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.

  • Letter
  • Published:

Effective breast cancer combination therapy targeting BACH1 and mitochondrial metabolism

Abstract

Mitochondrial metabolism is an attractive target for cancer therapy1,2. Reprogramming metabolic pathways could improve the ability of metabolic inhibitors to suppress cancers with limited treatment options, such as triple-negative breast cancer (TNBC)1,3. Here we show that BTB and CNC homology1 (BACH1)4, a haem-binding transcription factor that is increased in expression in tumours from patients with TNBC, targets mitochondrial metabolism. BACH1 decreases glucose utilization in the tricarboxylic acid cycle and negatively regulates transcription of electron transport chain (ETC) genes. BACH1 depletion by shRNA or degradation by hemin sensitizes cells to ETC inhibitors such as metformin5,6, suppressing growth of both cell line and patient-derived tumour xenografts. Expression of a haem-resistant BACH1 mutant in cells that express a short hairpin RNA for BACH1 rescues the BACH1 phenotype and restores metformin resistance in hemin-treated cells and tumours7. Finally, BACH1 gene expression inversely correlates with ETC gene expression in tumours from patients with breast cancer and in other tumour types, which highlights the clinical relevance of our findings. This study demonstrates that mitochondrial metabolism can be exploited by targeting BACH1 to sensitize breast cancer and potentially other tumour tissues to mitochondrial inhibitors.

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

Fig. 1: BACH1 inhibits mitochondrial genes in TNBC.
Fig. 2: BACH1 regulates mitochondrial metabolism.
Fig. 3: Hemin mimics shBACH1 through BACH1 degradation in TNBC.
Fig. 4: Combination treatment with hemin and metformin suppresses tumour growth through BACH1 inhibition.

Similar content being viewed by others

Data availability

All data are available from the authors upon reasonable request. Additional material including source data is available online.

References

  1. Vander Heiden, M. G. & DeBerardinis, R. J. Understanding the intersections between metabolism and cancer biology. Cell 168, 657–669 (2017).

    Article  Google Scholar 

  2. Weinberg, S. E. & Chandel, N. S. Targeting mitochondria metabolism for cancer therapy. Nat. Chem. Biol. 11, 9–15 (2015).

    Article  CAS  Google Scholar 

  3. DeBerardinis, R. J. & Chandel, N. S. Fundamentals of cancer metabolism. Sci. Adv. 2, e1600200 (2016).

    Article  ADS  Google Scholar 

  4. Oyake, T. et al. Bach proteins belong to a novel family of BTB-basic leucine zipper transcription factors that interact with MafK and regulate transcription through the NF-E2 site. Mol. Cell. Biol. 16, 6083–6095 (1996).

    Article  CAS  Google Scholar 

  5. Wheaton, W. W. et al. Metformin inhibits mitochondrial complex I of cancer cells to reduce tumorigenesis. eLife 3, e02242 (2014).

    Article  Google Scholar 

  6. Birsoy, K. et al. Metabolic determinants of cancer cell sensitivity to glucose limitation and biguanides. Nature 508, 108–112 (2014).

    Article  ADS  CAS  Google Scholar 

  7. Ogawa, K. et al. Heme mediates derepression of Maf recognition element through direct binding to transcription repressor Bach1. EMBO J. 20, 2835–2843 (2001).

    Article  CAS  Google Scholar 

  8. Bianchini, G., Balko, J. M., Mayer, I. A., Sanders, M. E. & Gianni, L. Triple-negative breast cancer: challenges and opportunities of a heterogeneous disease. Nat. Rev. Clin. Oncol. 13, 674–690 (2016).

    Article  CAS  Google Scholar 

  9. Dangi-Garimella, S. et al. Raf kinase inhibitory protein suppresses a metastasis signalling cascade involving LIN28 and let-7. EMBO J. 28, 347–358 (2009).

    Article  CAS  Google Scholar 

  10. Yun, J. et al. Signalling pathway for RKIP and Let-7 regulates and predicts metastatic breast cancer. EMBO J. 30, 4500–4514 (2011).

    Article  CAS  Google Scholar 

  11. Lee, U. et al. A prognostic gene signature for metastasis-free survival of triple negative breast cancer patients. PLoS ONE 8, e82125 (2013).

    Article  ADS  Google Scholar 

  12. Liang, Y. et al. Transcriptional network analysis identifies BACH1 as a master regulator of breast cancer bone metastasis. J. Biol. Chem. 287, 33533–33544 (2012).

    Article  CAS  Google Scholar 

  13. Sun, J. et al. Hemoprotein Bach1 regulates enhancer availability of heme oxygenase-1 gene. EMBO J. 21, 5216–5224 (2002).

    Article  CAS  Google Scholar 

  14. Cancer Genome Atlas Network. Comprehensive molecular portraits of human breast tumours. Nature 490, 61–70 (2012).

    Article  ADS  Google Scholar 

  15. Curtis, C. et al. The genomic and transcriptomic architecture of 2,000 breast tumours reveals novel subgroups. Nature 486, 346–352 (2012).

    Article  CAS  Google Scholar 

  16. Wang, Y. et al. Gene-expression profiles to predict distant metastasis of lymph-node-negative primary breast cancer. Lancet 365, 671–679 (2005).

    Article  CAS  Google Scholar 

  17. Schmidt, M. et al. The humoral immune system has a key prognostic impact in node-negative breast cancer. Cancer Res. 68, 5405–5413 (2008).

    Article  CAS  Google Scholar 

  18. Minn, A. J. et al. Genes that mediate breast cancer metastasis to lung. Nature 436, 518–524 (2005).

    Article  ADS  CAS  Google Scholar 

  19. Igarashi, K. et al. Multivalent DNA binding complex generated by small Maf and Bach1 as a possible biochemical basis for β-globin locus control region complex. J. Biol. Chem. 273, 11783–11790 (1998).

    Article  CAS  Google Scholar 

  20. Lee, J. et al. Network of mutually repressive metastasis regulators can promote cell heterogeneity and metastatic transitions. Proc. Natl Acad. Sci. USA 111, E364–E373 (2014).

    Article  CAS  Google Scholar 

  21. Barski, A. et al. High-resolution profiling of histone methylations in the human genome. Cell 129, 823–837 (2007).

    Article  CAS  Google Scholar 

  22. Liu, X., Romero, I. L., Litchfield, L. M., Lengyel, E. & Locasale, J. W. Metformin targets central carbon metabolism and reveals mitochondrial requirements in human cancers. Cell Metab. 24, 728–739 (2016).

    Article  CAS  Google Scholar 

  23. Schell, J. C. & Rutter, J. The long and winding road to the mitochondrial pyruvate carrier. Cancer Metab. 1, 6 (2013).

    Article  Google Scholar 

  24. Kim, J. W., Tchernyshyov, I., Semenza, G. L. & Dang, C. V. HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metab. 3, 177–185 (2006).

    Article  Google Scholar 

  25. Jitrapakdee, S., Vidal-Puig, A. & Wallace, J. C. Anaplerotic roles of pyruvate carboxylase in mammalian tissues. Cell. Mol. Life Sci. 63, 843–854 (2006).

    Article  CAS  Google Scholar 

  26. Gui, D. Y. et al. Environment dictates dependence on mitochondrial complex I for NAD+ and aspartate production and determines cancer cell sensitivity to metformin. Cell Metab. 24, 716–727 (2016).

    Article  CAS  Google Scholar 

  27. Murai, M. & Miyoshi, H. Current topics on inhibitors of respiratory complex I. Biochim. Biophys. Acta Bioenerg. 1857, 884–891 (2016).

    Article  CAS  Google Scholar 

  28. Dowling, R. J. O. et al. Metformin pharmacokinetics in mouse tumors: implications for human therapy. Cell Metab. 23, 567–568 (2016).

    Article  CAS  Google Scholar 

  29. Chandel, N. S. et al. Are metformin doses used in murine cancer models clinically relevant? Cell Metab. 23, 569–570 (2016).

