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:

SETD2 loss perturbs the kidney cancer epigenetic landscape to promote metastasis and engenders actionable dependencies on histone chaperone complexes

Abstract

SETD2 is a histone H3 lysine 36 (H3K36) trimethyltransferase that is mutated with high prevalence (13%) in clear cell renal cell carcinoma (ccRCC). Genomic profiling of primary ccRCC tumors reveals a positive correlation between SETD2 mutations and metastasis. However, whether and how SETD2 loss promotes metastasis remains unclear. In this study, we used a SETD2-mutant (SETD2MT) metastatic ccRCC human-derived cell line and xenograft models and showed that H3K36me3 restoration greatly reduced distant metastases of ccRCC in mice in a matrix metalloproteinase 1 (MMP1)-dependent manner. An integrated multiomics analysis using assay for transposase-accessible chromatin using sequencing (ATAC-seq), chromatin immunoprecipitation–sequencing (ChIP–seq) and RNA sequencing (RNA-seq) established a tumor suppressor model in which loss of SETD2-mediated H3K36me3 activates enhancers to drive oncogenic transcriptional output through regulation of chromatin accessibility. Furthermore, we uncovered mechanism-based therapeutic strategies for SETD2-deficient cancer through the targeting of specific histone chaperone complexes, including ASF1A/ASF1B and SPT16. Overall, SETD2 loss creates a permissive epigenetic landscape for cooperating oncogenic drivers to amplify transcriptional output, providing unique therapeutic opportunities.

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: SETD2 is highly mutated in metastatic ccRCC tumors, and restoration of H3K36me3 in SETD2MT ccRCC suppresses tumor metastasis.
Fig. 2: Loss of SETD2-mediated H3K36me3 induces a genome-wide increase in chromatin accessibility that correlates with increased oncogenic transcriptional output.
Fig. 3: Loss of SETD2-mediated H3K36me3 induces genome-wide epigenetic reprogramming and activation of enhancers.
Fig. 4: SETD2 loss of function induces MMP1 expression to promote ccRCC metastasis.
Fig. 5: Loss of SETD2-mediated H3K36me3 increases histone chaperone recruitment and sensitizes cancer cells to genetic inactivation of histone chaperones.
Fig. 6: Loss of SETD2-mediated H3K36me3 sensitizes cancer cells to inhibition of the FACT complex.

Similar content being viewed by others

Data availability

Raw ATAC-seq, RNA-seq and ChIP–seq sequencing data and the normalized tracks that support the findings of this study have been deposited in the Gene Expression Omnibus (GEO) database under GSE146583. Further information and requests for resources and reagents should be directed to the corresponding author. All unique/stable reagents generated in this study are available from the corresponding author with a completed materials transfer agreement. Source data are provided with this paper.

Code availability

Sequencing data processing and analysis was performed using open-source software packages. The RNA-seq analysis was conducted using Trimmomatic (v0.38), STAR (v2.7.1a), HTSeq (v0.11.2) and DESeq2 (v1.22.2), as reported in ref. 59. The ATAC-seq analysis was performed using Trimmomatic (v0.38), Bowtie2 (v2.3.4.3), MACS2 (v2.1.2) and IDR (v2.0.3) by using scripts in https://github.com/ENCODE-DCC/atac-seq-pipeline/tree/master/src with setting ‘–extsize 200–shift -100–nomodel’ for the peak calling step. ChIP–seq analysis was performed using BWA (v0.7.17-r1188), SAMtools (v1.9), Picard (v2.18.16), MACS2 (v2.1.2) and IDR (v2.0.3) by using scripts in https://github.com/ENCODE-DCC/chip-seq-pipeline2/tree/master/src. Each ATAC-seq and ChIP–seq peak was assigned to the closest gene as in https://github.com/hchintalapudi/ATAC-seq-ATAC-array/blob/777fbb6d9e281e7715ba15b7522c846f3b6234b2/As_Atlas_creation.R, and counts in these peaks were obtained using featureCounts (v1.6.4). The differential analyses were performed using gene counts for RNA-seq and peak counts for ATAC-seq and ChIP–seq following the DESeq2 (v1.22.2) pipeline (https://github.com/mikelove/DESeq2). Bedgraph files for each genomic dataset were generated using bedtools (v2.27.1) genomeCoverageBed scaled with (-scale) DESeq2 size factors (https://github.com/arq5x/bedtools/blob/master/docs/content/tools/genomecov.rst). Those scaled bedgraph files were converted to normalized bigwig tracks using UCSC bedgraph2bigwig. Heat maps and metaplots of ChIP–seq and ATAC-seq data were generated using deepTools (v3.1.1) plotHeatmap (https://github.com/deeptools/deepTools/blob/master/docs/content/tools/plotHeatmap.rst) and plotProfile (https://github.com/deeptools/deepTools/blob/master/docs/content/tools/plotProfile.rst) functions by providing peaks as bed files and normalized bigwig files.

