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:

Tbr1 instructs laminar patterning of retinal ganglion cell dendrites

This article has been updated

Abstract

Visual information is delivered to the brain by >40 types of retinal ganglion cells (RGCs). Diversity in this representation arises within the inner plexiform layer (IPL), where dendrites of each RGC type are restricted to specific sublaminae, limiting the interneuronal types that can innervate them. How such dendritic restriction arises is unclear. We show that the transcription factor Tbr1 is expressed by four mouse RGC types with dendrites in the outer IPL and is required for their laminar specification. Loss of Tbr1 results in elaboration of dendrites within the inner IPL, while misexpression in other cells retargets their neurites to the outer IPL. Two transmembrane molecules, Sorcs3 and Cdh8, act as effectors of the Tbr1-controlled lamination program. However, they are expressed in just one Tbr1+ RGC type, supporting a model in which a single transcription factor implements similar laminar choices in distinct cell types by recruiting partially non-overlapping effectors.

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: Expression of Tbr1 in four types of OFF-laminating RGCs.
Fig. 2: Tbr1 deletion results in ectopic dendritic lamination.
Fig. 3: Physiological defects in Tbr1 mutants.
Fig. 4: Ectopic expression of Tbr1 retargets neurites to the outer IPL.
Fig. 5: Cdh8 and Sorcs3 are Tbr1 targets in J-RGCs.
Fig. 6: Requirement of Cdh8 and Sorcs3 for laminar restriction in J-RGCs.
Fig. 7: Ectopic expression of Cdh8 or Sorcs3 retargets RGC dendrites to the outer IPL.
Fig. 8: Regulation of dendrite laminar identity by Tbr1.

Similar content being viewed by others

Change history

  • 15 August 2018

    In the version of this article initially published online, the links for GSE35077 and GSE71384 led to the wrong website. The error has been corrected in the print, PDF and HTML versions of this article.

References

  1. Sanes, J. R. & Yamagata, M. Formation of lamina-specific synaptic connections. Curr. Opin. Neurobiol. 9, 79–87 (1999).

    Article  PubMed  CAS  Google Scholar 

  2. Baier, H. Synaptic laminae in the visual system: molecular mechanisms forming layers of perception. Annu. Rev. Cell Dev. Biol. 29, 385–416 (2013).

    Article  PubMed  CAS  Google Scholar 

  3. Sanes, J. R. & Masland, R. H. The types of retinal ganglion cells: current status and implications for neuronal classification. Annu. Rev. Neurosci. 38, 221–246 (2015).

    Article  PubMed  CAS  Google Scholar 

  4. Famiglietti, E. V. Jr. & Kolb, H. Structural basis for ON-and OFF-center responses in retinal ganglion cells. Science 194, 193–195 (1976).

    Article  PubMed  Google Scholar 

  5. Roska, B. & Werblin, F. Vertical interactions across ten parallel, stacked representations in the mammalian retina. Nature 410, 583–587 (2001).

    Article  PubMed  CAS  Google Scholar 

  6. Krishnaswamy, A., Yamagata, M., Duan, X., Hong, Y. K. & Sanes, J. R. Sidekick 2 directs formation of a retinal circuit that detects differential motion. Nature 524, 466–470 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Yamagata, M. & Sanes, J. R. Dscam and Sidekick proteins direct lamina-specific synaptic connections in vertebrate retina. Nature 451, 465–469 (2008).

    Article  PubMed  CAS  Google Scholar 

  8. Yamagata, M. & Sanes, J. R. Expanding the Ig superfamily code for laminar specificity in retina: expression and role of contactins. J. Neurosci. 32, 14402–14414 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Peng, Y. R. et al. Satb1 regulates Contactin 5 to pattern dendrites of a mammalian retinal ganglion cell. Neuron 95, 869–883 (2017). e6.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  10. Duan, X., Krishnaswamy, A., De la Huerta, I. & Sanes, J. R. Type II cadherins guide assembly of a direction-selective retinal circuit. Cell 158, 793–807 (2014).

    Article  PubMed  CAS  Google Scholar 

  11. Matsuoka, R. L. et al. Transmembrane semaphorin signalling controls laminar stratification in the mammalian retina. Nature 470, 259–263 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Sun, L. O. et al. On and off retinal circuit assembly by divergent molecular mechanisms. Science 342, 1241974 (2013).

