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:

Massively parallel high-order combinatorial genetics in human cells

Abstract

The systematic functional analysis of combinatorial genetics has been limited by the throughput that can be achieved and the order of complexity that can be studied. To enable massively parallel characterization of genetic combinations in human cells, we developed a technology for rapid, scalable assembly of high-order barcoded combinatorial genetic libraries that can be quantified with high-throughput sequencing. We applied this technology, combinatorial genetics en masse (CombiGEM), to create high-coverage libraries of 1,521 two-wise and 51,770 three-wise barcoded combinations of 39 human microRNA (miRNA) precursors. We identified miRNA combinations that synergistically sensitize drug-resistant cancer cells to chemotherapy and/or inhibit cancer cell proliferation, providing insights into complex miRNA networks. More broadly, our method will enable high-throughput profiling of multifactorial genetic combinations that regulate phenotypes of relevance to biomedicine, biotechnology and basic science.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Strategy for assembling combinatorial genetic libraries and performing combinatorial miRNA screens.
Figure 2: High-coverage combinatorial miRNA libraries can be efficiently generated and delivered to human cells.
Figure 3: Two-wise combinatorial screen reveals miRNA combinations that confer docetaxel resistance or sensitization in cancer cells.
Figure 4: Three-wise combinatorial screens identify miRNA combinations modifying docetaxel sensitivity or proliferation in cancer cells.
Figure 5: High-throughput profiling of miRNA combinations reveals genetic interactions for modulation of docetaxel sensitivity and/or cell proliferation phenotypes.
Figure 6: miRNA combinations can modulate both docetaxel sensitivity and cancer cell proliferation.

Similar content being viewed by others

References

  1. Dixon, S.J., Costanzo, M., Baryshnikova, A., Andrews, B. & Boone, C. Systematic mapping of genetic interaction networks. Annu. Rev. Genet. 43, 601–625 (2009).

    Article  PubMed  CAS  Google Scholar 

  2. Vierbuchen, T. & Wernig, M. Molecular roadblocks for cellular reprogramming. Mol. Cell 47, 827–838 (2012).

    Article  PubMed  CAS  Google Scholar 

  3. Al-Lazikani, B., Banerji, U. & Workman, P. Combinatorial drug therapy for cancer in the post-genomic era. Nat. Biotechnol. 30, 679–692 (2012).

    Article  PubMed  CAS  Google Scholar 

  4. Zuk, O., Hechter, E., Sunyaev, S.R. & Lander, E.S. The mystery of missing heritability: Genetic interactions create phantom heritability. Proc. Natl. Acad. Sci. USA 109, 1193–1198 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Eichler, E.E. et al. Missing heritability and strategies for finding the underlying causes of complex disease. Nat. Rev. Genet. 11, 446–450 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  6. Manolio, T.A. et al. Finding the missing heritability of complex diseases. Nature 461, 747–753 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Johannessen, C.M. et al. A melanocyte lineage program confers resistance to MAP kinase pathway inhibition. Nature 504, 138–142 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  8. Voorhoeve, P.M. et al. A genetic screen implicates miRNA-372 and miRNA-373 as oncogenes in testicular germ cell tumors. Cell 124, 1169–1181 (2006).

    Article  PubMed  CAS  Google Scholar 

  9. Moffat, J. & Sabatini, D.M. Building mammalian signaling pathways with RNAi screens. Nat. Rev. Mol. Cell Biol. 7, 177–187 (2006).

    Article  PubMed  CAS  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  11. Koike-Yusa, H., Li, Y., Tan, E.-P., Velasco-Herrera, M.D.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  PubMed  CAS  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  13. 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  PubMed  CAS  Google Scholar 

  14. Metzker, M.L. Sequencing technologies - the next generation. Nat. Rev. Genet. 11, 31–46 (2010).

    Article  PubMed  CAS  Google Scholar 

  15. Yu, H. et al. Next-generation sequencing to generate interactome datasets. Nat. Methods 8, 478–480 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Bassik, M.C. et al. A systematic mammalian genetic interaction map reveals pathways underlying ricin susceptibility. Cell 152, 909–922 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Engler, C., Kandzia, R. & Marillonnet, S. A one pot, one step, precision cloning method with high throughput capability. PLoS ONE 3, e3647 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Gibson, D.G. et al. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 6, 343–345 (2009).