References

  1. Sun, X.-J. et al. Identification and characterization of a novel human histone H3 lysine 36-specific methyltransferase. J. Biol. Chem. 280, 35261–35271 (2005).

    Article  CAS  PubMed  Google Scholar 

  2. Kizer, K. O. et al. A novel domain in Set2 mediates RNA polymerase II interaction and couples histone H3 K36 methylation with transcript elongation. Mol. Cell. Biol. 25, 3305–3316 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Edmunds, J. W., Mahadevan, L. C. & Clayton, A. L. Dynamic histone H3 methylation during gene induction: HYPB/Setd2 mediates all H3K36 trimethylation. EMBO J. 27, 406–420 (2008).

    Article  CAS  PubMed  Google Scholar 

  4. Hu, M. et al. Histone H3 lysine 36 methyltransferase Hypb/Setd2 is required for embryonic vascular remodeling. Proc. Natl Acad. Sci. USA 107, 2956–2961 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Wagner, E. J. & Carpenter, P. B. Understanding the language of Lys36 methylation at histone H3. Nat. Rev. Mol. Cell Biol. 13, 115–126 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Venkatesh, S. et al. Set2 methylation of histone H3 lysine 36 suppresses histone exchange on transcribed genes. Nature 489, 452–455 (2012).

    Article  CAS  PubMed  Google Scholar 

  7. Luco, R. F. et al. Regulation of alternative splicing by histone modifications. Science 327, 996–1000 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Pfister, S. X. et al. SETD2-dependent histone H3K36 trimethylation is required for homologous recombination repair and genome stability. Cell Rep. 7, 2006–2018 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Carvalho, S. et al. SETD2 is required for DNA double-strand break repair and activation of the p53-mediated checkpoint. eLife 3, e02482 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Park, I. Y. et al. Dual chromatin and cytoskeletal remodeling by SETD2. Cell 166, 950–962 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Chen, K. et al. Methyltransferase SETD2-mediated methylation of STAT1 is critical for interferon antiviral activity. Cell 170, 492–506 (2017).

    Article  CAS  PubMed  Google Scholar 

  12. Hsieh, J. J. et al. Chromosome 3p loss-orchestrated VHL, HIF, and epigenetic deregulation in clear cell renal cell carcinoma. J. Clin. Oncol. 36, 3533–3539 (2018).

    Article  CAS  PubMed Central  Google Scholar 

  13. Cancer Genome Atlas Research Network. Comprehensive molecular characterization of clear cell renal cell carcinoma. Nature 499, 43–49 (2013).

  14. Hsieh, J. J. et al. Renal cell carcinoma. Nat. Rev. Dis. Primers 3, 17009 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Haase, V. H., Glickman, J. N., Socolovsky, M. & Jaenisch, R. Vascular tumors in livers with targeted inactivation of the von Hippel–Lindau tumor suppressor. Proc. Natl Acad. Sci. USA 98, 1583–1588 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Kleymenova, E. et al. Susceptibility to vascular neoplasms but no increased susceptibility to renal carcinogenesis in Vhl knockout mice. Carcinogenesis 25, 309–315 (2004).

