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.

  • Technical Report
  • Published:

Engineered AAVs for efficient noninvasive gene delivery to the central and peripheral nervous systems

Abstract

Adeno-associated viruses (AAVs) are commonly used for in vivo gene transfer. Nevertheless, AAVs that provide efficient transduction across specific organs or cell populations are needed. Here, we describe AAV-PHP.eB and AAV-PHP.S, capsids that efficiently transduce the central and peripheral nervous systems, respectively. In the adult mouse, intravenous administration of 1 × 1011 vector genomes (vg) of AAV-PHP.eB transduced 69% of cortical and 55% of striatal neurons, while 1 × 1012 vg of AAV-PHP.S transduced 82% of dorsal root ganglion neurons, as well as cardiac and enteric neurons. The efficiency of these vectors facilitates robust cotransduction and stochastic, multicolor labeling for individual cell morphology studies. To support such efforts, we provide methods for labeling a tunable fraction of cells without compromising color diversity. Furthermore, when used with cell-type-specific promoters and enhancers, these AAVs enable efficient and targetable genetic modification of cells throughout the nervous system of transgenic and non-transgenic animals.

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: Engineered AAV capsids for efficient transduction across the peripheral and central nervous systems.
Figure 2: AAV-PHP.eB transduces several CNS regions more efficiently than AAV-PHP.B.
Figure 3: AAV-PHP.S efficiently transduces peripheral neurons.
Figure 4: AAV-PHP.eB and S transfer multiple genomes per cell and enable tunable multicolor labeling.
Figure 5: AAV-PHP.eB can be used with gene regulatory elements to achieve cell-type-restricted gene expression throughout the brain.

Similar content being viewed by others

Accession codes

Primary accessions

NCBI Reference Sequence

References

  1. Hastie, E. & Samulski, R.J. Adeno-associated virus at 50: a golden anniversary of discovery, research, and gene therapy success—a personal perspective. Hum. Gene Ther. 26, 257–265 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Suzuki, K. et al. In vivo genome editing via CRISPR/Cas9 mediated homology-independent targeted integration. Nature 540, 144–149 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Yang, Y. et al. A dual AAV system enables the Cas9-mediated correction of a metabolic liver disease in newborn mice. Nat. Biotechnol. 34, 334–338 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Rajasethupathy, P., Ferenczi, E. & Deisseroth, K. Targeting neural circuits. Cell 165, 524–534 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Urban, D.J. & Roth, B.L. DREADDs (designer receptors exclusively activated by designer drugs): chemogenetic tools with therapeutic utility. Annu. Rev. Pharmacol. Toxicol. 55, 399–417 (2015).

    Article  CAS  PubMed  Google Scholar 

  6. Kim, C.K. et al. Simultaneous fast measurement of circuit dynamics at multiple sites across the mammalian brain. Nat. Methods 13, 325–328 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Van der Perren, A., Van den Haute, C. & Baekelandt, V. Viral vector-based models of Parkinson's disease. Curr. Top. Behav. Neurosci. 22, 271–301 (2015).

    Article  CAS  PubMed  Google Scholar 

  8. Hocquemiller, M., Giersch, L., Audrain, M., Parker, S. & Cartier, N. Adeno-associated virus-based gene therapy for CNS diseases. Hum. Gene Ther. 27, 478–496 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Samulski, R.J. & Muzyczka, N. AAV-mediated gene therapy for research and therapeutic purposes. Annu. Rev. Virol. 1, 427–451 (2014).

    Article  PubMed  CAS  Google Scholar 

  10. Bennett, J. et al. Safety and durability of effect of contralateral-eye administration of AAV2 gene therapy in patients with childhood-onset blindness caused by RPE65 mutations: a follow-on phase 1 trial. Lancet 388, 661–672 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Iyer, S.M. et al. Virally mediated optogenetic excitation and inhibition of pain in freely moving nontransgenic mice. Nat. Biotechnol. 32, 274–278 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Samad, O.A. et al. Virus-mediated shRNA knockdown of Nav1.3 in rat dorsal root ganglion attenuates nerve injury-induced neuropathic pain. Mol. Ther. 21, 49–56 (2013).

    Article  CAS  PubMed  Google Scholar 

  13. Chang, R.B., Strochlic, D.E., Williams, E.K., Umans, B.D. & Liberles, S.D. Vagal sensory neuron subtypes that differentially control breathing. Cell 161, 622–633 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Williams, E.K. et al. Sensory neurons that detect stretch and nutrients in the digestive system. Cell 166, 209–221 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Eldridge, M.A. et al. Chemogenetic disconnection of monkey orbitofrontal and rhinal cortex reversibly disrupts reward value. Nat. Neurosci. 19, 37–39 (2016).

    Article  CAS  PubMed  Google Scholar 

  16. Foust, K.D. et al. Intravascular AAV9 preferentially targets neonatal neurons and adult astrocytes. Nat. Biotechnol. 27, 59–65 (2009).

    Article  CAS  PubMed  Google Scholar 

  17. Choudhury, S.R. et al. Widespread central nervous system gene transfer and silencing after systemic delivery of novel AAV-AS vector. Mol. Ther. 24, 726–735 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Pulicherla, N. et al. Engineering liver-detargeted AAV9 vectors for cardiac and musculoskeletal gene transfer. Mol. Ther. 19, 1070–1078 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Marchiò, S., Sidman, R.L., Arap, W. & Pasqualini, R. Brain endothelial cell-targeted gene therapy of neurovascular disorders. EMBO Mol. Med. 8, 592–594 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Deverman, B.E. et al. Cre-dependent selection yields AAV variants for widespread gene transfer to the adult brain. Nat. Biotechnol. 34, 204–209 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Allen, W.E. et al. Global representations of goal-directed behavior in distinct cell types of mouse neocortex. Neuron 94, 891–907 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Hillier, D. et al. Causal evidence for retina-dependent and -independent visual motion computations in mouse cortex. Nat. Neurosci. http://dx.doi.org/10.1038/nn.4566 (2017).

  23. Betley, J.N. & Sternson, S.M. Adeno-associated viral vectors for mapping, monitoring, and manipulating neural circuits. Hum. Gene Ther. 22, 669–677 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Beier, K.T. et al. Circuit architecture of VTA dopamine neurons revealed by systematic input-output mapping. Cell 162, 622–634 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Cai, D., Cohen, K.B., Luo, T., Lichtman, J.W. & Sanes, J.R. Improved tools for the Brainbow toolbox. Nat. Methods 10, 540–547 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Gombash, S.E. et al. Intravenous AAV9 efficiently transduces myenteric neurons in neonate and juvenile mice. Front. Mol. Neurosci. 7, 81 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Inagaki, K. et al. Robust systemic transduction with AAV9 vectors in mice: efficient global cardiac gene transfer superior to that of AAV8. Mol. Ther. 14, 45–53 (2006).

    Article  CAS  PubMed  Google Scholar 

  28. Parekh, R. & Ascoli, G.A. Neuronal morphology goes digital: a research hub for cellular and system neuroscience. Neuron 77, 1017–1038 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Agha-Mohammadi, S. et al. Second-generation tetracycline-regulatable promoter: repositioned tet operator elements optimize transactivator synergy while shorter minimal promoter offers tight basal leakiness. J. Gene Med. 6, 817–828 (2004).

    Article  CAS  PubMed  Google Scholar 

  30. Raymond, C.S. & Soriano, P. High-efficiency FLP and PhiC31 site-specific recombination in mammalian cells. PLoS One 2, e162 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Gallia, G.L. & Khalili, K. Evaluation of an autoregulatory tetracycline regulated system. Oncogene 16, 1879–1884 (1998).

    Article  CAS  PubMed  Google Scholar 

  32. de Leeuw, C.N. et al. rAAV-compatible MiniPromoters for restricted expression in the brain and eye. Mol. Brain 9, 52 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Dimidschstein, J. et al. A viral strategy for targeting and manipulating interneurons across vertebrate species. Nat. Neurosci. 19, 1743–1749 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Fischer, G., Pan, B., Vilceanu, D., Hogan, Q.H. & Yu, H. Sustained relief of neuropathic pain by AAV-targeted expression of CBD3 peptide in rat dorsal root ganglion. Gene Ther. 21, 44–51 (2014).

    Article  CAS  PubMed  Google Scholar 

  35. Mingozzi, F. & High, K.A. Immune responses to AAV vectors: overcoming barriers to successful gene therapy. Blood 122, 23–36 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Tervo, D.G. et al. A designer AAV variant permits efficient retrograde access to projection neurons. Neuron 92, 372–382 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Nathanson, J.L. et al. Short promoters in viral vectors drive selective expression in mammalian inhibitory neurons, but do not restrict activity to specific inhibitory cell-types. Front. Neural Circuits 3, 19 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Jackson, K.L., Dayton, R.D., Deverman, B.E. & Klein, R.L. Better targeting, better efficiency for wide-scale neuronal transduction with the synapsin promoter and AAV-PHP.B. Front. Mol. Neurosci. 9, 116 (2016).

    PubMed  PubMed Central  Google Scholar 

  39. Micheva, K.D., Busse, B., Weiler, N.C., O'Rourke, N. & Smith, S.J. Single-synapse analysis of a diverse synapse population: proteomic imaging methods and markers. Neuron 68, 639–653 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Mikuni, T., Nishiyama, J., Sun, Y., Kamasawa, N. & Yasuda, R. High-throughput, high-resolution mapping of protein localization in mammalian brain by in vivo genome editing. Cell 165, 1803–1817 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Marshel, J.H., Mori, T., Nielsen, K.J. & Callaway, E.M. Targeting single neuronal networks for gene expression and cell labeling in vivo. Neuron 67, 562–574 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Wertz, A. et al. Single-cell-initiated monosynaptic tracing reveals layer-specific cortical network modules. Science 349, 70–74 (2015).

    Article  CAS  PubMed  Google Scholar 

  43. Schwarz, L.A. et al. Viral-genetic tracing of the input-output organization of a central noradrenaline circuit. Nature 524, 88–92 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Loulier, K. et al. Multiplex cell and lineage tracking with combinatorial labels. Neuron 81, 505–520 (2014).

    Article  CAS  PubMed  Google Scholar 

  45. Kitamura, K., Judkewitz, B., Kano, M., Denk, W. & Häusser, M. Targeted patch-clamp recordings and single-cell electroporation of unlabeled neurons in vivo. Nat. Methods 5, 61–67 (2008).

    Article  CAS  PubMed  Google Scholar 

  46. Haas, K., Sin, W.C., Javaherian, A., Li, Z. & Cline, H.T. Single-cell electroporation for gene transfer in vivo. Neuron 29, 583–591 (2001).

    Article  CAS  PubMed  Google Scholar 

  47. Young, P. et al. Single-neuron labeling with inducible Cre-mediated knockout in transgenic mice. Nat. Neurosci. 11, 721–728 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Badea, T.C. et al. New mouse lines for the analysis of neuronal morphology using CreER(T)/loxP-directed sparse labeling. PLoS One 4, e7859 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  49. Lin, M.Z. & Schnitzer, M.J. Genetically encoded indicators of neuronal activity. Nat. Neurosci. 19, 1142–1153 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Sofroniew, N.J., Flickinger, D., King, J. & Svoboda, K. A large field of view two-photon mesoscope with subcellular resolution for in vivo imaging. eLife 5, e14472 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Garcia, A.D., Doan, N.B., Imura, T., Bush, T.G. & Sofroniew, M.V. GFAP-expressing progenitors are the principal source of constitutive neurogenesis in adult mouse forebrain. Nat. Neurosci. 7, 1233–1241 (2004).

    Article  CAS  PubMed  Google Scholar 

  52. Vong, L. et al. Leptin action on GABAergic neurons prevents obesity and reduces inhibitory tone to POMC neurons. Neuron 71, 142–154 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Haddad, R. et al. Olfactory cortical neurons read out a relative time code in the olfactory bulb. Nat. Neurosci. 16, 949–957 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Goedhart, J. et al. Structure-guided evolution of cyan fluorescent proteins towards a quantum yield of 93%. Nat. Commun. 3, 751 (2012).

    Article  PubMed  CAS  Google Scholar 

  55. Shaner, N.C. et al. A bright monomeric green fluorescent protein derived from Branchiostoma lanceolatum. Nat. Methods 10, 407–409 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Lam, A.J. et al. Improving FRET dynamic range with bright green and red fluorescent proteins. Nat. Methods 9, 1005–1012 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Oh, M.S., Hong, S.J., Huh, Y. & Kim, K.S. Expression of transgenes in midbrain dopamine neurons using the tyrosine hydroxylase promoter. Gene Ther. 16, 437–440 (2009).

    Article  CAS  PubMed  Google Scholar 

  58. Gow, A., Friedrich, V.L. Jr. & Lazzarini, R.A. Myelin basic protein gene contains separate enhancers for oligodendrocyte and Schwann cell expression. J. Cell Biol. 119, 605–616 (1992).

    Article  CAS  PubMed  Google Scholar 

  59. Kügler, S., Kilic, E. & Bähr, M. Human synapsin 1 gene promoter confers highly neuron-specific long-term transgene expression from an adenoviral vector in the adult rat brain depending on the transduced area. Gene Ther. 10, 337–347 (2003).

    Article  PubMed  CAS  Google Scholar 

  60. Lee, Y., Messing, A., Su, M. & Brenner, M. GFAP promoter elements required for region-specific and astrocyte-specific expression. Glia 56, 481–493 (2008).

    Article  PubMed  Google Scholar 

  61. Gray, S.J. et al. Production of recombinant adeno-associated viral vectors and use in in vitro and in vivo administration. Curr. Protoc. Neurosci. 4, 4.17, http://dx.doi.org/10.1002/0471142301.ns0417s57 (2011).

    Google Scholar 

  62. Zolotukhin, S. et al. Recombinant adeno-associated virus purification using novel methods improves infectious titer and yield. Gene Ther. 6, 973–985 (1999).

    Article  CAS  PubMed  Google Scholar 

  63. Yang, B. et al. Single-cell phenotyping within transparent intact tissue through whole-body clearing. Cell 158, 945–958 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Hama, H. et al. ScaleS: an optical clearing palette for biological imaging. Nat. Neurosci. 18, 1518–1529 (2015).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank E. Mackey and K. Beadle for assistance with cloning and viral production, P. Anguiano for administrative assistance, and the entire Gradinaru group for discussions. We thank G. Stevens and V. Anand for their efforts in image analysis; P. Rajendran and S. Kalyanam at University of California, Los Angeles and C. Fowlkes at the University of California, Irvine, for discussions, the University of Pennsylvania vector core for the AAV2/9 Rep-Cap plasmid, and M. Brenner at the University of Alabama for the GfABC1D promoter. pAAV-Ef1a-DIO EYFP, pAAV-EF1a-Cre and pAAV-Ef1a-fDIO EYFP were gifts from K. Deisseroth (Addgene plasmids 27056, 55636 and 55641). pEMS2113 and pEMS2115 were gifts from E. Simpson (Addgene plasmids 49138 and 49140). This work was primarily supported by the National Institutes of Health (NIH) through grants to V.G.: Director's New Innovator DP2NS087949 and PECASE; SPARC OT2OD023848-01; National Institute on Aging R01AG047664; BRAIN U01NS090577; and National Institute of Mental Health (NIMH) R21MH103824. Additional funding included the Gordon and Betty Moore Foundation through grant GBMF2809 to the Caltech Programmable Molecular Technology Initiative (to V.G.), the Curci Foundation (to V.G.), the Hereditary Disease Foundation (to V.G. and B.E.D.), the Beckman Institute (to V.G. and B.E.D.) and Rosen Center (to C.L. and V.G.) at Caltech, NIH U01 MH109147 02S1 (to C.L. and V.G.), NIH NS085910 (to S.K.M. and V.G.), the Defense Advanced Research Projects Agency (DARPA) Biological Technologies Office (BTO; to V.G. and B.E.D.) and the Friedreich's Ataxia Research Alliance (FARA) and FARA Australasia (to B.E.D.). S.K.M and V.G. are Heritage Principal Investigators supported by the Heritage Medical Research Institute.

Author information

Authors and Affiliations

Authors

Contributions

K.Y.C. and B.E.D. designed and performed experiments, imaged samples and analyzed data. K.Y.C. prepared figures with input from B.E.D. and V.G. M.J.J. analyzed data, prepared figures and assisted with experiments and in manuscript preparation. B.B.Y. assisted with tissue processing, imaging and virus production. A.G. helped with image analysis. N.R. assisted with molecular cloning. W.-L.W. and L.S.-G. assisted in tissue processing. C.L. and S.K.M. assisted in experimental designs. K.Y.C., B.E.D. and V.G. wrote the manuscript with support from all authors. B.E.D. and V.G. conceived the project. V.G. supervised all aspects of the work.

Corresponding authors

Correspondence to Benjamin E Deverman or Viviana Gradinaru.

Ethics declarations

Competing interests

The California Institute of Technology has filed patent applications related to this work with B.E.D., K.Y.C. and V.G. listed as inventors. B.E.D. and V.G. receive research support from Voyager Therapeutics; this support was not used in preparation of this manuscript or for the studies described therein.

Integrated supplementary information

Supplementary Figure 1 AAV-PHP.eB transduces several brain regions more efficiently than AAV-PHP.B.

(a-c) The cumulative distribution plots show native GFP fluorescence, measured by the median NLS-GFP intensity within individual nuclei in the cortex (a), striatum (b) and within calbindin+ Purkinje cell bodies (c). In (a and b), ROIs were drawn manually around each DAPI+ nuclei. In (c), the ROI detection was automated (see Online Methods and Supplementary Fig. 2). (a-c) The dotted lines represent the 50% intersect, where 50% of the cells have the corresponding median intensity or lower. (c) The solid black vertical line displays the cutoff point used during the automated detection to classify whether a cell had been transduced with ssAAV-CAG-NLS-GFP. Cells with a median intensity below or equal to 0.1 were classified as non-transduced while those above 0.1 were classified as transduced. (a-c) n = 4 individual animals per group. P ≤ 0.0001 for each region, Kolmogorov-Smirnov test.

Supplementary Figure 2 An automated cell detection pipeline used for counting Purkinje cells in the cerebellum.

The pipeline used for automated cell body and nucleus detection. (a) To detect cell bodies from the original image, a morphological filter with a disk-shaped structural element is first used to blur thin fibrous processes (morphological filtering). A circular Hough transform was subsequently applied to the filtered image (circular object detection). Red circles indicate the objects detected, each of which corresponds to a single cell body. (b) In Fig. 2c, we further analyzed the GFP channel to detect the nucleus of each cell. First, the same area at each cell body region was cropped from the GFP image (cropped image of each cell body). The background of each cropped image, estimated by 2D Gaussian smoothing with a standard deviation of 5-fold the cell body size, was subtracted (background subtraction). After thresholding the image to make it binary (binarization), the 2-D convex hull of the object (convex hull) was obtained as the final nuclear region (NLS).

Supplementary Figure 3 Representative immunohistochemistry in the cortex and striatum after transduction with AAV-PHP.B or AAV-PHP.eB.

Single-plane confocal images of native GFP fluorescence (green) after immunohistochemistry (IHC) with DAPI (DNA labeling, blue), NeuN (neuronal marker, magenta) and S100 (glial marker, red) in (a) cortex or (b) striatum. These data are presented in support of Fig. 2. All scale bars are 50 μm. All imaging and display conditions for the GFP channel are matched across panel pairs to allow for comparisons. All images are from tissue extracted at three weeks post intravenous administration of ssAAV-CAG-NLS-GFP at 1 × 1011 vg to adult mice.

Supplementary Figure 4 AAV-PHP.S transduces peripheral organs through intravenous delivery in adult mice.

(a) Representative images of the liver, lungs and cardiac muscle transduced by AAV9 (top) or AAV-PHP.S (bottom). (b) Representative images of several layers of the stomach (longitudinal muscle, myenteric plexus, and circular muscle) transduced by AAV9 (top) or AAV-PHP.S (bottom). The myenteric plexus also contains muscle cells due to the uneven topography of the stomach. (c) Representative images of the submusocal plexus stained with neuronal marker PGP9.5 (red) and astrocyte marker S100b (blue) in the duodenum, jejunum, ileum, and proximal colon. (a-c) All images show native GFP fluorescence taken with confocal microscopy. Images in (a) are 20 μm maximum intensity projections (MIPs) and images in (b and c) are single-plane images. All imaging and display conditions for the GFP channel are matched across panel pairs to allow for comparisons. All scale bars are 50 μm. All images are from tissue extracted three weeks post intravenous administration of ssAAV-CAG-NLS-GFP at 1 × 1012 vg to adult mice. All experiments and imaging for direct comparisons were performed in parallel.

Supplementary Figure 5 The FLPo-FRT two-component system allows sparse gene expression.

(a) Schematic of the two viral vectors used for sparse gene expression with FlpO recombinase encoded in one viral vector and a transgene flanked by FRT sites on a second viral vector. In the basal state, the XFP is expressed in a double inverted orientation (fDIO) and upon the presence of FLPo is inverted into the orientation in line with the CAG promoter for expression. (b) Representative images of native mNeonGreen fluorescence two weeks after intravenous administration of ssAAV-PHP.B:CAG-fDIO-mNeonGreen at 1 × 1012 vg/mouse together with the indicated dose of ssAAV-PHP.B:EF1α-FLPo. (c) Magnified images of the cortex, striatum, and cerebellum. Scale bars are 1 mm (b) and 50 μm (c).

Supplementary Figure 6 The use of the tetracycline-inducible system with viral vectors provides tunable dense to sparse gene expression.

(a) A two-component vector system consisting of one virus, ssAAV-PHP.B:TRE-mNeonGreen, delivered at 1 × 1012 vg into adult mice, and a second virus, ssAAV-PHP.B:CAG-tTA, the dose of which was varied across mice. As a control, ssAAV-PHP.B:TRE-mNeonGreen was delivered alone to observe any basal expression from the ssAAV-PHP.B:TRE-mNeonGreen. All experimental, imaging and display conditions are matched across images. (b) Representative 20 μm MIPs of native mNeonGreen fluorescence from the cortex, striatum, and cerebellum at high and low activation doses. (c and d) Measurements of the shortest cell-to-cell distance (c) and cell density (d) in the cortex, striatum and cerebellum. n = 3 sections per group, mean ± SEM,unpaired t-test (***p ≤ 0.001; **p ≤ 0.01; *p ≤ 0.05; n.s., p ≥ 0.05). (e) A MIP image of a coronal section taken from the olfactory bulb of a Tbx21-Cre mouse transduced with vector-assisted spectral tracing (VAST) transgenes. The VAST cocktail consist of ssAAV-PHP.eB:TRE-DIO-3XFP (1 × 1012 vg total dose) and sAAV-PHP.eB:ihSyn1-tTA (1 × 1010 vg – high dose) or (1 × 108 vg – low dose). In (e, right) the inset box in gray represents a high magnification of the region of interest outlined. Scale bars are 1 mm (a) and 50 μm (b and e). All images are native GFP fluorescence taken with confocal microscopy.

Supplementary Figure 7 A positive feedback loop within the tTA inducer vector enhances gene expression.

Schematic of the two-component viral vector system without (a) and with (b) an inducible positive feedback loop with the viral vector containing the tTA. The schematic shows tTA protein (magenta hexagons) binding onto TRE elements (blue lines) to activate expression of (a) mNeonGreen or (b) mNeonGreen and the tTA containing vector. (c) Representative confocal images of sagittal brain sections transduced by the no-feedback vector and (d) positive-feedback tTA vector at 1 × 109 vg. Both tTA inducers were co-administered with 1 × 1012 vg of the mNeonGreen viral vector. Scale bar in (c) is 1 mm.

Supplementary Figure 8 Cell-type-restricted expression and sparse labeling in the enteric nervous system using AAV-PHP.S together with neuron-specific gene regulatory elements or Cre transgenic mice.

Images show XFP expression within the myenteric plexus. (a) ssAAV-PHP.S:hSyn1-3XFP was administered intravenously at a total dose of 1 × 1012 vg. (b) ssAAV-PHP.S:CAG-DIO-3XFP was administered intravenously at a total dose of 1 × 1012 vg to ChAT-IRES-Cre transgenic mice to achieve cell type-specific expression in cholinergic neurons. (c) Sparse labeling with the VAST system in the proximal colon of a ChAT-IRE-Cre transgenic using ssAAV-PHP.S-TET-DIO-3XFP at 1 × 1012 vg and ssAAV-PHP.S-ihSyn-DIO-tTA at 1 × 1010 vg. Color-coded square boxes display neurons traced manually in Imaris v8.3. (d) Rectangular boxes show higher magnification views highlighting how sparse multicolor labeling facilitates tracing of closely aligned processes. Double arrows mark traced processes. Scale bars in (a and d) are set to 50 μm and in (c) to 200 μm.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8 and Supplementary Tables 1–3 (PDF 3888 kb)

Supplementary Methods Checklist (PDF 347 kb)

Supplementary Software

Supplementary Software (ZIP 776 kb)

A video demonstrating the transduction of AAV-PHP.S across multiple layers of the small intestine after optical clearing with refractive index matching solution (RIMS).

Native GFP fluorescence (green) from ssAAV-PHP.S:CAG-NLS-GFP at 1 × 1012 vg/mouse is shown. S100b staining (magenta) is shown for contrast. (MOV 9284 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chan, K., Jang, M., Yoo, B. et al. Engineered AAVs for efficient noninvasive gene delivery to the central and peripheral nervous systems. Nat Neurosci 20, 1172–1179 (2017). https://doi.org/10.1038/nn.4593

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4593

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