Skip to main content
Log in

Plastin 3 is upregulated in iPSC-derived motoneurons from asymptomatic SMN1-deleted individuals

  • Original Article
  • Published:
Cellular and Molecular Life Sciences Aims and scope Submit manuscript

Abstract

Spinal muscular atrophy (SMA) is a devastating motoneuron (MN) disorder caused by homozygous loss of SMN1. Rarely, SMN1-deleted individuals are fully asymptomatic despite carrying identical SMN2 copies as their SMA III-affected siblings suggesting protection by genetic modifiers other than SMN2. High plastin 3 (PLS3) expression has previously been found in lymphoblastoid cells but not in fibroblasts of asymptomatic compared to symptomatic siblings. To find out whether PLS3 is also upregulated in MNs of asymptomatic individuals and thus a convincing SMA protective modifier, we generated induced pluripotent stem cells (iPSCs) from fibroblasts of three asymptomatic and three SMA III-affected siblings from two families and compared these to iPSCs from a SMA I patient and control individuals. MNs were differentiated from iPSC-derived small molecule neural precursor cells (smNPCs). All four genotype classes showed similar capacity to differentiate into MNs at day 8. However, SMA I-derived MN survival was significantly decreased while SMA III- and asymptomatic-derived MN survival was moderately reduced compared to controls at day 27. SMN expression levels and concomitant gem numbers broadly matched SMN2 copy number distribution; SMA I presented the lowest levels, whereas SMA III and asymptomatic showed similar levels. In contrast, PLS3 was significantly upregulated in mixed MN cultures from asymptomatic individuals pinpointing a tissue-specific regulation. Evidence for strong PLS3 accumulation in shaft and rim of growth cones in MN cultures from asymptomatic individuals implies an important role in neuromuscular synapse formation and maintenance. These findings provide strong evidence that PLS3 is a genuine SMA protective modifier.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Abbreviations

PLS3:

Plastin 3

SMA:

Spinal muscular atrophy

SMN1:

Survival motor neuron 1

SMN2:

Survival motor neuron 2

iPSC:

Induced pluripotent stem cell

CNS:

Central nervous system

LAAP:

l-ascorbic-acid-2-phosphate

MN:

Motoneuron

NSC:

Neural stem cell

PO:

Poly-l-ornithine

ROCK:

Rho-associated coiled-coil containing protein kinase

SeV:

Sendai virus

References

  1. Munsat TL, Davies KE (1992) International SMA consortium meeting. (26–28 June 1992, Bonn, Germany). Neuromuscul Disord 2:423–428

    Article  CAS  PubMed  Google Scholar 

  2. Pearn J (1978) Incidence, prevalence, and gene frequency studies of chronic childhood spinal muscular atrophy. J Med Genet 15:409–413

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Feldkotter M et al (2002) Quantitative analyses of SMN1 and SMN2 based on real-time lightCycler PCR: fast and highly reliable carrier testing and prediction of severity of spinal muscular atrophy. Am J Hum Genet 70:358–368

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Lefebvre S et al (1995) Identification and characterization of a spinal muscular atrophy-determining gene. Cell 80:155–165

    Article  CAS  PubMed  Google Scholar 

  5. Wirth B et al (2006) Mildly affected patients with spinal muscular atrophy are partially protected by an increased SMN2 copy number. Hum Genet 119:422–428

    Article  CAS  PubMed  Google Scholar 

  6. Lorson CL et al (1999) A single nucleotide in the SMN gene regulates splicing and is responsible for spinal muscular atrophy. Proc Natl Acad Sci USA 96:6307–6311

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Cartegni L, Krainer AR (2002) Disruption of an SF2/ASF-dependent exonic splicing enhancer in SMN2 causes spinal muscular atrophy in the absence of SMN1. Nat Genet 30:377–384

    Article  CAS  PubMed  Google Scholar 

  8. Liu Q, Dreyfuss G (1996) A novel nuclear structure containing the survival of motor neurons protein. EMBO J 15:3555–3565

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Li DK et al (2014) SMN control of RNP assembly: from post-transcriptional gene regulation to motor neuron disease. Semin Cell Dev Biol 32:22–29

    Article  CAS  PubMed  Google Scholar 

  10. Burghes AH, Beattie CE (2009) Spinal muscular atrophy: why do low levels of survival motor neuron protein make motor neurons sick? Nat Rev Neurosci 10:597–609

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Kariya S et al (2014) Requirement of enhanced survival motoneuron protein imposed during neuromuscular junction maturation. J Clin Invest 124:785–800

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Hahnen E et al (1995) Molecular analysis of candidate genes on chromosome 5q13 in autosomal recessive spinal muscular atrophy: evidence of homozygous deletions of the SMN gene in unaffected individuals. Hum Mol Genet 4:1927–1933

    Article  CAS  PubMed  Google Scholar 

  13. Wirth B et al (2013) How genetic modifiers influence the phenotype of spinal muscular atrophy and suggest future therapeutic approaches. Curr Opin Genet Dev 23:330–338

    Article  CAS  PubMed  Google Scholar 

  14. Oprea GE et al (2008) Plastin 3 is a protective modifier of autosomal recessive spinal muscular atrophy. Science 320:524–527

    Article  CAS  PubMed  Google Scholar 

  15. Lin CS et al (1993) Human plastin genes. Comparative gene structure, chromosome location, and differential expression in normal and neoplastic cells. J Biol Chem 268:2781–2792

    CAS  PubMed  Google Scholar 

  16. McWhorter ML et al (2003) Knockdown of the survival motor neuron (Smn) protein in zebrafish causes defects in motor axon outgrowth and pathfinding. J Cell Biol 162:919–931

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Lyon AN et al (2014) Calcium binding is essential for plastin 3 function in Smn-deficient motoneurons. Hum Mol Genet 23:1990–2004

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Ackermann B et al (2013) Plastin 3 ameliorates spinal muscular atrophy via delayed axon pruning and improves neuromuscular junction functionality. Hum Mol Genet 22:1328–1347

    Article  CAS  PubMed  Google Scholar 

  19. Takahashi K, Yamanaka S (2006) Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126:663–676

    Article  CAS  PubMed  Google Scholar 

  20. Yu J et al (2007) Induced pluripotent stem cell lines derived from human somatic cells. Science 318:1917–1920

    Article  CAS  PubMed  Google Scholar 

  21. Takahashi K et al (2007) Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131:861–872

    Article  CAS  PubMed  Google Scholar 

  22. Davis-Dusenbery BN et al (2014) How to make spinal motor neurons. Development 141:491–501

    Article  CAS  PubMed  Google Scholar 

  23. Ebert AD et al (2009) Induced pluripotent stem cells from a spinal muscular atrophy patient. Nature 457:277–280

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Amoroso MW et al (2013) Accelerated high-yield generation of limb-innervating motor neurons from human stem cells. J Neurosci 33:574–586

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Patani R et al (2011) Retinoid-independent motor neurogenesis from human embryonic stem cells reveals a medial columnar ground state. Nat Commun 2:214

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Wichterle H et al (2002) Directed differentiation of embryonic stem cells into motor neurons. Cell 110:385–397

    Article  CAS  PubMed  Google Scholar 

  27. Hu BY, Zhang SC (2009) Differentiation of spinal motor neurons from pluripotent human stem cells. Nat Protoc 4:1295–1304

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Li XJ et al (2005) Specification of motoneurons from human embryonic stem cells. Nat Biotechnol 23:215–221

    Article  PubMed  Google Scholar 

  29. Li XJ et al (2008) Directed differentiation of ventral spinal progenitors and motor neurons from human embryonic stem cells by small molecules. Stem Cells 26:886–893

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Reinhardt P et al (2013) Derivation and expansion using only small molecules of human neural progenitors for neurodegenerative disease modeling. PLoS One 8:e59252

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Helmken C et al (2003) Evidence for a modifying pathway in SMA discordant families: reduced SMN level decreases the amount of its interacting partners and Htra2-beta1. Hum Genet 114:11–21

    Article  CAS  PubMed  Google Scholar 

  32. Garbes L et al (2013) VPA response in SMA is suppressed by the fatty acid translocase CD36. Hum Mol Genet 22:398–407

    Article  CAS  PubMed  Google Scholar 

  33. Koch P et al (2011) Excitation-induced ataxin-3 aggregation in neurons from patients with Machado-Joseph disease. Nature 480:543–546

    CAS  PubMed  Google Scholar 

  34. Ban H et al (2011) Efficient generation of transgene-free human induced pluripotent stem cells (iPSCs) by temperature-sensitive Sendai virus vectors. Proc Natl Acad Sci USA 108:14234–14239

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Aasen T et al (2008) Efficient and rapid generation of induced pluripotent stem cells from human keratinocytes. Nat Biotechnol 26:1276–1284

    Article  CAS  PubMed  Google Scholar 

  36. Watanabe K et al (2007) A ROCK inhibitor permits survival of dissociated human embryonic stem cells. Nat Biotechnol 25:681–686

    Article  CAS  PubMed  Google Scholar 

  37. Itskovitz-Eldor J et al (2000) Differentiation of human embryonic stem cells into embryoid bodies compromising the three embryonic germ layers. Mol Med 6:88–95

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Borghese L et al (2010) Inhibition of notch signaling in human embryonic stem cell-derived neural stem cells delays G1/S phase transition and accelerates neuronal differentiation in vitro and in vivo. Stem Cells 28:955–964

    Article  CAS  PubMed  Google Scholar 

  39. Crawford TQ, Roelink H (2007) The notch response inhibitor DAPT enhances neuronal differentiation in embryonic stem cell-derived embryoid bodies independently of sonic hedgehog signaling. Dev Dyn 236:886–892

    Article  CAS  PubMed  Google Scholar 

  40. Fusaki N et al (2009) Efficient induction of transgene-free human pluripotent stem cells using a vector based on Sendai virus, an RNA virus that does not integrate into the host genome. Proc Jpn Acad Ser B Phys Biol Sci 85:348–362

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Smith KP et al (2009) Pluripotency: toward a gold standard for human ES and iPS cells. J Cell Physiol 220:21–29

    Article  CAS  PubMed  Google Scholar 

  42. Nishimura K et al (2011) Development of defective and persistent Sendai virus vector: a unique gene delivery/expression system ideal for cell reprogramming. J Biol Chem 286:4760–4771

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Kim JB et al (2009) Direct reprogramming of human neural stem cells by OCT4. Nature 461:649–653

    Article  CAS  PubMed  Google Scholar 

  44. Michalczyk K, Ziman M (2005) Nestin structure and predicted function in cellular cytoskeletal organisation. Histol Histopathol 20:665–671

    CAS  PubMed  Google Scholar 

  45. Alaynick WA et al (2011) SnapShot: spinal cord development. Cell 146(178–178):e171

    Google Scholar 

  46. Coovert DD et al (1997) The survival motor neuron protein in spinal muscular atrophy. Hum Mol Genet 6:1205–1214

    Article  CAS  PubMed  Google Scholar 

  47. Corti S et al (2012) Genetic correction of human induced pluripotent stem cells from patients with spinal muscular atrophy. Sci Transl Med 4:165ra162

    PubMed  PubMed Central  Google Scholar 

  48. Dunn KW et al (2011) A practical guide to evaluating colocalization in biological microscopy. Am J Physiol Cell Physiol 300:C723–C742

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Sareen D et al (2012) Inhibition of apoptosis blocks human motor neuron cell death in a stem cell model of spinal muscular atrophy. PLoS One 7:e39113

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Hsieh-Li HM et al (2000) A mouse model for spinal muscular atrophy. Nat Genet 24:66–70

    Article  CAS  PubMed  Google Scholar 

  51. Monani UR et al (2000) The human centromeric survival motor neuron gene (SMN2) rescues embryonic lethality in Smn(−/−) mice and results in a mouse with spinal muscular atrophy. Hum Mol Genet 9:333–339

    Article  CAS  PubMed  Google Scholar 

  52. Riessland M et al (2010) SAHA ameliorates the SMA phenotype in two mouse models for spinal muscular atrophy. Hum Mol Genet 19:1492–1506

    Article  CAS  PubMed  Google Scholar 

  53. William CM et al (2003) Regulation of motor neuron subtype identity by repressor activity of Mnx class homeodomain proteins. Development 130:1523–1536

    Article  CAS  PubMed  Google Scholar 

  54. Boza-Moran MG et al (2015) Decay in survival motor neuron and plastin 3 levels during differentiation of iPSC-derived human motor neurons. Sci Rep 5:11696

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Bernal S et al (2011) Plastin 3 expression in discordant spinal muscular atrophy (SMA) siblings. Neuromuscul Disord 21:413–419

    Article  PubMed  Google Scholar 

  56. Mentis GZ et al (2011) Early functional impairment of sensory-motor connectivity in a mouse model of spinal muscular atrophy. Neuron 69:453–467

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. le Hao T et al (2012) Survival motor neuron affects plastin 3 protein levels leading to motor defects. J Neurosci 32:5074–5084

    Article  CAS  PubMed Central  Google Scholar 

  58. Murray LM et al (2008) Selective vulnerability of motor neurons and dissociation of pre- and post-synaptic pathology at the neuromuscular junction in mouse models of spinal muscular atrophy. Hum Mol Genet 17:949–962

    Article  CAS  PubMed  Google Scholar 

  59. Kariya S et al (2008) Reduced SMN protein impairs maturation of the neuromuscular junctions in mouse models of spinal muscular atrophy. Hum Mol Genet 17:2552–2569

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Kong L et al (2009) Impaired synaptic vesicle release and immaturity of neuromuscular junctions in spinal muscular atrophy mice. J Neurosci 29:842–851

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Ruiz R et al (2010) Altered intracellular Ca2+ homeostasis in nerve terminals of severe spinal muscular atrophy mice. J Neurosci 30:849–857

    Article  CAS  PubMed  Google Scholar 

  62. Lowery LA, Van Vactor D (2009) The trip of the tip: understanding the growth cone machinery. Nat Rev Mol Cell Biol 10:332–343

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

We thank Dr John Dimos (iPierian) for generating the HGK13 and HGK16 iPSC clones. We thank M. Segschneider, C. Thiele, R. Konang and V. Poppe from the Institute of Reconstructive Neurobiology (RNB, Bonn, Germany) for excellent technical support and A. Leinhaas for carrying out the teratoma assays. Dr A. Schauss and I. Hensen of the Imaging Facility of the Cluster of Excellence in Cellular Stress Responses in Aging-associated Diseases (CECAD, Cologne, Germany) supported the confocal microscopy. This work has been supported by the European Community’s Seventh Framework Program FP7/2007-2013 under grant agreement no 2012-305121 (NeurOmics), the 7FP project SCR&Tox (HEALTH-F5-2010-26675), BIO.NRW (Project StemCellFactory, #005-1403-0106), the Center for Molecular Medicine Cologne (grant no C11), the Hertie Foundation and the Deutsche Forschungsgemeinschaft (Wi 945/13-1; Wi 945/16-1) to B.W.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Oliver Brüstle or Brunhilde Wirth.

Ethics declarations

Conflict of interest

No competing financial interests exist.

Additional information

L. Heesen and M. Peitz contributed equally to this work.

Electronic supplementary material

Below is the link to the electronic supplementary material.

18_2015_2084_MOESM1_ESM.tif

Supplementary material 1: Figure S1: Validation of human iPSCs from SMA III and asymptomatic siblings. (A) Typical flat colony morphology with dense centre in phase contrast was shown for one representative asymptomatic (HGK21.1) and one representative SMA III (HGK22.17) iPSC line. Pluripotency was proven by AP activity (blue) and via immunostaining with pluripotency surface markers SSEA-3, SSEA-4, TRA-1-60 and TRA-1-81 (all red). Nuclei were counterstained with DAPI (blue). Scale bars 200 µm, valid for all images. (B) Exemplary karyograms of human iPSC lines HGK21.1 (asymptomatic) and HGK22.17 (SMA III) with respective original fibroblast lines schematically displayed every chromosome with the corresponding B allele frequency (upper panel) and Log R ration (lower panel). (C) Semi-quantitative analysis of SeV nucleocapsid (NP) mRNA expression in representative iPSC lines in comparison to 18S rRNA expression as loading control. SeV vector activity vanished at later passages (> p7). Only the positive control (pos ctrl) depicted a strong NP expression (TIFF 62136 kb)

18_2015_2084_MOESM2_ESM.tif

Supplementary material 2: Figure S2: Validation of human iPSCs from SMA III and asymptomatic siblings. (A) Spontaneous undirected differentiation of human iPSCs delivered progeny of all three germ layers in representative cell lines showing endodermal cells (α-fetoprotein AFP, green), mesodermal cells (smooth muscle actin SMA, red) and ectodermal cells (β III-tubulin, red) upon immunocytochemical labelling. (B) Teratoma formation assay confirmed pluripotency in iPSCs by generating derivatives of all three germ layers in vivo. Immunohistologic haematoxylin/eosin (HE) staining in representative samples displayed endoderm (glandular epithelium), mesoderm (cartilage) and ectoderm (neural rosettes). Nuclei were counterstained with DAPI (blue) (A). Scale bars 200 µm (A and B), valid for all images (TIFF 42104 kb)

18_2015_2084_MOESM3_ESM.tif

Supplementary material 3: Validation of smNPCs. iPSC-derived smNPC lines clearly disclosed typical NSC markers PAX6 (green), SOX1 (red), SOX2 (green) and nestin (red) in four representative cell lines upon immunolabelling. Expression of anterior marker FORSE-1 (green) was visualised as well. smNPCs expressed characteristic rosette-markers PLZF (green) and ZO-1 (red), however, lacked the typical petal-like arrangement underscoring their early pre-rosette NSC fate. Nuclei were counterstained with DAPI (blue). Scale bar 100 µm, valid for all images (TIFF 30730 kb)

18_2015_2084_MOESM4_ESM.tif

Supplementary material 4 Immunocytochemistry of four smNPC-derived cultures subjected to MN differentiation. Cultures representing the respective phenotype classes (control 30 m-r12, SMA I patient HGK1, SMA III patient HGK27.13 and asymptomatic sister HGK28.9) proved correct pMN-regionalisation on day 8 of the differentiation protocol. Pan-NSC markers SOX2 (green) and nestin (red) were clearly visible while early CNS marker PAX6 (green) was moderately expressed. Cells broadly exhibited pMN-marker OLIG2 (green) and NKX6.1 (red) whose evident co-localisation confirmed true MN progenitor identity. Moreover, expression of HOXB4 (red) verified sufficient posteriorization. Presence of only few NKX2.2+ cells (green) marked correct domain boundaries meaning that cells were not too ventralised into domain p3. Insets showed corresponding pictures with stained nuclei. Nuclei were counterstained with DAPI (blue). Scale bar 100 µm, valid for all images (TIFF 15294 kb)

18_2015_2084_MOESM5_ESM.tif

Supplementary material 5: Quantification of MN marker expression in smNPC-derived MN cultures on day 8 and day 27. For each individual cell line, MN numbers are given for (A) HB9+ MNs on day 8, (B) ISL1+ MNs on day 8, (C) HB9+ MNs on day 27 and (D) ISL1+ MNs on day 27. Error bars represented ± SEM of two (controls 35 m-r1, 61f-s2, 51f-s6, 62 m-s4) to three (all other lines) independent MN differentiation experiments (A and B) and n = MN differentiation experiments as outlined in tables below (C and D). Significance was calculated relative to 30 m-r12 (Kruskal–Wallis test, p < 0.05 = *) (TIFF 2297 kb)

18_2015_2084_MOESM6_ESM.tif

Supplementary material 6: Determination of gem numbers in different cell populations. Quantification of individual cell lines was displayed in left hand panels in (A) fibroblasts, (B) iPSCs, (C) smNPCs, (D) early MN culture (day 8) and (E) late MN culture (day 27). Right hand panels show grouped phenotypes. Error bars represented ± SEM of duplicates in fibroblasts (n = 2) and triplicates in all other cell populations (n = 3). MN cultures displayed data of two to three independent MN differentiation experiments (n = 3). Significance was calculated relative to 30 m-r12 (individual cell lines) and grouped five healthy controls (phenotype groups), respectively (Kruskal-Wallis test, p < 0.05 = *, p < 0.01 = **, p < 0.001 = ***, p < 0.0001 = ****) (TIFF 4443 kb)

18_2015_2084_MOESM7_ESM.tif

Supplementary material 7 Quantitative analyses of SMN mRNA expression levels. (A) SMN mRNA expression occurred in a similar pattern in fibroblasts, iPSCs and smNPCs with controls at the top, SMA III and asymptomatic in the middle and SMA I at the lowest range, matching the SMN1/SMN2 copy numbers. In MN culture, SMA III and asymptomatic exhibited an elevated SMN level in comparison to control. Quantification of SMN mRNA expression in individual lines of (B) fibroblasts, (C) iPSCs, (D) smNPCs and (E) MN cultures (day 27) is shown. Expression was normalised to total RNA levels and significance calculated relative to 30 m-r12 (individual cell lines) and grouped five healthy controls (phenotype groups), respectively (Kruskal–Wallis test, p < 0.05 = *, p < 0.01 = **). Error bars represented ± SEM of duplicates (n = 2). MN culture displayed data of one to three MN differentiation experiments (TIFF 3619 kb)

18_2015_2084_MOESM8_ESM.tif

Supplementary material 8: Quantitative analyses of PLS3 mRNA expression (A) In fibroblasts, PLS3 mRNA expression levels hardly differed among the classes. Expression rates in iPSCs showed a wider spreading with asymptomatic at the top, SMA I and controls following closely and SMA III at lowest level. While asymptomatic prevailed as strongest expressing group underlining putative overexpression, PLS3 expression continuously declined in controls resulting in barely any detectable PLS3 mRNA in MN cultures. Expression in SMA I also dropped to levels equal to SMA III in MN cultures. Importantly, PLS3 levels in asymptomatic were significantly elevated in comparison to controls. Quantification of PLS3 mRNA expression in individual lines of (B) fibroblasts, (C) iPSCs, (D) smNPCs and (E) MN cultures (day 27) is shown. Expression was normalised to total RNA levels and significance computed relative to 30 m-r12 (individual cell lines) and grouped five healthy controls (phenotype groups), respectively (Kruskal–Wallis test, p < 0.05 = *, p < 0.01 = **). Error bars represented ± SEM of duplicates (n = 2). MN culture displayed data of one to three MN differentiation experiments (TIFF 3735 kb)

18_2015_2084_MOESM9_ESM.tif

Supplementary material 9 Quantitative analyses of SMN protein expression. Western blot analysis from healthy controls, SMA I- and SMA III-affected and asymptomatic discordant siblings in (A) fibroblasts, (B) iPSCs, (C) smNPCs, (D) MN cultures (day 27). Expression was normalised to β-actin protein levels and significance calculated relative to 30 m-r12 (individual cell lines, left hand panels) and grouped five healthy controls (phenotype groups, right hand panels), respectively (Kruskal-Wallis test). In fibroblasts, iPSCs and smNPCs error bars represented ± SEM of duplicates (n = 2). In MN culture (day 27) error bars represented ± SEM of two to three independent MN differentiation experiments (TIFF 4061 kb)

18_2015_2084_MOESM10_ESM.tif

Supplementary material 10 Quantitative analyses of PLS3 protein expression. Western blot analysis from healthy controls, SMA I- and SMA III-affected and asymptomatic discordant siblings in (A) fibroblasts, (B) iPSCs, (C) smNPCs and (D) MN culture (day 27). Note that all asymptomatic cell lines except for HGK16 strikingly elevated PLS3 expression in comparison to SMA III and controls. Expression was normalised to β-actin protein levels and significance calculated relative to 30 m-r12 (individual cell lines, left hand panels) and grouped five healthy controls (phenotype groups, right hand panels), respectively (Kruskal–Wallis test, p < 0.01 = **, p < 0.001 = ***, p < 0.0001 = ****). In fibroblasts, iPSCs and smNPCs error bars represented ± SEM of duplicates (n = 2). In MN culture (day 27) error bars represented ± SEM of two to three independent MN differentiation experiments (TIFF 3854 kb)

18_2015_2084_MOESM11_ESM.tif

Supplementary material 11 Protein expression levels in control cell lines. (A) Representative Western blots from male and female controls visualising SMN and PLS3 protein expression in MN culture on day 27. (B) Graphs visualise PLS3 expression on RNA and protein level in MN cultures on day 27 according to gender of controls. Note that no different PLS3 protein expression was detected between male and female controls. (C) PLS3 in vitro siRNA-mediated knock-down shows the upper band as the correct specific PLS3 signal (marked with an arrow). The lower band is unspecific (TIFF 2958 kb)

18_2015_2084_MOESM12_ESM.tif

Supplementary material 12: Analysis of growth cones in mixed MN cultures on day 27. (A) Growth cones of male control 30 m-r12, female control 51f-s6, SMA III HGK27.13 and asymptomatic HGK21.8 visualising PLS3 (green) and actin (red) in growth cones and MAPT (magenta) in axons. Growth cones were magnified (dashed lines) in PLS3/actin merge in insets. Scale bar 10 µm, valid for all images. (B) Average area size was measured by encircling every growth cone individually. In comparison to controls, average area size was increased in SMA III and asymptomatic yet only in asymptomatic the difference was significant. (C) Co-localisation of PLS3 and actin was calculated by Pearson’s coefficient. However, Pearson’s coefficient accounted for a comparable value (control: 0.52; SMA III: 0.52; asymptomatic: 0.47) among all three phenotype groups indicating a medium co-localisation between PLS3 and actin. Growth cones were analysed from four independent MN differentiation experiments (n = 4). In total, n = 28 (controls 30 m-r12, 35 m-r1, 51f-s6), n = 22 (SMA III HGK27.10, HGK27.13) and n = 20 (asymptomatic HGK21.1, HGK21.8, HGK28.9) growth cones were analysed. Significance was calculated relative to grouped three healthy controls (Kruskal-Wallis test, p < 0.05 = *) (TIFF 33,462 kb)

18_2015_2084_MOESM13_ESM.tif

Supplementary material 13: Overview table listed specific features of iPSC lines such as originating parental fibroblast line, SMA phenotype, age, sex, SMN1/SMN2 copy number, elevated (↑) PLS3 expression levels in blood, reprogramming technique and validation criteria. Note that numbers in HGK denote individual line and additional clone number of the same line (TIFF 2953 kb)

Supplementary material 14: Primer nucleotide sequences (TIFF 1860 kb)

18_2015_2084_MOESM15_ESM.tif

Supplementary material 15: List of primary antibodies and respective secondary antibodies (gt = goat, ms = mouse, rb = rabbit) (TIFF 2967 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Heesen, L., Peitz, M., Torres-Benito, L. et al. Plastin 3 is upregulated in iPSC-derived motoneurons from asymptomatic SMN1-deleted individuals. Cell. Mol. Life Sci. 73, 2089–2104 (2016). https://doi.org/10.1007/s00018-015-2084-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00018-015-2084-y

Keywords

Navigation