    Article  CAS  PubMed  Google Scholar 

  17. Rankin, E. B., Tomaszewski, J. E. & Haase, V. H. Renal cyst development in mice with conditional inactivation of the von Hippel–Lindau tumor suppressor. Cancer Res. 66, 2576–2583 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Kapitsinou, P. P. & Haase, V. H. The VHL tumor suppressor and HIF: insights from genetic studies in mice. Cell Death Differ. 15, 650–659 (2008).

    Article  CAS  PubMed  Google Scholar 

  19. Nargund, A. M. et al. The SWI/SNF protein PBRM1 restrains VHL-loss-driven clear cell renal cell carcinoma. Cell Rep. 18, 2893–2906 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Hakimi, A. A. et al. Adverse outcomes in clear cell renal cell carcinoma with mutations of 3p21 epigenetic regulators BAP1 and SETD2: a report by MSKCC and the KIRC TCGA research network. Clin. Cancer Res. 19, 3259–3267 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Turajlic, S. et al. Deterministic evolutionary trajectories influence primary tumor growth: TRACERx renal. Cell 173, 595–610 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Ho, T. H. et al. High-resolution profiling of histone h3 lysine 36 trimethylation in metastatic renal cell carcinoma. Oncogene 35, 1565 (2016).

    Article  CAS  PubMed  Google Scholar 

  23. Zehir, A. et al. Mutational landscape of metastatic cancer revealed from prospective clinical sequencing of 10,000 patients. Nat. Med. 23, 703–713 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Dong, Y. et al. Tumor xenografts of human clear cell renal cell carcinoma but not corresponding cell lines recapitulate clinical response to sunitinib: feasibility of using biopsy samples. Eur. Urol. Focus 3, 590–598 (2017).

    Article  PubMed  Google Scholar 

  25. Mandriota, S. J. et al. HIF activation identifies early lesions in VHL kidneys: evidence for site-specific tumor suppressor function in the nephron. Cancer Cell 1, 459–468 (2002).

    Article  CAS  PubMed  Google Scholar 

  26. Semenza, G. L. HIF-1 mediates metabolic responses to intratumoral hypoxia and oncogenic mutations. J. Clin. Invest. 123, 3664–3671 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Bos, P. D. et al. Genes that mediate breast cancer metastasis to the brain. Nature 459, 1005–1009 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Simon, J. M. et al. Variation in chromatin accessibility in human kidney cancer links H3K36 methyltransferase loss with widespread RNA processing defects. Genome Res. 24, 241–250 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Grant, C. E., Bailey, T. L. & Noble, W. S. FIMO: scanning for occurrences of a given motif. Bioinformatics 27, 1017–1018 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Denny, S. K. et al. Nfib promotes metastasis through a widespread increase in chromatin accessibility. Cell 166, 328–342 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Morrow, J. J. et al. Positively selected enhancer elements endow osteosarcoma cells with metastatic competence. Nat. Med. 24, 176–185 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Rodrigues, P. et al. NF-κB-dependent lymphoid enhancer co-option promotes renal carcinoma metastasis. Cancer Discov. 8, 850–865 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Kessenbrock, K., Plaks, V. & Werb, Z. Matrix metalloproteinases: regulators of the tumor microenvironment. Cell 141, 52–67 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Yeo, N. C. et al. An enhanced CRISPR repressor for targeted mammalian gene regulation. Nat. Methods 15, 611–616 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Hacker, K. E. et al. Structure/function analysis of recurrent mutations in SETD2 protein reveals a critical and conserved role for a set domain residue in maintaining protein stability and histone H3 Lys-36 trimethylation. J. Biol. Chem. 291, 21283–21295 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Burgess, R. J. & Zhang, Z. Histone chaperones in nucleosome assembly and human disease. Nat. Struct. Mol. Biol. 20, 14–22 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Venkatesh, S. & Workman, J. L. Histone exchange, chromatin structure and the regulation of transcription. Nat. Rev. Mol. Cell Biol. 16, 178–189 (2015).

    Article  CAS  PubMed  Google Scholar 

  38. Smolle, M., Workman, J. L. & Venkatesh, S. reSETting chromatin during transcription elongation. Epigenetics 8, 10–15 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Tsubota, T. et al. Histone H3-K56 acetylation is catalyzed by histone chaperone-dependent complexes. Mol. Cell 25, 703–712 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Carter, D. R. et al. Therapeutic targeting of the MYC signal by inhibition of histone chaperone FACT in neuroblastoma. Sci. Transl. Med. 7, 312ra176 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Zhang, L. et al. Multisite substrate recognition in Asf1-dependent acetylation of histone H3 K56 by Rtt109. Cell 174, 818–830 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Jeng, P. S., Inoue-Yamauchi, A., Hsieh, J. J. & Cheng, E. H. BH3-dependent and independent activation of BAX and BAK in mitochondrial apoptosis. Curr. Opin. Physiol. 3, 71–81 (2018).

  43. Cheng, E. H. et al. BCL-2, BCL-XL sequester BH3 domain-only molecules preventing BAX- and BAK-mediated mitochondrial apoptosis. Mol. Cell 8, 705–711 (2001).

    Article  CAS  PubMed  Google Scholar 

  44. Kim, H. et al. Hierarchical regulation of mitochondrion-dependent apoptosis by BCL-2 subfamilies. Nat. Cell Biol. 8, 1348–1358 (2006).

    Article  CAS  PubMed  Google Scholar 

  45. Kim, H. et al. Stepwise activation of BAX and BAK by tBID, BIM, and PUMA initiates mitochondrial apoptosis. Mol. Cell 36, 487–499 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Ren, D. et al. BID, BIM, and PUMA are essential for activation of the BAX- and BAK-dependent cell death program. Science 330, 1390–1393 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Chen, H. C. et al. An interconnected hierarchical model of cell death regulation by the BCL-2 family. Nat. Cell Biol. 17, 1270–1281 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Mar, B. G. et al. SETD2 alterations impair DNA damage recognition and lead to resistance to chemotherapy in leukemia. Blood 130, 2631–2641 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Guenther, M. G., Levine, S. S., Boyer, L. A., Jaenisch, R. & Young, R. A. A chromatin landmark and transcription initiation at most promoters in human cells. Cell 130, 77–88 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Wang, G. X. et al. ΔNp63 inhibits oxidative stress-induced cell death, including ferroptosis, and cooperates with the BCL-2 family to promote clonogenic survival. Cell Rep. 21, 2926–2939 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Sanjana, N. E., Shalem, O. & Zhang, F. Improved vectors and genome-wide libraries for CRISPR screening. Nat. Methods 11, 783–784 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Shi, J. et al. Discovery of cancer drug targets by CRISPR–Cas9 screening of protein domains. Nat. Biotechnol. 33, 661–667 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Lu, C. et al. Histone H3K36 mutations promote sarcomagenesis through altered histone methylation landscape. Science 352, 844–849 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. & Greenleaf, W. J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat. Methods 10, 1213–1218 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. 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 

  57. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Li, Q., Brown, J. B., Huang, H. & Bickel, P. J. Measuring reproducibility of high-throughput experiments. Ann. Appl. Stat. 5, 1752–1779 (2011).

    Article  Google Scholar 

  59. Philip, M. et al. Chromatin states define tumour-specific T cell dysfunction and reprogramming. Nature 545, 452–456 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  61. Anders, S., Pyl, P. T. & Huber, W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  63. Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl Acad. Sci. USA 102, 15545–15550 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Foroutan, M. et al. Single sample scoring of molecular phenotypes. BMC Bioinformatics 19, 404 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Colaprico, A. et al. TCGAbiolinks: an R/Bioconductor package for integrative analysis of TCGA data. Nucleic Acids Res. 44, e71 (2016).

    Article  PubMed  Google Scholar 

  66. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Subgroup, G. P. D. P. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  Google Scholar 

  68. Smyth, G. K., Shi, W. & Liao, Y. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30, 923–930 (2013).

    PubMed  Google Scholar 

  69. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Hinrichs, A. S., Zweig, A. S., Karolchik, D., Barber, G. & Kent, W. J. BigWig and BigBed: enabling browsing of large distributed datasets. Bioinformatics 26, 2204–2207 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  71. Richter, A. S. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We apologize to all the investigators whose research could not be appropriately cited due to space limitations. This work was supported by an NIH grant to E.H.C. (R01 CA223231) as well as an NCI Cancer Center Support Grant (P30 CA008748). Y.X. was supported by an NCI Predoctoral to Postdoctoral Fellow Transition Award (F99 CA234949).

Author information

Authors and Affiliations

Authors

Contributions

Y.X. designed and conducted experiments and analyzed the data. E.H.C. designed the research, analyzed the data and supervised the project. S.S., Y.W., S.H. and A.M.N. conducted some experiments. Y.L. analyzed some data. J.J.H. supervised the project. M.S. performed computational analyses. C.S.L. supervised the computational analyses.

Corresponding authors

Correspondence to James J. Hsieh, Christina S. Leslie or Emily H. Cheng.

Ethics declarations

Competing interests

J.J.H. has consulted for Eisai and BostonGene and has received clinical trial funding from Bristol Myers Squibb, Merck, AstraZeneca, Exelixis, Calithera and SillaJen. J.J.H. has received research funding from Merck, BostonGene and TScan. All other authors have no competing interests to declare.

Peer review

Peer review information

Nature Cancer thanks Jonathan Licht and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Restoration of H3K36me3 in SETD2 mutant ccRCC cells suppresses tumor metastasis.

a-b, Representative bioluminescence images of NOD.Cg-Prkdcscid Il2rgtm1Wjl/SzJ (NSG) mice injected with the indicated JHRCC12 cells into subrenal capsules of unilateral kidneys are shown in a and the quantification of bioluminescence is shown in b (mean ± s.d., n = 5 for H3K36me3-deficient and n = 4 for H3K36me3-proficient). n.s., not significant (two-tailed unpaired Student’s t-test). c, Representative gross images of the indicated organs in mice received subrenal capsule injection of the indicated JHRCC12 cells. Yellow arrowheads indicate metastatic tumors.

Source data

Extended Data Fig. 2 Summary of differentially expressed genes in H3K36me3 compared to H3K36me3+ JHRCC12 cells as well as in SETD2MT compared to SETD2WT ccRCC.

a, Venn diagram showing overlap of differentially upregulated genes (FDR < 0.05, log2(FC) > 0) in H3K36me3 compared to H3K36me3+ (SETD2∆N-transduced) JHRCC12 cells and differentially upregulated genes (FDR < 0.05, log2(FC) > 0) in SETD2MT compared to SETD2WT ccRCC from the TCGA-KIRC dataset. b, Venn diagram showing overlap of differentially downregulated genes (FDR < 0.05, log2(FC) < 0) in H3K36me3 compared to H3K36me3+ (SETD2∆N-transduced) JHRCC12 cells and differentially downregulated genes (FDR < 0.05, log2(FC) < 0) in SETD2MT compared to SETD2WT ccRCC from the TCGA-KIRC dataset. c, The open chromatin peaks comparing H3K36me3 with H3K36me3+ JHRCC12 cells were enriched for the binding motifs of STAT family transcription factors.

Extended Data Fig. 3 Genes that are upregulated in SETD2∆N-transduced compared to SETD2-deficient JHRCC12 cells have higher H3K36me3 levels than downregulated genes in SETD2∆N-transduced cells.

a, Normalized H3K36me3 ChIP-seq counts in SETD2-deficient JHRCC12 cells over gene bodies of stringently up- and down-regulated genes (FDR < 0.05 and log2(FC) > 1) in SETD2-proficient (SETD2∆N-transduced) compared to SETD2-deficient cells. b, Normalized H3K36me3 ChIP-seq counts in SETD2∆N-transduced JHRCC12 cells over gene bodies of stringently up- and down-regulated genes (FDR < 0.05 and log2(FC) > 1) in SETD2-proficient (SETD2∆N-transduced) compared to SETD2-deficient cells. Centers of the boxes indicate median values, the lower and upper hinges correspond to the first and third quartiles and the upper (lower) whiskers extend from the hinge to the largest (smallest) value no further than 1.5 times the distance between the first and third quartiles. P values were calculated using one-sided Wilcoxon rank sum tests. c, Cumulative distribution of H3K36me3 levels in SETD2∆N-transduced cells over gene bodies of significantly upregulated (red) or downregulated (blue) genes (FDR < 0.05) in SETD2-proficient (SETD2∆N-transduced) compared to SETD2-deficient cells. P values were calculated using one-sided KS test comparing H3K36me3 levels in differentially expressed genes to all genes. d, Metaplots showing the normalized average levels of histone marks across gene bodies of stringently upregulated genes (FDR < 0.05 and log2(FC) > 1, n = 174) comparing H3K36me3+ (SETD2∆N-transduced) with H3K36me3 (control) JHRCC12 cells by ChIP-seq. e, Metaplots showing the normalized average levels of histone marks across gene bodies of stringently downregulated genes (FDR < 0.05 and log2(FC) < -1, n = 194) comparing H3K36me3+ (SETD2∆N-transduced) with H3K36me3 (control) JHRCC12 cells by ChIP-seq. TSS, transcription start site; TES, transcription end site.

Extended Data Fig. 4 Loss of SETD2-mediated H3K36me3 induces genome-wide epigenetic changes.

a. Pie chart showing the percentage of differentially enriched ChIP-seq peaks (FDR < 0.05) for each histone mark in promoter, intronic, intergenic, and exonic regions comparing H3K36me3 with H3K36me3+ (SETD2∆N-transduced) JHRCC12 cells. b. Heatmap of differentially accessible ATAC-seq peaks (FDR < 0.05 and log2(FC) > 1) assigned to an upregulated gene, in a 5 kb window grouped by localization at promoter, intron, and intergenic regions (n = 1012). c. Heatmap of differentially accessible ATAC-seq peaks (FDR < 0.05 and log2(FC) > 1) assigned to a downregulated gene, in a 5 kb window grouped by localization at promoter, intron, and intergenic regions (n = 456).

Extended Data Fig. 5 Open chromatin regions that are not affected by the status of H3K36me3 show no differences in both enhancer and promoter marks.

a. Heatmaps for non-differential ATAC-seq peaks (FDR > 0.05; n = 24016) in 5 kb window grouped by localization at promoter, intron, and intergenic regions and heatmaps showing histone modifications in 5 kb window in the same regions of ATAC-seq peaks. b. Metapeak plots of non-differential ATAC-seq peaks (FDR > 0.05) in 5 kb window grouped by localization at promoter, intron, and intergenic regions and metapeak plots of histone modifications in 5 kb window in the same regions of ATAC-seq peaks.

Extended Data Fig. 6 Open chromatin regions that are not affected by the status of H3K36me3 show no differences in histone modifications.

a. Heatmap of non-differential ATAC-seq peaks assigned to an upregulated gene in a 5 kb window grouped by localization at promoter, intron, and intergenic regions (n = 6313) and heatmaps showing histone modifications in 5 kb window in the same regions of ATAC-seq peaks. b. Heatmap of non-differential ATAC-seq peaks assigned to a downregulated gene in a 5 kb window grouped by localization at promoter, intron, and intergenic regions (n = 7835) and heatmaps showing histone modifications in 5 kb window in the same regions of ATAC-seq peaks.

Extended Data Fig. 7 Loss of SETD2-mediated H3K36me3 induces a genome-wide increase of active/permissive histone marks that correlate with increased chromatin accessibility.

a, The extent of co-occurrence of two up/downregulated histone modification ChIP-seq or ATAC-seq peaks were assessed using Fisher’s exact test for each pairwise comparison of those peaks. The plot shows the odds ratio and Bonferroni adjusted P value from Fisher’s exact test. Enrichment ratios greater than 1 implies the peaks of interest are more likely to co-occur, whereas enrichment ratios less than 1 means that the peaks of interest are less likely to co-occur (*, P < 0.05; **, P < 0.01; ***, P < 0.001). b, The extent of co-occurrence of up- or downregulated genes with up- or downregulated histone modification ChIP-seq or ATAC-seq peaks were assessed using Fisher’s exact test for each pairwise comparison. The plot shows the odds ratio and Bonferroni adjusted P value from Fisher’s exact test. Enrichment ratios greater than 1 implies that up- (or down-) regulated genes are enriched in up- (or down-) regulated peaks, while enrichment ratios less than 1 implies that up- (or down-) regulated genes are depleted in up- (or down-) regulated peaks (*, P < 0.05; **, P < 0.01; ***, P < 0.001).

Extended Data Fig. 8 Comparison of ATAC-seq and ChIP-seq at the MMP1 locus.

ATAC-seq tracks and ChIP-seq tracks for the indicated histone marks at the MMP1 locus in H3K36me3 or H3K36me3+ (SETD2∆N-transduced) JHRCC12 cells.

Extended Data Fig. 9 Loss of SETD2-mediated H3K36me3 increases H3K56ac levels.

a. Metaplots showing the normalized distribution profiles of H3K56ac across gene bodies. b. Pie chart showing the percentage of differentially enriched ChIP-seq peaks for H3K56ac (FDR < 0.05; n = 11251) in promoter, intronic, intergenic, and exonic regions comparing H3K36me3 (control) with H3K36me3+ (SETD2∆N-transduced) JHRCC12 cells.

Extended Data Fig. 10 Loss of SETD2-mediated H3K36me3 sensitizes cancer cells to KO of both ASF1A and ASF1B but not the inhibitor of p300/CBP, CCS1477.

a-b, JHRCC12 cells infected with control retrovirus or retrovirus expressing SETD2ΔN were subjected to lentiviral CRISPR/Cas9-mediated KO of ASF1A (exon 3), ASF1B, or both ASF1A (exon 3) and ASF1B and analyzed by the indicated immunoblots in a. Cell death was quantified by annexin-V staining (mean ± s.d., n = 3) in b. **, P < 0.01 (two-tailed unpaired Student’s t-test). c, JHRCC12 cells infected with control retrovirus or retrovirus expressing SETD2ΔN were treated with the p300/CBP inhibitor CCS1477 at the indicated concentrations. Cell death was quantified by annexin-V staining (mean ± s.d., n = 3).

Source data

Supplementary information

Reporting Summary

Supplementary Tables

Supplementary Tables 1–7.

Source data

Source Data Fig. 1

Statistical source data for Fig. 1.

Source Data Fig. 1

Unprocessed western blots for Fig. 1.

Source Data Fig. 2

Unprocessed western blots for Fig. 2.

Source Data Fig. 4

Statistical source data for Fig. 4.

Source Data Fig. 4

Unprocessed western blots for Fig. 4.

Source Data Fig. 5

Statistical source data for Fig. 5.

Source Data Fig. 5

Unprocessed western blots for Fig. 5.

Source Data Fig. 6

Statistical source data for Fig. 6.

Source Data Fig. 6

Unprocessed western blots for Fig. 6.

Source Data Extended Data Fig. 1

Statistical source data for Extended Data Fig. 1.

Source Data Extended Data Fig. 10

Statistical source data for Extended Data Fig. 10.

Source Data Extended Data Fig. 10

Unprocessed western blots for Extended Data Fig. 10.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xie, Y., Sahin, M., Sinha, S. et al. SETD2 loss perturbs the kidney cancer epigenetic landscape to promote metastasis and engenders actionable dependencies on histone chaperone complexes. Nat Cancer 3, 188–202 (2022). https://doi.org/10.1038/s43018-021-00316-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s43018-021-00316-3

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer