Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Dopamine induces soluble α-synuclein oligomers and nigrostriatal degeneration

Abstract

Parkinson's disease (PD) is defined by the loss of dopaminergic neurons in the substantia nigra and the formation of Lewy body inclusions containing aggregated α-synuclein. Efforts to explain dopamine neuron vulnerability are hindered by the lack of dopaminergic cell death in α-synuclein transgenic mice. To address this, we manipulated both dopamine levels and α-synuclein expression. Nigrally targeted expression of mutant tyrosine hydroxylase with enhanced catalytic activity increased dopamine levels without damaging neurons in non-transgenic mice. In contrast, raising dopamine levels in mice expressing human A53T mutant α-synuclein induced progressive nigrostriatal degeneration and reduced locomotion. Dopamine elevation in A53T mice increased levels of potentially toxic α-synuclein oligomers, resulting in conformationally and functionally modified species. Moreover, in genetically tractable Caenorhabditis elegans models, expression of α-synuclein mutated at the site of interaction with dopamine prevented dopamine-induced toxicity. These data suggest that a unique mechanism links two cardinal features of PD: dopaminergic cell death and α-synuclein aggregation.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: TH-RREE lentiviral vector increases dopamine levels and causes hyperactivity in NonTg mice.
Figure 2: Dopamine-induced neurodegeneration of the SN is dependent on α-synuclein.
Figure 3: Dopaminergic neurodegeneration in A53T TH-RREE mice is progressive and ultimately leads to locomotor deficit.
Figure 4: Dopamine induces conformationally distinct α-synuclein oligomers in the mouse brain.
Figure 5: C. elegans expressing A53T α-synuclein lacking the site of interaction with dopamine are resistant to dopamine neurotoxicity.

Similar content being viewed by others

References

  1. Ehringer, H. & Hornykiewicz, O. Verteilung von noradrenalin und dopamin (3-hydroxytyramin) im gehirn des menschen und ihr verhalten bei erkrankungen des extrapyramidalen systems. [Distribution of noradrenaline and dopamine (3-hydroxytyramine) in the human brain and their behavior in diseases of the extrapyramidal system]. Klin. Wochenschr. 38, 1236–1239 (1960).

    CAS  PubMed  Google Scholar 

  2. Graham, D.G. Oxidative pathways for catecholamines in the genesis of neuromelanin and cytotoxic quinones. Mol. Pharmacol. 14, 633–643 (1978).

    CAS  PubMed  Google Scholar 

  3. Jenner, P. & Olanow, C.W. Oxidative stress and the pathogenesis of Parkinson's disease. Neurology 47 Suppl 3: S161–S170 (1996).

    CAS  PubMed  Google Scholar 

  4. Dexter, D.T. et al. Basal lipid peroxidation in substantia nigra is increased in Parkinson's disease. J. Neurochem. 52, 381–389 (1989).

    CAS  PubMed  Google Scholar 

  5. Zhang, J. et al. Parkinson's disease is associated with oxidative damage to cytoplasmic DNA and RNA in substantia nigra neurons. Am. J. Pathol. 154, 1423–1429 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Giasson, B.I. et al. Oxidative damage linked to neurodegeneration by selective alpha-synuclein nitration in synucleinopathy lesions. Science 290, 985–989 (2000).

    CAS  PubMed  Google Scholar 

  7. Hastings, T.G., Lewis, D.A. & Zigmond, M.J. Role of oxidation in the neurotoxic effects of intrastriatal dopamine injections. Proc. Natl. Acad. Sci. USA 93, 1956–1961 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Colebrooke, R.E. et al. Age-related decline in striatal dopamine content and motor performance occurs in the absence of nigral cell loss in a genetic mouse model of Parkinson's disease. Eur. J. Neurosci. 24, 2622–2630 (2006).

    PubMed  Google Scholar 

  9. Caudle, W.M. et al. Reduced vesicular storage of dopamine causes progressive nigrostriatal neurodegeneration. J. Neurosci. 27, 8138–8148 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Polymeropoulos, M.H. et al. Mutation in the alpha-synuclein gene identified in families with Parkinson's disease. Science 276, 2045–2047 (1997).

    CAS  PubMed  Google Scholar 

  11. Krüger, R. et al. Ala30Pro mutation in the gene encoding alpha-synuclein in Parkinson's disease. Nat. Genet. 18, 106–108 (1998).

    PubMed  Google Scholar 

  12. Singleton, A.B. et al. alpha-Synuclein locus triplication causes Parkinson's disease. Science 302, 841 (2003).

    CAS  PubMed  Google Scholar 

  13. Chartier-Harlin, M.-C. et al. Alpha-synuclein locus duplication as a cause of familial Parkinson's disease. Lancet 364, 1167–1169 (2004).

    CAS  PubMed  Google Scholar 

  14. Zarranz, J.J. et al. The new mutation, E46K, of alpha-synuclein causes Parkinson and Lewy body dementia. Ann. Neurol. 55, 164–173 (2004).

    CAS  PubMed  Google Scholar 

  15. Proukakis, C. et al. A novel α-synuclein missense mutation in Parkinson disease. Neurology 80, 1062–1064 (2013).

    PubMed  PubMed Central  Google Scholar 

  16. Lesage, S. et al. French Parkinson's Disease Genetics Study Group. G51D α-synuclein mutation causes a novel parkinsonian-pyramidal syndrome. Ann. Neurol. 73, 459–471 (2013).

    CAS  PubMed  Google Scholar 

  17. Pasanen, P. et al. Novel α-synuclein mutation A53E associated with atypical multiple system atrophy and Parkinson's disease-type pathology. Neurobiol. Aging 35, 2180.e1–2180.e5 (2014).

    CAS  Google Scholar 

  18. Spillantini, M.G. et al. Alpha-synuclein in Lewy bodies. Nature 388, 839–840 (1997).

    CAS  PubMed  Google Scholar 

  19. Baba, M. et al. Aggregation of alpha-synuclein in Lewy bodies of sporadic Parkinson's disease and dementia with Lewy bodies. Am. J. Pathol. 152, 879–884 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Iwai, A. et al. The precursor protein of non-A beta component of Alzheimer's disease amyloid is a presynaptic protein of the central nervous system. Neuron 14, 467–475 (1995).

    CAS  PubMed  Google Scholar 

  21. George, J.M., Jin, H., Woods, W.S. & Clayton, D.F. Characterization of a novel protein regulated during the critical period for song learning in the zebra finch. Neuron 15, 361–372 (1995).

    CAS  PubMed  Google Scholar 

  22. Abeliovich, A. et al. Mice lacking alpha-synuclein display functional deficits in the nigrostriatal dopamine system. Neuron 25, 239–252 (2000).

    CAS  PubMed  Google Scholar 

  23. Murphy, D.D., Rueter, S.M., Trojanowski, J.Q. & Lee, V.M. Synucleins are developmentally expressed, and alpha-synuclein regulates the size of the presynaptic vesicular pool in primary hippocampal neurons. J. Neurosci. 20, 3214–3220 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Conway, K.A., Rochet, J.C., Bieganski, R.M. & Lansbury, P.T. Jr. Kinetic stabilization of the alpha-synuclein protofibril by a dopamine-alpha-synuclein adduct. Science 294, 1346–1349 (2001).

    CAS  PubMed  Google Scholar 

  25. Norris, E.H. et al. Reversible inhibition of alpha-synuclein fibrillization by dopaminochrome-mediated conformational alterations. J. Biol. Chem. 280, 21212–21219 (2005).

    CAS  PubMed  Google Scholar 

  26. Mazzulli, J.R. et al. Cytosolic catechols inhibit alpha-synuclein aggregation and facilitate the formation of intracellular soluble oligomeric intermediates. J. Neurosci. 26, 10068–10078 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Mazzulli, J.R., Armakola, M., Dumoulin, M., Parastatidis, I. & Ischiropoulos, H. Cellular oligomerization of alpha-synuclein is determined by the interaction of oxidized catechols with a C-terminal sequence. J. Biol. Chem. 282, 31621–31630 (2007).

    CAS  PubMed  Google Scholar 

  28. Herrera, F.E. et al. Inhibition of alpha-synuclein fibrillization by dopamine is mediated by interactions with five C-terminal residues and with E83 in the NAC region. PLoS One 3, e3394 (2008).

    PubMed  PubMed Central  Google Scholar 

  29. Martinez-Vicente, M. et al. Dopamine-modified alpha-synuclein blocks chaperone-mediated autophagy. J. Clin. Invest. 118, 777–788 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Choi, B.K. et al. Large α-synuclein oligomers inhibit neuronal SNARE-mediated vesicle docking. Proc. Natl. Acad. Sci. USA 110, 4087–4092 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Nakashima, A. et al. The mutation of two amino acid residues in the N-terminus of tyrosine hydroxylase (TH) dramatically enhances the catalytic activity in neuroendocrine AtT-20 cells. J. Neurochem. 82, 202–206 (2002).

    CAS  PubMed  Google Scholar 

  32. Giasson, B.I. et al. Neuronal alpha-synucleinopathy with severe movement disorder in mice expressing A53T human alpha-synuclein. Neuron 34, 521–533 (2002).

    CAS  PubMed  Google Scholar 

  33. Tieu, K. et al. D-beta-hydroxybutyrate rescues mitochondrial respiration and mitigates features of Parkinson disease. J. Clin. Invest. 112, 892–901 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Tsika, E. et al. Distinct region-specific alpha-synuclein oligomers in A53T transgenic mice: implications for neurodegeneration. J. Neurosci. 30, 3409–3418 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Souza, J.M., Giasson, B.I., Chen, Q., Lee, V.M. & Ischiropoulos, H. Dityrosine cross-linking promotes formation of stable alpha -synuclein polymers. Implication of nitrative and oxidative stress in the pathogenesis of neurodegenerative synucleinopathies. J. Biol. Chem. 275, 18344–18349 (2000).

    CAS  PubMed  Google Scholar 

  36. Luk, K.C. et al. Pathological α-synuclein transmission initiates Parkinson-like neurodegeneration in nontransgenic mice. Science 338, 949–953 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Lakso, M. et al. Dopaminergic neuronal loss and motor deficits in Caenorhabditis elegans overexpressing human alpha-synuclein. J. Neurochem. 86, 165–172 (2003).

    CAS  PubMed  Google Scholar 

  38. Cao, S., Gelwix, C.C., Caldwell, K.A. & Caldwell, G.A. Torsin-mediated protection from cellular stress in the dopaminergic neurons of Caenorhabditis elegans. J. Neurosci. 25, 3801–3812 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Karpinar, D.P. et al. Pre-fibrillar alpha-synuclein variants with impaired beta-structure increase neurotoxicity in Parkinson's disease models. EMBO J. 28, 3256–3268 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Cao, P. et al. Alpha-synuclein disrupted dopamine homeostasis leads to dopaminergic neuron degeneration in Caenorhabditis elegans. PLoS One 5, e9312 (2010).

    PubMed  PubMed Central  Google Scholar 

  41. Fahn, S. et al. Levodopa and the progression of Parkinson's disease. N. Engl. J. Med. 351, 2498–2508 (2004).

    CAS  PubMed  Google Scholar 

  42. Olanow, C.W. Levodopa: effect on cell death and the natural history of Parkinson's disease. Mov. Disord. 30, 37–44 (2015).

    CAS  PubMed  Google Scholar 

  43. Lohr, K.M. et al. Increased vesicular monoamine transporter enhances dopamine release and opposes Parkinson disease-related neurodegeneration in vivo. Proc. Natl. Acad. Sci. USA 111, 9977–9982 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Rochet, J.C., Conway, K.A. & Lansbury, P.T. Jr. Inhibition of fibrillization and accumulation of prefibrillar oligomers in mixtures of human and mouse alpha-synuclein. Biochemistry 39, 10619–10626 (2000).

    CAS  PubMed  Google Scholar 

  45. Specht, C.G. & Schoepfer, R. Deletion of the alpha-synuclein locus in a subpopulation of C57BL/6J inbred mice. BMC Neurosci. 2, 11 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Mosharov, E.V. et al. Interplay between cytosolic dopamine, calcium and alpha-synuclein causes selective death of substantia nigra neurons. Neuron 62, 218–229 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Winner, B. et al. In vivo demonstration that alpha-synuclein oligomers are toxic. Proc. Natl. Acad. Sci. USA 108, 4194–4199 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Xu, J. et al. Dopamine-dependent neurotoxicity of alpha-synuclein: a mechanism for selective neurodegeneration in Parkinson disease. Nat. Med. 8, 600–606 (2002).

    CAS  PubMed  Google Scholar 

  49. Danzer, K.M. et al. Different species of alpha-synuclein oligomers induce calcium influx and seeding. J. Neurosci. 27, 9220–9232 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Cheng, H.-C., Ulane, C.M. & Burke, R.E. Clinical progression in Parkinson disease and the neurobiology of axons. Ann. Neurol. 67, 715–725 (2010).

    PubMed  PubMed Central  Google Scholar 

  51. Brichta, L. et al. Identification of neurodegenerative factors using translatome-regulatory network analysis. Nat. Neurosci. 18, 1325–1333 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Khasnavis, S., Ghosh, A., Roy, A. & Pahan, K. Castration induces Parkinson disease pathologies in young male mice via inducible nitric-oxide synthase. J. Biol. Chem. 288, 20843–20855 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Qian, Y., Forssberg, H. & Diaz Heijtz, R. Motor skill learning is associated with phase-dependent modifications in the striatal cAMP/PKA/DARPP-32 signaling pathway in rodents. PLoS One 10, e0140974 (2015).

    PubMed  PubMed Central  Google Scholar 

  54. Waxman, E.A., Duda, J.E. & Giasson, B.I. Characterization of antibodies that selectively detect alpha-synuclein in pathological inclusions. Acta Neuropathol. 116, 37–46 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Luk, K.C. et al. Intracerebral inoculation of pathological α-synuclein initiates a rapidly progressive neurodegenerative α-synucleinopathy in mice. J. Exp. Med. 209, 975–986 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Eales, K.L. et al. The MK2/3 cascade regulates AMPAR trafficking and cognitive flexibility. Nat. Commun. 5, 4701 (2014).

    CAS  PubMed  Google Scholar 

  57. Sun, Y. et al. The expression and significance of neuronal iconic proteins in podocytes. PLoS One 9, e93999 (2014).

    PubMed  PubMed Central  Google Scholar 

  58. Sarkar, S., Davies, J.E., Huang, Z., Tunnacliffe, A. & Rubinsztein, D.C. Trehalose, a novel mTOR-independent autophagy enhancer, accelerates the clearance of mutant huntingtin and alpha-synuclein. J. Biol. Chem. 282, 5641–5652 (2007).

    CAS  PubMed  Google Scholar 

  59. Mazzulli, J.R., Burbulla, L.F., Krainc, D. & Ischiropoulos, H. Detection of free and protein-bound ortho-quinones by near-infrared fluorescence. Anal. Chem. 88, 2399–2405 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Brenner, S. The genetics of Caenorhabditis elegans. Genetics 77, 71–94 (1974).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Berkowitz, L.A., Knight, A.L., Caldwell, G.A. & Caldwell, K.A. Generation of stable transgenic C. elegans using microinjection. J. Vis. Exp. 18, 833 (2008).

    Google Scholar 

  62. Hamamichi, S. et al. Hypothesis-based RNAi screening identifies neuroprotective genes in a Parkinson's disease model. Proc. Natl. Acad. Sci. USA 105, 728–733 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Volpicelli-Daley, L.A. et al. Exogenous α-synuclein fibrils induce Lewy body pathology leading to synaptic dysfunction and neuron death. Neuron 72, 57–71 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank T. Clarke and T. Pierson from the Wolfe lab for technical assistance with animals and for helpful advice on vector production. We thank H. Bennett from the Kalb lab for initial work on the dopamine neuron degeneration assay in C. elegans and input on worm breeding. We thank L. Spruce, H. Fazelinia and S. Seeholzer from the Children's Hospital of Philadelphia Proteomic Core facility for mass spectrometry. We thank V. Lee (University of Pennsylvania) and G. Miller (Emory University) for generously providing α-synuclein and VMAT2 antibodies, respectively, and S. Przedborski (Columbia University) for the use of stereology equipment and helpful feedback on the manuscript. Finally, we thank R. Lightfoot and members of the Ischiropoulos lab for productive discussions and technical support. This work was supported by grants from the US National Institutes of Health: AG013966 (H.I.), NS038690 (J.H.W.) and NS087077 and NS052325 (R.G.K.). D.E.M. was supported by the US National Institutes of Health Ruth L. Kirschstein National Research Service Award Individual Predoctoral Fellowship NS087779. The content of this work is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

Author information

Authors and Affiliations

Authors

Contributions

D.E.M., E.T., J.R.M., R.G.K. and H.I. conceived and designed the experiments. D.E.M., E.T., H.K., N.S.G., M.J.D., S.D., P.G., J.L.G. and V.X.T. performed the experiments and analyzed the data. D.E.M. wrote the paper, with important contributions from J.R.M., E.T., J.H.W., R.G.K., K.A.C., G.A.C. and H.I.

Corresponding author

Correspondence to Harry Ischiropoulos.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Increased TH expression and steady-state catecholamine concentrations in SH-SY5Y cells transduced with TH-RREE vector.

(a) Schematic representation of the lentiviral vectors used in this study. TH-RREE contained the gene encoding human tyrosine hydroxylase isoform 1 (hTH-1) with mutations R37E, R38E rendering TH insensitive to feedback inhibition by dopamine. The gene was absent from the empty vector control (CtrlVect), while all other elements were retained. (b) Western blot analysis of lysates showed TH protein only in TH-RREE transduced cells. Neural specific enolase (NSE) was used as a loading control. (n = 4 independent cultures per group). (c) TH immunofluorescence confirmed expression in TH-RREE transduced cells. Scale bar, 100 μm. (n = 5 independent cultures per group). (d) Catecholamines were detected in cells treated with TH-RREE lentivirus. ND, not detected. (n = 3 independent cultures per group except n = 6 independent cultures for DA TH-RREE and n = 6 independent cultures for DOPAC TH-RREE). Box plots show median, 25th and 75th percentiles, minimum and maximum values.

Supplementary Figure 2 TH-RREE vector increases TH expression in A53T mice.

(a) TH levels were increased in the striatum in A53T TH-RREE mice compared with A53T CtrlVect at 5 mpi. GAPDH was used as a loading control. (n = 3 mice per group, P = 0.0048, t = 5.657, d.f. = 4; two-tailed unpaired Student’s t test). (b) TH staining was increased in both the SN (left panels) and striatum (right panels) of A53T TH-RREE mice at 5 mpi. Cresyl Violet (Nissl) counterstain. Scale bar, 10 μm. (n = 4 mice per group). Box plots show median, 25th and 75th percentiles, minimum and maximum values. **P < 0.01.

Supplementary Figure 3 Postsynaptic markers remain unchanged by lentiviral vector treatment.

Striatal levels of the D1 postsynaptic receptor were unaltered in TH-RREE injected A53T and NonTg mice compared with CtrlVect-injected mice at 5 mpi (a-b). The postsynaptic signaling molecule DARPP-32 was also unaffected by viral treatment (a,c). All quantification was normalized to actin loading control. (D1R, n = 3 mice per group except n = 4 mice for NonTg CtrlVect, F(3,9) = 0.2561; DARPP-32, n = 3 mice per group, F(3,10) = 0.3835; one-way ANOVA with Tukey’s correction for multiple comparisons). Box plots show median, 25th and 75th percentiles, minimum and maximum values.

Supplementary Figure 4 Loss of striatal dopamine in A53T mice cannot be accounted for by formation of dopamine-protein adducts.

(a) Triton-soluble (S) and insoluble (P) extracts of striatal tissue from mice at 2.5 or 5 mpi were analyzed by SDS-PAGE and near-infrared fluorescence (nIRF). (b) Total levels of dopamine-protein adduct (combined soluble and insoluble nIRF signal) did not change either between injection groups or within groups over time. (n = 3 mice per group, F(3,8) = 1.763; one-way ANOVA with Tukey’s correction for multiple comparisons). Box plots show median, 25th and 75th percentiles, minimum and maximum values.

Supplementary Figure 5 Analysis of inclusion pathology and insoluble α-synuclein in A53T mice.

(a) Regardless of lentiviral treatment, α-synuclein inclusion pathology, as detected by Syn505 staining, was abundant in brainstem and absent in SN of A53T mice at 5 mpi. Scale bar, 20 μm. (n = 3 mice per group). (b-d) At 5 mpi, SN from A53T mice was sequentially extracted with buffers containing 1% Triton followed by 2% SDS. Triton-insoluble/SDS-soluble fractions from A53T TH-RREE and A53T CtrlVect contained similar levels of monomeric α-synuclein as assessed by LB509 (b), Syn211 (c), and SNL-4 (d) antibodies. Vimentin (Vim) was used as a loading control, and appears for Syn211 and SNL-4 as the same blot since the same membrane is shown. (LB509, n = 3 mice per group, P = 0.2605, t = 1.310, d.f. = 4; Syn211, n = 6 mice per group except n = 4 mice for CtrlVect, P = 0.7716, t = 0.3003, d.f. = 8; SNL-4, n = 6 mice per group except n = 4 mice for CtrlVect, P = 0.7355, t = 0.3498, d.f. = 8; two-tailed unpaired Student’s t test). Box plots show median, 25th and 75th percentiles, minimum and maximum values.

Supplementary Figure 6 Further characterization of α-synuclein and dopamine derived from A53T mice.

(a-b) Triton-soluble SN from A53T TH-RREE and A53T CtrlVect mice at 5 mpi was fractionated by size exclusion chromatography, and the resulting fractions were analyzed by western blot using α-synuclein antibodies Syn211 (C-terminal) (a) and SNL-4 (N-terminal) (b). Oligomer bands that uniquely appear in 122 Å fractions from TH-RREE mice are indicated by arrows. Neural specific enolase (NSE) was used as a loading control, and appears for Syn211 and SNL-4 as the same blot since the same membrane is shown. (n = 4 mice per group). (c) The SDS-PAGE gels were scanned for near-infrared fluorescence, which revealed oxidized dopamine (arrow) only in 72-122 Å fractions from A53T TH-RREE mice. (n = 4 mice per group).

Supplementary Figure 7 Biochemical and structural characterization of α-synuclein oligomers resulting from incubation with dopamine in vitro.

(a) Incubation of recombinant wild-type human α-synuclein under standard aggregation conditions (400 μM, 37°C, 1400 rpm) resulted in Thioflavin T (ThioT)-positive fibril formation, whereas ThioT signal was abolished in the presence of equimolar dopamine (DA). rfu, relative fluorescence units. The data are presented as mean ± s.e.m. (n = 3 independent experiments per group, P-DA day 4 versus +DA day 4 = 0.0045, P-DA day 6 versus +DA day 6 = 0.003, F(1,4) = 8.901; repeated measures two-way ANOVA with Bonferroni’s correction for multiple comparisons). (b) α-Synuclein fibrils exhibited characteristic β-sheet secondary structure by circular dichroism, while α-synuclein species resulting from interaction with dopamine maintained random coil structure on day 5 of incubation. Representative traces are shown. (n = 3 independent experiments per group). (c) Sedimentation analysis and coomassie blue staining revealed that over the 6-day incubation, in the absence of dopamine, soluble α-synuclein (supernatant) was lost while insoluble α-synuclein (pellet) accumulated. In the presence of dopamine, α-synuclein was detected as multiple soluble oligomer species. Similar to ex vivo oligomers extracted from A53T TH-RREE mice, in vitro generated oligomers were immunoreactive to Syn505 (N-terminal) and LB509 (C-terminal) α-synuclein antibodies. (n = 4 independent experiments per group). (d) Immunoelectron microscopy on α-synuclein from day 4 of incubation with dopamine confirmed both Syn505 and LB509 labeling indicating that dopamine-induced oligomers generated in vitro share common epitopes and potentially similar conformations to oligomers resulting from dopamine elevation in vivo. Scale bar, 20 nm. (n = 3 independent experiments per group). **P < 0.01.

Supplementary Figure 8 Dopamine-modified oligomers are not detected in motor cortex.

A53T TH-RREE and CtrlVect cortical tissue from 5 mpi was extracted with 1% Triton X-100 and the soluble fraction was further separated by size exclusion chromatography. Pooled fractions were analyzed by Western blotting using α-synuclein antibodies LB509 (a) and Syn211 (b). No differences in α-synuclein species were detected between TH-RREE and CtrlVect mice. Neural specific enolase (NSE) was used as a loading control. (n = 3 mice per group).

Supplementary Figure 9 α-Synuclein oligomers generated in vitro in the presence of dopamine become internalized and are toxic to primary neurons.

Hippocampal neurons at 7 days in vitro were exposed to 0.5 or 1 μM α-synuclein oligomer that had been prepared in the presence of dopamine (Olig), or 1 μM monomer (Mon), 1 μM dopamine that was incubated in parallel to oligomers but without α-synuclein (DA), or an equivalent dose of PBS. (a) Two weeks post-treatment with 1 μM Olig or PBS, cells were labeled with human α-synuclein antibodies in a two-stage protocol. Cells were live-incubated with Syn204 antibody to label extracellular α-synuclein. Following fixation and permeabilization, the cells were incubated with LB509 to label total α-synuclein. Confocal imaging showed puncta that were labeled by LB509 but not Syn204, indicating that exogenous oligomers had been internalized. These puncta were localized to neurites. Scale bar, 10 μm. (n = 3 independent cultures per group). (b-c) At two weeks post-treatments, cell viability was assayed using Calcein AM and propidium iodide (PI) dyes. The number of viable cells, defined as Calcein AM positive and PI negative, was significantly reduced by exposure to 1 μM Olig but was not affected by Mon or DA controls. Treatment with 0.5 μM Olig was insufficient to induce toxicity, indicating that the effect is dependent on dose. Scale bar, 100 μm. (n = 4 independent cultures per group except n = 3 independent cultures for PBS, n = 3 independent cultures for DA, and n = 5 independent cultures for 1 μM Olig, POlig 1 μM versus PBS = 0.0285, POlig 1 μM versus DA = 0.0308, POlig 1 μM versus Mon = 0.0098, F(4,14) = 6.385; one-way ANOVA with Tukey’s correction for multiple comparisons). Box plots show median, 25th and 75th percentiles, minimum and maximum values. *P < 0.05, **P < 0.01.

Supplementary Figure 10 Mass spectrometry-label free quantification of α-synuclein expression in C. elegans strains.

Parallel reaction monitoring mass spectrometry analysis of human A53T α-synuclein. Data depicts similar signal intensity of the primary fragmentation ion from each of the four unique peptides detected in the four C. elegans strains. The four peptides represent coverage of amino acids 11-21 and 33-43 of human α-synuclein. (n = 1 run per genotype).

Supplementary Figure 11 Full-length blots corresponding to cropped blots in the main figures.

(a) Full-length blot correspondng to Fig. 1a. TH (top arrow) and GAPDH (bottom arrow) were imaged in the same channel. (b) Full-length blots correspondng to Fig. 2e. For each genotype (NonTg, A53T), three blots are shown since DAT and Actin were imaged in separate channels. The left blot shows DAT (arrow), the middle shows Actin (arrow), and the right blot shows the merge of the two channels. (c) Full-length blots correspondng to Fig. 4a. Three blots are shown since α-synuclein and NSE were imaged in separate channels. The left blot shows α-synuclein (arrow), the middle shows NSE (arrow), and the right blot shows the merge of the two channels. (d) Full-length blots correspondng to Fig. 4c. Two blots are shown since α-synuclein and NSE were imaged in separate channels. The top blot shows NSE (arrow), and the bottom blot shows the merge of the two channels. The full-length α-synuclein blot appears in Fig. 4c. * denotes non-specific bands, or bands resulting from incubation with other antibodies not pertinent to the present experiment.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mor, D., Tsika, E., Mazzulli, J. et al. Dopamine induces soluble α-synuclein oligomers and nigrostriatal degeneration. Nat Neurosci 20, 1560–1568 (2017). https://doi.org/10.1038/nn.4641

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4641

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing