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:

A CRISPR-based screen for Hedgehog signaling provides insights into ciliary function and ciliopathies

Abstract

Primary cilia organize Hedgehog signaling and shape embryonic development, and their dysregulation is the unifying cause of ciliopathies. We conducted a functional genomic screen for Hedgehog signaling by engineering antibiotic-based selection of Hedgehog-responsive cells and applying genome-wide CRISPR-mediated gene disruption. The screen can robustly identify factors required for ciliary signaling with few false positives or false negatives. Characterization of hit genes uncovered novel components of several ciliary structures, including a protein complex that contains δ-tubulin and ε-tubulin and is required for centriole maintenance. The screen also provides an unbiased tool for classifying ciliopathies and showed that many congenital heart disorders are caused by loss of ciliary signaling. Collectively, our study enables a systematic analysis of ciliary function and of ciliopathies, and also defines a versatile platform for dissecting signaling pathways through CRISPR-based screening.

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: Development of a Hedgehog-pathway reporter-based screening strategy.
Fig. 2: Overview of genome-wide-screen results.
Fig. 3: Evaluation of screen performance.
Fig. 4: FAM92A and TTC23 as transition-zone and EvC-zone components.
Fig. 5: Insights into ciliopathies from previously uncharacterized screen hits.
Fig. 6: TEDC1 and TEDC2 form a tetrameric complex with δ-tubulin and ε-tubulin.
Fig. 7: The TED complex is required for centriole stability.

Similar content being viewed by others

References

  1. Goetz, S. C. & Anderson, K. V. The primary cilium: a signalling centre during vertebrate development. Nat. Rev. Genet. 11, 331–344 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Nachury, M. V. How do cilia organize signalling cascades? Phil. Trans. R. Soc. Lond. B 369, 20130465 (2014).

    Article  Google Scholar 

  3. Bangs, F. & Anderson, K. V. Primary cilia and mammalian Hedgehog signaling. Cold Spring Harb. Perspect. Biol. 9, a028175 (2017).

    Article  PubMed  Google Scholar 

  4. Braun, D. A. & Hildebrandt, F. Ciliopathies. Cold Spring Harb. Perspect. Biol. 9, a028191 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Akizu, N. et al. Mutations in CSPP1 lead to classical Joubert syndrome. Am. J. Hum. Genet. 94, 80–86 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Wu, F., Zhang, Y., Sun, B., McMahon, A. P. & Wang, Y. Hedgehog signaling: from basic biology to cancer therapy. Cell Chem. Biol. 24, 252–280 (2017).

    Article  CAS  PubMed  Google Scholar 

  7. Pak, E. & Segal, R. A. Hedgehog signal transduction: key players, oncogenic drivers, and cancer therapy. Dev. Cell 38, 333–344 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Roosing, S. et al. Functional genome-wide siRNA screen identifies KIAA0586 as mutated in Joubert syndrome. eLife 4, e06602 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Wheway, G. et al. An siRNA-based functional genomics screen for the identification of regulators of ciliogenesis and ciliopathy genes. Nat. Cell Biol. 17, 1074–1087 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Jacob, L. S. et al. Genome-wide RNAi screen reveals disease-associated genes that are common to Hedgehog and Wnt signaling. Sci. Signal. 4, ra4 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  11. Kim, J. et al. Functional genomic screen for modulators of ciliogenesis and cilium length. Nature 464, 1048–1051 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Shalem, O. et al. Genome-scale CRISPR-Cas9 knockout screening in human cells. Science 343, 84–87 (2014).

    Article  CAS  PubMed  Google Scholar 

  13. Koike-Yusa, H., Li, Y., Tan, E. P., Velasco-Herrera Mdel, C. & Yusa, K. Genome-wide recessive genetic screening in mammalian cells with a lentiviral CRISPR-guide RNA library. Nat. Biotechnol. 32, 267–273 (2014).

    Article  CAS  PubMed  Google Scholar 

  14. Wang, T., Wei, J. J., Sabatini, D. M. & Lander, E. S. Genetic screens in human cells using the CRISPR-Cas9 system. Science 343, 80–84 (2014).

    Article  CAS  PubMed  Google Scholar 

  15. Zhou, Y. et al. High-throughput screening of a CRISPR/Cas9 library for functional genomics in human cells. Nature 509, 487–491 (2014).

    Article  CAS  PubMed  Google Scholar 

  16. Deans, R. M. et al. Parallel shRNA and CRISPR–Cas9 screens enable antiviral drug target identification. Nat. Chem. Biol. 12, 361–366 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Wang, T. et al. Identification and characterization of essential genes in the human genome. Science 350, 1096–1101 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Wang, T. et al. Gene essentiality profiling reveals gene networks and synthetic lethal interactions with oncogenic Ras. Cell 168, 890–903.e15 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Hart, T. et al. High-resolution CRISPR screens reveal fitness genes and genotype-specific cancer liabilities. Cell 163, 1515–1526 (2015).

    Article  CAS  PubMed  Google Scholar 

  20. Tzelepis, K. et al. A CRISPR dropout screen identifies genetic vulnerabilities and therapeutic targets in acute myeloid leukemia. Cell Rep. 17, 1193–1205 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Orchard, R. C. et al. Discovery of a proteinaceous cellular receptor for a norovirus. Science 353, 933–936 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Park, R. J. et al. A genome-wide CRISPR screen identifies a restricted set of HIV host dependency factors. Nat. Genet. 49, 193–203 (2017).

    Article  CAS  PubMed  Google Scholar 

  23. Arroyo, J. D. et al. A genome-wide CRISPR death screen identifies genes essential for oxidative phosphorylation. Cell Metab. 24, 875–885 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Taipale, J. et al. Effects of oncogenic mutations in Smoothened and Patched can be reversed by cyclopamine. Nature 406, 1005–1009 (2000).

    Article  CAS  PubMed  Google Scholar 

  25. Morgens, D. W. et al. Genome-scale measurement of off-target activity using Cas9 toxicity in high-throughput screens. Nat. Commun. 8, 15178 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Morgens, D. W., Deans, R. M., Li, A. & Bassik, M. C. Systematic comparison of CRISPR/Cas9 and RNAi screens for essential genes. Nat. Biotechnol. 34, 634–636 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Hart, T., Brown, K. R., Sircoulomb, F., Rottapel, R. & Moffat, J. Measuring error rates in genomic perturbation screens: gold standards for human functional genomics. Mol. Syst. Biol. 10, 733 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Huang, W., Sherman, B. T. & Lempicki, R. A. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat. Protoc. 4, 44–57 (2009).

    Article  CAS  Google Scholar 

  29. Webb, T. R. et al. Diphthamide modification of eEF2 requires a J-domain protein and is essential for normal development. J. Cell Sci. 121, 3140–3145 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Chen, C. M. & Behringer, R. R. Ovca1 regulates cell proliferation, embryonic development, and tumorigenesis. Genes Dev. 18, 320–332 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Loucks, C. M. et al. Matching two independent cohorts validates DPH1 as a gene responsible for autosomal recessive intellectual disability with short stature, craniofacial, and ectodermal anomalies. Hum. Mutat. 36, 1015–1019 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Allen, B. L. et al. Overlapping roles and collective requirement for the coreceptors GAS1, CDO, and BOC in SHH pathway function. Dev. Cell 20, 775–787 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Izzi, L. et al. Boc and Gas1 each form distinct Shh receptor complexes with Ptch1 and are required for Shh-mediated cell proliferation. Dev. Cell 20, 788–801 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Brinkman, E. K., Chen, T., Amendola, M. & van Steensel, B. Easy quantitative assessment of genome editing by sequence trace decomposition. Nucleic Acids Res. 42, e168 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  35. Lee, Y. L. et al. Cby1 promotes Ahi1 recruitment to a ring-shaped domain at the centriole-cilium interface and facilitates proper cilium formation and function. Mol. Biol. Cell 25, 2919–2933 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Burke, M. C. et al. Chibby promotes ciliary vesicle formation and basal body docking during airway cell differentiation. J. Cell Biol. 207, 123–137 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Glazer, A. M. et al. The Zn finger protein Iguana impacts Hedgehog signaling by promoting ciliogenesis. Dev. Biol. 337, 148–156 (2010).

    Article  CAS  PubMed  Google Scholar 

  38. Li, F. Q. et al. BAR domain-containing FAM92 proteins interact with Chibby1 to facilitate ciliogenesis. Mol. Cell. Biol. 36, 2668–2680 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Dorn, K. V., Hughes, C. E. & Rohatgi, R. A Smoothened-Evc2 complex transduces the Hedgehog signal at primary cilia. Dev. Cell 23, 823–835 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Pusapati, G. V. et al. EFCAB7 and IQCE regulate hedgehog signaling by tethering the EVC-EVC2 complex to the base of primary cilia. Dev. Cell 28, 483–496 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Ruiz-Perez, V. L. & Goodship, J. A. Ellis-van Creveld syndrome and Weyers acrodental dysostosis are caused by cilia-mediated diminished response to hedgehog ligands. Am. J. Med. Genet. C. Semin. Med. Genet. 151C, 341–351 (2009).

    Article  CAS  PubMed  Google Scholar 

  42. Shaheen, R. et al. Genomic analysis of Meckel-Gruber syndrome in Arabs reveals marked genetic heterogeneity and novel candidate genes. Eur. J. Hum. Genet. 21, 762–768 (2013).

    Article  CAS  PubMed  Google Scholar 

  43. Shaheen, R. et al. Characterizing the morbid genome of ciliopathies. Genome Biol. 17, 242 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Kar, A., Phadke, S. R., Das Bhowmik, A. & Dalal, A. Whole exome sequencing reveals a mutation in ARMC9 as a cause of mental retardation, ptosis and polydactyly. Preprint at https://www.biorxiv.org/content/early/2017/02/17/109124 (2017).

    Article  CAS  PubMed  Google Scholar 

  45. Panigrahi, I., Phadke, S. R. & Agarwal, S. S. Mental retardation, ptosis and polydactyly: a new autosomal recessive syndrome? Clin. Dysmorphol. 11, 289–292 (2002).

    Article  PubMed  Google Scholar 

  46. Nager, A. R. et al. An actin network dispatches ciliary GPCRs into extracellular vesicles to modulate signaling. Cell 168, 252–263.e14 (2017).

    Article  CAS  PubMed  Google Scholar 

  47. Phua, S. C. et al. Dynamic remodeling of membrane composition drives cell cycle through primary cilia excision. Cell 168, 264–279.e15 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Van De Weghe, J. C. et al. Mutations in ARMC9, which encodes a basal body protein, cause Joubert syndrome in humans and ciliopathy phenotypes in zebrafish. Am. J. Hum. Genet. 101, 23–36 (2017).

    Article  Google Scholar 

  49. Xu, M. et al. Mutations in the spliceosome component CWC27 cause retinal degeneration with or without additional developmental anomalies. Am. J. Hum. Genet. 100, 592–604 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Sifrim, A. et al. Distinct genetic architectures for syndromic and nonsyndromic congenital heart defects identified by exome sequencing. Nat. Genet. 48, 1060–1065 (2016).

    Article  CAS  PubMed  Google Scholar 

  51. Oegema, R. et al. Human mutations in integrator complex subunits link transcriptome integrity to brain development. PLoS Genet. 13, e1006809 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Sollis, E. et al. Identification and functional characterization of de novo FOXP1 variants provides novel insights into the etiology of neurodevelopmental disorder. Hum. Mol. Genet. 25, 546–557 (2016).

    Article  CAS  PubMed  Google Scholar 

  53. Comartin, D. et al. CEP120 and SPICE1 cooperate with CPAP in centriole elongation. Curr. Biol. 23, 1360–1366 (2013).

    Article  CAS  PubMed  Google Scholar 

  54. Lin, Y. N. et al. CEP120 interacts with CPAP and positively regulates centriole elongation. J. Cell Biol. 202, 211–219 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Mahjoub, M. R., Xie, Z. & Stearns, T. Cep120 is asymmetrically localized to the daughter centriole and is essential for centriole assembly. J. Cell Biol. 191, 331–346 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Kanie, T. et al. The CEP19-RABL2 GTPase complex binds IFT-B to initiate intraflagellar transport at the ciliary base. Dev. Cell 42, 22–36.e12 (2017).

    Article  CAS  PubMed  Google Scholar 

  57. Mojarad, B. A. et al. CEP19 cooperates with FOP and CEP350 to drive early steps in the ciliogenesis programme. Open Biol. 7, 170114 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Nishijima, Y. et al. RABL2 interacts with the intraflagellar transport-B complex and CEP19 and participates in ciliary assembly. Mol. Biol. Cell 28, 1652–1666 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Gupta, G. D. et al. A dynamic protein interaction landscape of the human centrosome-cilium interface. Cell 163, 1484–1499 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Jakobsen, L. et al. Novel asymmetrically localizing components of human centrosomes identified by complementary proteomics methods. EMBO J. 30, 1520–1535 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Knorz, V. J. et al. Centriolar association of ALMS1 and likely centrosomal functions of the ALMS motif-containing proteins C10orf90 and KIAA1731. Mol. Biol. Cell 21, 3617–3629 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Izquierdo, D., Wang, W. J., Uryu, K. & Tsou, M. F. Stabilization of cartwheel-less centrioles for duplication requires CEP295-mediated centriole-to-centrosome conversion. Cell Rep. 8, 957–965 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Chang, P. & Stearns, T. Delta-tubulin and epsilon-tubulin: two new human centrosomal tubulins reveal new aspects of centrosome structure and function. Nat. Cell Biol. 2, 30–35 (2000).

    Article  CAS  PubMed  Google Scholar 

  64. Dupuis-Williams, P. et al. Functional role of epsilon-tubulin in the assembly of the centriolar microtubule scaffold. J. Cell Biol. 158, 1183–1193 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Dutcher, S. K., Morrissette, N. S., Preble, A. M., Rackley, C. & Stanga, J. Epsilon-tubulin is an essential component of the centriole. Mol. Biol. Cell 13, 3859–3869 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. O’Toole, E. T., Giddings, T. H., McIntosh, J. R. & Dutcher, S. K. Three-dimensional organization of basal bodies from wild-type and delta-tubulin deletion strains of Chlamydomonas reinhardtii. Mol. Biol. Cell 14, 2999–3012 (2003).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Ross, I., Clarissa, C., Giddings, T. H. Jr. & Winey, M. ε-tubulin is essential in Tetrahymena thermophila for the assembly and stability of basal bodies. J. Cell Sci. 126, 3441–3451 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Sharma, A. et al. Centriolar CPAP/SAS-4 imparts slow processive microtubule growth. Dev. Cell 37, 362–376 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Schmidt, T. I. et al. Control of centriole length by CPAP and CP110. Curr. Biol. 19, 1005–1011 (2009).

    Article  CAS  PubMed  Google Scholar 

  70. Kirkham, M., Müller-Reichert, T., Oegema, K., Grill, S. & Hyman, A. A. SAS-4 is a C. elegans centriolar protein that controls centrosome size. Cell 112, 575–587 (2003).

    Article  CAS  PubMed  Google Scholar 

  71. Hein, M. Y. et al. A human interactome in three quantitative dimensions organized by stoichiometries and abundances. Cell 163, 712–723 (2015).

    Article  CAS  PubMed  Google Scholar 

  72. Huttlin, E. L. et al. Architecture of the human interactome defines protein communities and disease networks. Nature 545, 505–509 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Wong, Y. L. et al. Reversible centriole depletion with an inhibitor of Polo-like kinase 4. Science 348, 1155–1160 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Aguirre, A. J. et al. Genomic copy number dictates a gene-independent cell response to CRISPR/Cas9 targeting. Cancer Discov. 6, 914–929 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Wang, J. T., Kong, D., Hoerner, C. R., Loncarek, J. & Stearns, T. Centriole triplet microtubules are required for stable centriole formation and inheritance in human cells. eLife 6, e29061 (2017).

    PubMed  PubMed Central  Google Scholar 

  76. Li, Y. et al. Global genetic analysis in mice unveils central role for cilia in congenital heart disease. Nature 521, 520–524 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. MacArthur, D. G. et al. Guidelines for investigating causality of sequence variants in human disease. Nature 508, 469–476 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Pimenta-Marques, A. et al. A mechanism for the elimination of the female gamete centrosome in Drosophila melanogaster. Science 353, aaf4866 (2016).

    Article  CAS  PubMed  Google Scholar 

  79. Zheng, X. et al. Molecular basis for CPAP-tubulin interaction in controlling centriolar and ciliary length. Nat. Commun. 7, 11874 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Bond, J. et al. A centrosomal mechanism involving CDK5RAP2 and CENPJ controls brain size. Nat. Genet. 37, 353–355 (2005).

    Article  CAS  PubMed  Google Scholar 

  81. Gilmore, E. C. & Walsh, C. A. Genetic causes of microcephaly and lessons for neuronal development. Wiley Interdiscip. Rev. Dev. Biol. 2, 461–478 (2013).

    Article  CAS  PubMed  Google Scholar 

  82. Liew, G. M. et al. The intraflagellar transport protein IFT27 promotes BBSome exit from cilia through the GTPase ARL6/BBS3. Dev. Cell 31, 265–278 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Frank-Kamenetsky, M. et al. Small-molecule modulators of Hedgehog signaling: identification and characterization of Smoothened agonists and antagonists. J. Biol. 1, 10 (2002).

    Article  PubMed  PubMed Central  Google Scholar 

  84. Chen, J. K., Taipale, J., Young, K. E., Maiti, T. & Beachy, P. A. Small molecule modulation of Smoothened activity. Proc. Natl. Acad. Sci. USA 99, 14071–14076 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Breslow, D. K., Koslover, E. F., Seydel, F., Spakowitz, A. J. & Nachury, M. V. An in vitro assay for entry into cilia reveals unique properties of the soluble diffusion barrier. J. Cell Biol. 203, 129–147 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  86. Langmead, B., Trapnell, C., Pop, M. & Salzberg, S. L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10, R25 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  87. Elias, J. E. & Gygi, S. P. Target-decoy search strategy for increased confidence in large-scale protein identifications by mass spectrometry. Nat. Methods 4, 207–214 (2007).

    Article  CAS  PubMed  Google Scholar 

  88. de Hoon, M. J., Imoto, S., Nolan, J. & Miyano, S. Open source clustering software. Bioinformatics 20, 1453–1454 (2004).

    Article  PubMed  Google Scholar 

  89. Saldanha, A. J. Java Treeview: extensible visualization of microarray data. Bioinformatics 20, 3246–3248 (2004).

    Article  CAS  PubMed  Google Scholar 

  90. Turk, E. et al. Zeta-tubulin is a member of a conserved tubulin module and is a component of the centriolar basal foot in multiciliated cells. Curr. Biol. 25, 2177–2183 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Altschul, S. F. et al. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25, 3389–3402 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Li, W. et al. The EMBL-EBI bioinformatics web and programmatic tools framework. Nucleic Acids Res. 43, W580–W584 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Perrière, G. & Gouy, M. WWW-query: an on-line retrieval system for biological sequence banks. Biochimie 78, 364–369 (1996).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We acknowledge members of the laboratories of J.K.C. and M.V.N. for advice and technical support; S. van Dorp for assistance with image quantification; P. Beachy (Stanford) for NIH-3T3 cells, 8×Gli-luciferase reporter plasmid, and ShhN-producing HEK293T cells; K. Anderson (Sloan Kettering) for antibody to SMO; G. Crabtree, K. C. Garcia, and A. Pyle for use of plate readers; N. Dimitrova for microscope use; C. Bustamante for use of an Illumina sequencer; R. Rohatgi (Stanford) for antibodies to EVC and IQCE and NIH-3T3 FlpIn cells; P. Jackson (Stanford) for IMCD3 FlpIn cells; J. Wang and T. Stearns (Stanford) for sharing unpublished results and cDNA for Cby1; and M. Scott for helpful discussions. This project was supported by NIH Pathway to Independence Award K99/R00 HD082280 (D.K.B.), Damon Runyon Dale F. Frey Award DFS-11-14 (D.K.B.), a seed grant from the Stanford Center for Systems Biology (D.K.B., S.H., and G.T.H.) and Stanford ChEM-H (M.C.B.), an NWO Rubicon Postdoctoral Fellowship (S.H.), National Science Foundation Graduate Research Fellowship DGE-114747 (D.W.M.), a Walter V. and Idun Berry Award (K.H.), and NIH T32 HG000044 (G.T.H.), DP2 HD084069 (M.C.B.), R01 GM113100 (J.K.C.), and R01 GM089933 (M.V.N.). Cell sorting/flow cytometry was done on instruments in the Stanford Shared FACS Facility, including an instrument supported by NIH shared instrument grant S10RR025518. Mass spectrometry analyses were conducted in the Vincent Coates Foundation Mass Spectrometry Laboratory, Stanford University Mass Spectrometry and the Stanford Cancer Institute Proteomics/Mass Spectrometry Shared Resource; these centers are supported by Award S10 RR027425 from the National Center for Research Resources and NIH P30 CA124435, respectively. We thank C. Carstens, B. Borgo, P. Sheffield, and L. Bruhn of Agilent Technologies for cilia-focused oligonucleotide sublibraries.

Author information

Authors and Affiliations

Authors

Contributions

D.K.B., S.H., J.K.C., and M.V.N. conceived the project with advice from M.C.B.; D.K.B. and S.H. developed the Hh-pathway reporter screening strategy with assistance from B.K.V.; and D.W.M., K.H., A.L., G.T.H., and M.C.B. provided functional genomics expertise, the genome-wide sgRNA library, and software for screen data analysis. D.K.B. conducted the genome-wide screen and screen data analysis with assistance from S.H. and A.R.K.; D.K.B., S.H., A.R.K., and M.C.K. functionally characterized hit genes of interest, analyzed data, and prepared figures. D.K.B., S.H., J.K.C., and M.V.N. wrote the manuscript with assistance from M.C.B.; D.K.B., S.H., G.T.H., M.C.B., J.K.C. and M.V.N. provided funding for the project.

Corresponding authors

Correspondence to David K. Breslow, James K. Chen or Maxence V. Nachury.

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 Validation of positive- and negative-control sgRNAs

. a) Lysates from cells transduced with the indicated sgRNAs were analyzed by western blot of the indicated proteins. GLI3-FL and GLI3-R indicate the positions of full-length and repressor forms of GLI3, respectively. Molecular weight marker is indicated by lane M. b) Lysates from cells transduced with the indicated sgRNAs were analyzed by western blot against IFT88 and loading control Importin β (Impβ). Representative blots are shown for one of two independent experiments. For full size blots, see Supplementary Figure 8.

Supplementary Figure 2 Overview of sgRNA library and genome-wide-screen results

. a) Description of sgRNA library used, indicating sub-libraries, number of sgRNAs per sub-library, and grouping of sub-libraries into four screen batches. b) Assessment of screen performance at detecting genes with growth phenotypes using 188 positive and 478 negative reference genes. Performance was determined by ROC curve (left) and precision-recall analysis (right), with the area under each curve (AUC) shown. Performance for various screens using different libraries are plotted, with the cell type and data source indicated for each. Dashed lines indicate performance of a random classification model. c) Relationship between growth and signaling phenotypes for genome-wide screen data. P-values were calculated using casTLE using 100,000 sgRNA permutations. See also Supplementary Table 3. Positive values indicate enrichment (increased growth or blasticidin resistance); negative values indicate depletion (reduced growth or blasticidin sensitivity). Genes shown in grey were filtered out of further analyses due to strong negative growth phenotypes.

Supplementary Figure 3 Evaluation of screen performance

. a) Comparison of hit gene detection rates across various screen datasets and functional classes of genes. The number of genes in each class is indicated above each group of bars. See Supplementary Table 4 for additional details. b) Comparison of screen replicates for 95 genes measured in two different screen batches. Note: a small amount of jitter was added during plotting so that overlapping data points can be resolved. c) Comparison of screen replicates for 263 genes analyzed for roles in response to ShhN-induced versus SAG-induced Hh signaling. Gas1 is highlighted in red as a prominent exception to the high level of concordance between screen results. P-values plotted in panels (b) and (c) were calculated using casTLE; see Methods and Supplementary Table 3.

Supplementary Figure 4 Characterization of FAM92A and TTC23 as transition-zone and EvC-zone components

. a) Analysis of Cas9-induced mutations in cell pools transduced with the indicated sgRNAs. Mutation frequencies were identified by TIDE sequencing (Tracking of Indels by Decomposition)34. Note that the sum of mutations for each gene is less than 100% because the TIDE algorithm only reports indels of up to 50 bp that can be identified with high confidence (P < 0.001). b) Signaling-induced blasticidin resistance is shown for 3T3-[Shh-BlastR;Cas9] cells transduced with the indicated sgRNAs. Bars show mean IC50 values from N=5 (Ctrl; same data as shown in Fig. 1c), N=4 (Fam92a sgRNA) or N=3 (Ttc23 sgRNA) independent experiments (indicated by circles). c) Transfected CBY1-GFP co-localizes with endogenous FAM92A at the transition zone of IMCD3 cells. Line plot at right shows relative intensity from base to tip. Scale bar: 1 μm. Representative image from five fields of view. d) Validation of the anti-FAM92A antibody. Staining reveals prominent transition zone staining in wildtype (WT) cells but not cells transduced with a Fam92a-targeting sgRNA. Scale bar: 5 μm. Representative image of three independent experiments (five fields of view in each). e) TTC23-LAP stably expressed in IMCD3 cells co-localizes with EVC. Line scans at right show relative intensity from base to tip. Scale bar: 1 μm. Representative image of three independent experiments (five fields of view in each). f) The frequency of ciliogenesis was assessed in 3T3-[Shh-BlastR;Cas9] cells transduced with the indicated sgRNAs. Bars show mean percentage of ciliated cells; dots show ciliated fraction in each independent experiment (at least 200 cells were analyzed in each of two independent experiments). g) Top: ciliary EVC fluorescence was measured for cells transduced with the indicated sgRNAs. The mean and standard error of the mean are plotted for N= 564, 277 and 285 cells expressing control (Ctrl), Ttc23-1 and Ttc23-2 sgRNAs, respectively. Data shown are from one of two representative experiments. Bottom: representative images showing EVC immunostaining in the indicated cells. Scale bars: 1 μm. h) Representative images showing ciliary IQCE staining in cells transduced with Ctrl and Ttc23-targeting sgRNAs. Scale bars: 5 μm. Representative images are shown from one of two independent experiments (five fields of view each).

Supplementary Figure 5 Functional characterization of TXNDC15 and ARMC9

. a) Distribution of cilium lengths measured for 3T3-[Shh-BlastR;Cas9] cells transduced with the indicated sgRNAs. Control (Ctrl) sgRNA cells were significantly different from cells with Txndc15-1 sgRNA (P=4.4x10−10) or with Txndc15-2 sgRNA (P=3.3x10−10). Significance determined by two-sided Kolmogorov-Smirnov test; N = 902, 899, 1048 and 1242 cilia for WT, Ctrl, Txndc15-1 and Txndc15-2 sgRNAs, respectively. Data shown are drawn from 3 independent experiments. b) Staining for ciliary markers ARL13B and acetylated tubulin (AcTub) was performed for cells transduced with the indicated sgRNAs. Integrated ciliary fluorescence is shown relative to that of Ctrl sgRNA. Lines show means of N=3 independent experiments (symbols; at least 200 cilia analyzed per condition). c) Staining for ciliary markers ARL13B, acetylated tubulin (AcTub), and polyglutamylated tubulin (polyGlu-tub) was performed for 3T3-[Shh-BlastR;Cas9] cells transduced with the indicated sgRNAs. Ciliary pixel intensity is shown relative to Ctrl sgRNA = 1. Lines show the mean of N=3 independent experiments (symbols; at least 200 cilia analyzed per condition). d) ARMC9-3xFLAG localization was assessed in NIH-3T3 FlpIn cells. Scale bar = 5 μm (1 μm for inset). Image shown is representative of three independent experiments (five fields of view in each). e) Plots of ARMC9-3xFLAG intensity along the length of the cilium from base (position 0) to tip (position 1.0) are shown for IMCD3 cells treated with ShhN in the presence or absence of Smo inhibitor vismodegib. The mean and standard deviation are plotted after normalizing the total intensity in each cilium to 1.0. f) Fluorescence intensity of GLI2 at the ciliary tip was measured for 3T3-[Shh-BlastR;Cas9] cells transduced with the indicated sgRNAs and treated with ShhN or left untreated. Mean intensity (relative to Ctrl sgRNA cells +ShhN) and standard error of the mean are shown for each of N=3 experiments (at least 250 cilia analyzed for each condition). g) Fluorescence intensity of ciliary SMO was measured for cells transduced with the indicated sgRNAs and treated with ShhN or left untreated. Mean and 95% confidence interval are shown (N = number of cilia analyzed, one representative replicate out of two independent experiments). h) Representative micrographs for ciliary GLI2, GLI3 (three independent experiments), and SMO (two independent experiments) are shown for the indicated cell lines and treatment conditions. Scale bars = 1 μm.

Supplementary Figure 6 Functional characterization of TEDC1 and TEDC2

. a) Proteins affinity purified from IMCD3 wildtype (WT) or TEDC2-LAP cells were analyzed by western blotting with the indicated antibodies. See also Supplementary Figure 8. b) For blots shown in Fig. 6c, non-specific bands in anti-FLAG blot (marked by asterisks) show comparable protein loading among samples. Result is representative of two independent experiments. c) Signaling-induced blasticidin resistance is shown for 3T3-[Shh-BlastR;Cas9] cells transduced with the indicated sgRNAs. Bars show mean IC50 values from N=5 (Ctrl; same data as shown in Fig. 1c), N=3 (Tedc1 sgRNAs) or N=2 (Tedc2 sgRNAs) independent experiments (circles). d) Signaling-induced regulation of GLI3 processing and expression of target gene GLI1 were analyzed by Western blot using cells transduced with the indicated sgRNAs. GLI3-FL and GLI3-R indicate the positions of full-length and repressor forms of GLI3, respectively. Representative blot from two independent experiments is shown. See also Supplementary Figure 8. e) Centrioles marked by centrin2 and γ-tubulin were counted in cells transduced with the indicated sgRNAs. Following sgRNA introduction, TEDC1 rescue cells were transduced to express sgRNA-resistant TEDC1-3xFLAG-T2A-GFP (rescued cells are marked by GFP). N = 100, 149, 76 for Ctrl sgRNA, Tedc1-2 sgRNA, and TEDC1 rescue, respectively; * indicates P = 9.0x10−34; ** indicates P = 1.7x10−43; N.S. indicates not significant (P > 0.05), two-sided Fisher’s exact test. One of two representative experiments. f). Centrioles were stained in the indicated cells using antibodies to centrin2 and γ-tubulin. Cilia were stained using an antibody to ARL13B. Representative results from one of three independent experiments are shown. g) TEDC1-3xFLAG localizes to centrioles in rescued cells. Representative results from one of two independent experiments are shown. Scale bars: 5 μm (2 μm for inset). h) For cells transduced with the indicated sgRNAs, centrioles were visualized by staining with antibodies to centrin3 and γ-tubulin. Insets show centrin3 staining in mitotic cells, marked by yellow arrowheads. Scale bars: 5 μm (2 μm for insets). Images shown are from one of two independent experiments (10 fields of view each).

Supplementary Figure 7 Evolutionary analysis of TED-complex components

. Tree diagrams showing sequence relationships for select proteins, with branch lengths scaled equally for all proteins. See also Supplementary Table 8.

Supplementary Figure 8 Full size western blots

. Uncropped western blots for the indicated Figures. For Fig. 6c and Supplementary Fig. 6b, lanes 1-3 refer to cells transfected with: TEDC1-6xMYC and TagRFP-3xFLAG (lane 1); TagRFP-6xMYC and TEDC2-3xFLAG (lane 2); and TEDC1-6xMYC and TEDC2-3xFLAG (lane 3). In other panels, labels “-“, “+”, and “M” refer to unstimulated cells, ShhN-stimulated cells, and molecular weight marker lanes, respectively.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8, Supplementary Note and Supplementary Tables 9 and 10

Life Sciences Reporting Summary

Supplementary Table 1

List of select sgRNA and primer sequences

Supplementary Table 2

List of sgRNA counts by deep sequencing of screen cell pools

Supplementary Table 3

casTLE output (gene scores) for genome-wide screen

Supplementary Table 4

Summary of screen data for select genes of interest

Supplementary Table 5

List of functional categories enriched among screen hits

Supplementary Table 6

List of proteins identified by mass spectrometry in affinity purifications

Supplementary Table 7

Clustered growth data from Aguirre et al

Supplementary Table 8

Evolutionary analysis of TED complex components

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Breslow, D.K., Hoogendoorn, S., Kopp, A.R. et al. A CRISPR-based screen for Hedgehog signaling provides insights into ciliary function and ciliopathies. Nat Genet 50, 460–471 (2018). https://doi.org/10.1038/s41588-018-0054-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-018-0054-7

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research