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:

Transneuronal propagation of mutant huntingtin contributes to non–cell autonomous pathology in neurons

An Author Correction to this article was published on 17 July 2018

This article has been updated

Abstract

In Huntington's disease (HD), whether transneuronal spreading of mutant huntingtin (mHTT) occurs and its contribution to non–cell autonomous damage in brain networks is largely unknown. We found mHTT spreading in three different neural network models: human neurons integrated in the neural network of organotypic brain slices of HD mouse model, an ex vivo corticostriatal slice model and the corticostriatal pathway in vivo. Transneuronal propagation of mHTT was blocked by two different botulinum neurotoxins, each known for specifically inactivating a single critical component of the synaptic vesicle fusion machinery. Moreover, healthy human neurons in HD mouse model brain slices displayed non–cell autonomous changes in morphological integrity that were more pronounced when these neurons bore mHTT aggregates. Altogether, our findings suggest that transneuronal propagation of mHTT might be an important and underestimated contributor to the pathophysiology of HD.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Maturation of hESC-derived neurons co-cultured in mouse wild-type OTBSs.
Figure 2: Non–cell autonomous morphological changes in hGFP neurons cultured in R6/2 OTBSs.
Figure 3: Transneuronal spreading of mHTT.
Figure 4: Cellular atrophy is associated with the presence of mHTT.
Figure 5: Propagation of mHTT aggregate pathology in mixed-genotype corticostriatal brain slice cultures.
Figure 6: Transneuronal propagation of mHTT in vivo.
Figure 7: Transneuronal propagation of mHTT requires synaptic machinery.

Similar content being viewed by others

Change history

  • 17 July 2018

    In the version of this article initially published, the catalog numbers for BoNT A and B were given in the Methods section as T0195 and T5644; the correct numbers are B8776 and B6403. The error has been corrected in the HTML and PDF versions of the article.

References

  1. The Huntington's Disease Collaborative Research Group. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington's disease chromosomes. Cell 72, 971–983 (1993).

    Google Scholar 

  2. Ross, C.A. & Tabrizi, S.J. Huntington's disease: from molecular pathogenesis to clinical treatment. Lancet Neurol. 10, 83–98 (2011).

    CAS  PubMed  Google Scholar 

  3. Landles, C. et al. Proteolysis of mutant huntingtin produces an exon 1 fragment that accumulates as an aggregated protein in neuronal nuclei in Huntington disease. J. Biol. Chem. 285, 8808–8823 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Sathasivam, K. et al. Aberrant splicing of HTT generates the pathogenic exon 1 protein in Huntington disease. Proc. Natl. Acad. Sci. USA 110, 2366–2370 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Mangiarini, L. et al. Exon 1 of the HD gene with an expanded CAG repeat is sufficient to cause a progressive neurological phenotype in transgenic mice. Cell 87, 493–506 (1996).

    CAS  PubMed  Google Scholar 

  6. Arrasate, M. & Finkbeiner, S. Protein aggregates in Huntington's disease. Exp. Neurol. 238, 1–11 (2012).

    CAS  PubMed  Google Scholar 

  7. Raymond, L.A. et al. Pathophysiology of Huntington's disease: time-dependent alterations in synaptic and receptor function. Neuroscience 198, 252–273 (2011).

    CAS  PubMed  Google Scholar 

  8. Hong, S.L. et al. Dysfunctional behavioral modulation of corticostriatal communication in the R6/2 mouse model of Huntington's disease. PLoS ONE 7, e47026 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Ehrlich, M.E. Huntington's disease and the striatal medium spiny neuron: cell-autonomous and non-cell-autonomous mechanisms of disease. Neurotherapeutics 9, 270–284 (2012).

    PubMed  PubMed Central  Google Scholar 

  10. Cepeda, C. et al. Transient and progressive electrophysiological alterations in the corticostriatal pathway in a mouse model of Huntington's disease. J. Neurosci. 23, 961–969 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Stack, E.C. & Ferrante, R.J. Huntington's disease: progress and potential in the field. Expert Opin. Investig. Drugs 16, 1933–1953 (2007).

    CAS  PubMed  Google Scholar 

  12. Gu, X. et al. Pathological cell-cell interactions elicited by a neuropathogenic form of mutant Huntingtin contribute to cortical pathogenesis in HD mice. Neuron 46, 433–444 (2005).

    CAS  PubMed  Google Scholar 

  13. Gray, M. et al. Full-length human mutant huntingtin with a stable polyglutamine repeat can elicit progressive and selective neuropathogenesis in BACHD mice. J. Neurosci. 28, 6182–6195 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Aguzzi, A. Cell biology: beyond the prion principle. Nature 459, 924–925 (2009).

    CAS  PubMed  Google Scholar 

  15. Guo, J.L. & Lee, V.M. Cell-to-cell transmission of pathogenic proteins in neurodegenerative diseases. Nat. Med. 20, 130–138 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Raj, A., Kuceyeski, A. & Weiner, M. A network diffusion model of disease progression in dementia. Neuron 73, 1204–1215 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Zhou, J., Gennatas, E.D., Kramer, J.H., Miller, B.L. & Seeley, W.W. Predicting regional neurodegeneration from the healthy brain functional connectome. Neuron 73, 1216–1227 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Ren, P.H. et al. Cytoplasmic penetration and persistent infection of mammalian cells by polyglutamine aggregates. Nat. Cell Biol. 11, 219–225 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Gousset, K. et al. Prions hijack tunnelling nanotubes for intercellular spread. Nat. Cell Biol. 11, 328–336 (2009).

    CAS  PubMed  Google Scholar 

  20. Constanzo, M. et al. Transfer of polyglutamine aggregates in neuronal cells occurs in tunneling nanotubes. J. Cell Sci. 126, 3678–3685 (2013).

    Google Scholar 

  21. Simpson, L.L. Identification of the major steps in botulinum toxin action. Annu. Rev. Pharmacol. Toxicol. 44, 167–193 (2004).

    CAS  PubMed  Google Scholar 

  22. Südhof, T.C. & Rizo, J. Synaptic vesicle exocytosis. Cold Spring Harb. Perspect. Biol. 3, a005637 (2011).

    PubMed  PubMed Central  Google Scholar 

  23. Letzkus, J.J. et al. A disinhibitory microcircuit for associative fear learning in the auditory cortex. Nature 480, 331–335 (2011).

    CAS  PubMed  Google Scholar 

  24. Svenningsson, P. et al. DARPP-32: an integrator of neurotransmission. Annu. Rev. Pharmacol. Toxicol. 44, 269–296 (2004).

    CAS  PubMed  Google Scholar 

  25. Greig, L.C., Woodworth, M.B., Galazo, M.J., Padmanabhan, H. & Macklis, J.D. Molecular logic of neocortical projection neuron specification, development and diversity. Nat. Rev. Neurosci. 14, 755–769 (2013).

    CAS  PubMed  Google Scholar 

  26. Yu, D.X., Marchetto, M.C. & Gage, F.H. How to make a hippocampal dentate gyrus granule neuron. Development 141, 2366–2375 (2014).

    CAS  PubMed  Google Scholar 

  27. Proenca, C.C. et al. Atg4b-dependent autophagic flux alleviates Huntington's disease progression. PLoS ONE 8, e68357 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Hickey, M.A. et al. Extensive early motor and non-motor behavioral deficits are followed by striatal neuronal loss in knock-in Huntington's disease mice. Neuroscience 157, 280–295 (2008).

    CAS  PubMed  Google Scholar 

  29. Li, H. et al. Ultrastructural localization and progressive formation of neuropil aggregates in Huntington's disease transgenic mice. Hum. Mol. Genet. 8, 1227–1236 (1999).

    CAS  PubMed  Google Scholar 

  30. Klapstein, G.J. et al. Electrophysiological and morphological changes in striatal spiny neurons in R6/2 Huntington's disease transgenic mice. J. Neurophysiol. 86, 2667–2677 (2001).

    CAS  PubMed  Google Scholar 

  31. Esposito, M.S., Capelli, P. & Arber, S. Brainstem nucleus MdV mediates skilled forelimb motor tasks. Nature 508, 351–356 (2014).

    CAS  PubMed  Google Scholar 

  32. Li, H., Wyman, T., Yu, Z.X., Li, S.H. & Li, X.J. Abnormal association of mutant huntingtin with synaptic vesicles inhibits glutamate release. Hum. Mol. Genet. 12, 2021–2030 (2003).

    CAS  PubMed  Google Scholar 

  33. Wickersham, I.R., Sullivan, H.A. & Seung, H.S. Axonal and subcellular labelling using modified rabies viral vectors. Nat. Commun. 4, 2332 (2013).

    PubMed  Google Scholar 

  34. Pivetta, C., Esposito, M.S., Sigrist, M. & Arber, S. Motor-circuit communication matrix from spinal cord to brainstem neurons revealed by developmental origin. Cell 156, 537–548 (2014).

    CAS  PubMed  Google Scholar 

  35. Foran, P.G. et al. Evaluation of the therapeutic usefulness of botulinum neurotoxin B, C1, E, and F compared with the long lasting type A. Basis for distinct durations of inhibition of exocytosis in central neurons. J. Biol. Chem. 278, 1363–1371 (2003).

    CAS  PubMed  Google Scholar 

  36. Vonsattel, J.P. Huntington disease models and human neuropathology: similarities and differences. Acta Neuropathol. 115, 55–69 (2008).

    PubMed  Google Scholar 

  37. Ma, L. et al. Human embryonic stem cell–derived GABA neurons correct locomotion deficits in quinolinic acid–lesioned mice. Cell Stem Cell 10, 455–464 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Cicchetti, F. et al. Neural transplants in patients with Huntington's disease undergo disease-like neuronal degeneration. Proc. Natl. Acad. Sci. USA 106, 12483–12488 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Aguzzi, A. & Rajendran, L. The transcellular spread of cytosolic amyloids, prions, and prionoids. Neuron 64, 783–790 (2009).

    CAS  PubMed  Google Scholar 

  40. Braak, H. et al. Staging of brain pathology related to sporadic Parkinson's disease. Neurobiol. Aging 24, 197–211 (2003).

    PubMed  Google Scholar 

  41. de Calignon, A. et al. Propagation of tau pathology in a model of early Alzheimer's disease. Neuron 73, 685–697 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Liu, L. et al. Trans-synaptic spread of tau pathology in vivo. PLoS ONE 7, e31302 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Dujardin, S. et al. Neuron-to-neuron wild-type Tau protein transfer through a trans-synaptic mechanism: relevance to sporadic tauopathies. Acta Neuropathol. Commun. 2, 14 (2014).

    PubMed  PubMed Central  Google Scholar 

  44. Neale, E.A., Bowers, L.M., Jia, M., Bateman, K.E. & Williamson, L.C. Botulinum neurotoxin A blocks synaptic vesicle exocytosis but not endocytosis at the nerve terminal. J. Cell Biol. 147, 1249–1260 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Pooler, A.M., Phillips, E.C., Lau, D.H., Noble, W. & Hanger, D.P. Physiological release of endogenous tau is stimulated by neuronal activity. EMBO Rep. 14, 389–394 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Yamada, K. et al. Neuronal activity regulates extracellular tau in vivo. J. Exp. Med. 211, 387–393 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Gutekunst, C.A. et al. Nuclear and neuropil aggregates in Huntington's disease: relationship to neuropathology. J. Neurosci. 19, 2522–2534 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Tabrizi, S.J. et al. Biological and clinical changes in premanifest and early stage Huntington's disease in the TRACK-HD study: the 12-month longitudinal analysis. Lancet Neurol. 10, 31–42 (2011).

    PubMed  Google Scholar 

  50. Thomson, J.A. et al. Embryonic stem cell lines derived from human blastocysts. Science 282, 1145–1147 (1998).

    CAS  PubMed  Google Scholar 

  51. Weiss, A. et al. Single-step detection of mutant huntingtin in animal and human tissues: a bioassay for Huntington's disease. Anal. Biochem. 395, 8–15 (2009).

    CAS  PubMed  Google Scholar 

  52. Gogolla, N., Galimberti, I., DePaola, V. & Caroni, P. Staining protocol for organotypic hippocampal slice cultures. Nat. Protoc. 1, 2452–2456 (2006).

    CAS  PubMed  Google Scholar 

  53. Tchorz, J.S. et al. A modified RMCE-compatible Rosa26 locus for the expression of transgenes from exogenous promoters. PLoS ONE 7, e30011 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Ko, J., Ou, S. & Patterson, P.H. New anti-huntingtin monoclonal antibodies: implications for huntingtin conformation and its binding proteins. Brain Res. Bull. 56, 319–329 (2001).

    CAS  PubMed  Google Scholar 

  55. Franklin, J. & Paxinos, G. Mouse Brain in Stereotaxic Coordinates (Academic Press, 2012).

  56. Baldo, B. et al. TR-FRET-based duplex immunoassay reveals an inverse correlation of soluble and aggregated mutant huntingtin in huntington's disease. Chem. Biol. 19, 264–275 (2012).

    CAS  PubMed  Google Scholar 

  57. Bobrowska, A., Donmez, G., Weiss, A., Guarente, L. & Bates, G. SIRT2 ablation has no effect on tubulin acetylation in brain, cholesterol biosynthesis or the progression of Huntington's disease phenotypes in vivo. PLoS ONE 7, e34805 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Di Giorgio, F.P., Carrasco, M.A., Siao, M.C., Maniatis, T. & Eggan, K. Non–cell autonomous effect of glia on motor neurons in an embryonic stem cell-based ALS model. Nat. Neurosci. 10, 608–614 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank A. Weiss (Promidis) for providing the mHTTQ72. D. Shimshek, T. Haffner and C. Stemmelen for help with stereotactic injection, C. Schöffel-Mattes for help with electrophysiological recording, T. Doll and I. Fruh for help with cell culture, P. Caroni, F.H. Gage and K. Eggan for comments on the paper, and R. Kuhn and G. Bilbe for providing valuable scientific input and continuous support. M.S.E. was supported by a long-term fellowship of the Human Frontier Science Program, and S.A. by a European Research Council Advanced Grant and the Swiss National Science Foundation. P.B. and A.L. are supported by the Novartis Research Foundation.

Author information

Authors and Affiliations

Authors

Contributions

E.P.-V., C.R., H.v.d.P. and F.P.D. designed the study. E.P.-V. and C.R. performed the experiments with contributions from S.F., D.B., C.G. and M.M. (generation of iPSC- and hESC-derived neurons) and from I.G. and M.B. (organotypic mouse brain slice technology). M.S.E. performed in vivo stereotaxic injections of lenti-mHTT. S.A. contributed to the design of the rabies monosynaptic tracing experiments and in vivo stereotactic injection of lenti-mHTT. P.B. performed optogenetic-based experiments and recordings from mixed-culture OTBS, with input and supervision from A.L. E.P.-V., C.R., H.v.d.P. and F.P.D. wrote the manuscript with input from T.B. and S.A.

Corresponding author

Correspondence to Francesco Paolo Di Giorgio.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Characterization of hGFP neurons

(a) Schematic representation of sequential experimental steps applied to differentiate human stem cells into neuronal precursors. Nucleofection with CAG-GFP reporter plasmid occurs before starting the co-culture with mouse OTBSs. (b) (top) Schematic representation of a corticostriatal OTBS. Dotted square indicates position of low magnification image. (bottom) Image shows hGPF-neurons (green), synaptophysin (red) and vGlut1 (purple) in mouse OTBS after 4 weeks of co-culture. (c) hGFP-neurons express human antigen (HA) and pan-neuronal marker Tuj1. (d) hGFP-neurons express Tuj1 and lack expression of astroglial marker GFAP. Histogram showing quantification of Tuj1 positive or GFAP positive GFP expressing cells in wt OTBSs at 4 weeks of co-culture (n = 52 cells, from 10 slices from 3 mice). (e) At 4 weeks of co-culture hGFP-neurons formed functional connections with mouse neurons with a probability of 60% (n=5 cells, from 3 sections, from 3 mice).

Supplementary Figure 2 hGFP neurons express brain region–specific neuronal cell-type markers in 6-week-old wild-type OTBSs.

(a) hGFP-neuron located in striatum and expressing the MSN marker DARPP-32. (b) hGFP-neuron located in cortex and expressing the upper cortical layer-specific neuronal marker SATB2. (c) hGFP-neuron located in a deep cortical layer and expressing the layer 6-specific marker Tbr1. (d) hGFP-neuron located in dentate gyrus (DG) and expressing Prox1, a marker for DG granule neurons.

Supplementary Figure 3 HD-like pathology in corticostriatal R6/2 OTBSs

(a) Representative images showing mHTT aggregate load in R6/2 OTBSs at 2 and 8 weeks of culture. Aggregates were detected using the mHTT-specific antibody EM48. (b) Quantification of mHTT aggregates per 0.8μm diameter surface area between weeks 2 and 8 in R6/2 OTBS striatum (left panel) and cortex (right panel). 2-8 weeks: n = 3 OTBSs. (c) Neurofilament (NF) staining in striatum (st) and cortex (cx) of wt and R6/2 OTBS at 6 weeks. (d) Semi-quantitative results showing greatly reduced neurofilament immunoreactivity in striatum but not cortex of 2, 3 and 6 weeks-old R6/2 OTBSs as compared to wt. (wt 2-6 weeks: n = 30 OTBSs; R6/2 2 weeks: n = 25 OTBSs, 3+6 weeks: n = 30 OTBSs). (e) Representative images showing co-expression of DARPP-32 and GAD65 in a hGFP-neuron located in striatum of a 6-weeks old OTBS. Data is presented as mean ± s.e.m. unpaired Student’s t-test: * p<0.05, ** p<0.01.

Supplementary Figure 4 mHTT aggregates are located inside hGFP neurons and localization gradually changes from cytoplasmatic to predominantly nuclear

(a) Low and high magnification of hGFP-neuron (green) and EM48+ mHTT aggregates before (pink) and after (red) Triton-X100 treatment. (b) Same sequence of images showing hGFP-neuron containing intracellular mHTT aggregates (red). The images below show xyz-cut illustrating intracellular location of mHTT aggregate. This experiment was successfully repeated three times. We did not encounter any issues with repeatability. (c) Representative images showing mHTT localization in hGFP-neurons after 3 and 6 weeks of co-culture in R6/2 OTBSs. (d) Bar diagram showing the percentages of hGFP-neurons in R6/2 OTBSs with cytoplasmic and nuclear mHTT aggregates at different time points. Upper graph: results for hGFP-neurons located in striatum. Lower graph: results for hGFP-neurons located in cortex. (striatum: wt 2 weeks: n = 112 cells, 3 weeks: n = 81 cells, 4 weeks: n = 125 cells, 6 weeks: n = 144 cells; 8 weeks: n = 13 cells; cortex: 2 weeks: n = 162 cells; 3 weeks: n = 137 cells, 4 weeks: n = 75 cells, 6 weeks: n = 116 cells, 8 weeks: n = 64 cells). (e) Scheme summarizing R6/2 OTBS-related non-cell autonomous morphological changes detected in hGFP-neurons with or without mHTT aggregates at 6 weeks of co-culture. hGFP-neuron in wt OTBSs (left), in R6/2 OTBS without mHTT aggregate (middle) and with mHTT aggregate (right). Morphological differences between hGFP-neurons cultured in wt versus R6/2 OTBSs, independent of the presence of mHTT aggregates, are indicated in light grey, whereas changes correlated with the presence of mHTT are shown in dark grey.

Supplementary Figure 5 Transneuronal spreading of mHTT to hGFP-iPSC–derived neurons

(a) hGFP-iPSC-derived neuron expressing human antigen (HA). (b) hGFP-iPSC-derived neuron expressing the pan-neuronal marker Tuj1. These experiments were successfully repeated three times. We did not encounter any issues with repeatability.

Supplementary Figure 6 hGFP neurons acquire mHTT aggregates after transplantation in R6/2 mice

(a,b) Percentages of hGFP-neurons with mHTT aggregates as normalized to the total number of hGFP-neurons at 4 and 8 weeks post grafting (a) and their cytoplasmic versus nuclear localization (b) (4 weeks: n = 426 cells from 12 sections from 4 mice; 8 weeks: n = 329 cells from 12 sections from 3 mice).. Box plots min to max show median (center line), 25-75th data percentiles (boxes) and data range (whiskers). Unpaired Student’s t-test: * p < 0.05.

(c) Representative images illustrate mHTT aggregates (arrow) accumulating in hGFP-neurons at 4 and 8 weeks after stereotactic injection into a R6/2 mouse brain.

Supplementary Figure 7 Propagation of HD-like pathology in mixed-genotype cultures

(a) Left panel: Schematic representation of different mixed-genotype cultures consisting of either wt cx and R6/2 st or R6/2 cx and R6/2 st. Middle panel: Anti-NF staining of the corresponding mixed-genotype cultures. Right panel: images illustrate the presence or absence of EM48+ mHTT aggregates in cx and st. (b) Left panel: Schematic representation of mixed-genotype cultures consisting of either wt cx and wt st, R6/2 cx and wt st, wt cx and R6/2 st. Right panels: representative images of the corresponding mixed-genotype cultures. These experiments were successfully repeated three times. We did not encounter any issues with repeatability.

Supplementary Figure 8 Colocalization of mHTT and clathrin and lack of mHTT in OTBS culture media

(a) Left panel: Images illustrate co-localization (arrows) of mHTT and clathrin; counterstained with DAPI. Right panel: Bar graph shows percentage of clathrin co-localizing with mHTT in hGFP-neurons normalized to total mHTT aggregates (n = 48 regions of interest from 6 OTBSs).(b) Mesoscale electrochemiluminescence detection of mHTT using 2B7 and MW1 antibodies did not reveal any trace of soluble huntingtin in the supernatant of R6/2 OTBS collected at 2, 3, 4, 5 and 6 weeks in culture (n = 3).

Supplementary Figure 9 Retrograde monosynaptic rabies viral tracing system

(a) Representative images show h-neuron expressing G(V5), TVA(GFP) and ΔGRabRFP(EnVA). (b) Images of a Tuj1 positive h-neuron (arrow H) expressing TVA(GFP), which is synaptically connected with a Tuj1 negative mouse neuron (arrow M). The mouse neuron expresses RFP after retrograde transfer of ΔGRabRFP(EnVA) from the infected human neuron. These experiments were successfully repeated three times. We did not encounter any issues with repeatability.

Supplementary Figure 10 BoNT/A blocks transneuronal propagation of Alexa 594–labeled mHTTQ72

(a) Experimental steps performed to assess transneuronal propagation of Alexa594-labeled mHTTQ72 from mouse cells to hGFP-neurons. (b) (left) Representative image illustrating uptake of exogenously supplied Alexa594-labeled mHTTQ72 in wt OTBS. Right: wt OTBS without mHTTQ72. (c) Representative xyz-cut of mHTTQ72 aggregates located either in close proximity to (yellow arrow) or inside (white arrow) a hGFP-neuron. (d) Bar diagram shows the percentage of hGFP-neurons containing mHTTQ72 aggregates normalized to the total number of hGFP-neurons in wt OTBSs infected with mHTTQ72, treated with or without BoNT/A. (w/o: n = 141 cells, with: n = 78 cells). Data is presented as mean ± s.e.m. unpaired Student’s t-test ** p<0.01.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pecho-Vrieseling, E., Rieker, C., Fuchs, S. et al. Transneuronal propagation of mutant huntingtin contributes to non–cell autonomous pathology in neurons. Nat Neurosci 17, 1064–1072 (2014). https://doi.org/10.1038/nn.3761

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.3761

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