    Article  PubMed  CAS  Google Scholar 

  13. Kim, I. J., Zhang, Y., Yamagata, M., Meister, M. & Sanes, J. R. Molecular identification of a retinal cell type that responds to upward motion. Nature 452, 478–482 (2008).

    Article  PubMed  CAS  Google Scholar 

  14. Rousso, D. L. et al. Two pairs of ON and OFF retinal ganglion cells are defined by intersectional patterns of transcription factor expression. Cell Rep. 15, 1930–1944 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Sweeney, N. T., James, K. N., Nistorica, A., Lorig-Roach, R. M. & Feldheim, D. A. Expression of transcription factors divides retinal ganglion cells into distinct classes. J. Comp. Neurol. 520, 633–655 (2017).

    Google Scholar 

  16. Mao, C. A. et al. T-box transcription regulator Tbr2 is essential for the formation and maintenance of Opn4/melanopsin-expressing intrinsically photosensitive retinal ganglion cells. J. Neurosci. 34, 13083–13095 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Sweeney, N. T., Tierney, H. & Feldheim, D. A. Tbr2 is required to generate a neural circuit mediating the pupillary light reflex. J. Neurosci. 34, 5447–5453 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Kay, J. N., Chu, M. W. & Sanes, J. R. MEGF10 and MEGF11 mediate homotypic interactions required for mosaic spacing of retinal neurons. Nature 483, 465–469 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  19. Rockhill, R. L., Euler, T. & Masland, R. H. Spatial order within but not between types of retinal neurons. Proc. Natl Acad. Sci. USA 97, 2303–2307 (2000).

    Article  PubMed  CAS  Google Scholar 

  20. Rodieck, R. W. The density recovery profile: a method for the analysis of points in the plane applicable to retinal studies. Vis. Neurosci. 6, 95–111 (1991).

    Article  PubMed  CAS  Google Scholar 

  21. Trenholm, S., Johnson, K., Li, X., Smith, R. G. & Awatramani, G. B. Parallel mechanisms encode direction in the retina. Neuron 71, 683–694 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Kim, I. J., Zhang, Y., Meister, M. & Sanes, J. R. Laminar restriction of retinal ganglion cell dendrites and axons: subtype-specific developmental patterns revealed with transgenic markers. J. Neurosci. 30, 1452–1462 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Krieger, B., Qiao, M., Rousso, D. L., Sanes, J. R. & Meister, M. Four alpha ganglion cell types in mouse retina: Function, structure, and molecular signatures. PLoS One 12, e0180091 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Pang, J. J., Gao, F. & Wu, S. M. Light-evoked excitatory and inhibitory synaptic inputs to ON and OFF alpha ganglion cells in the mouse retina. J. Neurosci. 23, 6063–6073 (2003).

    Article  PubMed  CAS  Google Scholar 

  25. Feng, G. et al. Imaging neuronal subsets in transgenic mice expressing multiple spectral variants of GFP. Neuron 28, 41–51 (2000).

    Article  PubMed  CAS  Google Scholar 

  26. Bleckert, A., Schwartz, G. W., Turner, M. H., Rieke, F. & Wong, R. O. Visual space is represented by nonmatching topographies of distinct mouse retinal ganglion cell types. Curr. Biol. 24, 310–315 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. Bassett, E. A. & Wallace, V. A. Cell fate determination in the vertebrate retina. Trends Neurosci. 35, 565–573 (2012).

    Article  PubMed  CAS  Google Scholar 

  28. Liu, J. & Sanes, J. R. Cellular and molecular analysis of dendritic morphogenesis in a retinal cell type that senses color contrast and ventral motion. J. Neurosci. 37, 12247–12262 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  29. Bulfone, A. et al. An olfactory sensory map develops in the absence of normal projection neurons or GABAergic interneurons. Neuron 21, 1273–1282 (1998).

