 10.1098/rstb.1999.0458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Bowling AC, Schulz JB, Brown RH, Jr., Beal MF. Superoxide Dismutase Activity, Oxidative Damage, and Mitochondrial Energy Metabolism in Familial and Sporadic Amyotrophic Lateral Sclerosis. J Neurochem. 1993;61:2322–2325. doi: 10.1111/j.1471-4159.1993.tb07478.x. [DOI] [PubMed] [Google Scholar]
  24. Brand AH, Perrimon N. Targeted Gene Expression As a Means of Altering Cell Fates and Generating Dominant Phenotypes. Development. 1993;118:401–415. doi: 10.1242/dev.118.2.401. [DOI] [PubMed] [Google Scholar]
  25. Brand S, Bourbon HM. The Developmentally-Regulated Drosophila Gene Rox8 Encodes an RRM-Type RNA Binding Protein Structurally Related to Human TIA-1-Type Nucleolysins. Nucleic Acids Res. 1993;21:3699–3704. doi: 10.1093/nar/21.16.3699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Bridges CB. Direct Proof Through Non-Disjunction That the Sex-Linked Genes of Drosophila Are Borne by the X-Chromosome. Science. 1914;40:107–109. doi: 10.1126/science.40.1020.107. [DOI] [PubMed] [Google Scholar]
  27. Bruijn LI, Becher MW, Lee MK, Anderson KL, Jenkins NA, Copeland NG, Sisodia SS, Rothstein JD, Borchelt DR, Price DL, Cleveland DW. ALS-Linked SOD1 Mutant G85R Mediates Damage to Astrocytes and Promotes Rapidly Progressive Disease With SOD1-Containing Inclusions. Neuron. 1997;18:327–338. doi: 10.1016/s0896-6273(00)80272-x. [DOI] [PubMed] [Google Scholar]
  28. Bruijn LI, Houseweart MK, Kato S, Anderson KL, Anderson SD, Ohama E, Reaume AG, Scott RW, Cleveland DW. Aggregation and Motor Neuron Toxicity of an ALS-Linked SOD1 Mutant Independent From Wild-Type SOD1. Science. 1998;281:1851–1854. doi: 10.1126/science.281.5384.1851. [DOI] [PubMed] [Google Scholar]
  29. Bucci C, Parton RG, Mather IH, Stunnenberg H, Simons K, Hoflack B, Zerial M. The Small GTPase Rab5 Functions As a Regulatory Factor in the Early Endocytic Pathway. Cell. 1992;70:715–728. doi: 10.1016/0092-8674(92)90306-w. [DOI] [PubMed] [Google Scholar]
  30. Buratti E, Brindisi A, Giombi M, Tisminetzky S, Ayala YM, Baralle FE. TDP-43 Binds Heterogeneous Nuclear Ribonucleoprotein A/B Through Its C-Terminal Tail: an Important Region for the Inhibition of Cystic Fibrosis Transmembrane Conductance Regulator Exon 9 Splicing. J Biol Chem. 2005;280:37572–37584. doi: 10.1074/jbc.M505557200. [DOI] [PubMed] [Google Scholar]
  31. Buratti E, Dork T, Zuccato E, Pagani F, Romano M, Baralle FE. Nuclear Factor TDP-43 and SR Proteins Promote in Vitro and in Vivo CFTR Exon 9 Skipping. EMBO J. 2001;20:1774–1784. doi: 10.1093/emboj/20.7.1774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Byrne S, Elamin M, Bede P, Shatunov A, Walsh C, Corr B, Heverin M, Jordan N, Kenna K, Lynch C, McLaughlin RL, Iyer PM, O'Brien C, Phukan J, Wynne B, Bokde AL, Bradley DG, Pender N, Al-Chalabi A, Hardiman O. Cognitive and Clinical Characteristics of Patients With Amyotrophic Lateral Sclerosis Carrying a C9orf72 Repeat Expansion: a Population-Based Cohort Study. Lancet Neurol. 2012;11:232–240. doi: 10.1016/S1474-4422(12)70014-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Byrne S, Walsh C, Lynch C, Bede P, Elamin M, Kenna K, McLaughlin R, Hardiman O. Rate of Familial Amyotrophic Lateral Sclerosis: a Systematic Review and Meta-Analysis. J Neurol Neurosurg Psychiatry. 2011;82:623–627. doi: 10.1136/jnnp.2010.224501. [DOI] [PubMed] [Google Scholar]
  34. Chai A, Withers J, Koh YH, Parry K, Bao H, Zhang B, Budnik V, Pennetta G. HVAPB, the Causative Gene of a Heterogeneous Group of Motor Neuron Diseases in Humans, Is Functionally Interchangeable With Its Drosophila Homologue DVAP-33A at the Neuromuscular Junction. Hum Mol Genet. 2008;17:266–280. doi: 10.1093/hmg/ddm303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Chang JC, Hazelett DJ, Stewart JA, Morton DB. Motor Neuron Expression of the Voltage-Gated Calcium Channel Cacophony Restores Locomotion Defects in a Drosophila, TDP-43 Loss of Function Model of ALS. Brain Res. 2013 doi: 10.1016/j.brainres.2013.11.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Charcot JM. De La Sclérose Latérale Amyotrophique. Progrès Médical. 1874;2:325. [Google Scholar]
  37. Charcot JM, Joffroy A. Deux Cas D'Atrophie Musculaire Progressive Avec Lesion De La Substance Grise Et Des Faisceaux Antero-Lateraux De La Moelle Epiniere. Arch Physiol Neurol Path. 1869;2:744–754. [Google Scholar]
  38. Chen HJ, Anagnostou G, Chai A, Withers J, Morris A, Adhikaree J, Pennetta G, de Belleroche JS. Characterization of the Properties of a Novel Mutation in VAPB in Familial Amyotrophic Lateral Sclerosis. J Biol Chem. 2010;285:40266–40281. doi: 10.1074/jbc.M110.161398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Chen Y, Yang M, Deng J, Chen X, Ye Y, Zhu L, Liu J, Ye H, Shen Y, Li Y, Rao EJ, Fushimi K, Zhou X, Bigio EH, Mesulam M, Xu Q, Wu JY. Expression of Human FUS Protein in Drosophila Leads to Progressive Neurodegeneration. Protein Cell. 2011;2:477–486. doi: 10.1007/s13238-011-1065-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Chio A, Logroscino G, Hardiman O, Swingler R, Mitchell D, Beghi E, Traynor BG. Prognostic Factors in ALS: A Critical Review. Amyotroph Lateral Scler. 2009;10:310–323. doi: 10.3109/17482960802566824. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Choksi DK, Roy B, Chatterjee S, Yusuff T, Bakhoum MF, Sengupta U, Ambegaokar S, Kayed R, Jackson GR. TDP-43 Phosphorylation by Casein Kinase I{Varepsilon} Promotes Oligomerization and Enhances Toxicity in Vivo. Hum Mol Genet. 2014;23:1025–1035. doi: 10.1093/hmg/ddt498. [DOI] [PubMed] [Google Scholar]
  42. Chou SM, Wang HS, Komai K. Colocalization of NOS and SOD1 in Neurofilament Accumulation Within Motor Neurons of Amyotrophic Lateral Sclerosis: an Immunohistochemical Study. J Chem Neuroanat. 1996a;10:249–258. doi: 10.1016/0891-0618(96)00137-8. [DOI] [PubMed] [Google Scholar]
  43. Chou SM, Wang HS, Taniguchi A. Role of SOD-1 and Nitric Oxide/Cyclic GMP Cascade on Neurofilament Aggregation in ALS/MND. J Neurol Sci. 1996b;139:16–26. doi: 10.1016/0022-510x(96)00090-1. Suppl. [DOI] [PubMed] [Google Scholar]
  44. Cooper AA, Gitler AD, Cashikar A, Haynes CM, Hill KJ, Bhullar B, Liu K, Xu K, Strathearn KE, Liu F, Cao S, Caldwell KA, Caldwell GA, Marsischky G, Kolodner RD, Labaer J, Rochet JC, Bonini NM, Lindquist S. Alpha-Synuclein Blocks ER-Golgi Traffic and Rab1 Rescues Neuron Loss in Parkinson's Models. Science. 2006;313:324–328. doi: 10.1126/science.1129462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Corrado L, Del BR, Castellotti B, Ratti A, Cereda C, Penco S, Soraru G, Carlomagno Y, Ghezzi S, Pensato V, Colombrita C, Gagliardi S, Cozzi L, Orsetti V, Mancuso M, Siciliano G, Mazzini L, Comi GP, Gellera C, Ceroni M, D'Alfonso S, Silani V. Mutations of FUS Gene in Sporadic Amyotrophic Lateral Sclerosis. J Med Genet. 2010;47:190–194. doi: 10.1136/jmg.2009.071027. [DOI] [PubMed] [Google Scholar]
  46. Couratier P, Hugon J, Sindou P, Vallat JM, Dumas M. Cell Culture Evidence for Neuronal Degeneration in Amyotrophic Lateral Sclerosis Being Linked to Glutamate AMPA/Kainate Receptors. Lancet. 1993;341:265–268. doi: 10.1016/0140-6736(93)92615-z. [DOI] [PubMed] [Google Scholar]
  47. Couthouis J, Hart MP, Erion R, King OD, Diaz Z, Nakaya T, Ibrahim F, Kim HJ, Mojsilovic-Petrovic J, Panossian S, Kim CE, Frackelton EC, Solski JA, Williams KL, Clay-Falcone D, Elman L, McCluskey L, Greene R, Hakonarson H, Kalb RG, Lee VM, Trojanowski JQ, Nicholson GA, Blair IP, Bonini NM, Van Deerlin VM, Mourelatos Z, Shorter J, Gitler AD. Evaluating the Role of the FUS/TLS-Related Gene EWSR1 in Amyotrophic Lateral Sclerosis. Hum Mol Genet. 2012;21:2899–2911. doi: 10.1093/hmg/dds116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Couthouis J, Hart MP, Shorter J, DeJesus-Hernandez M, Erion R, Oristano R, Liu AX, Ramos D, Jethava N, Hosangadi D, Epstein J, Chiang A, Diaz Z, Nakaya T, Ibrahim F, Kim HJ, Solski JA, Williams KL, Mojsilovic-Petrovic J, Ingre C, Boylan K, Graff-Radford NR, Dickson DW, Clay-Falcone D, Elman L, McCluskey L, Greene R, Kalb RG, Lee VM, Trojanowski JQ, Ludolph A, Robberecht W, Andersen PM, Nicholson GA, Blair IP, King OD, Bonini NM, Van D V, Rademakers R, Mourelatos Z, Gitler AD. A Yeast Functional Screen Predicts New Candidate ALS Disease Genes. Proc Natl Acad Sci U S A. 2011;108:20881–20890. doi: 10.1073/pnas.1109434108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Crow JP, Sampson JB, Zhuang Y, Thompson JA, Beckman JS. Decreased Zinc Affinity of Amyotrophic Lateral Sclerosis-Associated Superoxide Dismutase Mutants Leads to Enhanced Catalysis of Tyrosine Nitration by Peroxynitrite. J Neurochem. 1997;69:1936–1944. doi: 10.1046/j.1471-4159.1997.69051936.x. [DOI] [PubMed] [Google Scholar]
  50. Crozat A, Aman P, Mandahl N, Ron D. Fusion of CHOP to a Novel RNA-Binding Protein in Human Myxoid Liposarcoma. Nature. 1993;363:640–644. doi: 10.1038/363640a0. [DOI] [PubMed] [Google Scholar]
  51. Cushman-Nick M, Bonini NM, Shorter J. Hsp104 Suppresses Polyglutamine-Induced Degeneration Post Onset in a Drosophila MJD/SCA3 Model. PLoS Genet. 2013;9:e1003781. doi: 10.1371/journal.pgen.1003781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Daigle JG, Lanson NA, Jr., Smith RB, Casci I, Maltare A, Monaghan J, Nichols CD, Kryndushkin D, Shewmaker F, Pandey UB. RNA-Binding Ability of FUS Regulates Neurodegeneration, Cytoplasmic Mislocalization and Incorporation into Stress Granules Associated With FUS Carrying ALS-Linked Mutations. Hum Mol Genet. 2013;22:1193–1205. doi: 10.1093/hmg/dds526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. De Vos KJ, Morotz GM, Stoica R, Tudor EL, Lau KF, Ackerley S, Warley A, Shaw CE, Miller CC. VAPB Interacts With the Mitochondrial Protein PTPIP51 to Regulate Calcium Homeostasis. Hum Mol Genet. 2012;21:1299–1311. doi: 10.1093/hmg/ddr559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. DeJesus-Hernandez M, Kocerha J, Finch N, Crook R, Baker M, Desaro P, Johnston A, Rutherford N, Wojtas A, Kennelly K, Wszolek ZK, Graff-Radford N, Boylan K, Rademakers R. De Novo Truncating FUS Gene Mutation As a Cause of Sporadic Amyotrophic Lateral Sclerosis. Hum Mutat. 2010;31:E1377–E1389. doi: 10.1002/humu.21241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. DeJesus-Hernandez M, Mackenzie IR, Boeve BF, Boxer AL, Baker M, Rutherford NJ, Nicholson AM, Finch NA, Flynn H, Adamson J, Kouri N, Wojtas A, Sengdy P, Hsiung GY, Karydas A, Seeley WW, Josephs KA, Coppola G, Geschwind DH, Wszolek ZK, Feldman H, Knopman DS, Petersen RC, Miller BL, Dickson DW, Boylan KB, Graff-Radford NR, Rademakers R. Expanded GGGGCC Hexanucleotide Repeat in Noncoding Region of C9ORF72 Causes Chromosome 9p-Linked FTD and ALS. Neuron. 2011;72:245–256. doi: 10.1016/j.neuron.2011.09.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Del BR, Tiloca C, Pensato V, Corrado L, Ratti A, Ticozzi N, Corti S, Castellotti B, Mazzini L, Soraru G, Cereda C, D'Alfonso S, Gellera C, Comi GP, Silani V. Novel Optineurin Mutations in Patients With Familial and Sporadic Amyotrophic Lateral Sclerosis. J Neurol Neurosurg Psychiatry. 2011;82:1239–1243. doi: 10.1136/jnnp.2011.242313. [DOI] [PubMed] [Google Scholar]