    Article  PubMed  CAS  Google Scholar 

  19. Zelcbuch, L. et al. Spanning high-dimensional expression space using ribosome-binding site combinatorics. Nucleic Acids Res. 41, e98 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Yoo, A.S. et al. MicroRNA-mediated conversion of human fibroblasts to neurons. Nature 476, 228–231 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Brown, B.D. & Naldini, L. Exploiting and antagonizing microRNA regulation for therapeutic and experimental applications. Nat. Rev. Genet. 10, 578–585 (2009).

    Article  PubMed  CAS  Google Scholar 

  22. Honma, K. et al. RPN2 gene confers docetaxel resistance in breast cancer. Nat. Med. 14, 939–948 (2008).

    Article  PubMed  CAS  Google Scholar 

  23. Creighton, C.J. et al. Proteomic and transcriptomic profiling reveals a link between the PI3K pathway and lower estrogen-receptor (ER) levels and activity in ER+ breast cancer. Breast Cancer Res. 12, R40 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Blower, P.E. et al. MicroRNA expression profiles for the NCI-60 cancer cell panel. Mol. Cancer Ther. 6, 1483–1491 (2007).

    Article  PubMed  CAS  Google Scholar 

  25. Patnaik, S.K. et al. Expression of microRNAs in the NCI-60 cancer cell-lines. PLoS ONE 7, e49918 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Gholami, A.M. et al. Global proteome analysis of the NCI-60 cell line panel. Cell Reports 4, 609–620 (2013).

    Article  PubMed  CAS  Google Scholar 

  27. Hsu, S.-D. et al. MiRTarBase update 2014: an information resource for experimentally validated miRNA-target interactions. Nucleic Acids Res. 42, D78–D85 (2014).

    Article  PubMed  CAS  Google Scholar 

  28. Strezoska, Ž. et al. Optimized PCR conditions and increased shRNA fold representation improve reproducibility of pooled shRNA screens. PLoS ONE 7, e42341 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  29. Bhattacharya, R. et al. MiR-15a and MiR-16 control Bmi-1 expression in ovarian cancer. Cancer Res. 69, 9090–9095 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Cheng, A.A., Ding, H. & Lu, T.K. Enhanced killing of antibiotic-resistant bacteria enabled by massively parallel combinatorial genetics. Proc. Natl. Acad. Sci. USA 111, 12462–12467 (2014).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  31. Kampmann, M., Bassik, M.C. & Weissman, J.S. Integrated platform for genome-wide screening and construction of high-density genetic interaction maps in mammalian cells. Proc. Natl. Acad. Sci. USA 110, E2317–E2326 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Pierce, S.E., Davis, R.W., Nislow, C. & Giaever, G. Genome-wide analysis of barcoded Saccharomyces cerevisiae gene-deletion mutants in pooled cultures. Nat. Protoc. 2, 2958–2974 (2007).

    Article  PubMed  CAS  Google Scholar 

  33. Storey, J.D. & Tibshirani, R. Statistical significance for genomewide studies. Proc. Natl. Acad. Sci. USA 100, 9440–9445 (2003).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  34. Xia, L. et al. miR-15b and miR-16 modulate multidrug resistance by targeting BCL2 in human gastric cancer cells. Int. J. Cancer 123, 372–379 (2008).

    Article  PubMed  CAS  Google Scholar 

  35. Kastl, L., Brown, I. & Schofield, A.C. MiRNA-34a is associated with docetaxel resistance in human breast cancer cells. Breast Cancer Res. Treat. 131, 445–454 (2012).

    Article  PubMed  CAS  Google Scholar 

  36. Krek, A. et al. Combinatorial microRNA target predictions. Nat. Genet. 37, 495–500 (2005).

    Article  PubMed  CAS  Google Scholar 

  37. Jacobsen, A. et al. Analysis of microRNA-target interactions across diverse cancer types. Nat. Struct. Mol. Biol. 20, 1325–1332 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Lewis, B.P., Burge, C.B. & Bartel, D.P. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 120, 15–20 (2005).

    Article  PubMed  CAS  Google Scholar 

  39. Guillamot, M. et al. Cdc14b regulates mammalian RNA polymerase II and represses cell cycle transcription. Sci. Rep. 1, 189 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Huntzinger, E. & Izaurralde, E. Gene silencing by microRNAs: contributions of translational repression and mRNA decay. Nat. Rev. Genet. 12, 99–110 (2011).

    Article  PubMed  CAS  Google Scholar 

  41. McLornan, D.P., List, A. & Mufti, G.J. Applying synthetic lethality for the selective targeting of cancer. N. Engl. J. Med. 371, 1725–1735 (2014).

    Article  PubMed  CAS  Google Scholar 

  42. The Cancer Genome Atlas Research Network. et al. The Cancer Genome Atlas Pan-Cancer analysis project. Nat. Genet. 45, 1113–1120 (2013).

  43. Andersson, R. et al. An atlas of active enhancers across human cell types and tissues. Nature 507, 455–461 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. FANTOM Consortium and the RIKEN PMI and CLST (DGT). et al. A promoter-level mammalian expression atlas. Nature 507, 462–470 (2014).

  45. Gandhi, S. & Wood, N.W. Genome-wide association studies: the key to unlocking neurodegeneration? Nat. Neurosci. 13, 789–794 (2010).

    Article  PubMed  CAS  Google Scholar 

  46. Peden, J.F. & Farrall, M. Thirty-five common variants for coronary artery disease: the fruits of much collaborative labour. Hum. Mol. Genet. 20, R198–R205 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Gobeil, S., Zhu, X., Doillon, C.J. & Green, M.R. A genome-wide shRNA screen identifies GAS1 as a novel melanoma metastasis suppressor gene. Genes Dev. 22, 2932–2940 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Park, J. et al. RAS–MAPK–MSK1 pathway modulates ataxin 1 protein levels and toxicity in SCA1. Nature 498, 325–331 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  49. Chia, N.-Y. et al. A genome-wide RNAi screen reveals determinants of human embryonic stem cell identity. Nature 468, 316–320 (2010).

    Article  PubMed  CAS  Google Scholar 

  50. Ebert, M.S., Neilson, J.R. & Sharp, P.A. MicroRNA sponges: competitive inhibitors of small RNAs in mammalian cells. Nat. Methods 4, 721–726 (2007).

    Article  PubMed  CAS  Google Scholar 

  51. Guttman, M. et al. lincRNAs act in the circuitry controlling pluripotency and differentiation. Nature 477, 295–300 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  52. Gilbert, L.A. et al. Genome-scale CRISPR-mediated control of gene repression and activation. Cell 159, 647–661 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Hsu, P.D., Lander, E.S. & Zhang, F. Development and applications of CRISPR-Cas9 for genome engineering. Cell 157, 1262–1278 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  54. Gaj, T., Gersbach, C.A. & Barbas, C.F. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 31, 397–405 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Zhu, F. et al. DICE, an efficient system for iterative genomic editing in human pluripotent stem cells. Nucleic Acids Res. 42, e34 (2014).

    Article  PubMed  CAS  Google Scholar 

  56. Bruno, I.G. et al. Identification of a microRNA that activates gene expression by repressing nonsense-mediated RNA decay. Mol. Cell 42, 500–510 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Klein, M.E. et al. Homeostatic regulation of MeCP2 expression by a CREB-induced microRNA. Nat. Neurosci. 10, 1513–1514 (2007).

    Article  PubMed  CAS  Google Scholar 

  58. Miyamichi, K. et al. Cortical representations of olfactory input by trans-synaptic tracing. Nature 472, 191–196 (2011).

    Article  PubMed  CAS  Google Scholar 

  59. Ren, Y. et al. Targeted tumor-penetrating siRNA nanocomplexes for credentialing the ovarian cancer oncogene ID4. Sci. Transl. Med. 4, 147ra112 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Drees, B.L. et al. Derivation of genetic interaction networks from quantitative phenotype data. Genome Biol. 6, R38 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  61. Tong, A.H.Y. et al. Global mapping of the yeast genetic interaction network. Science 303, 808–813 (2004).

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

We thank members of the Lu laboratory and H. Ding for helpful discussions. We thank S. Patnaik of the Roswell Park Cancer Institute for processing miRNA expression data from the ArrayExpress database of European Bioinformatics Institute, the Massachusetts Institute of Technology BioMicroCenter for technical support on Illumina HiSeq, J. Weis for assisting computational analysis of next-generation sequencing data and C. Cui for technical assistance on cell viability assays. This work was supported by the US National Institutes of Health (DP2 OD008435 and P50 GM098792), the Office of Naval Research (N00014-13-1-0424), the Ellison Foundation New Scholar in Aging Award, the Defense Advanced Research Projects Agency and the Defense Threat Reduction Agency (HDTRA1-15-1-0050). A.S.L.W. was supported by the Croucher Foundation. The pAWp6 vector backbone (pFUGW-UBCp-GFP) was a gift from L. Nissim of the T.K. Lu laboratory, MIT; miR-128 was a gift from M.F. Wilkinson, University of California San Diego, USA and miR-132 was a gift from R.H. Goodman, Oregon Health and Science University, USA; HOSE 11-12 and HOSE 17-1 cells were gifts from G.S.W. Tsao, University of Hong Kong, Hong Kong; OVCAR8 and OVCAR8-ADR cells were gifts from S.N. Bhatia, MIT, and T. Ochiya, Japanese National Cancer Center Research Institute, Japan, respectively.

Author information

Authors and Affiliations

Authors

Contributions

A.S.L.W., G.C.G.C., A.A.C. and T.K.L. conceived the work. A.S.L.W. and G.C.G.C. performed experiments. A.S.L.W., G.C.G.C., A.A.C. and O.P. performed computational analyses on next-generation sequencing data. A.S.L.W., G.C.G.C. and T.K.L. designed the experiments, interpreted and analyzed the data. A.S.L.W. and T.K.L. wrote the paper.

Corresponding author

Correspondence to Timothy K Lu.

Ethics declarations

Competing interests

T.K.L., A.W. and G.C. have filed a provisional patent application (U.S. Provisional Application No.: 62/102,255) on this work.

Integrated supplementary information

Supplementary Figure 1 Lentiviral Delivery of Combinatorial miRNA Expression Constructs Provides Efficient Target Gene Repression.

a, Design for lentiviral combinatorial miRNA expression and sensor constructs. Single or multiple miRNA precursor sequences arranged in tandem were placed downstream of a GFP gene to monitor expression driven by a CMV promoter in a lentiviral vector. Sensors harboring four repeats of the miRNA target sequence were cloned in the 3’UTR of a RFP gene expressed from an UBC promoter to report on miRNA activity. The constructs were delivered by lentiviruses to HEK293T cells and then analyzed for GFP and RFP expression using flow cytometry. b, Repression of RFP reporter activity by miRNA expression. Lentiviral constructs harboring a miRNA, the cognate sensor, or both were introduced into HEK293T cells. c, Combinatorial miRNA expression constructs effectively repressed RFP reporters containing the cognate miRNA sensors. Lentiviral constructs harboring two-wise or three-wise miRNA combinations, with or without the cognate sensors, were introduced into HEK293T cells. d, Limited cross-reactivity between miRNAs and non-cognate sensors. Lentiviral constructs harboring miRNAs paired with different sensors were delivered into HEK293T cells. The percentages of RFP-positive cells within GFP-positive cell populations were determined by flow cytometry. Data represent mean ± s.d. (n = 3).

Supplementary Figure 2 Efficient Lentiviral Delivery of a Dual-Fluorescent Protein Reporter Construct in Human Cells.

a, Strategy for testing lentiviral delivery of a dual-fluorescent protein reporter construct in human cells. Lentiviruses generated for vectors containing a GFP gene expressed from a CMV or UBC promoter, or a single vector encoding RFP and GFP genes under the UBC and CMV promoters, respectively, were delivered to HEK293T cells for analysis of GFP and RFP expression. b, c, Lentiviral delivery and expression of the dual-fluorescent protein reporter construct in human cells. (b) Fluorescent microscopy revealed that RFP and GFP were expressed in UBCp-RFP-CMVp-GFP virus-infected cells, whereas only GFP was expressed in cells infected with UBCp-GFP and CMVp-GFP lentiviruses. Scale bar denotes 400 μm. (c) Flow cytometry was used to measure cell populations positive for RFP and GFP fluorescence. Over 97 percent of UBCp-RFP-CMVp-GFP virus-infected HEK293T cells were positive for both RFP and GFP, and similar percentages of UBCp-GFP or CMVp-GFP virus-infected cells were GFP-positive.

Supplementary Figure 3 Identification of the Exponential Phase During PCR for CombiGEM Barcode Amplification.

a, b, Procedures for identifying the transition point from exponential to linear phase during PCR for CombiGEM barcode amplification. The one-wise miRNA vector library pooled-assembled in E. coli (a) and the genomic DNA isolated from human breast cancer cells (MCF7) infected with the two-wise library (b) were used as templates in replicate PCR reactions, and the barcodes representing each miRNA combinations were amplified using primers targeting the sequences located outside the barcode region. PCR products were collected from the reactions stopped at cycles between 10 to 20 (a) or 19 to 28 (b), and were then diluted as templates for quantitative PCR reactions. The mean difference of threshold cycle (Ct) between cycles was determined. Error bars indicate s.d. from triplicates. Primer efficiencies were estimated to be 102% (a) and 100% (b) respectively. PCR cycle numbers highlighted in magenta (a) and green (b) were used in unbiased barcode amplification for subsequent Illumina sequencing. c, d, Agarose gel analyses of the amplified PCR products with indicated cycle numbers from (a) and (b) are shown in (c) and (d), respectively.

Supplementary Figure 4 High Reproducibility of Barcode Quantitation in Biological Replicates for Combinatorial miRNA Screens.

a, b, Scatter plots showing high correlation between barcode representations (log2 number of normalized barcode counts) between two biological replicates for both docetaxel (25 nM)-treated or vehicle-treated OVCAR8-ADR cells infected with the two-wise (a) or three-wise (b) miRNA combinatorial libraries. R is Pearson correlation coefficient.

Supplementary Figure 5 Consistent Fold Changes of Barcodes among Same miRNA Combinations Arranged in Different Orders in the Expression Constructs.

a-c, The coefficient of variation (CV; defined as s.d./mean of the fold changes of normalized barcode counts for docetaxel (25 nM)-treated versus four-day vehicle-treated (a, b) and four-day versus one-day cultured cells (c)) was determined for the same two-wise (a) or three-wise (b, c) combination arranged in different orders (i.e., [A,B] vs. [B,A] for two-wise combinations; [A,B,C], [A,C,B], [B,A,C], [B,C,A], [C,A,B], and [C,B,A] for three-wise combinations). 92% of two-wise miRNA combinations had a CV of <0.2 in the drug-sensitivity screen (a), while 95% and 98% of three-wise combinations showed a CV of <0.2 in the drug-sensitivity (b) and cell-proliferation (c) screens respectively.

Supplementary Figure 6 Docetaxel Dose-Response Curves for the OVCAR8 Cell Line and the Docetaxel-Resistant OVCAR8-ADR Cell Line.

OVCAR8 cells and OVCAR8-ADR cells (OVCAR8’s docetaxel-resistant derivative) cells were treated with docetaxel at indicated doses for three days and subjected to the MTT assay. Cell viabilities were compared to their respective no drug controls. The OVCAR8-ADR cell line has a ~3-fold higher IC50 than the parental OVCAR8 cell line. Data represent mean ± s.d. (n = 3).

Supplementary Figure 7 Log2 Fold-Changes in Barcode Representation Between Biological Replicates for All Individual Combinations in the Pooled Screens.

a, b (upper panel), Log2 fold-changes in biological replicate 1 is plotted against replicate 2 for mean values of normalized barcode counts for docetaxel (25 nM)-treated versus vehicle-treated OVCAR8-ADR cells (a), and for relative cell viability at day 4 versus day 1 (b), for each three-wise combination. a, b (lower panel), Distributions of differences in the log2 fold change between two biological replicates at a bin size of 0.1 are shown. A majority of three-wise combinations (88-90%) had <0.3 log2 fold-change differences. R is the Pearson correlation coefficient.

Supplementary Figure 8 Three-dimensional Plots Depicting the Docetaxel-Sensitizing (a) and Proliferation-Modulating (b) Effects of Three-Wise miRNA Combinations.

The log2 ratios of the normalized barcode counts for docetaxel-treated versus four-day vehicle-treated OVCAR8-ADR cells (a) or four-day versus one-day cultured cells (b) were determined for all three-wise miRNA combinations, and were presented as colored bubbles. miRNA combinations with drug-resistance and drug-sensitization effects (a) have log2 ratios >0 and <0, respectively, and those with pro-proliferation and anti-proliferation effects (b) have log2 ratios >0 and <0, respectively. Each two-dimensional plane was arranged in the same hierarchically clustered order as in (Fig. 5a-c), and the additional third miRNA element is labeled. All log2 ratios shown were determined from the mean of two biological replicates.

Supplementary Figure 9 Definitions of Genetic Interactions (GIs) in This Study.

Synergistic or buffering interactions have positive and negative GI scores, respectively, as described for case 1 to 7 in this figure. Positive and negative phenotypes have fold-changes of normalized barcode reads of >1 and <1 respectively, while no phenotype corresponds to a fold-change = 1. For miRNAs [A] and [B] with individual phenotypes “A” and “B”, the expected phenotype for the two-wise combination [A,B] is (“A” + “B” - 1) according to the additive model. Deviation was calculated by subtracting the expected phenotype from observed phenotype (i.e., Observed phenotype – Expected phenotype).

Supplementary Figure 10 GI Scores Between Biological Replicates for All Individual Combinations in the Pooled Screens.

a-c (upper panel), GI scores in biological replicate 1 are plotted against replicate 2 for docetaxel (25 nM)-treated versus vehicle-treated OVCAR8-ADR cells for each two-wise combination (a) and three-wise miRNA combination (b) respectively, and for relative cell viability at day 4 versus day 1 for each three-wise combination (c). a-c (lower panel), Distributions of differences in the GI scores between two biological replicates at a bin size of 0.1 are shown. A majority of combinations (80-93%) had GI score differences of <0.2.

Supplementary Figure 11 Synergistic Interactions between the miR-16-1/15a Cluster, miR-128b, and the let-7e/miR-99b Cluster Modulate Cell Proliferation Phenotypes.

a, GI scores for a given three-wise miRNA combination [A,B,C] are plotted and compared to other combinations that harbor two of the same miRNAs (in the first two positions) and every other miRNA member in our library (denoted as X in the third position). GI scores of each three-wise miRNA combination were calculated as described in Online Methods , and represent the interaction between the additional third miRNA and the two-wise miRNA combinations. GI scores were determined for the three possible permutations (i.e., [A,B,C], [B,C,A], and [A,C,B], see Online Methods ). In this example, A, B, and C represent the miR-16-1/15a cluster, miR-128b, and the let-7e/miR-99b cluster respectively, and X represents all 39 library members. b-d, GI maps based on cell-proliferation phenotypes for all three-wise miRNA combinations harboring the miR-16-1/15a cluster, miR-128b, and/or the let-7e/miR-99b cluster are displayed in the same order as for the GI map in Fig. 5a for easy comparison, with the additional third miRNA element in the three-wise combination being labeled above each GI map. The combinations for which no GIs were measured are indicated in gray. The combinations with GI at a Q-value of <0.003 are boxed.

Supplementary Figure 12 Three-Wise miRNA Combinations Display Distinct Docetaxel Sensitivity and Anti-Proliferation Phenotypes.

Fold changes of normalized barcode counts for docetaxel (25 nM)-treated versus vehicle-treated OVCAR8-ADR cells (y-axis) and also for four-day versus one-day cultured cells (x-axis) were plotted for all three-wise miRNA combinations. Each data point represents the mean of two biological replicates.

Supplementary Figure 13 Combinatorial Expression of the miR-16-1/15a Cluster, miR-128b, and the let-7e/miR-99b Cluster Inhibits Colony Formation by Viable OVCAR8-ADR Cells.

a, b, ~10,000 OVCAR8-ADR cells infected with each indicated miRNA combinations were treated with 25 nM of docetaxel for three days, and were cultured for another eleven days. Cells were stained with crystal violet to visualize colony formation for quantification. Representative images are shown in (a). The number of colonies for each sample was determined (b). The maximum number of discrete colonies that could be reliably counted was ~500 per well, and thus, samples with more than 500 colonies are presented as >500 colonies. Data represent mean ± s.d. (n = 3; *P < 0.05).

Supplementary Figure 14 High Consistency between Pooled Screens and Validation Data for Individual Hits.

For each two-wise and three-wise miRNA combination, the fold-change in the normalized barcode count for docetaxel (25 nM)-treated versus vehicle-treated OVCAR8-ADR cells, or four-day versus one-day cultured cells, obtained from the pooled screening data (‘Screen phenotype’) was plotted against its relative cell viability compared to vector control determined from the individual drug-sensitivity or cell-proliferation assays, respectively (‘Validation phenotype’) (R = 0.899). Data for the screening data are the mean of two biological replicates while the individual validation data represent the mean of three independent experiments. R is the Pearson correlation coefficient.

Supplementary Figure 15 Combinatorial Expression of the miR-16/15a Cluster, miR-128b, and/or the let-7e/miR-99b Cluster Reduce mRNA Levels of Targeted Genes in OVCAR8-ADR Cells.

a, qRT-PCR quantification of relative mRNA levels in OVCAR8-ADR cells expressing the miR-16/15a cluster or co-expressing the miR-16/15a cluster, miR-128b, and the let-7e/miR-99b cluster. Measured mRNA levels were normalized to GAPDH mRNA levels, and data represent mean ± s.d. (n = 3). mRNAs that were predicted or validated to contain conserved sites matching the seed region of the corresponding miRNAs using TargetScan and miRTarBase are shaded in orange in the table below the graph. Significant differences in the mRNA levels of CCND1, CCND3, CCNE1 and CHEK1 in cells expressing the miR-16/15a cluster (orange #; P < 0.05) or co-expressing the miR-16/15a cluster, miR-128b, and the let-7e/miR-99b cluster (purple #; P < 0.05) were determined by comparison with vector-control-infected cells. The asterisk (*P < 0.05) represents a significant difference in mRNA levels between cells expressing the miR-16/15a cluster versus co-expressing the miR-16/15a cluster, miR-128b, and the let-7e/miR-99b cluster. b, Relative mRNA levels of CDC14B in cells expressing various combinations of the miR-16/15a cluster, miR-128b, and/or the let-7e/miR-99b cluster. The mRNA level of CDC14B was significantly reduced in cells co-expressing the let-7e/miR-99b cluster and miR-128b, or the triple-combination expressing the miR-16/15a cluster, miR-128b, and the let-7e/miR-99b cluster. Data represent mean ± s.d. (n = 9; *P < 0.05). c, Summary diagram illustrating the potential roles of the miR-16/15a cluster, miR-128b, and the let-7e/miR-99b cluster in regulating the mRNA levels of multiple downstream targets. These genes may represent targets for future investigation into mechanisms that can modulate docetaxel resistance and/or proliferation phenotypes in OVCAR8-ADR cells.

Supplementary information

Supplementary Figures

Supplementary Figures 1-15 (PDF 2714 kb)

Supplementary Text and Figures

Supplementary Figures 16, 17 and legends; Supplementary Tables 1, 2, 4–14 (PDF 6798 kb)

Supplementary Table 3 (XLSX 225 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wong, A., Choi, G., Cheng, A. et al. Massively parallel high-order combinatorial genetics in human cells. Nat Biotechnol 33, 952–961 (2015). https://doi.org/10.1038/nbt.3326

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nbt.3326

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