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:

Toxicity after AAV delivery of RNAi expression constructs into nonhuman primate brain

Abstract

RNA interference (RNAi) for spinocerebellar ataxia type 1 can prevent and reverse behavioral deficits and neuropathological readouts in mouse models, with safety and benefit lasting over many months. The RNAi trigger, expressed from adeno-associated virus vectors (AAV.miS1), also corrected misregulated microRNAs (miRNA) such as miR150. Subsequently, we showed that the delivery method was scalable, and that AAV.miS1 was safe in short-term pilot nonhuman primate (NHP) studies. To advance the technology to patients, investigational new drug (IND)-enabling studies in NHPs were initiated. After AAV.miS1 delivery to deep cerebellar nuclei, we unexpectedly observed cerebellar toxicity. Both small-RNA-seq and studies using AAVs devoid of miRNAs showed that this was not a result of saturation of the endogenous miRNA processing machinery. RNA-seq together with sequencing of the AAV product showed that, despite limited amounts of cross-packaged material, there was substantial inverted terminal repeat (ITR) promoter activity that correlated with neuropathologies. ITR promoter activity was reduced by altering the miS1 expression context. The surprising contrast between our rodent and NHP findings highlight the need for extended safety studies in multiple species when assessing new therapeutics for human application.

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

Fig. 1: AAV.miS1 delivery to NHP cerebella causes neurological deficits.
Fig. 2: Molecular and histological readouts after AAV.miS1 delivery.
Fig. 3: AAV.miS1 does not cause small RNA dysregulation.
Fig. 4: RNA-seq shows robust immune response in AAV.miS1 dosed animals.
Fig. 5: Transcripts derived from stuffer and/or packaged backbone are sufficient to induce toxicity.

Similar content being viewed by others

Data availability

Sequencing datasets generated as a part of this manuscript can be accessed using NCBI Gene Expression Omnibus (GEO) at accession number: GSE182666. The following individual files in GSE182666are associated with the indicated figure: Fig. 3, GSM5534318, GSM5534319, GSM5534320, GSM5534321, GSM5534322, GSM5534323; Fig. 4, GSM5534296, GSM5534297, GSM5534298, GSM5534299, GSM5534300, GSM5534301, GSM5534302, GSM5534303, GSM5534304, GSM5534305; Fig. 5, GSM5534306, GSM5534307, GSM5534308, GSM5534309, GSM5534310, GSM5534311, GSM5534312, GSM5534313, GSM5534314, GSM5534315, GSM5534316, GSM5534317; Extended Data Fig. 5, GSM5534324, GSM5534325, GSM5534326, GSM5534327, GSM5534328, GSM5534329. Raw data are available as Supplementary Data for graphs shown in Figs. 1c,f, 2i–n and 5c and Extended Data Figs. 2c, 3b–d, 4b–e,g and 6b,d–f. The original uncropped gel is given in the Supplementary Data for Extended Data Fig. 3a. The following public datasets used: Ensemble, Rhesus macaque genome (Mmul10), http://ftp.ensembl.org/pub/release-104/fasta/macaca_mulatta/dna/Macaca_mulatta.Mmul_10.dna.toplevel.fa.gz for Figs. 35 and Extended Data Fig. 5; Ensembl, (GRCh38), http://ftp.ensembl.org/pub/release-104/fasta/homo_sapiens/dna/Homo_sapiens.GRCh38.dna.toplevel.fa.gz for Fig. 4 and Extended Data Figs. 4 and 5. All vectors presented in this work are available on request with approval from the CHOP Office of Technology Transfer. Source data are provided with this paper.

Code availability

The software tools generated as a part of this study are archived at https://github.com/DavidsonLabCHOP/Keiser_NatMed_2021.

References

  1. Gonzalez-Alegre, P. Recent advances in molecular therapies for neurological disease: triplet repeat disorders. Hum. Mol. Genet. 28, R80–R87 (2019).

    Article  CAS  Google Scholar 

  2. Oz, G. et al. In vivo monitoring of recovery from neurodegeneration in conditional transgenic SCA1 mice. Exp. Neurol. 232, 290–298 (2011).

    Article  Google Scholar 

  3. Zu, T. et al. Recovery from polyglutamine-induced neurodegeneration in conditional SCA1 transgenic mice. J. Neurosci. 24, 8853–8861 (2004).

    Article  CAS  Google Scholar 

  4. Matilla, A. et al. Mice lacking ataxin-1 display learning deficits and decreased hippocampal paired-pulse facilitation. J. Neurosci. 18, 5508–5516 (1998).

    Article  CAS  Google Scholar 

  5. Keiser, M. S., Boudreau, R. L. & Davidson, B. L. Broad therapeutic benefit after RNAi expression vector delivery to deep cerebellar nuclei: implications for spinocerebellar ataxia type 1 therapy. Mol. Ther. 22, 588–595 (2014).

    Article  CAS  Google Scholar 

  6. Keiser, M. S., Geoghegan, J. C., Boudreau, R. L., Lennox, K. A. & Davidson, B. L. RNAi or overexpression: alternative therapies for spinocerebellar ataxia type 1. Neurobiol. Dis. 56, 6–13 (2013).

    Article  CAS  Google Scholar 

  7. Keiser, M. S., Monteys, A. M., Corbau, R., Gonzalez-Alegre, P. & Davidson, B. L. RNAi prevents and reverses phenotypes induced by mutant human ataxin-1. Ann. Neurol. 80, 754–765 (2016).

    Article  CAS  Google Scholar 

  8. Xia, H. et al. RNAi suppresses polyglutamine-induced neurodegeneration in a model of spinocerebellar ataxia. Nat. Med. 10, 816–820 (2004).

    Article  CAS  Google Scholar 

  9. Nitschke, L. et al. miR760 regulates ATXN1 levels via interaction with its 5′ untranslated region. Genes Dev. 34, 1147–1160 (2020).

    Article  CAS  Google Scholar 

  10. Friedrich, J. et al. Antisense oligonucleotide-mediated ataxin-1 reduction prolongs survival in SCA1 mice and reveals disease-associated transcriptome profiles.JCI Insight. 3, e123193 (2018).

    Article  Google Scholar 

  11. O’Callaghan, B. et al. Antisense oligonucleotide therapeutic approach for suppression of ataxin-1 expression: a safety assessment. Mol. Ther. Nucleic Acids 21, 1006–1016 (2020).

    Article  Google Scholar 

  12. Keiser, M. S., Kordower, J. H., Gonzalez-Alegre, P. & Davidson, B. L. Broad distribution of ataxin 1 silencing in rhesus cerebella for spinocerebellar ataxia type 1 therapy. Brain 138, 3555–3566 (2015).

    Article  Google Scholar 

  13. Monteys, A. M. et al. Single nucleotide seed modification restores in vivo tolerability of a toxic artificial miRNA sequence in the mouse brain. Nucleic Acids Res. 42, 13315–13327 (2014).

    Article  CAS  Google Scholar 

  14. Rodriguez-Lebron, E., Liu, G., Keiser, M., Behlke, M. A. & Davidson, B. L. Altered Purkinje cell miRNA expression and SCA1 pathogenesis. Neurobiol. Dis. 54, 456–463 (2013).

    Article  CAS  Google Scholar 

  15. Boudreau, R. L., Martins, I. & Davidson, B. L. Artificial microRNAs as siRNA shuttles: improved safety as compared to shRNAs in vitro and in vivo. Mol. Ther. 17, 169–175 (2009).

    Article  CAS  Google Scholar 

  16. van Gestel, M. A. et al. shRNA-induced saturation of the microRNA pathway in the rat brain. Gene Ther. 21, 205–211 (2014).

    Article  Google Scholar 

  17. Radukic, M. T., Brandt, D., Haak, M., Muller, K. M. & Kalinowski, J. Nanopore sequencing of native adeno-associated virus (AAV) single-stranded DNA using a transposase-based rapid protocol. NAR Genom. Bioinform. 2, lqaa074 (2020).

    Article  Google Scholar 

  18. Crespo-Barreto, J., Fryer, J. D., Shaw, C. A., Orr, H. T. & Zoghbi, H. Y. Partial loss of ataxin-1 function contributes to transcriptional dysregulation in spinocerebellar ataxia type 1 pathogenesis. PLoS Genet. 6, e1001021 (2010).

    Article  Google Scholar 

  19. Chadeuf, G., Ciron, C., Moullier, P. & Salvetti, A. Evidence for encapsidation of prokaryotic sequences during recombinant adeno-associated virus production and their in vivo persistence after vector delivery. Mol. Ther. 12, 744–753 (2005).

    Article  CAS  Google Scholar 

  20. Tai, P. W. L. et al. Adeno-associated virus genome population sequencing achieves full vector genome resolution and reveals human-vector chimeras. Mol. Ther. Methods Clin. Dev. 9, 130–141 (2018).

    Article  CAS  Google Scholar 

  21. Burright, E. N. et al. SCA1 transgenic mice: a model for neurodegeneration caused by an expanded CAG trinucleotide repeat. Cell 82, 937–948 (1995).

    Article  CAS  Google Scholar 

  22. Flotte, T. R. et al. Gene expression from adeno-associated virus vectors in airway epithelial cells. Am. J. Respir. Cell Mol. Biol. 7, 349–356 (1992).

    Article  CAS  Google Scholar 

  23. Earley, L. F. et al. Adeno-associated virus serotype-specific inverted terminal repeat sequence role in vector transgene expression. Hum. Gene Ther. 31, 151–162 (2020).

    Article  CAS  Google Scholar 

  24. Ayuso, E. et al. High AAV vector purity results in serotype- and tissue-independent enhancement of transduction efficiency. Gene Ther. 17, 503–510 (2010).

    Article  CAS  Google Scholar 

  25. Maguire, A. M. et al. Age-dependent effects of RPE65 gene therapy for Leber’s congenital amaurosis: a phase 1 dose-escalation trial. Lancet 374, 1597–1605 (2009).

    Article  CAS  Google Scholar 

  26. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  Google Scholar 

  27. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  Google Scholar 

  28. Li, H. Minimap2: pairwise alignment for nucleotide sequences. Bioinformatics 34, 3094–3100 (2018).

    Article  CAS  Google Scholar 

  29. Chen, Y. H., Keiser, M. S. & Davidson, B. L. Adeno-associated virus production, purification, and titering. Curr. Protoc. Mouse Biol. 8, e56 (2018).

    Article  Google Scholar 

  30. Keiser, M. S., Chen, Y. H. & Davidson, B. L. Techniques for intracranial stereotaxic injections of adeno-associated viral vectors in adult mice. Curr. Protoc. Mouse Biol. 8, e57 (2018).

    Article  Google Scholar 

Download references

Acknowledgements

This work was funded by the NIH NS094355 (M.K., P.G.-A., T.J.L., B.L.D.), NIH T32 HG009495 (P.T.R.) and the Children’s Hospital of Philadelphia Research Institute.

Author information

Authors and Affiliations

Authors

Contributions

M.S.K. designed the study, performed experiments, evaluated the data and wrote and edited the paper. P.T.R. designed, performed and analyzed all small RNA and RNA sequencing and wrote and edited the paper. C.M.Y. designed, performed and analyzed the nanopore sequencing and edited the paper. E.M.C. designed constructs, performed rodent work and edited the paper. G.R.S. and A.L.M. assisted with all NHP and rodent work. J.M.S. read and assessed surgical MRIs. Y.H.C. performed antibody analyses and prepared test article to keep study staff blinded. R.L.W. read and assessed the in-life imaging studies. E.R. graded histopathology and microscopic image acquisition. T.J.L. developed the neurosurgical approaches. P.G.-A. designed the study, evaluated data and edited the manuscript. B.L.D. designed the study, evaluated the data and wrote and edited the manuscript.

Corresponding author

Correspondence to Beverly L. Davidson.

Ethics declarations

Competing interests

B.L.D. is a founder of Spark Therapeutics and Spirovant Sciences. She serves an advisory role and/or receives sponsored research support for her laboratory from Roche, NBIR, Homology Medicines, Triplet Therapeutics, Resilience, Intellia Therapeutics, Spirovant Sciences, Panorama Medicines and Voyager Therapeutics. P.G.-A. has consulted for Eisai Therapeutics, Spark Therapeutics and NeuExcell Therapeutics. The remaining authors declare no competing interests.

Additional information

Peer review information Nature Medicine thanks H.T. Orr, L. Naldini and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Jerome Staal was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Immunohistochemistry and lesion scores in AAV.miS1 treated animals.

Representative images from vehicle- (a) and AAV.miS1-injected (b) sagittal cerebellar sections (10 µm thick) immunostained for glial fibrillary acidic protein (GFAP). N = 4 or 6 animals per group. Granule cell layer (GCL) and molecular layers (ML) are identified. Scale bar = 100 µm. c, Tissue lesion scoring parameters and associated heat chart (See Extended Data Table 2). d, Quantitation and statistics of lesion scores. Each dot represents a single animal. Data are represented as mean ± SEM (N = 2, 3, 4 or 7 animals per group), significance was determined by one-way ANOVA followed by Dunnett’s multiple comparison post hoc for each location (** P(3E12vg) = 0.0052; **P(1E13vg) = 0.0079 relative to vehicle). e-f, Representative images (N = 10 independent 20X objective fields) from cerebellar sections immunostained for Calbindin in vehicle- (e; N = 4 animals) and AAV.miS1-injected (f; N = 6 animals) used to quantify Purkinje cell counts throughout the cerebellum. Scale bar = 300 µm.

Extended Data Fig. 2 Non-coding stuffer sequence does not exhibit promoter activity in HEK293 cells.

a, Cartoon map of AAV.miS1 proviral plasmid. Bracketed regions below indicate locations of unexpected RNA-seq reads. Bracketed regions above represent the full or partial stuffer regions tested for activity. b, Construct series designed to test promoter activity of sequences near the 3’ ITR. Proviral plasmid segments were cloned upstream Renilla luciferase coding sequence. c, Renilla luciferase activity in HEK293 cells at 24 hrs post-transfection. Activity was normalized to Firefly luciferase activity encoded within the same plasmid.

Source data

Extended Data Fig. 3 Probe based ddPCR assays.

a, Assays were designed to quantify cross-packaged sequence upstream of the left ITR (5’ backbone (BB)), downstream of the right ITR (3’ BB), or within the cargo (stuffer sequence) of AAV.miS1, to reflect transcripts identified by RNA-seq (See Fig. 4). b, Copies of vector cargo (stuffer sequence) and cross-packaged sequence measured from punches taken from the intermediate lobule of NHPs. c, Cargo, 5‘- and 3’- transcript levels isolated from the superficial medial punches of animals injected with high dose AAV.miS1. d, Mock RT cDNA for the cargo sequence (stuffer for AAV.miS1 and EF1α for AAV.IntmiS1) was quantified in parallel to confirm that expression was not arising from contaminating vector DNA. All data are represented as mean ± SEM (N = 3 animals per group).

Source data

Extended Data Fig. 4 Testing of AAV.IntmiS1.

AAV1.IntmiS1 was injected bilaterally into B05 SCA1 mice cerebella (N = 4 or 6) and tissues collected 3 weeks later. a, Representative agarose gels of cDNAs following PCR using primers flanking the miRNA-containing intron (N = 3 biological replicates). b-d, mRNA levels of miS1 (***P = 0.0003; ****P < 0.0001), (b) hATXN1L (**P = 0.001; ****P < 0.0001), (c) and hATXN1 (**P = 0.0060; ****P, 0.0001), (d) in mice cerebella following AAV.IntmiS1 injection at the indicated doses. All data are represented as mean ± SD (N = 4 and 6 animals per treatment group), with significance determined by one-way ANOVA followed by Dunnett’s multiple comparisons post hoc. e) Relative expression of miS1 as quantified by stem-loop qPCR in WT mice injected bilaterally with AAV.miS1 or AAV.IntmiS1 into the striatum. Data are represented as mean ± SD, (N = 3 animals per group). f) Cartoons of the packaged AAV genome and plasmid sequences for assessing transcript levels by ddPCR. Green arrowheads depict the probe locations for assessing transcripts arising from the 5’ ITR (5’ backbone (BB)), the 3’ ITR (3’ BB), or the EF1α promoter of AAV.IntmiS1. g, ddPCR quantification of RNA isolated from striatum of mice injected with AAV.miS1 or AAV.IntmiS1. Data are represented as mean ± SEM (N = 3 animals per group).

Source data

Extended Data Fig. 5 AAV.IntmiS1 does not cause small RNA dysregulation.

a, Differential expression analysis on small RNASeq data obtained from medial deep cerebellar tissue punches from AAV.IntmiS1 and empty capsid treated NHPs. b,c, Heatmaps of the top 30 most highly detected (b) or 11 specific neuronal miRNAs (c). * denotes 28 day in-life animal.

Extended Data Fig. 6 Study 2 histological and molecular readouts.

a, Quantitation of lesion scores from cerebellar sections of NHPs injected with the indicated AAVs (Empty, miSCA7, Stuffer, IntmiS1). Each dot represents a single animal. b, Human ATXN1L levels normalized to endogenous GAPDH as assessed by RT-qPCR. c, Total read counts of human ATXN1L by RNA-seq. d, miS1 levels in AAV.IntmiS1 treated animals, normalized to endogenous U6 RNA and relative to empty capsid treated animals. e, Ionized calcium binding adapter molecule 1 (IBA1) mRNA levels. f, Glial fibrillary acidic protein (GFAP) mRNA levels. Data are represented as mean ± SEM (N = 3 animals per group). There was no significant difference as measured by two-way ANOVA followed by a Dunnett’s multiple comparisons post hoc.

Source data

Extended Data Table 1 Study designs and demographics
Extended Data Table 2 Lesion scoring parameters
Extended Data Table 3 Study 1 and 2 top 30 most abundant miRNAs
Extended Data Table 4 Top differentially upregulated genes from Study 1 and 2 for GO terms ‘Immune System Process’ and ‘Immune Response’

Supplementary information

Source data

Source Data Fig. 1

Statistical Source Data

Source Data Fig. 2

Statistical Source Data

Source Data Fig. 5

Statistical Source Data

Source Data Extended Data Fig. 2

Statistical Source Data

Source Data Extended Data Fig. 3

Unprocessed gel

Source Data Extended Data Fig. 3

Statistical Source Data

Source Data Extended Data Fig. 4

Statistical Source Data

Source Data Extended Data Fig. 6

Statistical Source Data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Keiser, M.S., Ranum, P.T., Yrigollen, C.M. et al. Toxicity after AAV delivery of RNAi expression constructs into nonhuman primate brain. Nat Med 27, 1982–1989 (2021). https://doi.org/10.1038/s41591-021-01522-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41591-021-01522-3

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