  57. Deng HX, Chen W, Hong ST, Boycott KM, Gorrie GH, Siddique N, Yang Y, Fecto F, Shi Y, Zhai H, Jiang H, Hirano M, Rampersaud E, Jansen GH, Donkervoort S, Bigio EH, Brooks BR, Ajroud K, Sufit RL, Haines JL, Mugnaini E, Pericak-Vance MA, Siddique T. Mutations in UBQLN2 Cause Dominant X-Linked Juvenile and Adult-Onset ALS and ALS/Dementia. Nature. 2011;477:211–215. doi: 10.1038/nature10353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Deng HX, Shi Y, Furukawa Y, Zhai H, Fu R, Liu E, Gorrie GH, Khan MS, Hung WY, Bigio EH, Lukas T, Dal Canto MC, O'Halloran TV, Siddique T. Conversion to the Amyotrophic Lateral Sclerosis Phenotype Is Associated With Intermolecular Linked Insoluble Aggregates of SOD1 in Mitochondria. Proc Natl Acad Sci U S A. 2006;103:7142–7147. doi: 10.1073/pnas.0602046103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Dewil M, dela Cruz VF, Van Den Bosch L, Robberecht W. Inhibition of P38 Mitogen Activated Protein Kinase Activation and Mutant SOD1(G93A)-Induced Motor Neuron Death. Neurobiol Dis. 2007;26:332–341. doi: 10.1016/j.nbd.2006.12.023. [DOI] [PubMed] [Google Scholar]
  60. Diaper DC, Adachi Y, Lazarou L, Greenstein M, Simoes FA, Di DA, Solomon DA, Lowe S, Alsubaie R, Cheng D, Buckley S, Humphrey DM, Shaw CE, Hirth F. Drosophila TDP-43 Dysfunction in Glia and Muscle Cells Cause Cytological and Behavioural Phenotypes That Characterize ALS and FTLD. Hum Mol Genet. 2013a;22:3883–3893. doi: 10.1093/hmg/ddt243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Diaper DC, Adachi Y, Sutcliffe B, Humphrey DM, Elliott CJ, Stepto A, Ludlow ZN, Vanden Broeck L, Callaerts P, Dermaut B, Al-Chalabi A, Shaw CE, Robinson IM, Hirth F. Loss and Gain of Drosophila TDP-43 Impair Synaptic Efficacy and Motor Control Leading to Age-Related Neurodegeneration by Loss-of-Function Phenotypes. Hum Mol Genet. 2013b;22:1539–1557. doi: 10.1093/hmg/ddt005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Dormann D, Madl T, Valori CF, Bentmann E, Tahirovic S, Abou-Ajram C, Kremmer E, Ansorge O, Mackenzie IR, Neumann M, Haass C. Arginine Methylation Next to the PY-NLS Modulates Transportin Binding and Nuclear Import of FUS. EMBO J. 2012;31:4258–4275. doi: 10.1038/emboj.2012.261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Du K, Arai S, Kawamura T, Matsushita A, Kurokawa R. TLS and PRMT1 Synergistically Coactivate Transcription at the Survivin Promoter Through TLS Arginine Methylation. Biochem Biophys Res Commun. 2011;404:991–996. doi: 10.1016/j.bbrc.2010.12.097. [DOI] [PubMed] [Google Scholar]
  64. Elden AC, Kim HJ, Hart MP, Chen-Plotkin AS, Johnson BS, Fang X, Armakola M, Geser F, Greene R, Lu MM, Padmanabhan A, Clay-Falcone D, McCluskey L, Elman L, Juhr D, Gruber PJ, Rub U, Auburger G, Trojanowski JQ, Lee VM, Van Deerlin VM, Bonini NM, Gitler AD. Ataxin-2 Intermediate-Length Polyglutamine Expansions Are Associated With Increased Risk for ALS. Nature. 2010;466:1069–1075. doi: 10.1038/nature09320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Elia AJ, Parkes TL, Kirby K, St George-Hyslop P, Boulianne GL, Phillips JP, Hilliker AJ. Expression of Human FALS SOD in Motorneurons of Drosophila. Free Radic Biol Med. 1999;26:1332–1338. doi: 10.1016/s0891-5849(98)00333-5. [DOI] [PubMed] [Google Scholar]
  66. Estes PS, Boehringer A, Zwick R, Tang JE, Grigsby B, Zarnescu DC. Wild-Type and A315T Mutant TDP-43 Exert Differential Neurotoxicity in a Drosophila Model of ALS. Hum Mol Genet. 2011;20:2308–2321. doi: 10.1093/hmg/ddr124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Estes PS, Daniel SG, McCallum AP, Boehringer AV, Sukhina AS, Zwick RA, Zarnescu DC. Motor Neurons and Glia Exhibit Specific Individualized Responses to TDP-43 Expression in a Drosophila Model of Amyotrophic Lateral Sclerosis. Dis Model Mech. 2013;6:721–733. doi: 10.1242/dmm.010710. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Fackelmayer FO. Protein Arginine Methyltransferases: Guardians of the Arg? Trends Biochem Sci. 2005;30:666–671. doi: 10.1016/j.tibs.2005.10.002. [DOI] [PubMed] [Google Scholar]
  69. Fang F, Valdimarsdottir U, Bellocco R, Ronnevi LO, Sparen P, Fall K, Ye W. Amyotrophic Lateral Sclerosis in Sweden, 1991-2005. Arch Neurol. 2009;66:515–519. doi: 10.1001/archneurol.2009.13. [DOI] [PubMed] [Google Scholar]
  70. Farny NG, Kedersha NL, Silver PA. Metazoan Stress Granule Assembly Is Mediated by P-EIF2alpha-Dependent and -Independent Mechanisms. RNA. 2009;15:1814–1821. doi: 10.1261/rna.1684009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Fecto F, Yan J, Vemula SP, Liu E, Yang Y, Chen W, Zheng JG, Shi Y, Siddique N, Arrat H, Donkervoort S, Ajroud-Driss S, Sufit RL, Heller SL, Deng HX, Siddique T. SQSTM1 Mutations in Familial and Sporadic Amyotrophic Lateral Sclerosis. Arch Neurol. 2011;68:1440–1446. doi: 10.1001/archneurol.2011.250. [DOI] [PubMed] [Google Scholar]
  72. Feiguin F, Godena VK, Romano G, D'Ambrogio A, Klima R, Baralle FE. Depletion of TDP-43 Affects Drosophila Motoneurons Terminal Synapsis and Locomotive Behavior. FEBS Lett. 2009;583:1586–1592. doi: 10.1016/j.febslet.2009.04.019. [DOI] [PubMed] [Google Scholar]
  73. Fernandez-Funez P, Nino-Rosales ML, de GB, She WC, Luchak JM, Martinez P, Turiegano E, Benito J, Capovilla M, Skinner PJ, McCall A, Canal I, Orr HT, Zoghbi HY, Botas J. Identification of Genes That Modify Ataxin-1-Induced Neurodegeneration. Nature. 2000;408:101–106. doi: 10.1038/35040584. [DOI] [PubMed] [Google Scholar]
  74. Ferri A, Cozzolino M, Crosio C, Nencini M, Casciati A, Gralla EB, Rotilio G, Valentine JS, Carri MT. Familial ALS-Superoxide Dismutases Associate With Mitochondria and Shift Their Redox Potentials. Proc Natl Acad Sci U S A. 2006;103:13860–13865. doi: 10.1073/pnas.0605814103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Fiesel FC, Voigt A, Weber SS, Van den Haute C, Waldenmaier A, Gorner K, Walter M, Anderson ML, Kern JV, Rasse TM, Schmidt T, Springer W, Kirchner R, Bonin M, Neumann M, Baekelandt V, Alunni-Fabbroni M, Schulz JB, Kahle PJ. Knockdown of Transactive Response DNA-Binding Protein (TDP-43) Downregulates Histone Deacetylase 6. EMBO J. 2010;29:209–221. doi: 10.1038/emboj.2009.324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Forrest S, Chai A, Sanhueza M, Marescotti M, Parry K, Georgiev A, Sahota V, Mendez-Castro R, Pennetta G. Increased Levels of Phosphoinositides Cause Neurodegeneration in a Drosophila Model of Amyotrophic Lateral Sclerosis. Hum Mol Genet. 2013;22:2689–2704. doi: 10.1093/hmg/ddt118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Fridovich I. Superoxide Dismutases. Adv Enzymol Relat Areas Mol Biol. 1986;58:61–97. doi: 10.1002/9780470123041.ch2. [DOI] [PubMed] [Google Scholar]
  78. Gijselinck I, Van LT, van der Zee J, Sleegers K, Philtjens S, Kleinberger G, Janssens J, Bettens K, Van CC, Pereson S, Engelborghs S, Sieben A, De JP, Vandenberghe R, Santens P, De BJ, Maes G, Baumer V, Dillen L, Joris G, Cuijt I, Corsmit E, Elinck E, Van DJ, Vermeulen S, Van den Broeck M, Vaerenberg C, Mattheijssens M, Peeters K, Robberecht W, Cras P, Martin JJ, De Deyn PP, Cruts M, Van BC. A C9orf72 Promoter Repeat Expansion in a Flanders-Belgian Cohort With Disorders of the Frontotemporal Lobar Degeneration-Amyotrophic Lateral Sclerosis Spectrum: a Gene Identification Study. Lancet Neurol. 2012;11:54–65. doi: 10.1016/S1474-4422(11)70261-7. [DOI] [PubMed] [Google Scholar]
  79. Gordon PH. Amyotrophic Lateral Sclerosis: An Update for 2013 Clinical Features, Pathophysiology, Management and Therapeutic Trials. Aging Dis. 2013;4:295–310. doi: 10.14336/AD.2013.0400295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Gregory JM, Barros TP, Meehan S, Dobson CM, Luheshi LM. The Aggregation and Neurotoxicity of TDP-43 and Its ALS-Associated 25 KDa Fragment Are Differentially Affected by Molecular Chaperones in Drosophila. PLoS One. 2012;7:e31899. doi: 10.1371/journal.pone.0031899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Grozinger CM, Hassig CA, Schreiber SL. Three Proteins Define a Class of Human Histone Deacetylases Related to Yeast Hda1p. Proc Natl Acad Sci U S A. 1999;96:4868–4873. doi: 10.1073/pnas.96.9.4868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Hadano S, Hand CK, Osuga H, Yanagisawa Y, Otomo A, Devon RS, Miyamoto N, Showguchi-Miyata J, Okada Y, Singaraja R, Figlewicz DA, Kwiatkowski T, Hosler BA, Sagie T, Skaug J, Nasir J, Brown RH, Jr., Scherer SW, Rouleau GA, Hayden MR, Ikeda JE. A Gene Encoding a Putative GTPase Regulator Is Mutated in Familial Amyotrophic Lateral Sclerosis 2. Nat Genet. 2001;29:166–173. doi: 10.1038/ng1001-166. [DOI] [PubMed] [Google Scholar]
  83. Han SM, Tsuda H, Yang Y, Vibbert J, Cottee P, Lee SJ, Winek J, Haueter C, Bellen HJ, Miller MA. Secreted VAPB/ALS8 Major Sperm Protein Domains Modulate Mitochondrial Localization and Morphology Via Growth Cone Guidance Receptors. Dev Cell. 2012;22:348–362. doi: 10.1016/j.devcel.2011.12.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Hanson KA, Kim SH, Wassarman DA, Tibbetts RS. Ubiquilin Modifies TDP-43 Toxicity in a Drosophila Model of Amyotrophic Lateral Sclerosis (ALS) J Biol Chem. 2010;285:11068–11072. doi: 10.1074/jbc.C109.078527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Hardiman O, van den Berg LH, Kiernan MC. Clinical Diagnosis and Management of Amyotrophic Lateral Sclerosis. Nat Rev Neurol. 2011;7:639–649. doi: 10.1038/nrneurol.2011.153. [DOI] [PubMed] [Google Scholar]
  86. HARMAN D. Aging: a Theory Based on Free Radical and Radiation Chemistry. J Gerontol. 1956;11:298–300. doi: 10.1093/geronj/11.3.298. [DOI] [PubMed] [Google Scholar]
  87. Hasegawa M, Arai T, Nonaka T, Kametani F, Yoshida M, Hashizume Y, Beach TG, Buratti E, Baralle F, Morita M, Nakano I, Oda T, Tsuchiya K, Akiyama H. Phosphorylated TDP-43 in Frontotemporal Lobar Degeneration and Amyotrophic Lateral Sclerosis. Ann Neurol. 2008;64:60–70. doi: 10.1002/ana.21425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Hazelett DJ, Chang JC, Lakeland DL, Morton DB. Comparison of Parallel High-Throughput RNA Sequencing Between Knockout of TDP-43 and Its Overexpression Reveals Primarily Nonreciprocal and Nonoverlapping Gene Expression Changes in the Central Nervous System of Drosophila. G3 (Bethesda ) 2012;2:789–802. doi: 10.1534/g3.112.002998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. He F, Krans A, Freibaum BD, Taylor JP, Todd PK. TDP-43 Suppresses CGG Repeat-Induced Neurotoxicity Through Interactions With HnRNP A2/B1. Hum Mol Genet. 2014 doi: 10.1093/hmg/ddu216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Higashiyama H, Hirose F, Yamaguchi M, Inoue YH, Fujikake N, Matsukage A, Kakizuka A. Identification of Ter94, Drosophila VCP, As a Modulator of Polyglutamine-Induced Neurodegeneration. Cell Death Differ. 2002;9:264–273. doi: 10.1038/sj.cdd.4400955. [DOI] [PubMed] [Google Scholar]
  91. Hindle S, Afsari F, Stark M, Middleton CA, Evans GJ, Sweeney ST, Elliott CJ. Dopaminergic Expression of the Parkinsonian Gene LRRK2-G2019S Leads to Non-Autonomous Visual Neurodegeneration, Accelerated by Increased Neural Demands for Energy. Hum Mol Genet. 2013;22:2129–2140. doi: 10.1093/hmg/ddt061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Hirano A, Donnenfeld H, Sasaki S, Nakano I. Fine Structural Observations of Neurofilamentous Changes in Amyotrophic Lateral Sclerosis. J Neuropathol Exp Neurol. 1984a;43:461–470. doi: 10.1097/00005072-198409000-00001. [DOI] [PubMed] [Google Scholar]
  93. Hirano A, Nakano I, Kurland LT, Mulder DW, Holley PW, Saccomanno G. Fine Structural Study of Neurofibrillary Changes in a Family With Amyotrophic Lateral Sclerosis. J Neuropathol Exp Neurol. 1984b;43:471–480. doi: 10.1097/00005072-198409000-00002. [DOI] [PubMed] [Google Scholar]
  94. Holasek SS, Wengenack TM, Kandimalla KK, Montano C, Gregor DM, Curran GL, Poduslo JF. Activation of the Stress-Activated MAP Kinase, P38, but Not JNK in Cortical Motor Neurons During Early Presymptomatic Stages of Amyotrophic Lateral Sclerosis in Transgenic Mice. Brain Res. 2005;1045:185–198. doi: 10.1016/j.brainres.2005.03.037. [DOI] [PubMed] [Google Scholar]
  95. Huang C, Tong J, Bi F, Wu Q, Huang B, Zhou H, Xia XG. Entorhinal Cortical Neurons Are the Primary Targets of FUS Mislocalization and Ubiquitin Aggregation in FUS Transgenic Rats. Hum Mol Genet. 2012;21:4602–4614. doi: 10.1093/hmg/dds299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Huang C, Zhou H, Tong J, Chen H, Liu YJ, Wang D, Wei X, Xia XG. FUS Transgenic Rats Develop the Phenotypes of Amyotrophic Lateral Sclerosis and Frontotemporal Lobar Degeneration. PLoS Genet. 2011;7:e1002011. doi: 10.1371/journal.pgen.1002011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Hubbert C, Guardiola A, Shao R, Kawaguchi Y, Ito A, Nixon A, Yoshida M, Wang XF, Yao TP. HDAC6 Is a Microtubule-Associated Deacetylase. Nature. 2002;417:455–458. doi: 10.1038/417455a. [DOI] [PubMed] [Google Scholar]
  98. Huisman MH, de Jong SW, van Doormaal PT, Weinreich SS, Schelhaas HJ, van der Kooi AJ, de VM, Veldink JH, van den Berg LH. Population Based Epidemiology of Amyotrophic Lateral Sclerosis Using Capture-Recapture Methodology. J Neurol Neurosurg Psychiatry. 2011;82:1165–1170. doi: 10.1136/jnnp.2011.244939. [DOI] [PubMed] [Google Scholar]
  99. Hutchison S, LeBel C, Blanchette M, Chabot B. Distinct Sets of Adjacent Heterogeneous Nuclear Ribonucleoprotein (HnRNP) A1/A2 Binding Sites Control 5' Splice Site Selection in the HnRNP A1 MRNA Precursor. J Biol Chem. 2002;277:29745–29752. doi: 10.1074/jbc.M203633200. [DOI] [PubMed] [Google Scholar]
  100. Ibrahim F, Maragkakis M, Alexiou P, Maronski MA, Dichter MA, Mourelatos Z. Identification of in Vivo, Conserved, TAF15 RNA Binding Sites Reveals the Impact of TAF15 on the Neuronal Transcriptome. Cell Rep. 2013;3:301–308. doi: 10.1016/j.celrep.2013.01.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Igaz LM, Kwong LK, Chen-Plotkin A, Winton MJ, Unger TL, Xu Y, Neumann M, Trojanowski JQ, Lee VM. Expression of TDP-43 C-Terminal Fragments in Vitro Recapitulates Pathological Features of TDP-43 Proteinopathies. J Biol Chem. 2009;284:8516–8524. doi: 10.1074/jbc.M809462200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Igaz LM, Kwong LK, Lee EB, Chen-Plotkin A, Swanson E, Unger T, Malunda J, Xu Y, Winton MJ, Trojanowski JQ, Lee VM. Dysregulation of the ALS-Associated Gene TDP-43 Leads to Neuronal Death and Degeneration in Mice. J Clin Invest. 2011;121:726–738. doi: 10.1172/JCI44867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Igaz LM, Kwong LK, Xu Y, Truax AC, Uryu K, Neumann M, Clark CM, Elman LB, Miller BL, Grossman M, McCluskey LF, Trojanowski JQ, Lee VM. Enrichment of C-Terminal Fragments in TAR DNA-Binding Protein-43 Cytoplasmic Inclusions in Brain but Not in Spinal Cord of Frontotemporal Lobar Degeneration and Amyotrophic Lateral Sclerosis. Am J Pathol. 2008;173:182–194. doi: 10.2353/ajpath.2008.080003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Ihara R, Matsukawa K, Nagata Y, Kunugi H, Tsuji S, Chihara T, Kuranaga E, Miura M, Wakabayashi T, Hashimoto T, Iwatsubo T. RNA Binding Mediates Neurotoxicity in the Transgenic Drosophila Model of TDP-43 Proteinopathy. Hum Mol Genet. 2013;22:4474–4484. doi: 10.1093/hmg/ddt296. [DOI] [PubMed] [Google Scholar]
  105. Iida A, Hosono N, Sano M, Kamei T, Oshima S, Tokuda T, Kubo M, Nakamura Y, Ikegawa S. Optineurin Mutations in Japanese Amyotrophic Lateral Sclerosis. J Neurol Neurosurg Psychiatry. 2012;83:233–235. doi: 10.1136/jnnp.2010.234963. [DOI] [PubMed] [Google Scholar]
  106. Imbert G, Saudou F, Yvert G, Devys D, Trottier Y, Garnier JM, Weber C, Mandel JL, Cancel G, Abbas N, Durr A, Didierjean O, Stevanin G, Agid Y, Brice A. Cloning of the Gene for Spinocerebellar Ataxia 2 Reveals a Locus With High Sensitivity to Expanded CAG/Glutamine Repeats. Nat Genet. 1996;14:285–291. doi: 10.1038/ng1196-285. [DOI] [PubMed] [Google Scholar]
  107. Inukai Y, Nonaka T, Arai T, Yoshida M, Hashizume Y, Beach TG, Buratti E, Baralle FE, Akiyama H, Hisanaga S, Hasegawa M. Abnormal Phosphorylation of Ser409/410 of TDP-43 in FTLD-U and ALS. FEBS Lett. 2008;582:2899–2904. doi: 10.1016/j.febslet.2008.07.027. [DOI] [PubMed] [Google Scholar]
  108. Islam R, Kumimoto EL, Bao H, Zhang B. ALS-linked SOD1 in glial cells enhances ß-N-Methylamino L-Alanine (BMAA)-induced toxicity in Drosophila. 2012. F1000 Research. [DOI] [PMC free article] [PubMed]
  109. Jackson GR, Salecker I, Dong X, Yao X, Arnheim N, Faber PW, MacDonald ME, Zipursky SL. Polyglutamine-Expanded Human Huntingtin Transgenes Induce Degeneration of Drosophila Photoreceptor Neurons. Neuron. 1998;21:633–642. doi: 10.1016/s0896-6273(00)80573-5. [DOI] [PubMed] [Google Scholar]
  110. Jin P, Zarnescu DC, Zhang F, Pearson CE, Lucchesi JC, Moses K, Warren ST. RNA-Mediated Neurodegeneration Caused by the Fragile X Premutation RCGG Repeats in Drosophila. Neuron. 2003;39:739–747. doi: 10.1016/s0896-6273(03)00533-6. [DOI] [PubMed] [Google Scholar]
  111. Johnson JO, Mandrioli J, Benatar M, Abramzon Y, Van Deerlin VM, Trojanowski JQ, Gibbs JR, Brunetti M, Gronka S, Wuu J, Ding J, McCluskey L, Martinez-Lage M, Falcone D, Hernandez DG, Arepalli S, Chong S, Schymick JC, Rothstein J, Landi F, Wang YD, Calvo A, Mora G, Sabatelli M, Monsurro MR, Battistini S, Salvi F, Spataro R, Sola P, Borghero G, Galassi G, Scholz SW, Taylor JP, Restagno G, Chio A, Traynor BJ. Exome Sequencing Reveals VCP Mutations As a Cause of Familial ALS. Neuron. 2010;68:857–864. doi: 10.1016/j.neuron.2010.11.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Johnson JO, Pioro EP, Boehringer A, Chia R, Feit H, Renton AE, Pliner HA, Abramzon Y, Marangi G, Winborn BJ, Gibbs JR, Nalls MA, Morgan S, Shoai M, Hardy J, Pittman A, Orrell RW, Malaspina A, Sidle KC, Fratta P, Harms MB, Baloh RH, Pestronk A, Weihl CC, Rogaeva E, Zinman L, Drory VE, Borghero G, Mora G, Calvo A, Rothstein JD, Drepper C, Sendtner M, Singleton AB, Taylor JP, Cookson MR, Restagno G, Sabatelli M, Bowser R, Chio A, Traynor BJ. Mutations in the Matrin 3 Gene Cause Familial Amyotrophic Lateral Sclerosis. Nat Neurosci. 2014;17:664–666. doi: 10.1038/nn.3688. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Johnston CA, Stanton BR, Turner MR, Gray R, Blunt AH, Butt D, Ampong MA, Shaw CE, Leigh PN, Al-Chalabi A. Amyotrophic Lateral Sclerosis in an Urban Setting: a Population Based Study of Inner City London. J Neurol. 2006;253:1642–1643. doi: 10.1007/s00415-006-0195-y. [DOI] [PubMed] [Google Scholar]
  114. Jones CT, Shaw PJ, Chari G, Brock DJ. Identification of a Novel Exon 4 SOD1 Mutation in a Sporadic Amyotrophic Lateral Sclerosis Patient. Mol Cell Probes. 1994a;8:329–330. doi: 10.1006/mcpr.1994.1046. [DOI] [PubMed] [Google Scholar]
  115. Jones CT, Swingler RJ, Brock DJ. Identification of a Novel SOD1 Mutation in an Apparently Sporadic Amyotrophic Lateral Sclerosis Patient and the Detection of Ile113Thr in Three Others. Hum Mol Genet. 1994b;3:649–650. doi: 10.1093/hmg/3.4.649. [DOI] [PubMed] [Google Scholar]
  116. Jones MA, Amr S, Ferebee A, Huynh P, Rosenfeld JA, Miles MF, Davies AG, Korey CA, Warrick JM, Shiang R, Elsea SH, Girirajan S, Grotewiel M. Genetic Studies in Drosophila and Humans Support a Model for the Concerted Function of CISD2, PPT1 and CLN3 in Disease. Biol Open. 2014 doi: 10.1242/bio.20147559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Joyce PI, Fratta P, Fisher EM, Acevedo-Arozena A. SOD1 and TDP-43 Animal Models of Amyotrophic Lateral Sclerosis: Recent Advances in Understanding Disease Toward the Development of Clinical Treatments. Mamm Genome. 2011;22:420–448. doi: 10.1007/s00335-011-9339-1. [DOI] [PubMed] [Google Scholar]
  118. Kabashi E, Valdmanis PN, Dion P, Spiegelman D, McConkey BJ, Vande VC, Bouchard JP, Lacomblez L, Pochigaeva K, Salachas F, Pradat PF, Camu W, Meininger V, Dupre N, Rouleau GA. TARDBP Mutations in Individuals With Sporadic and Familial Amyotrophic Lateral Sclerosis. Nat Genet. 2008;40:572–574. doi: 10.1038/ng.132. [DOI] [PubMed] [Google Scholar]
  119. Kanekura K, Nishimoto I, Aiso S, Matsuoka M. Characterization of Amyotrophic Lateral Sclerosis-Linked P56S Mutation of Vesicle-Associated Membrane Protein-Associated Protein B (VAPB/ALS8) J Biol Chem. 2006;281:30223–30233. doi: 10.1074/jbc.M605049200. [DOI] [PubMed] [Google Scholar]
  120. Kato S, Hayashi H, Nakashima K, Nanba E, Kato M, Hirano A, Nakano I, Asayama K, Ohama E. Pathological Characterization of Astrocytic Hyaline Inclusions in Familial Amyotrophic Lateral Sclerosis. Am J Pathol. 1997;151:611–620. [PMC free article] [PubMed] [Google Scholar]
  121. Kawaguchi Y, Kovacs JJ, McLaurin A, Vance JM, Ito A, Yao TP. The Deacetylase HDAC6 Regulates Aggresome Formation and Cell Viability in Response to Misfolded Protein Stress. Cell. 2003;115:727–738. doi: 10.1016/s0092-8674(03)00939-5. [DOI] [PubMed] [Google Scholar]
  122. Kawahara Y, Ito K, Sun H, Aizawa H, Kanazawa I, Kwak S. Glutamate Receptors: RNA Editing and Death of Motor Neurons. Nature. 2004;427:801. doi: 10.1038/427801a. [DOI] [PubMed] [Google Scholar]
  123. Kawasaki F, Zou B, Xu X, Ordway RW. Active Zone Localization of Presynaptic Calcium Channels Encoded by the Cacophony Locus of Drosophila. J Neurosci. 2004;24:282–285. doi: 10.1523/JNEUROSCI.3553-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Kazemi-Esfarjani P, Benzer S. Genetic Suppression of Polyglutamine Toxicity in Drosophila. Science. 2000;287:1837–1840. doi: 10.1126/science.287.5459.1837. [DOI] [PubMed] [Google Scholar]
  125. Kim EK, Choi EJ. Pathological Roles of MAPK Signaling Pathways in Human Diseases. Biochim Biophys Acta. 2010;1802:396–405. doi: 10.1016/j.bbadis.2009.12.009. [DOI] [PubMed] [Google Scholar]
  126. Kim HJ, Kim NC, Wang YD, Scarborough EA, Moore J, Diaz Z, MacLea KS, Freibaum B, Li S, Molliex A, Kanagaraj AP, Carter R, Boylan KB, Wojtas AM, Rademakers R, Pinkus JL, Greenberg SA, Trojanowski JQ, Traynor BJ, Smith BN, Topp S, Gkazi AS, Miller J, Shaw CE, Kottlors M, Kirschner J, Pestronk A, Li YR, Ford AF, Gitler AD, Benatar M, King OD, Kimonis VE, Ross ED, Weihl CC, Shorter J, Taylor JP. Mutations in Prion-Like Domains in HnRNPA2B1 and HnRNPA1 Cause Multisystem Proteinopathy and ALS. Nature. 2013;495:467–473. doi: 10.1038/nature11922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Kim HJ, Raphael AR, Ladow ES, McGurk L, Weber RA, Trojanowski JQ, Lee VM, Finkbeiner S, Gitler AD, Bonini NM. Therapeutic Modulation of EIF2alpha Phosphorylation Rescues TDP-43 Toxicity in Amyotrophic Lateral Sclerosis Disease Models. Nat Genet. 2014;46:152–160. doi: 10.1038/ng.2853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Kim SH, Shanware NP, Bowler MJ, Tibbetts RS. Amyotrophic Lateral Sclerosis-Associated Proteins TDP-43 and FUS/TLS Function in a Common Biochemical Complex to Co-Regulate HDAC6 MRNA. J Biol Chem. 2010;285:34097–34105. doi: 10.1074/jbc.M110.154831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Kim SH, Zhan L, Hanson KA, Tibbetts RS. High-Content RNAi Screening Identifies the Type 1 Inositol Triphosphate Receptor As a Modifier of TDP-43 Localization and Neurotoxicity. Hum Mol Genet. 2012;21:4845–4856. doi: 10.1093/hmg/dds321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Klichko VI, Radyuk SN, Orr WC. CuZn-SOD Promoter-Driven Expression in the Drosophila Central Nervous System. Neurobiol Aging. 1999;20:537–543. doi: 10.1016/s0197-4580(99)00090-1. [DOI] [PubMed] [Google Scholar]
  131. Ko HS, Uehara T, Tsuruma K, Nomura Y. Ubiquilin Interacts With Ubiquitylated Proteins and Proteasome Through Its Ubiquitin-Associated and Ubiquitin-Like Domains. FEBS Lett. 2004;566:110–114. doi: 10.1016/j.febslet.2004.04.031. [DOI] [PubMed] [Google Scholar]
  132. Kuhnlein P, Sperfeld AD, Vanmassenhove B, Van D V, Lee VM, Trojanowski JQ, Kretzschmar HA, Ludolph AC, Neumann M. Two German Kindreds With Familial Amyotrophic Lateral Sclerosis Due to TARDBP Mutations. Arch Neurol. 2008;65:1185–1189. doi: 10.1001/archneur.65.9.1185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Kumimoto EL, Fore TR, Zhang B. Transcriptome Profiling Following Neuronal and Glial Expression of ALS-Linked SOD1 in Drosophila. G3 (Bethesda ) 2013 doi: 10.1534/g3.113.005850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Kwiatkowski TJ, Jr., Bosco DA, Leclerc AL, Tamrazian E, Vanderburg CR, Russ C, Davis A, Gilchrist J, Kasarskis EJ, Munsat T, Valdmanis P, Rouleau GA, Hosler BA, Cortelli P, de Jong PJ, Yoshinaga Y, Haines JL, Pericak-Vance MA, Yan J, Ticozzi N, Siddique T, McKenna-Yasek D, Sapp PC, Horvitz HR, Landers JE, Brown RH., Jr. Mutations in the FUS/TLS Gene on Chromosome 16 Cause Familial Amyotrophic Lateral Sclerosis. Science. 2009;323:1205–1208. doi: 10.1126/science.1166066. [DOI] [PubMed] [Google Scholar]
  135. Kwon I, Xiang S, Kato M, Wu L, Theodoropoulos P, Wang T, Kim J, Yun J, Xie Y, McKnight SL. Poly-Dipeptides Encoded by the C9orf72 Repeats Bind Nucleoli, Impede RNA Biogenesis, and Kill Cells. Science. 2014;345:1139–1145. doi: 10.1126/science.1254917. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Labelle Y, Zucman J, Stenman G, Kindblom LG, Knight J, Turc-Carel C, Dockhorn-Dworniczak B, Mandahl N, Desmaze C, Peter M. Oncogenic Conversion of a Novel Orphan Nuclear Receptor by Chromosome Translocation. Hum Mol Genet. 1995;4:2219–2226. doi: 10.1093/hmg/4.12.2219. [DOI] [PubMed] [Google Scholar]
  137. Lagier-Tourenne C, Cleveland DW. Rethinking ALS: the FUS About TDP-43. Cell. 2009;136:1001–1004. doi: 10.1016/j.cell.2009.03.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Lagier-Tourenne C, Polymenidou M, Hutt KR, Vu AQ, Baughn M, Huelga SC, Clutario KM, Ling SC, Liang TY, Mazur C, Wancewicz E, Kim AS, Watt A, Freier S, Hicks GG, Donohue JP, Shiue L, Bennett CF, Ravits J, Cleveland DW, Yeo GW. Divergent Roles of ALS-Linked Proteins FUS/TLS and TDP-43 Intersect in Processing Long Pre-MRNAs. Nat Neurosci. 2012;15:1488–1497. doi: 10.1038/nn.3230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Lanson NA, Jr., Maltare A, King H, Smith R, Kim JH, Taylor JP, Lloyd TE, Pandey UB. A Drosophila Model of FUS-Related Neurodegeneration Reveals Genetic Interaction Between FUS and TDP-43. Hum Mol Genet. 2011;20:2510–2523. doi: 10.1093/hmg/ddr150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Lattante S, Rouleau GA, Kabashi E. TARDBP and FUS Mutations Associated With Amyotrophic Lateral Sclerosis: Summary and Update. Hum Mutat. 2013;34:812–826. doi: 10.1002/humu.22319. [DOI] [PubMed] [Google Scholar]
  141. Law WJ, Cann KL, Hicks GG. TLS, EWS and TAF15: a Model for Transcriptional Integration of Gene Expression. Brief Funct Genomic Proteomic. 2006;5:8–14. doi: 10.1093/bfgp/ell015. [DOI] [PubMed] [Google Scholar]
  142. Leblond CS, Kaneb HM, Dion PA, Rouleau GA. Dissection of Genetic Factors Associated With Amyotrophic Lateral Sclerosis. Exp Neurol. 2014 doi: 10.1016/j.expneurol.2014.04.013. [DOI] [PubMed] [Google Scholar]
  143. Lee CT, Chiu YW, Wang KC, Hwang CS, Lin KH, Lee IT, Tsai CP. Riluzole and Prognostic Factors in Amyotrophic Lateral Sclerosis Long-Term and Short-Term Survival: a Population-Based Study of 1149 Cases in Taiwan. J Epidemiol. 2013;23:35–40. doi: 10.2188/jea.JE20120119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Lee HJ, Kim S, Pelletier J, Kim J. Stimulation of HTAFII68 (NTD)-Mediated Transactivation by V-Src. FEBS Lett. 2004;564:188–198. doi: 10.1016/S0014-5793(04)00314-X. [DOI] [PubMed] [Google Scholar]
  145. Li HY, Yeh PA, Chiu HC, Tang CY, Tu BP. Hyperphosphorylation As a Defense Mechanism to Reduce TDP-43 Aggregation. PLoS One. 2011;6:e23075. doi: 10.1371/journal.pone.0023075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Li LB, Yu Z, Teng X, Bonini NM. RNA Toxicity Is a Component of Ataxin-3 Degeneration in Drosophila. Nature. 2008;453:1107–1111. doi: 10.1038/nature06909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Li X, Overton IM, Baines RA, Keegan LP, O'Connell MA. The ADAR RNA Editing Enzyme Controls Neuronal Excitability in Drosophila Melanogaster. Nucleic Acids Res. 2014;42:1139–1151. doi: 10.1093/nar/gkt909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Li Y, Ray P, Rao EJ, Shi C, Guo W, Chen X, Woodruff EA, III, Fushimi K, Wu JY. A Drosophila Model for TDP-43 Proteinopathy. Proc Natl Acad Sci U S A. 2010;107:3169–3174. doi: 10.1073/pnas.0913602107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Li Z, Lu Y, Xu XL, Gao FB. The FTD/ALS-Associated RNA-Binding Protein TDP-43 Regulates the Robustness of Neuronal Specification Through MicroRNA-9a in Drosophila. Hum Mol Genet. 2013;22:218–225. doi: 10.1093/hmg/dds420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Lin MJ, Cheng CW, Shen CK. Neuronal Function and Dysfunction of Drosophila DTDP. PLoS One. 2011;6:e20371. doi: 10.1371/journal.pone.0020371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Ling SC, Albuquerque CP, Han JS, Lagier-Tourenne C, Tokunaga S, Zhou H, Cleveland DW. ALS-Associated Mutations in TDP-43 Increase Its Stability and Promote TDP-43 Complexes With FUS/TLS. Proc Natl Acad Sci U S A. 2010;107:13318–13323. doi: 10.1073/pnas.1008227107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Liu J, Lillo C, Jonsson PA, Vande VC, Ward CM, Miller TM, Subramaniam JR, Rothstein JD, Marklund S, Andersen PM, Brannstrom T, Gredal O, Wong PC, Williams DS, Cleveland DW. Toxicity of Familial ALS-Linked SOD1 Mutants From Selective Recruitment to Spinal Mitochondria. Neuron. 2004;43:5–17. doi: 10.1016/j.neuron.2004.06.016. [DOI] [PubMed] [Google Scholar]
  153. Lloyd TE, Taylor JP. Flightless Flies: Drosophila Models of Neuromuscular Disease. Ann N Y Acad Sci. 2010;1184:e1–20. doi: 10.1111/j.1749-6632.2010.05432.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Logroscino G, Traynor BJ, Hardiman O, Chio' A, Couratier P, Mitchell JD, Swingler RJ, Beghi E. Descriptive Epidemiology of Amyotrophic Lateral Sclerosis: New Evidence and Unsolved Issues. J Neurol Neurosurg Psychiatry. 2008;79:6–11. doi: 10.1136/jnnp.2006.104828. [DOI] [PubMed] [Google Scholar]
  155. Logroscino G, Traynor BJ, Hardiman O, Chio A, Mitchell D, Swingler RJ, Millul A, Benn E, Beghi E. Incidence of Amyotrophic Lateral Sclerosis in Europe. J Neurol Neurosurg Psychiatry. 2010;81:385–390. doi: 10.1136/jnnp.2009.183525. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Lorenzetti D, Bohlega S, Zoghbi HY. The Expansion of the CAG Repeat in Ataxin-2 Is a Frequent Cause of Autosomal Dominant Spinocerebellar Ataxia. Neurology. 1997;49:1009–1013. doi: 10.1212/wnl.49.4.1009. [DOI] [PubMed] [Google Scholar]
  157. Lu Y, Ferris J, Gao FB. Frontotemporal Dementia and Amyotrophic Lateral Sclerosis-Associated Disease Protein TDP-43 Promotes Dendritic Branching. Mol Brain. 2009;2:30. doi: 10.1186/1756-6606-2-30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Lyons TJ, Liu H, Goto JJ, Nersissian A, Roe JA, Graden JA, Cafe C, Ellerby LM, Bredesen DE, Gralla EB, Valentine JS. Mutations in Copper-Zinc Superoxide Dismutase That Cause Amyotrophic Lateral Sclerosis Alter the Zinc Binding Site and the Redox Behavior of the Protein. Proc Natl Acad Sci U S A. 1996;93:12240–12244. doi: 10.1073/pnas.93.22.12240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Mackenzie IR, Bigio EH, Ince PG, Geser F, Neumann M, Cairns NJ, Kwong LK, Forman MS, Ravits J, Stewart H, Eisen A, McClusky L, Kretzschmar HA, Monoranu CM, Highley JR, Kirby J, Siddique T, Shaw PJ, Lee VM, Trojanowski JQ. Pathological TDP-43 Distinguishes Sporadic Amyotrophic Lateral Sclerosis From Amyotrophic Lateral Sclerosis With SOD1 Mutations. Ann Neurol. 2007;61:427–434. doi: 10.1002/ana.21147. [DOI] [PubMed] [Google Scholar]
  160. Majounie E, Renton AE, Mok K, Dopper EG, Waite A, Rollinson S, Chio A, Restagno G, Nicolaou N, Simon-Sanchez J, van Swieten JC, Abramzon Y, Johnson JO, Sendtner M, Pamphlett R, Orrell RW, Mead S, Sidle KC, Houlden H, Rohrer JD, Morrison KE, Pall H, Talbot K, Ansorge O, Hernandez DG, Arepalli S, Sabatelli M, Mora G, Corbo M, Giannini F, Calvo A, Englund E, Borghero G, Floris GL, Remes AM, Laaksovirta H, McCluskey L, Trojanowski JQ, Van Deerlin VM, Schellenberg GD, Nalls MA, Drory VE, Lu CS, Yeh TH, Ishiura H, Takahashi Y, Tsuji S, Le B I, Brice A, Drepper C, Williams N, Kirby J, Shaw P, Hardy J, Tienari PJ, Heutink P, Morris HR, Pickering-Brown S, Traynor BJ. Frequency of the C9orf72 Hexanucleotide Repeat Expansion in Patients With Amyotrophic Lateral Sclerosis and Frontotemporal Dementia: a Cross-Sectional Study. Lancet Neurol. 2012;11:323–330. doi: 10.1016/S1474-4422(12)70043-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Martin I, Jones MA, Grotewiel M. Manipulation of Sod1 Expression Ubiquitously, but Not in the Nervous System or Muscle, Impacts Age-Related Parameters in Drosophila. FEBS Lett. 2009;583:2308–2314. doi: 10.1016/j.febslet.2009.06.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Martini A, La SR, Janssen H, Bilhou-Nabera C, Corveleyn A, Somers R, Aventin A, Foa R, Hagemeijer A, Mecucci C, Marynen P. Recurrent Rearrangement of the Ewing's Sarcoma Gene, EWSR1, or Its Homologue, TAF15, With the Transcription Factor CIZ/NMP4 in Acute Leukemia. Cancer Res. 2002;62:5408–5412. [PubMed] [Google Scholar]
  163. Maruyama H, Morino H, Ito H, Izumi Y, Kato H, Watanabe Y, Kinoshita Y, Kamada M, Nodera H, Suzuki H, Komure O, Matsuura S, Kobatake K, Morimoto N, Abe K, Suzuki N, Aoki M, Kawata A, Hirai T, Kato T, Ogasawara K, Hirano A, Takumi T, Kusaka H, Hagiwara K, Kaji R, Kawakami H. Mutations of Optineurin in Amyotrophic Lateral Sclerosis. Nature. 2010;465:223–226. doi: 10.1038/nature08971. [DOI] [PubMed] [Google Scholar]
  164. May WA, Lessnick SL, Braun BS, Klemsz M, Lewis BC, Lunsford LB, Hromas R, Denny CT. The Ewing's Sarcoma EWS/FLI-1 Fusion Gene Encodes a More Potent Transcriptional Activator and Is a More Powerful Transforming Gene Than FLI-1. Mol Cell Biol. 1993;13:7393–7398. doi: 10.1128/mcb.13.12.7393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. McGuire SE, Roman G, Davis RL. Gene Expression Systems in Drosophila: a Synthesis of Time and Space. Trends Genet. 2004;20:384–391. doi: 10.1016/j.tig.2004.06.012. [DOI] [PubMed] [Google Scholar]
  166. Miguel L, Avequin T, Delarue M, Feuillette S, Frebourg T, Campion D, Lecourtois M. Accumulation of Insoluble Forms of FUS Protein Correlates With Toxicity in Drosophila. Neurobiol Aging. 2012;33:1008–1015. doi: 10.1016/j.neurobiolaging.2011.10.008. [DOI] [PubMed] [Google Scholar]
  167. Miguel L, Frebourg T, Campion D, Lecourtois M. Both Cytoplasmic and Nuclear Accumulations of the Protein Are Neurotoxic in Drosophila Models of TDP-43 Proteinopathies. Neurobiol Dis. 2011;41:398–406. doi: 10.1016/j.nbd.2010.10.007. [DOI] [PubMed] [Google Scholar]
  168. Min KT, Benzer S. Spongecake and Eggroll: Two Hereditary Diseases in Drosophila Resemble Patterns of Human Brain Degeneration. Curr Biol. 1997;7:885–888. doi: 10.1016/s0960-9822(06)00378-2. [DOI] [PubMed] [Google Scholar]
  169. Miskiewicz K, Jose LE, Yeshaw WM, Valadas JS, Swerts J, Munck S, Feiguin F, Dermaut B, Verstreken P. HDAC6 Is a Bruchpilot Deacetylase That Facilitates Neurotransmitter Release. Cell Rep. 2014;8:94–102. doi: 10.1016/j.celrep.2014.05.051. [DOI] [PubMed] [Google Scholar]
  170. Mizielinska S, Gronke S, Niccoli T, Ridler CE, Clayton EL, Devoy A, Moens T, Norona FE, Woollacott IO, Pietrzyk J, Cleverley K, Nicoll AJ, Pickering-Brown S, Dols J, Cabecinha M, Hendrich O, Fratta P, Fisher EM, Partridge L, Isaacs AM. C9orf72 Repeat Expansions Cause Neurodegeneration in Drosophila Through Arginine-Rich Proteins. Science. 2014;345:1192–1194. doi: 10.1126/science.1256800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Mohan RD, Dialynas G, Weake VM, Liu J, Martin-Brown S, Florens L, Washburn MP, Workman JL, Abmayr SM. Loss of Drosophila Ataxin-7, a SAGA Subunit, Reduces H2B Ubiquitination and Leads to Neural and Retinal Degeneration. Genes Dev. 2014;28:259–272. doi: 10.1101/gad.225151.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Moran MT, Tare M, Kango-Singh M, Singh A. Homeotic Gene Teashirt (Tsh) Has a Neuroprotective Function in Amyloid-Beta 42 Mediated Neurodegeneration. PLoS One. 2013;8:e80829. doi: 10.1371/journal.pone.0080829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Mori K, Weng SM, Arzberger T, May S, Rentzsch K, Kremmer E, Schmid B, Kretzschmar HA, Cruts M, Van BC, Haass C, Edbauer D. The C9orf72 GGGGCC Repeat Is Translated into Aggregating Dipeptide-Repeat Proteins in FTLD/ALS. Science. 2013;339:1335–1338. doi: 10.1126/science.1232927. [DOI] [PubMed] [Google Scholar]
  174. Mortara P, Chio A, Rosso MG, Leone M, Schiffer D. Motor Neuron Disease in the Province of Turin, Italy, 1966-1980. Survival Analysis in an Unselected Population. J Neurol Sci. 1984;66:165–173. doi: 10.1016/0022-510x(84)90004-2. [DOI] [PubMed] [Google Scholar]
  175. Moustaqim-Barrette A, Lin YQ, Pradhan S, Neely GG, Bellen HJ, Tsuda H. The Amyotrophic Lateral Sclerosis 8 Protein, VAP, Is Required for ER Protein Quality Control. Hum Mol Genet. 2014;23:1975–1989. doi: 10.1093/hmg/ddt594. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Muller HJ. Genetic Variability, Twin Hybrids and Constant Hybrids, in a Case of Balanced Lethal Factors. Genetics. 1918;3:422–499. doi: 10.1093/genetics/3.5.422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Muller HJ. Artificial Transmutation of the Gene. Science. 1927;66:84–87. doi: 10.1126/science.66.1699.84. [DOI] [PubMed] [Google Scholar]
  178. Myers EW, Sutton GG, Delcher AL, Dew IM, Fasulo DP, Flanigan MJ, Kravitz SA, Mobarry CM, Reinert KH, Remington KA, Anson EL, Bolanos RA, Chou HH, Jordan CM, Halpern AL, Lonardi S, Beasley EM, Brandon RC, Chen L, Dunn PJ, Lai Z, Liang Y, Nusskern DR, Zhan M, Zhang Q, Zheng X, Rubin GM, Adams MD, Venter JC. A Whole-Genome Assembly of Drosophila. Science. 2000;287:2196–2204. doi: 10.1126/science.287.5461.2196. [DOI] [PubMed] [Google Scholar]
  179. Nakano Y, Hirayama K, Terao K. Hepatic Ultrastructural Changes and Liver Dysfunction in Amyotrophic Lateral Sclerosis. Arch Neurol. 1987;44:103–106. doi: 10.1001/archneur.1987.00520130079022. [DOI] [PubMed] [Google Scholar]
  180. Neumann M, Rademakers R, Roeber S, Baker M, Kretzschmar HA, Mackenzie IR. A New Subtype of Frontotemporal Lobar Degeneration With FUS Pathology. Brain. 2009;132:2922–2931. doi: 10.1093/brain/awp214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Neumann M, Sampathu DM, Kwong LK, Truax AC, Micsenyi MC, Chou TT, Bruce J, Schuck T, Grossman M, Clark CM, McCluskey LF, Miller BL, Masliah E, Mackenzie IR, Feldman H, Feiden W, Kretzschmar HA, Trojanowski JQ, Lee VM. Ubiquitinated TDP-43 in Frontotemporal Lobar Degeneration and Amyotrophic Lateral Sclerosis. Science. 2006;314:130–133. doi: 10.1126/science.1134108. [DOI] [PubMed] [Google Scholar]
  182. Nishimura AL, Mitne-Neto M, Silva HC, Oliveira JR, Vainzof M, Zatz M. A Novel Locus for Late Onset Amyotrophic Lateral Sclerosis/Motor Neurone Disease Variant at 20q13. J Med Genet. 2004a;41:315–320. doi: 10.1136/jmg.2003.013029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Nishimura AL, Mitne-Neto M, Silva HC, Richieri-Costa A, Middleton S, Cascio D, Kok F, Oliveira JR, Gillingwater T, Webb J, Skehel P, Zatz M. A Mutation in the Vesicle-Trafficking Protein VAPB Causes Late-Onset Spinal Muscular Atrophy and Amyotrophic Lateral Sclerosis. Am J Hum Genet. 2004b;75:822–831. doi: 10.1086/425287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Niwa J, Ishigaki S, Hishikawa N, Yamamoto M, Doyu M, Murata S, Tanaka K, Taniguchi N, Sobue G. Dorfin Ubiquitylates Mutant SOD1 and Prevents Mutant SOD1-Mediated Neurotoxicity. J Biol Chem. 2002;277:36793–36798. doi: 10.1074/jbc.M206559200. [DOI] [PubMed] [Google Scholar]
  185. Novoa I, Zeng H, Harding HP, Ron D. Feedback Inhibition of the Unfolded Protein Response by GADD34-Mediated Dephosphorylation of EIF2alpha. J Cell Biol. 2001;153:1011–1022. doi: 10.1083/jcb.153.5.1011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Nucifora FC, Jr., Li SH, Danoff S, Ullrich A, Ross CA. Molecular Cloning of a CDNA for the Human Inositol 1,4,5-Trisphosphate Receptor Type 1, and the Identification of a Third Alternatively Spliced Variant. Brain Res Mol Brain Res. 1995;32:291–296. doi: 10.1016/0169-328x(95)00089-b. [DOI] [PubMed] [Google Scholar]
  187. O'Toole O, Traynor BJ, Brennan P, Sheehan C, Frost E, Corr B, Hardiman O. Epidemiology and Clinical Features of Amyotrophic Lateral Sclerosis in Ireland Between 1995 and 2004. J Neurol Neurosurg Psychiatry. 2008;79:30–32. doi: 10.1136/jnnp.2007.117788. [DOI] [PubMed] [Google Scholar]
  188. Otomo A, Hadano S, Okada T, Mizumura H, Kunita R, Nishijima H, Showguchi-Miyata J, Yanagisawa Y, Kohiki E, Suga E, Yasuda M, Osuga H, Nishimoto T, Narumiya S, Ikeda JE. ALS2, a Novel Guanine Nucleotide Exchange Factor for the Small GTPase Rab5, Is Implicated in Endosomal Dynamics. Hum Mol Genet. 2003;12:1671–1687. doi: 10.1093/hmg/ddg184. [DOI] [PubMed] [Google Scholar]
  189. Ou SH, Wu F, Harrich D, Garcia-Martinez LF, Gaynor RB. Cloning and Characterization of a Novel Cellular Protein, TDP-43, That Binds to Human Immunodeficiency Virus Type 1 TAR DNA Sequence Motifs. J Virol. 1995;69:3584–3596. doi: 10.1128/jvi.69.6.3584-3596.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Oyanagi K, Yamazaki M, Takahashi H, Watabe K, Wada M, Komori T, Morita T, Mizutani T. Spinal Anterior Horn Cells in Sporadic Amyotrophic Lateral Sclerosis Show Ribosomal Detachment From, and Cisternal Distention of the Rough Endoplasmic Reticulum. Neuropathol Appl Neurobiol. 2008;34:650–658. doi: 10.1111/j.1365-2990.2008.00941.x. [DOI] [PubMed] [Google Scholar]
  191. Pandey UB, Nichols CD. Human Disease Models in Drosophila Melanogaster and the Role of the Fly in Therapeutic Drug Discovery. Pharmacol Rev. 2011;63:411–436. doi: 10.1124/pr.110.003293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Pandey UB, Nie Z, Batlevi Y, McCray BA, Ritson GP, Nedelsky NB, Schwartz SL, DiProspero NA, Knight MA, Schuldiner O, Padmanabhan R, Hild M, Berry DL, Garza D, Hubbert CC, Yao TP, Baehrecke EH, Taylor JP. HDAC6 Rescues Neurodegeneration and Provides an Essential Link Between Autophagy and the UPS. Nature. 2007;447:859–863. doi: 10.1038/nature05853. [DOI] [PubMed] [Google Scholar]
  193. Parker JD, Parker KM, Sohal BH, Sohal RS, Keller L. Decreased Expression of Cu-Zn Superoxide Dismutase 1 in Ants With Extreme Lifespan. Proc Natl Acad Sci U S A. 2004;101:3486–3489. doi: 10.1073/pnas.0400222101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Parkes TL, Elia AJ, Dickinson D, Hilliker AJ, Phillips JP, Boulianne GL. Extension of Drosophila Lifespan by Overexpression of Human SOD1 in Motorneurons. Nat Genet. 1998;19:171–174. doi: 10.1038/534. [DOI] [PubMed] [Google Scholar]
  195. Paronetto MP, Minana B, Valcarcel J. The Ewing Sarcoma Protein Regulates DNA Damage-Induced Alternative Splicing. Mol Cell. 2011;43:353–368. doi: 10.1016/j.molcel.2011.05.035. [DOI] [PubMed] [Google Scholar]
  196. Pasinelli P, Belford ME, Lennon N, Bacskai BJ, Hyman BT, Trotti D, Brown RH., Jr. Amyotrophic Lateral Sclerosis-Associated SOD1 Mutant Proteins Bind and Aggregate With Bcl-2 in Spinal Cord Mitochondria. Neuron. 2004;43:19–30. doi: 10.1016/j.neuron.2004.06.021. [DOI] [PubMed] [Google Scholar]
  197. Peng IF, Wu CF. Drosophila Cacophony Channels: a Major Mediator of Neuronal Ca2+ Currents and a Trigger for K+ Channel Homeostatic Regulation. J Neurosci. 2007;27:1072–1081. doi: 10.1523/JNEUROSCI.4746-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Pennetta G, Hiesinger PR, Fabian-Fine R, Meinertzhagen IA, Bellen HJ. Drosophila VAP-33A Directs Bouton Formation at Neuromuscular Junctions in a Dosage-Dependent Manner. Neuron. 2002;35:291–306. doi: 10.1016/s0896-6273(02)00769-9. [DOI] [PubMed] [Google Scholar]
  199. Phillips JP, Tainer JA, Getzoff ED, Boulianne GL, Kirby K, Hilliker AJ. Subunit-Destabilizing Mutations in Drosophila Copper/Zinc Superoxide Dismutase: Neuropathology and a Model of Dimer Dysequilibrium. Proc Natl Acad Sci U S A. 1995;92:8574–8578. doi: 10.1073/pnas.92.19.8574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Plaitakis A, Caroscio JT. Abnormal Glutamate Metabolism in Amyotrophic Lateral Sclerosis. Ann Neurol. 1987;22:575–579. doi: 10.1002/ana.410220503. [DOI] [PubMed] [Google Scholar]
  201. Pugliatti M, Parish LD, Cossu P, Leoni S, Ticca A, Saddi MV, Ortu E, Traccis S, Borghero G, Puddu R, Chio A, Pirina P. Amyotrophic Lateral Sclerosis in Sardinia, Insular Italy, 1995-2009. J Neurol. 2013;260:572–579. doi: 10.1007/s00415-012-6681-5. [DOI] [PubMed] [Google Scholar]
  202. Pulst SM, Nechiporuk A, Nechiporuk T, Gispert S, Chen XN, Lopes-Cendes I, Pearlman S, Starkman S, Orozco-Diaz G, Lunkes A, DeJong P, Rouleau GA, Auburger G, Korenberg JR, Figueroa C, Sahba S. Moderate Expansion of a Normally Biallelic Trinucleotide Repeat in Spinocerebellar Ataxia Type 2. Nat Genet. 1996;14:269–276. doi: 10.1038/ng1196-269. [DOI] [PubMed] [Google Scholar]
  203. Qiu H, Lee S, Shang Y, Wang WY, Au KF, Kamiya S, Barmada SJ, Finkbeiner S, Lui H, Carlton CE, Tang AA, Oldham MC, Wang H, Shorter J, Filiano AJ, Roberson ED, Tourtellotte WG, Chen B, Tsai LH, Huang EJ. ALS-Associated Mutation FUS-R521C Causes DNA Damage and RNA Splicing Defects. J Clin Invest. 2014;124:981–999. doi: 10.1172/JCI72723. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Rabbitts TH, Forster A, Larson R, Nathan P. Fusion of the Dominant Negative Transcription Regulator CHOP With a Novel Gene FUS by Translocation T(12;16) in Malignant Liposarcoma. Nat Genet. 1993;4:175–180. doi: 10.1038/ng0693-175. [DOI] [PubMed] [Google Scholar]
  205. Ratnaparkhi A, Lawless GM, Schweizer FE, Golshani P, Jackson GR. A Drosophila Model of ALS: Human ALS-Associated Mutation in VAP33A Suggests a Dominant Negative Mechanism. PLoS One. 2008;3:e2334. doi: 10.1371/journal.pone.0002334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Reiter LT, Potocki L, Chien S, Gribskov M, Bier E. A Systematic Analysis of Human Disease-Associated Gene Sequences in Drosophila Melanogaster. Genome Res. 2001;11:1114–1125. doi: 10.1101/gr.169101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Ren J, Jegga AG, Zhang M, Deng J, Liu J, Gordon CB, Aronow BJ, Lu LJ, Zhang B, Ma J. A Drosophila Model of the Neurodegenerative Disease SCA17 Reveals a Role of RBP-J/Su(H) in Modulating the Pathological Outcome. Hum Mol Genet. 2011;20:3424–3436. doi: 10.1093/hmg/ddr251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Renton AE, Chio A, Traynor BJ. State of Play in Amyotrophic Lateral Sclerosis Genetics. Nat Neurosci. 2014;17:17–23. doi: 10.1038/nn.3584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Renton AE, Majounie E, Waite A, Simon-Sanchez J, Rollinson S, Gibbs JR, Schymick JC, Laaksovirta H, van Swieten JC, Myllykangas L, Kalimo H, Paetau A, Abramzon Y, Remes AM, Kaganovich A, Scholz SW, Duckworth J, Ding J, Harmer DW, Hernandez DG, Johnson JO, Mok K, Ryten M, Trabzuni D, Guerreiro RJ, Orrell RW, Neal J, Murray A, Pearson J, Jansen IE, Sondervan D, Seelaar H, Blake D, Young K, Halliwell N, Callister JB, Toulson G, Richardson A, Gerhard A, Snowden J, Mann D, Neary D, Nalls MA, Peuralinna T, Jansson L, Isoviita VM, Kaivorinne AL, Holtta-Vuori M, Ikonen E, Sulkava R, Benatar M, Wuu J, Chio A, Restagno G, Borghero G, Sabatelli M, Heckerman D, Rogaeva E, Zinman L, Rothstein JD, Sendtner M, Drepper C, Eichler EE, Alkan C, Abdullaev Z, Pack SD, Dutra A, Pak E, Hardy J, Singleton A, Williams NM, Heutink P, Pickering-Brown S, Morris HR, Tienari PJ, Traynor BJ. A Hexanucleotide Repeat Expansion in C9ORF72 Is the Cause of Chromosome 9p21-Linked ALS-FTD. Neuron. 2011;72:257–268. doi: 10.1016/j.neuron.2011.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Rincon-Limas DE, Jensen K, Fernandez-Funez P. Drosophila Models of Proteinopathies: the Little Fly That Could. Curr Pharm Des. 2012;18:1108–1122. doi: 10.2174/138161212799315894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  211. Ritson GP, Custer SK, Freibaum BD, Guinto JB, Geffel D, Moore J, Tang W, Winton MJ, Neumann M, Trojanowski JQ, Lee VM, Forman MS, Taylor JP. TDP-43 Mediates Degeneration in a Novel Drosophila Model of Disease Caused by Mutations in VCP/P97. J Neurosci. 2010;30:7729–7739. doi: 10.1523/JNEUROSCI.5894-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Rogelj B, Easton LE, Bogu GK, Stanton LW, Rot G, Curk T, Zupan B, Sugimoto Y, Modic M, Haberman N, Tollervey J, Fujii R, Takumi T, Shaw CE, Ule J. Widespread Binding of FUS Along Nascent RNA Regulates Alternative Splicing in the Brain. Sci Rep. 2012;2:603. doi: 10.1038/srep00603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Rorth P, Szabo K, Bailey A, Laverty T, Rehm J, Rubin GM, Weigmann K, Milan M, Benes V, Ansorge W, Cohen SM. Systematic Gain-of-Function Genetics in Drosophila. Development. 1998;125:1049–1057. doi: 10.1242/dev.125.6.1049. [DOI] [PubMed] [Google Scholar]
  214. Rosen DR, Siddique T, Patterson D, Figlewicz DA, Sapp P, Hentati A, Donaldson D, Goto J, O'Regan JP, Deng HX. Mutations in Cu/Zn Superoxide Dismutase Gene Are Associated With Familial Amyotrophic Lateral Sclerosis. Nature. 1993;362:59–62. doi: 10.1038/362059a0. [DOI] [PubMed] [Google Scholar]
  215. Rothstein JD, Martin LJ, Kuncl RW. Decreased Glutamate Transport by the Brain and Spinal Cord in Amyotrophic Lateral Sclerosis. N Engl J Med. 1992;326:1464–1468. doi: 10.1056/NEJM199205283262204. [DOI] [PubMed] [Google Scholar]
  216. Rothstein JD, Tsai G, Kuncl RW, Clawson L, Cornblath DR, Drachman DB, Pestronk A, Stauch BL, Coyle JT. Abnormal Excitatory Amino Acid Metabolism in Amyotrophic Lateral Sclerosis. Ann Neurol. 1990;28:18–25. doi: 10.1002/ana.410280106. [DOI] [PubMed] [Google Scholar]
  217. Rothstein JD, Van KM, Levey AI, Martin LJ, Kuncl RW. Selective Loss of Glial Glutamate Transporter GLT-1 in Amyotrophic Lateral Sclerosis. Ann Neurol. 1995;38:73–84. doi: 10.1002/ana.410380114. [DOI] [PubMed] [Google Scholar]
  218. Rowland LP. Diagnosis of Amyotrophic Lateral Sclerosis. J Neurol Sci. 1998;160(Suppl 1):S6–24. doi: 10.1016/s0022-510x(98)00193-2. [DOI] [PubMed] [Google Scholar]
  219. Rubin GM, Lewis EB. A Brief History of Drosophila's Contributions to Genome Research. Science. 2000;287:2216–2218. doi: 10.1126/science.287.5461.2216. [DOI] [PubMed] [Google Scholar]
  220. Rubin GM, Spradling AC. Genetic Transformation of Drosophila With Transposable Element Vectors. Science. 1982;218:348–353. doi: 10.1126/science.6289436. [DOI] [PubMed] [Google Scholar]
  221. Rutherford NJ, Zhang YJ, Baker M, Gass JM, Finch NA, Xu YF, Stewart H, Kelley BJ, Kuntz K, Crook RJ, Sreedharan J, Vance C, Sorenson E, Lippa C, Bigio EH, Geschwind DH, Knopman DS, Mitsumoto H, Petersen RC, Cashman NR, Hutton M, Shaw CE, Boylan KB, Boeve B, Graff-Radford NR, Wszolek ZK, Caselli RJ, Dickson DW, Mackenzie IR, Petrucelli L, Rademakers R. Novel Mutations in TARDBP (TDP-43) in Patients With Familial Amyotrophic Lateral Sclerosis. PLoS Genet. 2008;4:e1000193. doi: 10.1371/journal.pgen.1000193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Sabatelli M, Moncada A, Conte A, Lattante S, Marangi G, Luigetti M, Lucchini M, Mirabella M, Romano A, Del GA, Bisogni G, Doronzio PN, Rossini PM, Zollino M. Mutations in the 3' Untranslated Region of FUS Causing FUS Overexpression Are Associated With Amyotrophic Lateral Sclerosis. Hum Mol Genet. 2013;22:4748–4755. doi: 10.1093/hmg/ddt328. [DOI] [PubMed] [Google Scholar]
  223. Sanhueza M, Zechini L, Gillespie T, Pennetta G. Gain-of-Function Mutations in the ALS8 Causative Gene VAPB Have Detrimental Effects on Neurons and Muscles. Biol Open. 2014;3:59–71. doi: 10.1242/bio.20137070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  224. Sanpei K, Takano H, Igarashi S, Sato T, Oyake M, Sasaki H, Wakisaka A, Tashiro K, Ishida Y, Ikeuchi T, Koide R, Saito M, Sato A, Tanaka T, Hanyu S, Takiyama Y, Nishizawa M, Shimizu N, Nomura Y, Segawa M, Iwabuchi K, Eguchi I, Tanaka H, Takahashi H, Tsuji S. Identification of the Spinocerebellar Ataxia Type 2 Gene Using a Direct Identification of Repeat Expansion and Cloning Technique, DIRECT. Nat Genet. 1996;14:277–284. doi: 10.1038/ng1196-277. [DOI] [PubMed] [Google Scholar]
  225. Sasaki S, Iwata M. Ultrastructural Change of Synapses of Betz Cells in Patients With Amyotrophic Lateral Sclerosis. Neurosci Lett. 1999;268:29–32. doi: 10.1016/s0304-3940(99)00374-2. [DOI] [PubMed] [Google Scholar]
  226. Sasayama H, Shimamura M, Tokuda T, Azuma Y, Yoshida T, Mizuno T, Nakagawa M, Fujikake N, Nagai Y, Yamaguchi M. Knockdown of the Drosophila Fused in Sarcoma (FUS) Homologue Causes Deficient Locomotive Behavior and Shortening of Motoneuron Terminal Branches. PLoS One. 2012;7:e39483. doi: 10.1371/journal.pone.0039483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  227. Scaramuzzino C, Monaghan J, Milioto C, Lanson NA, Jr., Maltare A, Aggarwal T, Casci I, Fackelmayer FO, Pennuto M, Pandey UB. Protein Arginine Methyltransferase 1 and 8 Interact With FUS to Modify Its Sub-Cellular Distribution and Toxicity in Vitro and in Vivo. PLoS One. 2013;8:e61576. doi: 10.1371/journal.pone.0061576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  228. Sephton CF, Cenik C, Kucukural A, Dammer EB, Cenik B, Han Y, Dewey CM, Roth FP, Herz J, Peng J, Moore MJ, Yu G. Identification of Neuronal RNA Targets of TDP-43-Containing Ribonucleoprotein Complexes. J Biol Chem. 2011;286:1204–1215. doi: 10.1074/jbc.M110.190884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  229. Shahidullah M, Le Marchand SJ, Fei H, Zhang J, Pandey UB, Dalva MB, Pasinelli P, Levitan IB. Defects in Synapse Structure and Function Precede Motor Neuron Degeneration in Drosophila Models of FUS-Related ALS. J Neurosci. 2013;33:19590–19598. doi: 10.1523/JNEUROSCI.3396-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  230. Shaw PJ, Forrest V, Ince PG, Richardson JP, Wastell HJ. CSF and Plasma Amino Acid Levels in Motor Neuron Disease: Elevation of CSF Glutamate in a Subset of Patients. Neurodegeneration. 1995;4:209–216. doi: 10.1006/neur.1995.0026. [DOI] [PubMed] [Google Scholar]
  231. Shibata N, Asayama K, Hirano A, Kobayashi M. Immunohistochemical Study on Superoxide Dismutases in Spinal Cords From Autopsied Patients With Amyotrophic Lateral Sclerosis. Dev Neurosci. 1996a;18:492–498. doi: 10.1159/000111445. [DOI] [PubMed] [Google Scholar]
  232. Shibata N, Hirano A, Kobayashi M, Sasaki S, Kato T, Matsumoto S, Shiozawa Z, Komori T, Ikemoto A, Umahara T. Cu/Zn Superoxide Dismutase-Like Immunoreactivity in Lewy Body-Like Inclusions of Sporadic Amyotrophic Lateral Sclerosis. Neurosci Lett. 1994;179:149–152. doi: 10.1016/0304-3940(94)90956-3. [DOI] [PubMed] [Google Scholar]
  233. Shibata N, Hirano A, Kobayashi M, Siddique T, Deng HX, Hung WY, Kato T, Asayama K. Intense Superoxide Dismutase-1 Immunoreactivity in Intracytoplasmic Hyaline Inclusions of Familial Amyotrophic Lateral Sclerosis With Posterior Column Involvement. J Neuropathol Exp Neurol. 1996b;55:481–490. doi: 10.1097/00005072-199604000-00011. [DOI] [PubMed] [Google Scholar]
  234. Shieh SY, Bonini NM. Genes and Pathways Affected by CAG-Repeat RNA-Based Toxicity in Drosophila. Hum Mol Genet. 2011;20:4810–4821. doi: 10.1093/hmg/ddr420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  235. Shimamura M, Kyotani A, Azuma Y, Yoshida H, Binh NT, Mizuta I, Yoshida T, Mizuno T, Nakagawa M, Tokuda T, Yamaguchi M. Genetic Link Between Cabeza, a Drosophila Homologue of Fused in Sarcoma (FUS), and the EGFR Signaling Pathway. Exp Cell Res. 2014;326:36–45. doi: 10.1016/j.yexcr.2014.06.004. [DOI] [PubMed] [Google Scholar]
  236. Shulman JM, Feany MB. Genetic Modifiers of Tauopathy in Drosophila. Genetics. 2003;165:1233–1242. doi: 10.1093/genetics/165.3.1233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. Simonsen A, Cumming RC, Lindmo K, Galaviz V, Cheng S, Rusten TE, Finley KD. Genetic Modifiers of the Drosophila Blue Cheese Gene Link Defects in Lysosomal Transport With Decreased Life Span and Altered Ubiquitinated-Protein Profiles. Genetics. 2007;176:1283–1297. doi: 10.1534/genetics.106.065011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  238. Spradling AC, Rubin GM. Transposition of Cloned P Elements into Drosophila Germ Line Chromosomes. Science. 1982;218:341–347. doi: 10.1126/science.6289435. [DOI] [PubMed] [Google Scholar]
  239. Spreux-Varoquaux O, Bensimon G, Lacomblez L, Salachas F, Pradat PF, Le FN, Marouan A, Dib M, Meininger V. Glutamate Levels in Cerebrospinal Fluid in Amyotrophic Lateral Sclerosis: a Reappraisal Using a New HPLC Method With Coulometric Detection in a Large Cohort of Patients. J Neurol Sci. 2002;193:73–78. doi: 10.1016/s0022-510x(01)00661-x. [DOI] [PubMed] [Google Scholar]
  240. Sreedharan J, Blair IP, Tripathi VB, Hu X, Vance C, Rogelj B, Ackerley S, Durnall JC, Williams KL, Buratti E, Baralle F, de BJ, Mitchell JD, Leigh PN, Al-Chalabi A, Miller CC, Nicholson G, Shaw CE. TDP-43 Mutations in Familial and Sporadic Amyotrophic Lateral Sclerosis. Science. 2008;319:1668–1672. doi: 10.1126/science.1154584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  241. Steffan JS, Bodai L, Pallos J, Poelman M, McCampbell A, Apostol BL, Kazantsev A, Schmidt E, Zhu YZ, Greenwald M, Kurokawa R, Housman DE, Jackson GR, Marsh JL, Thompson LM. Histone Deacetylase Inhibitors Arrest Polyglutamine-Dependent Neurodegeneration in Drosophila. Nature. 2001;413:739–743. doi: 10.1038/35099568. [DOI] [PubMed] [Google Scholar]
  242. Stolow DT, Haynes SR. Cabeza, a Drosophila Gene Encoding a Novel RNA Binding Protein, Shares Homology With EWS and TLS, Two Genes Involved in Human Sarcoma Formation. Nucleic Acids Res. 1995;23:835–843. doi: 10.1093/nar/23.5.835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  243. Sturtevant AH. The Linear Arrangement of Six Sex-Linked Factors in Drosophila, As Shown by Their Mode of Association. Journal of Experimental Zoology. 1913;14:43–59. [Google Scholar]
  244. Sturtevant AH. A History of Genetics. Cold Spring Harbor Laboratory Press; New York: 1965. [Google Scholar]
  245. Sun J, Tower J. FLP Recombinase-Mediated Induction of Cu/Zn-Superoxide Dismutase Transgene Expression Can Extend the Life Span of Adult Drosophila Melanogaster Flies. Mol Cell Biol. 1999;19:216–228. doi: 10.1128/mcb.19.1.216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  246. Swerdlow RH, Parks JK, Cassarino DS, Trimmer PA, Miller SW, Maguire DJ, Sheehan JP, Maguire RS, Pattee G, Juel VC, Phillips LH, Tuttle JB, Bennett JP, Jr., Davis RE, Parker WD., Jr. Mitochondria in Sporadic Amyotrophic Lateral Sclerosis. Exp Neurol. 1998;153:135–142. doi: 10.1006/exnr.1998.6866. [DOI] [PubMed] [Google Scholar]
  247. Takayama Y, Itoh RE, Tsuyama T, Uemura T. Age-Dependent Deterioration of Locomotion in Drosophila Melanogaster Deficient in the Homologue of Amyotrophic Lateral Sclerosis 2. Genes Cells. 2014 doi: 10.1111/gtc.12146. [DOI] [PubMed] [Google Scholar]
  248. Takuma H, Kwak S, Yoshizawa T, Kanazawa I. Reduction of GluR2 RNA Editing, a Molecular Change That Increases Calcium Influx Through AMPA Receptors, Selective in the Spinal Ventral Gray of Patients With Amyotrophic Lateral Sclerosis. Ann Neurol. 1999;46:806–815. doi: 10.1002/1531-8249(199912)46:6<806::aid-ana2>3.0.co;2-s. [DOI] [PubMed] [Google Scholar]
  249. Tan AY, Manley JL. TLS Inhibits RNA Polymerase III Transcription. Mol Cell Biol. 2010;30:186–196. doi: 10.1128/MCB.00884-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  250. Tan CF, Eguchi H, Tagawa A, Onodera O, Iwasaki T, Tsujino A, Nishizawa M, Kakita A, Takahashi H. TDP-43 Immunoreactivity in Neuronal Inclusions in Familial Amyotrophic Lateral Sclerosis With or Without SOD1 Gene Mutation. Acta Neuropathol. 2007;113:535–542. doi: 10.1007/s00401-007-0206-9. [DOI] [PubMed] [Google Scholar]
  251. Tollervey JR, Curk T, Rogelj B, Briese M, Cereda M, Kayikci M, Konig J, Hortobagyi T, Nishimura AL, Zupunski V, Patani R, Chandran S, Rot G, Zupan B, Shaw CE, Ule J. Characterizing the RNA Targets and Position-Dependent Splicing Regulation by TDP-43. Nat Neurosci. 2011;14:452–458. doi: 10.1038/nn.2778. [DOI] [PMC free article] [PubMed] [Google Scholar]
  252. Tortarolo M, Veglianese P, Calvaresi N, Botturi A, Rossi C, Giorgini A, Migheli A, Bendotti C. Persistent Activation of P38 Mitogen-Activated Protein Kinase in a Mouse Model of Familial Amyotrophic Lateral Sclerosis Correlates With Disease Progression. Mol Cell Neurosci. 2003;23:180–192. doi: 10.1016/s1044-7431(03)00022-8. [DOI] [PubMed] [Google Scholar]
  253. Tradewell ML, Yu Z, Tibshirani M, Boulanger MC, Durham HD, Richard S. Arginine Methylation by PRMT1 Regulates Nuclear-Cytoplasmic Localization and Toxicity of FUS/TLS Harbouring ALS-Linked Mutations. Hum Mol Genet. 2012;21:136–149. doi: 10.1093/hmg/ddr448. [DOI] [PubMed] [Google Scholar]
  254. Trotti D, Aoki M, Pasinelli P, Berger UV, Danbolt NC, Brown RH, Jr., Hediger MA. Amyotrophic Lateral Sclerosis-Linked Glutamate Transporter Mutant Has Impaired Glutamate Clearance Capacity. J Biol Chem. 2001;276:576–582. doi: 10.1074/jbc.M003779200. [DOI] [PubMed] [Google Scholar]
  255. Trotti D, Rolfs A, Danbolt NC, Brown RH, Jr., Hediger MA. SOD1 Mutants Linked to Amyotrophic Lateral Sclerosis Selectively Inactivate a Glial Glutamate Transporter. Nat Neurosci. 1999;2:427–433. doi: 10.1038/8091. [DOI] [PubMed] [Google Scholar]
  256. Tsuda H, Han SM, Yang Y, Tong C, Lin YQ, Mohan K, Haueter C, Zoghbi A, Harati Y, Kwan J, Miller MA, Bellen HJ. The Amyotrophic Lateral Sclerosis 8 Protein VAPB Is Cleaved, Secreted, and Acts As a Ligand for Eph Receptors. Cell. 2008;133:963–977. doi: 10.1016/j.cell.2008.04.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  257. Turner BJ, Talbot K. Transgenics, Toxicity and Therapeutics in Rodent Models of Mutant SOD1-Mediated Familial ALS. Prog Neurobiol. 2008;85:94–134. doi: 10.1016/j.pneurobio.2008.01.001. [DOI] [PubMed] [Google Scholar]
  258. Uranishi H, Tetsuka T, Yamashita M, Asamitsu K, Shimizu M, Itoh M, Okamoto T. Involvement of the Pro-Oncoprotein TLS (Translocated in Liposarcoma) in Nuclear Factor-Kappa B P65-Mediated Transcription As a Coactivator. J Biol Chem. 2001;276:13395–13401. doi: 10.1074/jbc.M011176200. [DOI] [PubMed] [Google Scholar]
  259. Van Deerlin VM, Leverenz JB, Bekris LM, Bird TD, Yuan W, Elman LB, Clay D, Wood EM, Chen-Plotkin AS, Martinez-Lage M, Steinbart E, McCluskey L, Grossman M, Neumann M, Wu IL, Yang WS, Kalb R, Galasko DR, Montine TJ, Trojanowski JQ, Lee VM, Schellenberg GD, Yu CE. TARDBP Mutations in Amyotrophic Lateral Sclerosis With TDP-43 Neuropathology: a Genetic and Histopathological Analysis. Lancet Neurol. 2008;7:409–416. doi: 10.1016/S1474-4422(08)70071-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  260. van BM, van Es MA, Koppers M, van RW, Medic J, Schelhaas HJ, van der Kooi AJ, de VM, Veldink JH, van den Berg LH. VAPB and C9orf72 Mutations in 1 Familial Amyotrophic Lateral Sclerosis Patient. Neurobiol Aging. 2012;33:2950–2954. doi: 10.1016/j.neurobiolaging.2012.07.004. [DOI] [PubMed] [Google Scholar]
  261. Vance C, Rogelj B, Hortobagyi T, De Vos KJ, Nishimura AL, Sreedharan J, Hu X, Smith B, Ruddy D, Wright P, Ganesalingam J, Williams KL, Tripathi V, Al-Saraj S, Al-Chalabi A, Leigh PN, Blair IP, Nicholson G, de BJ, Gallo JM, Miller CC, Shaw CE. Mutations in FUS, an RNA Processing Protein, Cause Familial Amyotrophic Lateral Sclerosis Type 6. Science. 2009;323:1208–1211. doi: 10.1126/science.1165942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  262. Vanden Broeck L, Naval-Sanchez M, Adachi Y, Diaper D, Dourlen P, Chapuis J, Kleinberger G, Gistelinck M, Van BC, Lambert JC, Hirth F, Aerts S, Callaerts P, Dermaut B. TDP-43 Loss-of-Function Causes Neuronal Loss Due to Defective Steroid Receptor-Mediated Gene Program Switching in Drosophila. Cell Rep. 2013;3:160–172. doi: 10.1016/j.celrep.2012.12.014. [DOI] [PubMed] [Google Scholar]
  263. Verdel A, Curtet S, Brocard MP, Rousseaux S, Lemercier C, Yoshida M, Khochbin S. Active Maintenance of MHDA2/MHDAC6 Histone-Deacetylase in the Cytoplasm. Curr Biol. 2000;10:747–749. doi: 10.1016/s0960-9822(00)00542-x. [DOI] [PubMed] [Google Scholar]
  264. Voigt A, Herholz D, Fiesel FC, Kaur K, Muller D, Karsten P, Weber SS, Kahle PJ, Marquardt T, Schulz JB. TDP-43-Mediated Neuron Loss in Vivo Requires RNA-Binding Activity. PLoS One. 2010;5:e12247. doi: 10.1371/journal.pone.0012247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Volkening K, Leystra-Lantz C, Yang W, Jaffee H, Strong MJ. Tar DNA Binding Protein of 43 KDa (TDP-43), 14-3-3 Proteins and Copper/Zinc Superoxide Dismutase (SOD1) Interact to Modulate NFL MRNA Stability. Implications for Altered RNA Processing in Amyotrophic Lateral Sclerosis (ALS) Brain Res. 2009;1305:168–182. doi: 10.1016/j.brainres.2009.09.105. [DOI] [PubMed] [Google Scholar]
  266. Wang IF, Wu LS, Shen CK. TDP-43: an Emerging New Player in Neurodegenerative Diseases. Trends Mol Med. 2008a;14:479–485. doi: 10.1016/j.molmed.2008.09.001. [DOI] [PubMed] [Google Scholar]
  267. Wang JW, Brent JR, Tomlinson A, Shneider NA, McCabe BD. The ALS-Associated Proteins FUS and TDP-43 Function Together to Affect Drosophila Locomotion and Life Span. J Clin Invest. 2011;121:4118–4126. doi: 10.1172/JCI57883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  268. Wang L, Deng HX, Grisotti G, Zhai H, Siddique T, Roos RP. Wild-Type SOD1 Overexpression Accelerates Disease Onset of a G85R SOD1 Mouse. Hum Mol Genet. 2009;18:1642–1651. doi: 10.1093/hmg/ddp085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. Wang X, Arai S, Song X, Reichart D, Du K, Pascual G, Tempst P, Rosenfeld MG, Glass CK, Kurokawa R. Induced NcRNAs Allosterically Modify RNA-Binding Proteins in Cis to Inhibit Transcription. Nature. 2008b;454:126–130. doi: 10.1038/nature06992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. Warrick JM, Chan HY, Gray-Board, Chai Y, Paulson HL, Bonini NM. Suppression of Polyglutamine-Mediated Neurodegeneration in Drosophila by the Molecular Chaperone HSP70. Nat Genet. 1999;23:425–428. doi: 10.1038/70532. [DOI] [PubMed] [Google Scholar]
  271. Warrick JM, Paulson HL, Gray-Board, Bui QT, Fischbeck KH, Pittman RN, Bonini NM. Expanded Polyglutamine Protein Forms Nuclear Inclusions and Causes Neural Degeneration in Drosophila. Cell. 1998;93:939–949. doi: 10.1016/s0092-8674(00)81200-3. [DOI] [PubMed] [Google Scholar]
  272. Watanabe M, Dykes-Hoberg M, Culotta VC, Price DL, Wong PC, Rothstein JD. Histological Evidence of Protein Aggregation in Mutant SOD1 Transgenic Mice and in Amyotrophic Lateral Sclerosis Neural Tissues. Neurobiol Dis. 2001;8:933–941. doi: 10.1006/nbdi.2001.0443. [DOI] [PubMed] [Google Scholar]
  273. Watson MR, Lagow RD, Xu K, Zhang B, Bonini NM. A Drosophila Model for Amyotrophic Lateral Sclerosis Reveals Motor Neuron Damage by Human SOD1. J Biol Chem. 2008;283:24972–24981. doi: 10.1074/jbc.M804817200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  274. Watts GD, Wymer J, Kovach MJ, Mehta SG, Mumm S, Darvish D, Pestronk A, Whyte MP, Kimonis VE. Inclusion Body Myopathy Associated With Paget Disease of Bone and Frontotemporal Dementia Is Caused by Mutant Valosin-Containing Protein. Nat Genet. 2004;36:377–381. doi: 10.1038/ng1332. [DOI] [PubMed] [Google Scholar]
  275. Wegorzewska I, Bell S, Cairns NJ, Miller TM, Baloh RH. TDP-43 Mutant Transgenic Mice Develop Features of ALS and Frontotemporal Lobar Degeneration. Proc Natl Acad Sci U S A. 2009;106:18809–18814. doi: 10.1073/pnas.0908767106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  276. Wiedemann FR, Manfredi G, Mawrin C, Beal MF, Schon EA. Mitochondrial DNA and Respiratory Chain Function in Spinal Cords of ALS Patients. J Neurochem. 2002;80:616–625. doi: 10.1046/j.0022-3042.2001.00731.x. [DOI] [PubMed] [Google Scholar]
  277. Wiedemann FR, Winkler K, Kuznetsov AV, Bartels C, Vielhaber S, Feistner H, Kunz WS. Impairment of Mitochondrial Function in Skeletal Muscle of Patients With Amyotrophic Lateral Sclerosis. J Neurol Sci. 1998;156:65–72. doi: 10.1016/s0022-510x(98)00008-2. [DOI] [PubMed] [Google Scholar]
  278. Wils H, Kleinberger G, Janssens J, Pereson S, Joris G, Cuijt I, Smits V, Ceuterick-de GC, Van BC, Kumar-Singh S. TDP-43 Transgenic Mice Develop Spastic Paralysis and Neuronal Inclusions Characteristic of ALS and Frontotemporal Lobar Degeneration. Proc Natl Acad Sci U S A. 2010;107:3858–3863. doi: 10.1073/pnas.0912417107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  279. Winton MJ, Van Deerlin VM, Kwong LK, Yuan W, Wood EM, Yu CE, Schellenberg GD, Rademakers R, Caselli R, Karydas A, Trojanowski JQ, Miller BL, Lee VM. A90V TDP-43 Variant Results in the Aberrant Localization of TDP-43 in Vitro. FEBS Lett. 2008;582:2252–2256. doi: 10.1016/j.febslet.2008.05.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  280. Wishart TM, Rooney TM, Lamont DJ, Wright AK, Morton AJ, Jackson M, Freeman MR, Gillingwater TH. Combining Comparative Proteomics and Molecular Genetics Uncovers Regulators of Synaptic and Axonal Stability and Degeneration in Vivo. PLoS Genet. 2012;8:e1002936. doi: 10.1371/journal.pgen.1002936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  281. Wu CH, Fallini C, Ticozzi N, Keagle PJ, Sapp PC, Piotrowska K, Lowe P, Koppers M, McKenna-Yasek D, Baron DM, Kost JE, Gonzalez-Perez P, Fox AD, Adams J, Taroni F, Tiloca C, Leclerc AL, Chafe SC, Mangroo D, Moore MJ, Zitzewitz JA, Xu ZS, van den Berg LH, Glass JD, Siciliano G, Cirulli ET, Goldstein DB, Salachas F, Meininger V, Rossoll W, Ratti A, Gellera C, Bosco DA, Bassell GJ, Silani V, Drory VE, Brown RH, Jr., Landers JE. Mutations in the Profilin 1 Gene Cause Familial Amyotrophic Lateral Sclerosis. Nature. 2012a;488:499–503. doi: 10.1038/nature11280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  282. Wu LS, Cheng WC, Shen CK. Targeted Depletion of TDP-43 Expression in the Spinal Cord Motor Neurons Leads to the Development of Amyotrophic Lateral Sclerosis-Like Phenotypes in Mice. J Biol Chem. 2012b;287:27335–27344. doi: 10.1074/jbc.M112.359000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  283. Xia R, Liu Y, Yang L, Gal J, Zhu H, Jia J. Motor Neuron Apoptosis and Neuromuscular Junction Perturbation Are Prominent Features in a Drosophila Model of Fus-Mediated ALS. Mol Neurodegener. 2012;7:10. doi: 10.1186/1750-1326-7-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  284. Xu Z, Poidevin M, Li X, Li Y, Shu L, Nelson DL, Li H, Hales CM, Gearing M, Wingo TS, Jin P. Expanded GGGGCC Repeat RNA Associated With Amyotrophic Lateral Sclerosis and Frontotemporal Dementia Causes Neurodegeneration. Proc Natl Acad Sci U S A. 2013;110:7778–7783. doi: 10.1073/pnas.1219643110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  285. Yadav S, Tapadia MG. Neurodegeneration Caused by Polyglutamine Expansion Is Regulated by P-Glycoprotein in Drosophila Melanogaster. Genetics. 2013;195:857–870. doi: 10.1534/genetics.113.155077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Yamaguchi A, Kitajo K. The Effect of PRMT1-Mediated Arginine Methylation on the Subcellular Localization, Stress Granules, and Detergent-Insoluble Aggregates of FUS/TLS. PLoS One. 2012;7:e49267. doi: 10.1371/journal.pone.0049267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  287. Yamanaka K, Vande VC, Eymard-Pierre E, Bertini E, Boespflug-Tanguy O, Cleveland DW. Unstable Mutants in the Peripheral Endosomal Membrane Component ALS2 Cause Early-Onset Motor Neuron Disease. Proc Natl Acad Sci U S A. 2003;100:16041–16046. doi: 10.1073/pnas.2635267100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  288. Yang L, Embree LJ, Tsai S, Hickstein DD. Oncoprotein TLS Interacts With Serine-Arginine Proteins Involved in RNA Splicing. J Biol Chem. 1998;273:27761–27764. doi: 10.1074/jbc.273.43.27761. [DOI] [PubMed] [Google Scholar]
  289. Yang Y, Hentati A, Deng HX, Dabbagh O, Sasaki T, Hirano M, Hung WY, Ouahchi K, Yan J, Azim AC, Cole N, Gascon G, Yagmour A, Ben-Hamida M, Pericak-Vance M, Hentati F, Siddique T. The Gene Encoding Alsin, a Protein With Three Guanine-Nucleotide Exchange Factor Domains, Is Mutated in a Form of Recessive Amyotrophic Lateral Sclerosis. Nat Genet. 2001;29:160–165. doi: 10.1038/ng1001-160. [DOI] [PubMed] [Google Scholar]
  290. Yang Z, Huh SU, Drennan JM, Kathuria H, Martinez JS, Tsuda H, Hall MC, Clemens JC. Drosophila Vap-33 Is Required for Axonal Localization of Dscam Isoforms. J Neurosci. 2012;32:17241–17250. doi: 10.1523/JNEUROSCI.2834-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  291. Zhan L, Hanson KA, Kim SH, Tare A, Tibbetts RS. Identification of Genetic Modifiers of TDP-43 Neurotoxicity in Drosophila. PLoS One. 2013;8:e57214. doi: 10.1371/journal.pone.0057214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  292. Zhang YJ, Jansen-West K, Xu YF, Gendron TF, Bieniek KF, Lin WL, Sasaguri H, Caulfield T, Hubbard J, Daughrity L, Chew J, Belzil VV, Prudencio M, Stankowski JN, Castanedes-Casey M, Whitelaw E, Ash PE, DeTure M, Rademakers R, Boylan KB, Dickson DW, Petrucelli L. Aggregation-Prone C9FTD/ALS Poly(GA) RAN-Translated Proteins Cause Neurotoxicity by Inducing ER Stress. Acta Neuropathol. 2014;128:505–524. doi: 10.1007/s00401-014-1336-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  293. Zhang YJ, Xu YF, Cook C, Gendron TF, Roettges P, Link CD, Lin WL, Tong J, Castanedes-Casey M, Ash P, Gass J, Rangachari V, Buratti E, Baralle F, Golde TE, Dickson DW, Petrucelli L. Aberrant Cleavage of TDP-43 Enhances Aggregation and Cellular Toxicity. Proc Natl Acad Sci U S A. 2009;106:7607–7612. doi: 10.1073/pnas.0900688106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  294. Zhou H, Huang C, Chen H, Wang D, Landel CP, Xia PY, Bowser R, Liu YJ, Xia XG. Transgenic Rat Model of Neurodegeneration Caused by Mutation in the TDP Gene. PLoS Genet. 2010;6:e1000887. doi: 10.1371/journal.pgen.1000887. [DOI] [PMC free article] [PubMed] [Google Scholar]
  295. Zinszner H, Sok J, Immanuel D, Yin Y, Ron D. TLS (FUS) Binds RNA in Vivo and Engages in Nucleo-Cytoplasmic Shuttling. J Cell Sci. 1997;110:1741–1750. doi: 10.1242/jcs.110.15.1741. Pt 15. [DOI] [PubMed] [Google Scholar]
  296. Zu T, Gibbens B, Doty NS, Gomes-Pereira M, Huguet A, Stone MD, Margolis J, Peterson M, Markowski TW, Ingram MA, Nan Z, Forster C, Low WC, Schoser B, Somia NV, Clark HB, Schmechel S, Bitterman PB, Gourdon G, Swanson MS, Moseley M, Ranum LP. Non-ATG-Initiated Translation Directed by Microsatellite Expansions. Proc Natl Acad Sci U S A. 2011;108:260–265. doi: 10.1073/pnas.1013343108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  297. Zu T, Liu Y, Banez-Coronel M, Reid T, Pletnikova O, Lewis J, Miller TM, Harms MB, Falchook AE, Subramony SH, Ostrow LW, Rothstein JD, Troncoso JC, Ranum LP. RAN Proteins and RNA Foci From Antisense Transcripts in C9ORF72 ALS and Frontotemporal Dementia. Proc Natl Acad Sci U S A. 2013;110:E4968–E4977. doi: 10.1073/pnas.1315438110. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES