Skip to main content
Frontiers in Cellular Neuroscience logoLink to Frontiers in Cellular Neuroscience
. 2015 Jun 16;9:225. doi: 10.3389/fncel.2015.00225

Calcium dysregulation links ALS defective proteins and motor neuron selective vulnerability

Sónia S Leal 1,2, Cláudio M Gomes 1,2,*
PMCID: PMC4468822  PMID: 26136661

Abstract

More than 20 distinct gene loci have so far been implicated in amyotrophic lateral sclerosis (ALS), a fatal neurodegenerative disorder characterized by progressive neurodegeneration of motor neurons (MN) and death. Most of this distinct set of ALS-related proteins undergoes toxic deposition specifically in MN for reasons which remain unclear. Here we overview a recent body of evidence indicative that mutations in ALS-related proteins can disrupt fundamental Ca2+ signalling pathways in MN, and that Ca2+ itself impacts both directly or indirectly in many ALS critical proteins and cellular processes that result in MN neurodegeneration. We argue that the inherent vulnerability of MN to dysregulation of intracellular Ca2+ is deeply associated with discriminating pathogenicity and aberrant crosstalk of most of the critical proteins involved in ALS. Overall, Ca2+ deregulation in MN is at the cornerstone of different ALS processes and is likely one of the factors contributing to the selective susceptibility of these cells to this particular neurodegenerative disease.

Keywords: neurodegenerative diseases, calcium homeostasis, ALS, proteinopathies, SOD1

Sporadic and Familial ALS Aggregates Share Identical Proteins

Amyotrophic lateral sclerosis (ALS) is a fatal neurodegenerative disease characterized by the selective degeneration of motor neurons (MN) in the spinal cord, brainstem and cerebral cortex (Rowland and Shneider, 2001). Most cases of ALS are sporadic (sALS) with no known genetic linkage, while approximately 10% are associated with familial forms (fALS), presenting mutations in over 20 genes encoding for distinct proteins with varied functions (Table 1). Despite the heterogeneous genetics of fALS and unknown etiology of sALS and its clear multifactorial character, most ALS patients present similar phenotypes with formation of cytoplasmic proteinaceous aggregates in the affected MN (Al-Chalabi et al., 2012). Notably, many proteins involved in fALS forms are also found in sALS toxic aggregates (Maekawa et al., 2009; Deng et al., 2010; Forsberg et al., 2010; Blokhuis et al., 2013) and found to cross-talk and impact on each other in ALS pathology (Kanekura et al., 2004; Volkening et al., 2009; Tudor et al., 2010; Nihei et al., 2012; Pokrishevsky et al., 2012; Stoica et al., 2014; Osaka et al., 2015). This suggests that apart from mutations, additional chemical and/or biological factors influence the selective involvement of these proteins also in sALS neurodegeneration. In agreement, most of the proteins which are involved in ALS are ordinarily expressed in many distinct cell tissues other than the nervous system (e.g., SOD1, FUS, TDP-43, VAPB, matrin-3, ataxin-2, alsin) but are only found to generate toxicity among MN. The selective vulnerability of such cells suggests that environmental triggers within those neurons are mandatory for the onset of ALS.

Table 1.

Heterogeneity of fALS causative genes.

Protein (% prevalence) Gene/Protein Function
FTD/ALS (40–50%) (DeJesus-Hernandez et al., 2011) C9ORF72 Suggested to regulate endosomal trafficking and autophagy
ALS 1 (20%) (Rosen et al., 1993) SOD (SOD1) Antioxidant; scavenging of superoxide
ALS 10 (5%) (Sreedharan et al., 2008) TARDBP (TDP-43) RNA and DNA regulation
ALS 6 (4%) (Vance et al., 2009) FUS (FUS) RNA and DNA regulation
ALS 13 (>1%) (Elden et al., 2010) ATXN2 (Ataxin-2) Suggested to regulate RNA processing
Other (less prevalent fALS) (Chen et al., 2004; Nishimura et al., 2004; Johnson et al., 2010; Maruyama et al., 2010; Wu et al., 2012; Iguchi et al., 2013; Su et al., 2014) ALS2, ANG, CHMP2B, DCTN1, EWSR1, ERBB4, FIG4, hnRNPA1, MATR3 (Matrin-3), OPTN, PFN1, SETX, SIGMAR1, SQSTM1, SPG11, TAF15, UBQ2N2, VAPB, VCP Protein degradation; ER-Golgi pathways; Trafficking; Endosomal sorting complexes required for transport; DNA and RNA processing; Actin dynamics; Mitogenesis and differentiation

Proteins are indicated in parenthesis for some cases: SOD1, Superoxide Dismutase 1; TDP-43, TAR DNA-binding protein 43; FUS, Fused in Sarcoma protein.

Calcium Dysregulation in ALS-Affected Motor Neurons—The Factual Case of ALS1

A particular feature that distinguishes ALS affected MN from other cells relates to their inherent vulnerability to Ca2+ overload; indeed, these neurons highly express Ca2+ permeable AMPA receptors (Williams et al., 1997; Shaw and Eggett, 2000; Van Den Bosch et al., 2000; Vandenberghe et al., 2000; Guatteo et al., 2007) concurrently with a low Ca2+ buffering capacity due to endogenous low expression of Ca2+ buffering proteins (CaBPs) such as parvalbumin and calbindin (Alexianu et al., 1994; Palecek et al., 1999; Jaiswal, 2013) albeit the presence of EF-hand Ca2+ binding proteins in MN (Migheli et al., 1999; Zhang et al., 2014). This combination of inherent physiological features of MN to manage Ca2+ levels, though essential for normal functioning (von Lewinski and Keller, 2005), are likely a predisposition risk for the systematic intracellular Ca2+ overload that is detected in ALS1 affected MN (Siklós et al., 1996, 1998; Kruman et al., 1999; Grosskreutz et al., 2010; Kawamata and Manfredi, 2010). Interestingly, the levels of calretinin and parvalbumin in MN axons is found further decreased in ALS patients (Hayashi et al., 2013), thus establishing an increased deficit in MN Ca2+ buffering capability under pathological conditions.

In fact, a direct outcome of the low expression of CaBPs in MN is that mitochondria are likely to assume a major role in buffering calcium in these cells. In agreement, it might not be a coincidence that in ALS1, mutated SOD1 was shown to abnormally accumulate in the mitochondrial intermembrane space (Jaarsma et al., 2001; Liu et al., 2004), affecting mitochondrial function leading to disturbance of Ca2+ homeostasis and ERMCC cycle (Jaiswal and Keller, 2009; Lautenschläger et al., 2013). Moreover, studies on cellular and animal ALS1 models have shown that Ca2+ overload is linked with SOD1 aggregation (Tateno et al., 2004; Tradewell et al., 2011). Indeed, we have recently shown that Ca2+ can bind to SOD1 immature states promoting its aggregation (Leal et al., 2013; Estácio et al., 2015) thus establishing an additional ALS1 pathological pathway for the impact of Ca2+ overload on SOD1 toxic deposition. In agreement, a decrease of intracellular Ca2+ overload in ALS1 models through AMPA channel antagonists or overexpression of calcium-buffering proteins has been shown to reduce SOD1 toxic aggregation and neurodegeneration (Beers et al., 2001; Van Damme et al., 2003; Tateno et al., 2004; Tortarolo et al., 2006; Yin et al., 2007; Parone et al., 2013).

Crossways of ALS Critical Proteins with Calcium

In spite of the scarcity of available data regarding more recently discovered fALS models linked with systematic Ca2+ deregulation, compelling evidence argues that many critical proteins involved in ALS (other than SOD1), are directly or indirectly involved with Ca2+. It is noteworthy that in such instances, ALS-associated mutations in many of these ALS critical proteins happen to potentiate Ca2+ deregulation and/or result in an increased vulnerability to the effects of Ca2+. For example, VAP-B which is involved in ALS8, is a ER-membrane MAM protein directly engaged in Ca2+ exchange between the ER and mitochondria (De Vos et al., 2012). The ALS VAPBP56S variant was shown to disrupt Ca2+ homeostasis leading to a perturbation of the anterograde mitochondrial axonal transport and affecting the Miro1/kinesin-1 interaction with tubulin (Mórotz et al., 2012). Alsin, a protein implicated in juvenile ALS2, is involved in endossome/membrane trafficking that undergoes Ca2+ dependent binding to the NCS regulating neurocalcin alpha protein (Masutani et al., 2008). This suggests that alsin membrane binding might be a Ca2+ dependent process, and therefore passible to become affected by dysregulation of Ca2+ levels. Moreover, alsin is also found to play a role in AMPAR trafficking, where ALS2 mutations lead to distinct subcellular GRIP1 localization and reduction of the calcium-impermeable GluR2 containing AMPA receptors, thus likely rendering neurons susceptible to deviant Ca2+ influxes (Lai et al., 2006). Matrin 3, a multifunctional nuclear matrix protein involved in ALS21, is suggested to be regulated through a Ca2+ dependent interaction with CaM (Valencia et al., 2007) and is therefore likely to be affected by deregulated Ca2+ levels. As only very recently matrin 3 has been implicated in ALS (Johnson et al., 2014), future studies will be needed to clarify this possibility. Ataxin-2, which is involved in ALS13 is an ubiquitous cytoplasmic protein proposed to induce defects in the ER–Golgi pathway and disrupt Ca2+ signalling (van den Heuvel et al., 2014). The possibility that ataxin-2 influences ER–Golgi function is inferred from a recent study where it was shown that intermediate-length polyQ expansions in ataxin-2 mutants enhance FUS-induced ER stress and Golgi fragmentation (Farg et al., 2013). The suggestion that the ALS13 related intermediate-length polyQ expansions in ataxin-2 can lead to Ca2+ signalling disruption derives from the causal association that polyQ-expanded forms of ataxin-2, or huntingtin and ataxin-3 in other diseases, associates with the C-terminal domain of the intracellular calcium release channel receptor—InsP3R1 and enhance InsP3R1-mediated calcium release in neurons (Tang et al., 2003; Chen et al., 2008; Liu et al., 2009). Thus, it is tempting to speculate that intermediate-length polyQ expansions in ataxin-2 may also cause it to associate with InsP3R1 and thereby modulate calcium signaling, though this was not yet shown. In ALS6, the FUS protein leads to CAMK2N2 up-regulation (Convertini et al., 2013), an inhibitor of CAMKII that regulates neuronal synaptic plasticity through phosphorylation of AMPA receptors. Given that it has been shown that abnormal CAMKII inhibition by small molecules and peptides results in dysregulation of Ca2+/glutamate signalling (Ashpole et al., 2012), it may be hypothesized that CAMKII inhibitor—CAMK2N2 up-regulation by mutant FUS will also result in Ca2+ dysregulation in ALS6. Moreover, deregulated Ca2+ levels can activate the Ca2+-dependent calpain protease that cleaves TDP-43 at the C-terminal, generating aggregation prone N-terminal segments that are found misallocated in the majority of ALS patients (even in those that do not carry TDP-43 mutations associated with ALS10) driving TDP-43 toxicity across ALS pathology (Aggad et al., 2014; Yamashita and Kwak, 2014). In addition, TDP-43 is simultaneously found to interact with other critical proteins in ALS namely matrin-3 (Johnson et al., 2014), ataxin-2 (Nihei et al., 2012), VAPB (Stoica et al., 2014), and SOD1 (Volkening et al., 2009) and is therefore rather tempting to conjecture about a tight interrelation between Ca2+ dyshomeostasis and the involvement of critical proteins in ALS.

ALS Toxic Processes and the Role of Calcium

In fact, major pathological processes in ALS involving excitotoxicity and the ER-mitochondria Ca2+ cycle are deeply connected and potentially trigger or/and are enhanced by intracellular Ca2+ deregulation: (a) Glutamatergic excitotoxicity is tough to be mediated by an excessive influx of extracellular ions, including Ca2+, resulting in elevated intracellular Ca2+ levels that can activate cytoplasmatic Ca2+-dependent apoptotic proteins (e.g., calcineurin, calpain) which promote cell death (Wang et al., 1999; Kim et al., 2002); (b) elevated intracellular levels of Ca2+ also lead to mitochondrial Ca2+ overload, that is deeply interconnected with mitochondrial dysfunction resulting in ROS production, oxidative stress and eventually to apoptosis or necrosis (Kawamata and Manfredi, 2010; Cozzolino and Carrì, 2012); and (c) depletion of Ca2+ levels in the ER which is suggested to occur via a persistent shift of Ca2+ from the ER to the mitochondria due to deregulated ER MCC leads to protein folding dysfunction and proteasome impairment, resulting in ER stress and apoptosis (Prell et al., 2013; Tadic et al., 2014). Mutations in critical proteins associated with ALS actually seem to increase the susceptibility for these toxic processes to occur. For example, misfolded and aggregated SOD1 mutants localized within the mitochondrial membrane of spinal cord MN cause dysfunction in oxidative phosphorylation and bind aberrantly to Bcl-2, generating toxicity (Jung et al., 2002; Mattiazzi et al., 2002; Liu et al., 2004; Vande Velde et al., 2008; Pedrini et al., 2010); also, the ALS linked P56S mutation in VAPB, or th A4V, G85R and G93A SOD1 mutations leads to toxic protein aggregation and ER stress (Prosser et al., 2008; Kim et al., 2010; Atkin et al., 2014).

Conclusions

Overall, we here argue that Ca2+ deregulation seems to establish a converging point for major ALS dysfunctional pathways and critical associated proteins, and can therefore be a key environmental factor to better understand ALS etiology and its pathomechanisms. However, we do not intend to ground that all proteins implicated in ALS will necessarily lead to Ca2+ deregulation; rather, we seek to discuss that processes involving Ca2+ could directly or indirectly (e.g., via Ca2+ effects on processes dependent of other divalent cations) account for their mutual involvement in ALS. Interestingly, the so-called “calcium hypothesis” establishing a close link between Ca2+ deregulation and neurodegeneration, has also been suggested to play a central role in other neurodegenerative disorders such as Alzheimer’s, Ataxia, Parkinson’s and Huntington Diseases (Bezprozvanny, 2010; Kasumu and Bezprozvanny, 2012), where Ca2+ channels and proteins involved in neuronal Ca2+ signalling systems are likely potential targets for therapeutic strategies (Zundorf and Reiser, 2011).

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

This work was supported by Fundação para a Ciência e a Tecnologia through grant PTDC/QUI-BIQ/117789/2010 (to CMG), post-doctoral fellowship SFRH/BPD/47477/2008 (to SSL), strategic grant PEst-OE/EQB/LA0004/2011 (to ITQB-Laboratόrio Associado) and grant UID/MULTI/04046/2013 from FCT/MCTES/PIDDAC (to BioISI).

References

  1. Aggad D., Vérièpe J., Tauffenberger A., Parker J. A. (2014). TDP-43 toxicity proceeds via calcium dysregulation and necrosis in aging Caenorhabditis elegans motor neurons. J. Neurosci. 34, 12093–12103. 10.1523/jneurosci.2495-13.2014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Al-Chalabi A., Jones A., Troakes C., King A., Al-Sarraj S., van den Berg L. H. (2012). The genetics and neuropathology of amyotrophic lateral sclerosis. Acta Neuropathol. 124, 339–352. 10.1007/s00401-012-1022-4 [DOI] [PubMed] [Google Scholar]
  3. Alexianu M. E., Ho B. K., Mohamed A. H., La Bella V., Smith R. G., Appel S. H. (1994). The role of calcium-binding proteins in selective motoneuron vulnerability in amyotrophic lateral sclerosis. Ann. Neurol. 36, 846–858. 10.1002/ana.410360608 [DOI] [PubMed] [Google Scholar]
  4. Ashpole N. M., Song W., Brustovetsky T., Engleman E. A., Brustovetsky N., Cummins T. R., et al. (2012). Calcium/calmodulin-dependent protein kinase II (CaMKII) inhibition induces neurotoxicity via dysregulation of glutamate/calcium signaling and hyperexcitability. J. Biol. Chem. 287, 8495–8506. 10.1074/jbc.m111.323915 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Atkin J. D., Farg M. A., Soo K. Y., Walker A. K., Halloran M., Turner B. J., et al. (2014). Mutant SOD1 inhibits ER-Golgi transport in amyotrophic lateral sclerosis. J. Neurochem. 129, 190–204. 10.1111/jnc.12493 [DOI] [PubMed] [Google Scholar]
  6. Beers D. R., Ho B. K., Siklós L., Alexianu M. E., Mosier D. R., Mohamed A. H., et al. (2001). Parvalbumin overexpression alters immune-mediated increases in intracellular calcium and delays disease onset in a transgenic model of familial amyotrophic lateral sclerosis. J. Neurochem. 79, 499–509. 10.1046/j.1471-4159.2001.00582.x [DOI] [PubMed] [Google Scholar]
  7. Bezprozvanny I. B. (2010). Calcium signaling and neurodegeneration. Acta Naturae 2, 72–82. [PMC free article] [PubMed] [Google Scholar]
  8. Blokhuis A. M., Groen E. J., Koppers M., van den Berg L. H., Pasterkamp R. J. (2013). Protein aggregation in amyotrophic lateral sclerosis. Acta Neuropathol. 125, 777–794. 10.1007/s00401-013-1125-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Chen Y. Z., Bennett C. L., Huynh H. M., Blair I. P., Puls I., Irobi J., et al. (2004). DNA/RNA helicase gene mutations in a form of juvenile amyotrophic lateral sclerosis (ALS4). Am. J. Hum. Genet. 74, 1128–1135. 10.1086/421054 [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Chen X., Tang T. S., Tu H., Nelson O., Pook M., Hammer R., et al. (2008). Deranged calcium signaling and neurodegeneration in spinocerebellar ataxia type 3. J. Neurosci. 28, 12713–12724. 10.1523/JNEUROSCI.3909-08.2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Convertini P., Zhang J., de la Grange P., Hayward L. J., Zhu H., Stamm S. (2013). Genome wide array analysis indicates that an amyotrophic lateral sclerosis mutation of FUS causes an early increase of CAMK2N2 in vitro. Biochim. Biophys. Acta 1832, 1129–1135. 10.1016/j.bbadis.2013.03.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Cozzolino M., Carrì M. T. (2012). Mitochondrial dysfunction in ALS. Prog. Neurobiol. 97, 54–66. 10.1016/j.pneurobio.2011.06.003 [DOI] [PubMed] [Google Scholar]
  13. DeJesus-Hernandez M., Mackenzie I. R., Boeve B. F., Boxer A. L., Baker M., Rutherford N. J., et al. (2011). Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron 72, 245–256. 10.1016/j.neuron.2011.09.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Deng H. X., Zhai H., Bigio E. H., Yan J., Fecto F., Ajroud K., et al. (2010). FUS-immunoreactive inclusions are a common feature in sporadic and non-SOD1 familial amyotrophic lateral sclerosis. Ann. Neurol. 67, 739–748. 10.1002/ana.22051 [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. De Vos K. J., Mórotz G. M., Stoica R., Tudor E. L., Lau K. F., Ackerley S., et al. (2012). VAPB interacts with the mitochondrial protein PTPIP51 to regulate calcium homeostasis. Hum. Mol. Genet. 21, 1299–1311. 10.1093/hmg/ddr559 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Elden A. C., Kim H. J., Hart M. P., Chen-Plotkin A. S., Johnson B. S., Fang X., et al. (2010). Ataxin-2 intermediate-length polyglutamine expansions are associated with increased risk for ALS. Nature 466, 1069–1075. 10.1038/nature09320 [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Estácio S. G., Leal S. S., Cristóvão J. S., Faísca P. F., Gomes C. M. (2015). Calcium binding to gatekeeper residues flanking aggregation-prone segments underlies non-fibrillar amyloid traits in superoxide dismutase 1 (SOD1). Biochim. Biophys. Acta 1854, 118–126. 10.1016/j.bbapap.2014.11.005 [DOI] [PubMed] [Google Scholar]
  18. Farg M. A., Soo K. Y., Warraich S. T., Sundaramoorthy V., Blair I. P., Atkin J. D. (2013). Ataxin-2 interacts with FUS and intermediate-length polyglutamine expansions enhance FUS-related pathology in amyotrophic lateral sclerosis. Hum. Mol. Genet. 22, 717–728. 10.1093/hmg/dds479 [DOI] [PubMed] [Google Scholar]
  19. Forsberg K., Jonsson P. A., Andersen P. M., Bergemalm D., Graffmo K. S., Hultdin M., et al. (2010). Novel antibodies reveal inclusions containing non-native SOD1 in sporadic ALS patients. PLoS One 5:e11552. 10.1371/journal.pone.0011552 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Grosskreutz J., Van Den Bosch L., Keller B. U. (2010). Calcium dysregulation in amyotrophic lateral sclerosis. Cell Calcium 47, 165–174. 10.1016/j.ceca.2009.12.002 [DOI] [PubMed] [Google Scholar]
  21. Guatteo E., Carunchio I., Pieri M., Albo F., Canu N., Mercuri N. B., et al. (2007). Altered calcium homeostasis in motor neurons following AMPA receptor but not voltage-dependent calcium channels’ activation in a genetic model of amyotrophic lateral sclerosis. Neurobiol. Dis. 28, 90–100. 10.1016/j.nbd.2007.07.002 [DOI] [PubMed] [Google Scholar]
  22. Hayashi S., Amari M., Okamoto K. (2013). Loss of calretinin- and parvalbumin-immunoreactive axons in anterolateral columns beyond the corticospinal tracts of amyotrophic lateral sclerosis spinal cords. J. Neurol. Sci. 331, 61–66. 10.1016/j.jns.2013.05.008 [DOI] [PubMed] [Google Scholar]
  23. Iguchi Y., Katsuno M., Ikenaka K., Ishigaki S., Sobue G. (2013). Amyotrophic lateral sclerosis: an update on recent genetic insights. J. Neurol. 260, 2917–2927. 10.1007/s00415-013-7112-y [DOI] [PubMed] [Google Scholar]
  24. Jaarsma D., Rognoni F., van Duijn W., Verspaget H. W., Haasdijk E. D., Holstege J. C. (2001). CuZn superoxide dismutase (SOD1) accumulates in vacuolated mitochondria in transgenic mice expressing amyotrophic lateral sclerosis-linked SOD1 mutations. Acta Neuropathol. 102, 293–305. 10.1007/s004010100399 [DOI] [PubMed] [Google Scholar]
  25. Jaiswal M. K. (2013). Calcium, mitochondria and the pathogenesis of ALS: the good, the bad and the ugly. Front. Cell. Neurosci. 7:199. 10.3389/fncel.2013.00199 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Jaiswal M. K., Keller B. U. (2009). Cu/Zn superoxide dismutase typical for familial amyotrophic lateral sclerosis increases the vulnerability of mitochondria and perturbs Ca2+ homeostasis in SOD1G93A mice. Mol. Pharmacol. 75, 478–489. 10.1124/mol.108.050831 [DOI] [PubMed] [Google Scholar]
  27. Johnson J. O., Mandrioli J., Benatar M., Abramzon Y., Van Deerlin V. M., Trojanowski J. Q., et al. (2010). Exome sequencing reveals VCP mutations as a cause of familial ALS. Neuron 68, 857–864. 10.1016/j.neuron.2010.11.036 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Johnson J. O., Pioro E. P., Boehringer A., Chia R., Feit H., Renton A. E., et al. (2014). Mutations in the Matrin 3 gene cause familial amyotrophic lateral sclerosis. Nat. Neurosci. 17, 664–666. 10.1038/nn.3688 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Jung C., Higgins C. M., Xu Z. (2002). Mitochondrial electron transport chain complex dysfunction in a transgenic mouse model for amyotrophic lateral sclerosis. J. Neurochem. 83, 535–545. 10.1046/j.1471-4159.2002.01112.x [DOI] [PubMed] [Google Scholar]
  30. Kanekura K., Hashimoto Y., Niikura T., Aiso S., Matsuoka M., Nishimoto I. (2004). Alsin, the product of ALS2 gene, suppresses SOD1 mutant neurotoxicity through RhoGEF domain by interacting with SOD1 mutants. J. Biol. Chem. 279, 19247–19256. 10.1074/jbc.m313236200 [DOI] [PubMed] [Google Scholar]
  31. Kasumu A., Bezprozvanny I. (2012). Deranged calcium signaling in Purkinje cells and pathogenesis in spinocerebellar ataxia 2 (SCA2) and other ataxias. Cerebellum 11, 630–639. 10.1007/s12311-010-0182-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Kawamata H., Manfredi G. (2010). Mitochondrial dysfunction and intracellular calcium dysregulation in ALS. Mech. Ageing Dev. 131, 517–526. 10.1016/j.mad.2010.05.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Kim M. J., Jo D. G., Hong G. S., Kim B. J., Lai M., Cho D. H., et al. (2002). Calpain-dependent cleavage of cain/cabin1 activates calcineurin to mediate calcium-triggered cell death. Proc. Natl. Acad. Sci. U S A 99, 9870–9875. 10.1073/pnas.152336999 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Kim S., Leal S. S., Ben Halevy D., Gomes C. M., Lev S. (2010). Structural requirements for VAP-B oligomerization and their implication in amyotrophic lateral sclerosis-associated VAP-B(P56S) neurotoxicity. J. Biol. Chem. 285, 13839–13849. 10.1074/jbc.m109.097345 [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Kruman I. I., Pedersen W. A., Springer J. E., Mattson M. P. (1999). ALS-linked Cu/Zn-SOD mutation increases vulnerability of motor neurons to excitotoxicity by a mechanism involving increased oxidative stress and perturbed calcium homeostasis. Exp. Neurol. 160, 28–39. 10.1006/exnr.1999.7190 [DOI] [PubMed] [Google Scholar]
  36. Lai C., Xie C., McCormack S. G., Chiang H. C., Michalak M. K., Lin X., et al. (2006). Amyotrophic lateral sclerosis 2-deficiency leads to neuronal degeneration in amyotrophic lateral sclerosis through altered AMPA receptor trafficking. J. Neurosci. 26, 11798–11806. 10.1523/jneurosci.2084-06.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Lautenschläger J., Prell T., Ruhmer J., Weidemann L., Witte O. W., Grosskreutz J. (2013). Overexpression of human mutated G93A SOD1 changes dynamics of the ER mitochondria calcium cycle specifically in mouse embryonic motor neurons. Exp. Neurol. 247, 91–100. 10.1016/j.expneurol.2013.03.027 [DOI] [PubMed] [Google Scholar]
  38. Leal S. S., Cardoso I., Valentine J. S., Gomes C. M. (2013). Calcium ions promote superoxide dismutase 1 (SOD1) aggregation into non-fibrillar amyloid: a link to toxic effects of calcium overload in amyotrophic lateral sclerosis (ALS)? J. Biol. Chem. 288, 25219–25228. 10.1074/jbc.M113.470740 [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Liu J., Lillo C., Jonsson P. A., Vande Velde C., Ward C. M., Miller T. M., et al. (2004). Toxicity of familial ALS-linked SOD1 mutants from selective recruitment to spinal mitochondria. Neuron 43, 5–17. 10.1016/j.neuron.2004.06.016 [DOI] [PubMed] [Google Scholar]
  40. Liu J., Tang T. S., Tu H., Nelson O., Herndon E., Huynh D. P., et al. (2009). Deranged calcium signaling and neurodegeneration in spinocerebellar ataxia type 2. J. Neurosci. 29, 9148–9162. 10.1523/JNEUROSCI.0660-09.2009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Maekawa S., Leigh P. N., King A., Jones E., Steele J. C., Bodi I., et al. (2009). TDP-43 is consistently co-localized with ubiquitinated inclusions in sporadic and Guam amyotrophic lateral sclerosis but not in familial amyotrophic lateral sclerosis with and without SOD1 mutations. Neuropathology 29, 672–683. 10.1111/j.1440-1789.2009.01029.x [DOI] [PubMed] [Google Scholar]
  42. Maruyama H., Morino H., Ito H., Izumi Y., Kato H., Watanabe Y., et al. (2010). Mutations of optineurin in amyotrophic lateral sclerosis. Nature 465, 223–226. 10.1038/nature08971 [DOI] [PubMed] [Google Scholar]
  43. Masutani T., Taguchi K., Kumanogoh H., Nakamura S., Maekawa S. (2008). Molecular interaction of neurocalcin alpha with alsin (ALS2). Neurosci. Lett. 438, 26–28. 10.1016/j.neulet.2008.04.066 [DOI] [PubMed] [Google Scholar]
  44. Mattiazzi M., D’Aurelio M., Gajewski C. D., Martushova K., Kiaei M., Beal M. F., et al. (2002). Mutated human SOD1 causes dysfunction of oxidative phosphorylation in mitochondria of transgenic mice. J. Biol. Chem. 277, 29626–29633. 10.1074/jbc.m203065200 [DOI] [PubMed] [Google Scholar]
  45. Migheli A., Cordera S., Bendotti C., Atzori C., Piva R., Schiffer D. (1999). S-100beta protein is upregulated in astrocytes and motor neurons in the spinal cord of patients with amyotrophic lateral sclerosis. Neurosci. Lett. 261, 25–28. 10.1016/s0304-3940(98)01001-5 [DOI] [PubMed] [Google Scholar]
  46. Mórotz G. M., De Vos K. J., Vagnoni A., Ackerley S., Shaw C. E., Miller C. C. (2012). Amyotrophic lateral sclerosis-associated mutant VAPBP56S perturbs calcium homeostasis to disrupt axonal transport of mitochondria. Hum. Mol. Genet. 21, 1979–1988. 10.1093/hmg/dds011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Nihei Y., Ito D., Suzuki N. (2012). Roles of ataxin-2 in pathological cascades mediated by TAR DNA-binding protein 43 (TDP-43) and Fused in Sarcoma (FUS). J. Biol. Chem. 287, 41310–41323. 10.1074/jbc.m112.398099 [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Nishimura A. L., Mitne-Neto M., Silva H. C., Richieri-Costa A., Middleton S., Cascio D., et al. (2004). A mutation in the vesicle-trafficking protein VAPB causes late-onset spinal muscular atrophy and amyotrophic lateral sclerosis. Am. J. Hum. Genet. 75, 822–831. 10.1086/425287 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Osaka M., Ito D., Yagi T., Nihei Y., Suzuki N. (2015). Evidence of a link between ubiquilin 2 and optineurin in amyotrophic lateral sclerosis. Hum. Mol. Genet. 24, 1617–1629. 10.1093/hmg/ddu575 [DOI] [PubMed] [Google Scholar]
  50. Palecek J., Lips M. B., Keller B. U. (1999). Calcium dynamics and buffering in motoneurones of the mouse spinal cord. J. Physiol. 520(Pt. 2), 485–502. 10.1111/j.1469-7793.1999.00485.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Parone P. A., Da Cruz S., Han J. S., McAlonis-Downes M., Vetto A. P., Lee S. K., et al. (2013). Enhancing mitochondrial calcium buffering capacity reduces aggregation of misfolded SOD1 and motor neuron cell death without extending survival in mouse models of inherited amyotrophic lateral sclerosis. J. Neurosci. 33, 4657–4671. 10.1523/jneurosci.1119-12.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Pedrini S., Sau D., Guareschi S., Bogush M., Brown R. H., Jr., Naniche N., et al. (2010). ALS-linked mutant SOD1 damages mitochondria by promoting conformational changes in Bcl-2. Hum. Mol. Genet. 19, 2974–2986. 10.1093/hmg/ddq202 [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Pokrishevsky E., Grad L. I., Yousefi M., Wang J., Mackenzie I. R., Cashman N. R. (2012). Aberrant localization of FUS and TDP43 is associated with misfolding of SOD1 in amyotrophic lateral sclerosis. PLoS One 7:e35050. 10.1371/journal.pone.0035050 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Prell T., Lautenschläger J., Grosskreutz J. (2013). Calcium-dependent protein folding in amyotrophic lateral sclerosis. Cell Calcium 54, 132–143. 10.1016/j.ceca.2013.05.007 [DOI] [PubMed] [Google Scholar]
  55. Prosser D. C., Tran D., Gougeon P. Y., Verly C., Ngsee J. K. (2008). FFAT rescues VAPA-mediated inhibition of ER-to-Golgi transport and VAPB-mediated ER aggregation. J. Cell. Sci. 121, 3052–3061. 10.1242/jcs.028696 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Rosen D. R., Siddique T., Patterson D., Figlewicz D. A., Sapp P., Hentati A., et al. (1993). Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature 362, 59–62. 10.1038/362059a0 [DOI] [PubMed] [Google Scholar]
  57. Rowland L. P., Shneider N. A. (2001). Amyotrophic lateral sclerosis. N. Engl. J. Med. 344, 1688–1700. 10.1056/NEJM200105313442207 [DOI] [PubMed] [Google Scholar]
  58. Shaw P. J., Eggett C. J. (2000). Molecular factors underlying selective vulnerability of motor neurons to neurodegeneration in amyotrophic lateral sclerosis. J. Neurol. 247(Suppl. 1), I17–I27. 10.1007/bf03161151 [DOI] [PubMed] [Google Scholar]
  59. Siklós L., Engelhardt J. I., Alexianu M. E., Gurney M. E., Siddique T., Appel S. H. (1998). Intracellular calcium parallels motoneuron degeneration in SOD-1 mutant mice. J. Neuropathol. Exp. Neurol. 57, 571–587. 10.1097/00005072-199806000-00005 [DOI] [PubMed] [Google Scholar]
  60. Siklós L., Engelhardt J., Harati Y., Smith R. G., Joó F., Appel S. H. (1996). Ultrastructural evidence for altered calcium in motor nerve terminals in amyotropic lateral sclerosis. Ann. Neurol. 39, 203–216. 10.1002/ana.410390210 [DOI] [PubMed] [Google Scholar]
  61. Sreedharan J., Blair I. P., Tripathi V. B., Hu X., Vance C., Rogelj B., et al. (2008). TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis. Science 319, 1668–1672. 10.1126/science.1154584 [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Stoica R., De Vos K. J., Paillusson S., Mueller S., Sancho R. M., Lau K. F., et al. (2014). ER-mitochondria associations are regulated by the VAPB-PTPIP51 interaction and are disrupted by ALS/FTD-associated TDP-43. Nat. Commun. 5:3996. 10.1038/ncomms4996 [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Su X. W., Broach J. R., Connor J. R., Gerhard G. S., Simmons Z. (2014). Genetic heterogeneity of amyotrophic lateral sclerosis: implications for clinical practice and research. Muscle Nerve 49, 786–803. 10.1002/mus.24198 [DOI] [PubMed] [Google Scholar]
  64. Tadic V., Prell T., Lautenschlaeger J., Grosskreutz J. (2014). The ER mitochondria calcium cycle and ER stress response as therapeutic targets in amyotrophic lateral sclerosis. Front. Cell. Neurosci. 8:147. 10.3389/fncel.2014.00147 [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Tang T. S., Tu H., Chan E. Y., Maximov A., Wang Z., Wellington C. L., et al. (2003). Huntingtin and huntingtin-associated protein 1 influence neuronal calcium signaling mediated by inositol-(1,4,5) triphosphate receptor type 1. Neuron 39, 227–239. 10.1016/S0896-6273(03)00366-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Tateno M., Sadakata H., Tanaka M., Itohara S., Shin R. M., Miura M., et al. (2004). Calcium-permeable AMPA receptors promote misfolding of mutant SOD1 protein and development of amyotrophic lateral sclerosis in a transgenic mouse model. Hum. Mol. Genet. 13, 2183–2196. 10.1093/hmg/ddh246 [DOI] [PubMed] [Google Scholar]
  67. Tortarolo M., Grignaschi G., Calvaresi N., Zennaro E., Spaltro G., Colovic M., et al. (2006). Glutamate AMPA receptors change in motor neurons of SOD1G93A transgenic mice and their inhibition by a noncompetitive antagonist ameliorates the progression of amytrophic lateral sclerosis-like disease. J. Neurosci. Res. 83, 134–146. 10.1002/jnr.20715 [DOI] [PubMed] [Google Scholar]
  68. Tradewell M. L., Cooper L. A., Minotti S., Durham H. D. (2011). Calcium dysregulation, mitochondrial pathology and protein aggregation in a culture model of amyotrophic lateral sclerosis: mechanistic relationship and differential sensitivity to intervention. Neurobiol. Dis. 42, 265–275. 10.1016/j.nbd.2011.01.016 [DOI] [PubMed] [Google Scholar]
  69. Tudor E. L., Galtrey C. M., Perkinton M. S., Lau K. F., De Vos K. J., Mitchell J. C., et al. (2010). Amyotrophic lateral sclerosis mutant vesicle-associated membrane protein-associated protein-B transgenic mice develop TAR-DNA-binding protein-43 pathology. Neuroscience 167, 774–785. 10.1016/j.neuroscience.2010.02.035 [DOI] [PubMed] [Google Scholar]
  70. Valencia C. A., Ju W., Liu R. (2007). Matrin 3 is a Ca2+/calmodulin-binding protein cleaved by caspases. Biochem. Biophys. Res. Commun. 361, 281–286. 10.1016/j.bbrc.2007.06.156 [DOI] [PubMed] [Google Scholar]
  71. Vance C., Rogelj B., Hortobágyi T., De Vos K. J., Nishimura A. L., Sreedharan J., et al. (2009). Mutations in FUS, an RNA processing protein, cause familial amyotrophic lateral sclerosis type 6. Science 323, 1208–1211. 10.1126/science.1165942 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Van Damme P., Leyssen M., Callewaert G., Robberecht W., Van Den Bosch L. (2003). The AMPA receptor antagonist NBQX prolongs survival in a transgenic mouse model of amyotrophic lateral sclerosis. Neurosci. Lett. 343, 81–84. 10.1016/s0304-3940(03)00314-8 [DOI] [PubMed] [Google Scholar]
  73. Vandenberghe W., Robberecht W., Brorson J. R. (2000). AMPA receptor calcium permeability, GluR2 expression and selective motoneuron vulnerability. J. Neurosci. 20, 123–132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Van Den Bosch L., Vandenberghe W., Klaassen H., Van Houtte E., Robberecht W. (2000). Ca(2+)-permeable AMPA receptors and selective vulnerability of motor neurons. J. Neurol. Sci. 180, 29–34. 10.1016/s0022-510x(00)00414-7 [DOI] [PubMed] [Google Scholar]
  75. van den Heuvel D. M., Harschnitz O., Van Den Berg L. H., Pasterkamp R. J. (2014). Taking a risk: a therapeutic focus on ataxin-2 in amyotrophic lateral sclerosis? Trends Mol. Med. 20, 25–35. 10.1016/j.molmed.2013.09.001 [DOI] [PubMed] [Google Scholar]
  76. Vande Velde C., Miller T. M., Cashman N. R., Cleveland D. W. (2008). Selective association of misfolded ALS-linked mutant SOD1 with the cytoplasmic face of mitochondria. Proc. Natl. Acad. Sci. U S A 105, 4022–4027. 10.1073/pnas.0712209105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Volkening K., Leystra-Lantz C., Yang W., Jaffee H., Strong M. J. (2009). 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. 1305, 168–182. 10.1016/j.brainres.2009.09.105 [DOI] [PubMed] [Google Scholar]
  78. von Lewinski F., Keller B. U. (2005). Ca2+, mitochondria and selective motoneuron vulnerability: implications for ALS. Trends Neurosci. 28, 494–500. 10.1016/j.tins.2005.07.001 [DOI] [PubMed] [Google Scholar]
  79. Wang H. G., Pathan N., Ethell I. M., Krajewski S., Yamaguchi Y., Shibasaki F., et al. (1999). Ca2+-induced apoptosis through calcineurin dephosphorylation of BAD. Science 284, 339–343. 10.1126/science.284.5412.339 [DOI] [PubMed] [Google Scholar]
  80. Williams T. L., Day N. C., Ince P. G., Kamboj R. K., Shaw P. J. (1997). Calcium-permeable alpha-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid receptors: a molecular determinant of selective vulnerability in amyotrophic lateral sclerosis. Ann. Neurol. 42, 200–207. 10.1002/ana.410420211 [DOI] [PubMed] [Google Scholar]
  81. Wu C. H., Fallini C., Ticozzi N., Keagle P. J., Sapp P. C., Piotrowska K., et al. (2012). Mutations in the profilin 1 gene cause familial amyotrophic lateral sclerosis. Nature 488, 499–503. 10.1038/nature11280 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Yamashita T., Kwak S. (2014). The molecular link between inefficient GluA2 Q/R site-RNA editing and TDP-43 pathology in motor neurons of sporadic amyotrophic lateral sclerosis patients. Brain Res. 1584, 28–38. 10.1016/j.brainres.2013.12.011 [DOI] [PubMed] [Google Scholar]
  83. Yin H. Z., Tang D. T., Weiss J. H. (2007). Intrathecal infusion of a Ca(2+)-permeable AMPA channel blocker slows loss of both motor neurons and of the astrocyte glutamate transporter, GLT-1 in a mutant SOD1 rat model of ALS. Exp. Neurol. 207, 177–185. 10.1016/j.expneurol.2007.07.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Zhang M. D., Tortoriello G., Hsueh B., Tomer R., Ye L., Mitsios N., et al. (2014). Neuronal calcium-binding proteins 1/2 localize to dorsal root ganglia and excitatory spinal neurons and are regulated by nerve injury. Proc. Natl. Acad. Sci. U S A 111, E1149–E1158. 10.1073/pnas.1402318111 [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Zundorf G., Reiser G. (2011). Calcium dysregulation and homeostasis of neural calcium in the molecular mechanisms of neurodegenerative diseases provide multiple targets for neuroprotection. Antioxid. Redox Signal. 14, 1275–1288. 10.1089/ars.2010.3359 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Frontiers in Cellular Neuroscience are provided here courtesy of Frontiers Media SA

RESOURCES