    Article  PubMed  CAS  Google Scholar 

  30. Hong, Y. K., Kim, I. J. & Sanes, J. R. Stereotyped axonal arbors of retinal ganglion cell subsets in the mouse superior colliculus. J. Comp. Neurol. 519, 1691–1711 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Kay, J. N. et al. Retinal ganglion cells with distinct directional preferences differ in molecular identity, structure, and central projections. J. Neurosci. 31, 7753–7762 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Notwell, J. H. et al. TBR1 regulates autism risk genes in the developing neocortex. Genome Res. 26, 1013–1022 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Willnow, T. E., Petersen, C. M. & Nykjaer, A. VPS10P-domain receptors - regulators of neuronal viability and function. Nat. Rev. Neurosci. 9, 899–909 (2008).

    Article  PubMed  CAS  Google Scholar 

  34. Takai, Y., Ikeda, W., Ogita, H. & Rikitake, Y. The immunoglobulin-like cell adhesion molecule nectin and its associated protein afadin. Annu. Rev. Cell Dev. Biol. 24, 309–342 (2008).

    Article  PubMed  CAS  Google Scholar 

  35. Beaudoin, G. M. III et al. Afadin, a Ras/Rap effector that controls cadherin function, promotes spine and excitatory synapse density in the hippocampus. J. Neurosci. 32, 99–110 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Mihalas, A. B. & Hevner, R. F. Control of neuronal development by T-box genes in the brain. Curr. Top. Dev. Biol. 122, 279–312 (2017).

    Article  PubMed  CAS  Google Scholar 

  37. Hevner, R. F. et al. Tbr1 regulates differentiation of the preplate and layer 6. Neuron 29, 353–366 (2001).

    Article  PubMed  CAS  Google Scholar 

  38. Han, W. et al. TBR1 directly represses Fezf2 to control the laminar origin and development of the corticospinal tract Proc. Natl Acad. Sci. USA 108, 3041–3046 (2011).

    Article  PubMed  Google Scholar 

  39. McKenna, W. L. et al. Tbr1 and Fezf2 regulate alternate corticofugal neuronal identities during neocortical development. J. Neurosci. 31, 549–564 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Huang, T. N. et al. Tbr1 haploinsufficiency impairs amygdalar axonal projections and results in cognitive abnormality. Nat. Neurosci. 17, 240–247 (2014).

    Article  PubMed  CAS  Google Scholar 

  41. Englund, C. et al. Pax6, Tbr2, and Tbr1 are expressed sequentially by radial glia, intermediate progenitor cells, and postmitotic neurons in developing neocortex. J. Neurosci. 25, 247–251 (2005).

    Article  PubMed  CAS  Google Scholar 

  42. Shimoyama, Y., Tsujimoto, G., Kitajima, M. & Natori, M. Identification of three human type-II classic cadherins and frequent heterophilic interactions between different subclasses of type-II classic cadherins. Biochem. J. 349, 159–167 (2000).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Hermey, G. et al. The three sorCS genes are differentially expressed and regulated by synaptic activity. J. Neurochem. 88, 1470–1476 (2004).

    Article  PubMed  CAS  Google Scholar 

  44. Oetjen, S., Mahlke, C., Hermans-Borgmeyer, I. & Hermey, G. Spatiotemporal expression analysis of the growth factor receptor SorCS3. J. Comp. Neurol. 522, 3386–3402 (2014).

    Article  PubMed  CAS  Google Scholar 

  45. Breiderhoff, T. et al. Sortilin-related receptor SORCS3 is a postsynaptic modulator of synaptic depression and fear extinction. PLoS One 8, e75006 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Christiansen, G. B. et al. The sorting receptor SorCS3 is a stronger regulator of glutamate receptor functions compared to GABAergic mechanisms in the hippocampus. Hippocampus 27, 235–248 (2017).

    Article  PubMed  CAS  Google Scholar 

  47. Hobert, O. Terminal selectors of neuronal identity. Curr. Top. Dev. Biol. 116, 455–475 (2016).

    Article  PubMed  Google Scholar 

  48. Huang, T. N. & Hsueh, Y. P. Brain-specific transcriptional regulator T-brain-1 controls brain wiring and neuronal activity in autism spectrum disorders. Front. Neurosci. 9, 406 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  49. De Rubeis, S. et al. Synaptic, transcriptional and chromatin genes disrupted in autism. Nature 515, 209–215 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Rodríguez, C. I. et al. High-efficiency deleter mice show that FLPe is an alternative to Cre-loxP. Nat. Genet. 25, 139–140 (2000).

    Article  PubMed  Google Scholar 

  51. Buffelli, M. et al. Genetic evidence that relative synaptic efficacy biases the outcome of synaptic competition. Nature 424, 430–434 (2003).

    Article  PubMed  CAS  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  PubMed  PubMed Central  CAS  Google Scholar 

  53. Madisen, L. et al. Transgenic mice for intersectional targeting of neural sensors and effectors with high specificity and performance. Neuron 85, 942–958 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  54. Wichterle, H., Lieberam, I., Porter, J. A. & Jessell, T. M. Directed differentiation of embryonic stem cells into motor neurons. Cell 110, 385–397 (2002).

    Article  PubMed  CAS  Google Scholar 

  55. Huberman, A. D. et al. Genetic identification of an On-Off direction-selective retinal ganglion cell subtype reveals a layer-specific subcortical map of posterior motion. Neuron 62, 327–334 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  56. Furuta, Y., Lagutin, O., Hogan, B. L. & Oliver, G. C. Retina- and ventral forebrain-specific Cre recombinase activity in transgenic mice. Genesis 26, 130–132 (2000).

    Article  PubMed  CAS  Google Scholar 

  57. Suzuki, S. C. et al. Cadherin-8 is required for the first relay synapses to receive functional inputs from primary sensory afferents for cold sensation. J. Neurosci. 27, 3466–3476 (2007).

    Article  PubMed  CAS  Google Scholar 

  58. 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  PubMed Central  CAS  PubMed  Google Scholar 

  59. Buhusi, M. et al. ALCAM regulates mediolateral retinotopic mapping in the superior colliculus. J. Neurosci. 29, 15630–15641 (2009).

    Article  PubMed  CAS  Google Scholar 

  60. Matsuda, T. & Cepko, C. L. Controlled expression of transgenes introduced by in vivo electroporation. Proc. Natl Acad. Sci. USA 104, 1027–1032 (2007).

    Article  PubMed  CAS  Google Scholar 

  61. Dhande, O. S. & Crair, M. C. Transfection of mouse retinal ganglion cells by in vivo electroporation. J. Vis. Exp. 50, 2678 (2011).

    Google Scholar 

  62. Trapnell, C. et al. Differential gene and transcript expression analysis of RNA-seq experiments with TopHat and Cufflinks. Nat. Protoc. 7, 562–578 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  63. Robinson, M. D., McCarthy, D. J. & Smyth, G. K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26, 139–140 (2010).

    Article  PubMed  CAS  Google Scholar 

  64. Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  65. McLean, C. Y. et al. GREAT improves functional interpretation of cis-regulatory regions. Nat. Biotechnol. 28, 495–501 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  66. Picelli, S. et al. Full-length RNA-seq from single cells using Smart-seq2. Nat. Protoc. 9, 171–181 (2014).

    Article  PubMed  CAS  Google Scholar 

  67. Adachi, H. et al. Stage-specific reference genes significant for quantitative PCR during mouse retinal development. Genes Cells 20, 625–635 (2015).

    Article  PubMed  CAS  Google Scholar 

  68. Duan, X. et al. Subtype-specific regeneration of retinal ganglion cells following axotomy: effects of osteopontin and mTOR signaling. Neuron 85, 1244–1256 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  69. Choi, J. H. et al. Optimization of AAV expression cassettes to improve packaging capacity and transgene expression in neurons. Mol. Brain 7, 17 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  70. Euler, T. et al. Eyecup scope–optical recordings of light stimulus-evoked fluorescence signals in the retina. Pflugers Arch. 457, 1393–1414 (2009).

    Article  PubMed  CAS  Google Scholar 

  71. Baden, T. et al. The functional diversity of retinal ganglion cells in the mouse. Nature 529, 345–350 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

We thank Z. He and C. Wang at Boston Children’s Hospital Viral Core for their generous support (5P30EY012196). This work was supported by NIH grants NS29169 and EY022073 to J.R.S., NS34661 to J.L.R., and MH094589 to B.C. J.L. was funded by an Agency for Science, Technology and Research (A*STAR) fellowship from Singapore.

Author information

Authors and Affiliations

Authors

Contributions

J.L. and J.R.S. conceived the study, planned experiments, analyzed data, and wrote the paper. J.L. performed all experiments unless otherwise stated. J.D.S.R. performed and analyzed calcium imaging experiments. A.K. built instrumentation, wrote stimuli, and analyzed calcium imaging experiments. M.A.L. performed experiments on afadin, Cdh8, and axonal projections. S.P. generated cDNA libraries for Tbr1 wild-type and mutant J-RGCs. J.L.R.R. and B.C. generated conditional Tbr1 mutant mice.

Corresponding author

Correspondence to Joshua R. Sanes.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Integrated supplementary information

Supplementary Figure 1 Morphological characterization of Tbr1-RGCs.

(a) Whole mounts of retina stained for Tbr1 plus Tbr2, FoxP2, FoxP1 or Satb1. (b) Effective radius and cell density of each Tbr1-RGC type. n=4 fields per retina, 3 retinas per type, each retina from a different animal. Each point indicates mean + standard error for both x and y axes. Scale = 25µm. (c-f) Distribution of somata of each Tbr1-RGC type. (g) Proportions of each Tbr1-RGC type as a fraction of all RGCs (RBPMS-positive cells) in a retina. (h) Whole mount of a retina stained with Tbr1, Brn3c, calretinin, Brn3b and Opn. All Tbr1+ cells express one of these four markers. Open triangles, closed triangles and arrows indicate Tbr1+ cells expressing Brn3c/calretinin, Brn3b or Opn respectively. Scale=25µm. (i-l) Whole mounts of each Tbr1-RGC type and the markers they express. Scale=25µm. (m, n) Whole mount showing that YFP-positive RGCs in the TYW3 and Hb9-GFP transgenic lines are Tbr1-negative. Scale=25µm. (o) Coverage factor, defined as product of mean soma density (from b) and mean dendritic field area (from p). Bars and brackets indicate mean + standard error. (p) Dendritic field area of each Tbr1-RGC type. Top and bottom whiskers represent 10th and 90th percentiles. Upper and lower box limits represent 25th and 75th percentiles. Center line marks median. n=19, 87, 19, 36 cells for J-RGC, α-OFF-s-RGC, Tbr1-S1-RGC and Tbr1-S2-RGC respectively. Each experiment in a, c-f and h-n was repeated independently in three animals with similar results.

Supplementary Figure 2 Embryonic expression of Tbr1.

(a) Retinal cross-sections at indicated embryonic ages stained for Tbr1, Brn3b/c. Tbr1 (yellow arrowheads) appears in RGCs by E17.5. (b-c) Retinal cross-sections at E17.5, showing Tbr1, Tbr2 and FoxP2 protein in non-overlapping populations. Scale=20µm. Each experiment in a-c was repeated independently in three animals with similar results.

Supplementary Figure 3 Generation and analysis of a conditional Tbr1 mutant.

(a) Tbr1 gene and targeting vector used to generate the Tbr1loxP allele. The wildtype Tbr1 allele has six known exons (numbered black boxes, 1–6); the initiation codon is in exon 1, and the termination codon is in exon 6. White boxes indicate the 5’ and 3’ UTR. Red arrowheads correspond to the location of LoxP sites; black box with ‘F’ inside are Frt sites. Germline mutants were mated to a Flp-expressing transgenic to remove the Neomycin expression cassette (grey box with Neo inside). Upon Cre recombination, exons 2 and 3 are deleted to generate Tbr1 mutant allele. (b) Cross-section showing normal retinal architecture in the absence of Tbr1. RBPMS marks all RGCs; Syt2 marks primarily type II bipolar cells, VGlut3 marks a S3-laminating amacrine cell type, Calbindin (Calb) marks horizontal cells and subsets of amacrine and ganglion cells. Scale=25μm. Experiment was repeated independently in three animals with similar results.

Supplementary Figure 4 Morphological and molecular features unaffected by loss of Tbr1.

(a-b) Dendritic area and length in control and Tbr1J/J J-RGCs. n=10 cells from 3 retinas per genotype, each retina from a different animal. Bars represent mean ± standard error. Circles represent individual cells. (c) Proportions of ventrally asymmetric J-RGCs in control versus Tbr1J/J retinas. n=132–710 control cells and 77–299 Tbr1J/J cells, 4 retinas per genotype, each retina from a different animal. Bars represent mean+ standard error, circles represent individual retina. p=0.62, 0.40 and 0.39 for area, length and proportion of ventrally asymmetric J-RGCs, two-tailed Student’s t-test, ns=not significant. (d) Axonal projections of Tbr1+/J and Tbr1J/J J-RGCs into the superior colliculus. Scale=100µm. (e) Cross-sections of Tbr1J/J J-RGCs (arrowheads) showing no loss of markers normally expressed by J-RGCs (in yellow box) and no gain in markers for other RGC types. Each box outlined in black encloses corresponding images. Pcsk2 was identified from the J-RGC RNAseq data. All J-RGCs shown here were labeled by tamoxifen at P0, except cells in the first row, which were labeled by tamoxifen at E14.5. Scale = 25µm. (f-g) Dendritic area and length of α-OFF-s-RGCs in control and Tbr1ret/ret retinas. n=15 and 13 cells from 6 control and 9 Tbr1ret/ret retinas respectively. Two-tailed Student’s t-test, p=0.75 and 0.18 for area and length respectively. Bars represent mean ± standard error. Circles represent individual cells. Each experiment in d-e was repeated independently in three animals with similar results.

Supplementary Figure 5 Effects of early versus late deletion of Tbr1 on J-RGC dendrites.

(a) Experimental timeline for early (magenta) and late (red) ablation of Tbr1. Green arrowhead marks when J-RGC dendrites become restricted to S128. (b) P4 retinal cross-sections from controls and Tbr1J/J retinas following deletion of Tbr1 at P0. Scale=25µm. (c) Distribution of fluorescence intensities of J-RGC dendrites from images such as those in b. Gray arrowheads indicate S2 and S4 as labeled by VAchT immunostaining. p=0.00018, Cochran-Armitage test. n = 11 and 18 sections from control and Tbr1J/J animals respectively, 3 animals per genotype. (d) P20 retinal cross-sections from control and Tbr1J/J retinas where Tbr1 is ablated after laminar restriction. Scale=25µm. (e) Distribution of fluorescence intensities of J-RGC dendrites from images such as those in d. p=0.096, Cochran-Armitage test. n = 8 and 14 sections from 3 control and 4 Tbr1J/J animals respectively.

Supplementary Figure 6 Physiology of Tbr1-RGCs in control and Tbr1ret/ret.

(a-d) Averaged responses (in black) to 2-second full-field flashes (in red) for 22, 18, 13 and 15 J-RGCs, α-OFF-s-RGCs, Tbr1-S1-RGCs and Tbr1-S2-RGCs respectively, as identified by Tbr1 immunostaining. x and y axes are time in seconds and Δf/f respectively. J-RGCs give ON-OFF responses as shown previously in Figure 2d from Kim et al. 200813; it appears to be due to a strong surround. Tbr1-S1-RGCs also showed ON-OFF responses to full field flashes. Tbr1-S2-RGCs were poorly responsive with the stimulus parameters used. α-OFF-s-RGCs gave strong OFF responses (data re-plotted from Figure 3c for comparison). (e-g) Polar plots of maximum calcium responses evoked by bars moving in 8 directions. Outermost gray circle is Δf/f=1. e, Tbr1+ Opn+ control α-OFF-s-RGCs. f, Tbr2- Brn3c- Opn+ control α-OFF-s-RGCs. g, Tbr2- Brn3c- Opn+ Tbr1ret/ret α-OFF-s-RGCs. (h-i) Distribution of z-score (h) and quality index (i) for all RGCs recorded from the retinas shown in Figures 3b-m (n= 1083 RGCs from 3 control and 3030 RGCs from 6 mutant retinas). These statistics were calculated as described in Methods. Responsiveness of control and mutant RGCs were similar by both measures.

Supplementary Figure 7 Ectopic expression of Tbr1 in the retina.

(a) En face view of Tbr1-misexpressing RGCs showing a variety of dendritic arborization patterns, contrasting with the similarity in laminar restriction shown in Figure 4. Arrowheads mark Tbr1-overexpressing RGCs. Note that XFP-positive cells include Brn3bhigh, Brn3blow and Brn3b-negative types. Scale=50μm. Experiment was repeated independently in two animals with similar results. (b-d) Tbr1 mis-expressing cells in the inner nuclear layer showing robust overexpression of Tbr1. These cells are amacrine cells (Rbpms-negative, AP2-positive). Scale=25μm. Experiment was repeated independently in three animals with similar results.

Supplementary Figure 8 Identification and validation of Tbr1 target genes.

(a) Integrative Genomics Viewer plots showing alignment of Tbr1 ChIP-seq peaks32 to the Cdh8 (top panel) and Sorcs3 (bottom panel) genomic loci. For each gene, the top row represents viewing window in kilobases and the gene locus. Bottom two blue traces are peaks from n = 2 Tbr1 ChIP-seq replicates. Peaks associated with Cdh8 and Sorcs3, but not Jam2, Alcam or Smo, were within 50kb upstream of the annotated transcriptional start site, and may therefore be associated with regulatory regions. (b-c) Retinal cross-sections from control and Tbr1ret/ret stained for Alcam and Neo1, showing no difference. Scale=25µm. (d) Expression of lacZ within the ganglion cell layer (GCL) of Cdh8lacZ retinas from P3 to P7. Scale=50µm. (e) Retinal cross-section stained for Sorcs3 and Protein kinase C (PKC), a rod bipolar cell marker. Yellow arrows mark co-localization of Sorcs3 with dendrites of rod bipolar cells. Scale=25µm. (f) Expression of Sorcs3 in retinal cross-sections from P4 to P16. DR, animals dark-reared from birth. Upregulation of Sorcs3 in RGCs after eye-opening appears to be light-dependent. Scale=25µm. (g) Whole mounts from a Cdh8lacZ; TYW7 retina. RGCs labeled in this line, including α-OFF-s-RGCs, do not express lacZ. Scale =25µm. Each experiment in b-g was repeated independently in three animals with similar results.

Supplementary Figure 9 Transcriptomic analysis of wild-type versus Tbr1-mutant J-RGCs.

(a) Average CPM of genes encoding cell-surface molecules that are significantly different between wildtype and mutant J-RGCs (two-tailed Fisher’s Exact Test, p < 0.001). The larger the -log10(p-value), the greater the statistical significance. n= 5 samples from 4 Tbr1+/+ animals and 6 samples from 4 Tbr1J/J animals. (b) Average CPM of markers normally expressed by J-RGCs (highlighted green) and of markers expressed by other specific RGC types (highlighted magenta). Tbr1 mutant J-RGCs retained expression of markers normally expressed by wildtype J-RGCs and did not gain expression of markers for other RGC types. Bars and brackets in a-b indicate average CPM and standard error. (c) Heatmap showing relative microarray values across different groups of RGCs for the differentially expressed genes in a. Most of these genes are present in other non-Tbr1 RGC types, except for Jam2, Sorcs3 and Cdh8 (yellow box).

Supplementary Figure 10 Cdh8 and Sorcs3 loss-of-function mutations.

(a) Cross-sections of OFF alpha RGCs labeled by TYW7; Cdh8+/lacZ and TYW7; Cdh8lacZ/lacZ retinas. (b) Distribution of GFP intensities from OFF alpha RGC dendrites along IPL depth. n=10 and 8 sections from Cdh8+/lacZ and Cdh8lacZ/lacZ animals respectively, 3 animals per genotype. p=0.12, Cochran-Armitage test. Scale=25µm. Lines and brackets indicate mean and standard error. (c) DNA construct for AAV encoding short-hairpin RNA against Sorcs3 (shSorcs3). (d-e) Whole mount view of GCL of a retina infected with control AAV expressing Td-Tomato (TdT), or AAV encoding shSORCS3. Sorcs3 protein is greatly reduced in shSORCS3-treated retinas. Scale=25µm. Each experiment in d-e was repeated independently in 3 and 5 animals respectively with similar results.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, J., Reggiani, J.D.S., Laboulaye, M.A. et al. Tbr1 instructs laminar patterning of retinal ganglion cell dendrites. Nat Neurosci 21, 659–670 (2018). https://doi.org/10.1038/s41593-018-0127-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-018-0127-z

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