    Article  CAS  Google Scholar 

  30. Klil-Drori, A. J., Azoulay, L. & Pollak, M. N. Cancer, obesity, diabetes, and antidiabetic drugs: is the fog clearing? Nat. Rev. Clin. Oncol. 14, 85–99 (2017).

    Article  CAS  Google Scholar 

  31. Emami Riedmaier, A., Fisel, P., Nies, A. T., Schaeffeler, E. & Schwab, M. Metformin and cancer: from the old medicine cabinet to pharmacological pitfalls and prospects. Trends Pharmacol. Sci. 34, 126–135 (2013).

    Article  Google Scholar 

  32. Liang, H. & Ward, W. F. PGC-1α: a key regulator of energy metabolism. Adv. Physiol. Educ. 30, 145–151 (2006).

    Article  Google Scholar 

  33. Zenke-Kawasaki, Y. et al. Heme induces ubiquitination and degradation of the transcription factor Bach1. Mol. Cell. Biol. 27, 6962–6971 (2007).

    Article  CAS  Google Scholar 

  34. Dhar, G. J., Bossenmaier, I., Petryka, Z. J., Cardinal, R. & Watson, C. J. Effects of hematin in hepatic porphyria. Further studies. Ann. Intern. Med. 83, 20–30 (1975).

    Article  CAS  Google Scholar 

  35. Suzuki, H. et al. Heme regulates gene expression by triggering Crm1-dependent nuclear export of Bach1. EMBO J. 23, 2544–2553 (2004).

    Article  CAS  Google Scholar 

  36. Zhang, X. et al. A renewable tissue resource of phenotypically stable, biologically and ethnically diverse, patient-derived human breast cancer xenograft models. Cancer Res. 73, 4885–4897 (2013).

    Article  CAS  Google Scholar 

  37. Efron, B. & Tibshirani, R. On testing the significance of sets of genes. Ann. Appl. Stat. 1, 107–129 (2007).

    Article  MathSciNet  Google Scholar 

  38. Cerami, E. et al. The cBio cancer genomics portal: an open platform for exploring multidimensional cancer genomics data. Cancer Discov. 2, 401–404 (2012).

    Article  Google Scholar 

  39. Gao, J. et al. Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Sci. Signal. 6, pl1 (2013).

    Article  Google Scholar 

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

    Article  Google Scholar 

  41. Young, M. D., Wakefield, M. J., Smyth, G. K. & Oshlack, A. Gene ontology analysis for RNA-seq: accounting for selection bias. Genome Biol. 11, R14 (2010).

    Article  Google Scholar 

  42. Wan, Q. et al. BioXpress: an integrated RNA-seq-derived gene expression database for pan-cancer analysis. Database 2015, bav019 (2015).

    Article  Google Scholar 

  43. Wu, J. et al. BCIP: a gene-centered platform for identifying potential regulatory genes in breast cancer. Sci. Rep. 7, 45235 (2017).

    Article  ADS  CAS  Google Scholar 

  44. Louie, S. M. et al. GSTP1 Is a driver of triple-negative breast cancer cell metabolism and pathogenicity. Cell Chem. Biol. 23, 567–578 (2016).

    Article  CAS  Google Scholar 

  45. Benjamin, D. I. et al. Ether lipid generating enzyme AGPS alters the balance of structural and signaling lipids to fuel cancer pathogenicity. Proc. Natl Acad. Sci. USA 110, 14912–14917 (2013).

    Article  ADS  CAS  Google Scholar 

  46. Gao, X. et al. Serine availability influences mitochondrial dynamics and function through lipid metabolism. Cell Reports 22, 3507–3520 (2018).

    Article  CAS  Google Scholar 

  47. Liu, J. & Locasale, J. W. Metabolomics: A primer. Trends Biochem. Sci. 42, 274–284 (2017).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank G. Greene and members of his laboratory for sharing the PDX tumors. This study was supported by NIH R01CA184494 (M.R.R.), R01GM121735 (M.R.R.), NIH R01CA172667 (D.K.N.), DoD Breakthrough Breast Cancer BC161588 (J. Lee), NIH 1R01AI131267 (M.G.B.), 1R56ES028149 (M.G.B.) and NIH R01CA193256 (J.W.L.). We also thank G. Balazsi and members of the Rosner laboratory for careful reading of the manuscript.

Reviewer information

Nature thanks Matthew J. Ellis, Kazuhiko Igarashi and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

J. Lee and M.R.R. contributed overall experimental design, J. Lee. performed most experiments, A.E.Y. contributed bioinformatics analyses, J.P.W. contributed to BACH1(mut) experiments, C.F. contributed to gene array analyses, P.C.H., C.K. and M.G.B. contributed Seahorse experiments, E.A.G. and D.K.N. contributed mass spectrometry analysis of metabolites, J. Liu, S.M.S. and J.W.L. contributed tracing analysis, J.Y., F.D.R. and P.T. contributed to cell culture or xenograft experiments, M.E. and D.C.R. contributed to PDX experiments, J.A. contributed biostatistics analyses, M.R.R. and J. Lee supervised the project, analysed data and wrote the manuscript with assistance from J.W.L. All authors reviewed the manuscript.

Corresponding author

Correspondence to Marsha Rich Rosner.

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.

Extended data figures and tables

Extended Data Fig. 1 BACH1 expression is high in patients with TNBC and suppresses expression of ETC genes at their promoter.

a, Left, BACH1 expression levels (determined by RNA-seq) with respect to relative DNA copy-number alterations in TCGA breast cancers (n = 1105). Middle, BACH1 expression (RNA-seq) in TNBC (n = 83) or basal (n = 98) breast cancers compared to non-TNBC (n = 734) or non-basal (n = 424) breast cancers using Pam50 classification of TCGA data. Right, breast cancer subtypes classified by Pam50 (n = 522 total, n = 98 basal, n = 58 HER2-enriched, n = 231 luminal-A, n = 127 luminal-B, n = 8 normal-like). Two-tailed t-test. b, BACH1 expression levels (by RNA-seq) in patients with TNBC compared to patients that did not have TNBC, using the datasets of patients with breast cancer of METABRIC (n = 2509), GSE2034 (n = 286) and GSE11121 (n = 200). Two-tailed t-test. c, Gene Ontology terms as determined by gene set analysis for cell components that are positively correlated with BACH1 depletion based on microarray analysis of BM1-shBACH1 cell transcripts. n = 3 biologically independent samples, FDR-corrected P < 0.05. d, Left, relative mRNA levels of mitochondrial inner membrane genes in MB436-shBACH1 cells (two shBACH1 vectors, clone 1, clone 2) compared to the wild type control (MB436-shCont). Data are mean ± s.e.m., n = 3 biologically independent samples, two-tailed t-test. Right, representative western blots of mitochondrial genes using MB436-shBACH1 or control cell lysates. Each experiment was repeated independently three times with similar results. Band density quantification is shown below the blots. e, Schematic showing proximal BACH1 binding on the promoter regions of mitochondrial membrane genes. TSS, transcription start site. Arrows, primers used for ChIP-PCR. f, ChIP assays showing relative fold enrichment of BACH1 recruitment to the HMOX1 promoter using BACH1-depleted TNBC (BM1 and MB436) or control cells. g, h, ChIP assays showing fold enrichment of BACH1 and H3K27me3 recruitment to the mitochondrial membrane genes in low-BACH1-expressing MB468 and MB436 cells. For ChIP assays in fh, data are mean ± s.e.m., n = 3 biologically independent samples, two-tailed t-test. i, KEGG pathways demonstrating the negative correlation between BACH1 expression and oxidative phosphorylation in all patients with breast cancer (n = 1,105, left) and patients with TNBC (n = 119, right). FDR values (−log10(FDR)) are generated in the R package GOseq using the default program Wallenius P values with Benjamini−Hochberg-corrected P values. j, Expression of ETC genes (COX15, ATP5D and ATP5G2 (also known as ATP5MC2)) in TNBC compared to tumours from patients that did not have TNBC using multiple breast cancer datasets: METABRIC (TNBC n = 319, non-TNBC n = 1661), GSE2034 (TNBC n = 54, non-TNBC n = 232) and GSE11121 (TNBC n = 33, non-TNBC n = 150). P values are determined by two-tailed t-test. k, Co-expression plots of UQCRC1 or ATP5D and BACH1 in TCGA breast cancer (n = 1105) or TNBC (n = 115) dataset. Pearson’s and Spearman’s correlation coefficients are shown.

Source data

Extended Data Fig. 2 BACH1 depletion increases mitochondrial metabolism.

a, Measurement of OCR and ECAR in BM1 or MB436 cells expressing control or shBACH1. Data are mean ± s.e.m., n = 6 biologically independent samples, unpaired two-tailed t-test. b, Relative abundance of steady-state metabolites in BM1-shBACH1 or control cells cultured with DMEM (glucose, 10 mM) measured by mass spectrometry. Pyr, pyruvate; Lac, lactate. Data are mean ± s.e.m., n = 5 biologically independent samples, two-tailed t-test. c, d, Fractional isotopic incorporation of [U-13C6]-glucose (c) or [U-13C5]-glutamine (d) into the metabolites in glycolysis and the TCA cycle are shown. Data are mean ± s.e.m., n = 4 biologically independent samples, two-tailed t-test. M indicates number of carbons labelled. Fraction is ratio of isotopologues to sum of all isotopologues. e, Relative mRNA and protein levels of PDH and PDK genes in MB436-shBACH1 cells compared to controls. qRT–PCR data are mean ± s.e.m., n = 3 biologically independent samples, two-tailed t-test. Representative images of western blots are shown. Band density quantification is shown below the blots. Each experiment was repeated independently three times with similar results. f, ChIP assays showing fold enrichment of BACH1 recruitment to promoters of PDK genes using MB436 and MB468 cells. Data are mean ± s.e.m., n = 3 biological replicates per cell line, two-tailed t-test. g, Relative mRNA levels of pyruvate carboxylase (PC) in shBACH1 cells compared to control. Data are mean ± s.e.m., n = 3 biologically independent samples. NS, not significant by two-tailed student’s t-test.

Source data

Extended Data Fig. 3 BACH1 levels determine response to ETC inhibitor treatment in breast cancer cells.

a, Cellular growth (per cent confluency) of BACH1-depleted cells (BM1-shBACH1 or MB436-shBACH1) or their controls treated with vehicle (veh), metformin (met), rotenone (rot) or antimycin A (ant). b, c, Relative cell viability (%) of BACH1-depleted cells (BM1-shBACH1 or MB436-shBACH1) or their controls treated with vehicle, metformin, rotenone or antimycin A. d, e, Cellular growth (per cent confluency) (d) or cell viability (%) (e) of low-BACH1 (MB468), medium-BACH1 (MB436) or high-BACH1 (BM1)-expressing TNBC cells treated with vehicle (control), 1 mM metformin, or 1–10 mM metformin. f, Cell viability (%) of non-malignant mammary epithelial cells (MCF10A and 184A1) treated with vehicle, metformin, rotenone or antimycin A. For cell viability and growth assays in a and d, values are mean ± s.e.m., n = 6 biologically independent samples, unpaired two-tailed t-test. Arrow indicates the time at which inhibitors were added. For cell viability assays in b, c, e and f, cells were incubated for 48 h after addition of inhibitors and stained with CaAM for 1 h.

Source data

Extended Data Fig. 4 Rescue of BACH1-depleted TNBC cells from metformin treatment.

a, Cellular growth (per cent confluency) of BM1-shBACH1 or control cells treated with vehicle or metformin in growth medium containing glucose (1 mM) and supplemented with or without pyruvate (2.5 mM). b, Relative NAD+/NADH ratios in BACH1-depleted BM1 cells treated with pyruvate (2.5 mM) and/or metformin (5 mM) for 24 h. Data are mean ± s.e.m., n = 3 biologically independent samples, two-tailed t-test. c, Representative western blots of COX15, UQCRC1 and α-tubulin using BM1-shBACH1 cell lysates transfected with siCOX15 (150 nM), siUQCRC1 (150 nM), and siControl (150 nM). Each experiment was repeated independently two times with similar results. d, Cellular growth (per cent confluence) of BM1-shBACH1 cells transfected with siRNA for COX15 and/or UQCRC1 and treated with vehicle or metformin (10 mM). For a and d, data are mean ± s.e.m., n = 6 biologically independent samples, unpaired two-tailed t-test between vehicle-treated and metformin (10 mM)-treated group. e, Relative OCT1 (also known as SLC22A1, left), PPARG and PGC1A (also known as PPARGC1A, right) mRNA levels in BM1-shBACH1 cells. Data are mean ± s.e.m., n = 3 biologically independent samples.

Source data

Extended Data Fig. 5 Hemin treatment of cells expressing wild-type BACH1 or hemin-resistant BACH1(mut).

a, Left, cellular growth (per cent confluence) of BM1 and MB436 cells treated with hemin (10, 20, 40 or 80 μM) as indicated. Right, representative western blots of BACH1 from MB436 cells after treatment with hemin (10–40 μM) for 24 h (see also Fig. 3d). Each experiment was repeated independently three times with similar results. b, Relative mRNA levels of mitochondrial membrane genes in MB436 cells treated with vehicle or hemin (20 μM) for 48 h. and representative western blots. Data are mean ± s.e.m., n = 3 biologically independent samples, two-tailed t-test. Band density quantification is shown below the blots. c, Measurement of OCR or ECAR of BM1 cells treated with vehicle or hemin (full data from Fig. 3b). d, Cell viability (%) of BM1 cells treated with vehicle, hemin (20 μM) or ETC inhibitors (metformin, rotenone or antimycin A) for 48 h. e, Cell viability (%) or cell growth (per cent confluence) of MB436 cells treated with vehicle, hemin (20 μM) or metformin (1 mM). f, Representative western blots from MB436-shBACH1 cells transiently transfected with Bach1mut (100 ng) and treated with vehicle or hemin (10, 20, 40 or 80 μM) for 48 h. Each experiment was repeated independently three times with similar results. g, Measurement of OCR in BM1 or MB436 cells stably expressing shControl, shBACH1 or shBACH1 + Bach1mut vectors. h, Relative mRNA levels of mitochondrial genes in BM1-shBACH1 cells, shCont cells, or BM1-shBACH1 cells transfected with BACH1(mut). Data are mean ± s.e.m., n = 3 biologically independent samples, two-tailed t-test. i, Left, cell viability (%) of BM1-shBACH1 cells transfected with BACH1(mut) and then treated with hemin (20 μM) or vehicle for 48 h. Right, representative western blots showing BACH1(mut) from cells treated with vehicle or hemin. j, Measurement of OCR and ECAR of BM1-shBACH1 cells expressing BACH1(mut) pre-treated with hemin. Conditions for OCR and ECAR, and statistics are the same as in Extended Data Fig. 2a. Data are mean ± s.e.m., n = 6 biologically independent samples. k, Cell viability (%) of MB436 cells stably expressing shRNA-resistant BACH1(WT), BACH1(mut) or shCont vectors treated with vehicle, hemin (20 μM) or metformin (5 mM) for 48 h. Representative western blots of BACH1 expression are shown. Each experiment was repeated independently three times with similar results. For growth and viability assays in a, d, e, i, j and k, data are mean ± s.e.m., n = 4 biologically independent samples), unpaired two-tailed t-test.

Source data

Extended Data Fig. 6 Metformin suppresses growth of BACH1-depleted breast tumours.

a, Tumour weights and volumes of mice injected with MB436-shBACH1 or control cells (left, n = 6–7 per group) or BM1-shBACH1 or control cells (right, n = 8–10 per group) and treated with vehicle or metformin. Data are mean ± s.e.m., unpaired two-tailed t-test. b, Tumour images of representative mice in each treatment group of mouse models. Scale bar, 1 cm. c, Primary tumour (%) indicates the ratio of mice with tumours or tumour-free upon metformin treatment compared to the total number of mice per treatment group at the end of experiment. d, Body weights of mice monitored before and after treatment of hemin and metformin. Arrow indicates initiation of hemin (H) or metformin (M) treatment. e, Relative mRNA expression of PDK and PDH mRNAs in tumours from MB436-shBACH1 xenograft mice by qRT–PCR. Data are mean ± s.e.m., n = 2 per group. f, Representative western blots of total PDH, BACH1 and mitochondrial membrane proteins (COX15, SLC25A15, NDUFA9) using MB436-shBACH1 or control tumour lysates. Each experiment was repeated independently three times with similar results. g, Lung metastases from mice with MB436-shBACH1 or control xenograft tumours. Lung tissues sectioned and H & E-stained to visualize and count lung metastases in mice. n = 5 mice per group. Data are mean ± s.e.m., two-tailed unpaired t-test. h, Representative lung metastasis images. Arrow indicates tumour metastases with a scale bar (1000 μm).

Source data

Extended Data Fig. 7 Combination treatment using hemin and metformin suppresses growth of tumours through BACH1 in multiple TNBC mouse models.

a, Left, monitoring of BACH1 degradation by hemin treatment assayed by western blotting using tumour lysates. Mice (n = 2 per treatment group) injected with BM1 cells (2 × 106 cells) for 4 weeks to form tumours were treated with 25 mg kg−1 or 50 mg kg−1 hemin for the indicated times. Right, representative western blots showing relative BACH1 expression using tumours from mice treated with hemin for the experiments (see Fig. 4a–d). Western blotting experiments were repeated at least twice with similar results. b, Schematic depicting experimental plans with time line for cancer cell injection, hemin treatment (50 mg kg−1 day−1) or metformin treatment (200–300 mg kg−1 day−1) using TNBC mouse models. c, Representative western blots showing relative BACH1 expression from xenograft tumour models using tumours derived from BM1 cells, MB436 cells or two independent patients (n = 2 biologically independent samples). Western blotting experiments were repeated twice with similar results. d, Relative tumour volumes of BM1 or PDX (no. 4195) mouse xenograft monitored weekly during treated with vehicle, hemin or metformin. Tumour volume data are mean ± s.e.m., two-way ANOVA with multiple comparisons. BM1 tumours (vehicle n = 10, hemin n = 10, metformin n = 10, hemin + metformin n = 5) or PDX tumours (vehicle n = 9, hemin n = 10, metformin n = 8, hemin + metformin n = 7). e, Tumour weights, collected and measured at the end of the treatment using hemin and metformin of MB436, BM1-xenograft, or two PDX models (no. 2147 and no. 4195). Data are mean ± s.e.m. with P values using unpaired two-tailed t-test. f, Representative tumour images from each treatment group of MB436, BM1-xenograft or two PDX models are shown. Scale bar. 1 cm. g, Representative western blots of BACH1 using tumour lysates from mice xenografted with BM1-shCont, BM1-shBACH1 expressing BACH1(mut) or BM1-shBACH1 expressing wild-type BACH1. h, Tumour weights from mice xenografted with MB436-shBACH1 cells expressing BACH1(mut) and treated with vehicle, hemin or metformin. Data are mean ± s.e.m., two-tailed t-test. i, j, Tumour growth of BM1 BACH1(mut) (vehicle n = 5, hemin n = 4, metformin n = 5, hemin + metformin n = 5) or wild-type BACH1 xenografts (vehicle n = 5, hemin n = 4, metformin n = 4, hemin + metformin n = 5) treated with vehicle or hemin and representative tumour images from each treatment group of mice. Scale bar, 1 cm. Representative western blots showing BACH1 expression in multiple mouse tumour lysates.

Source data

Extended Data Fig. 8 BACH1 expression in multiple cancer types.

a, Distribution of BACH1 expression in TNBC. Clinical and RNA-seq data associated with the TCGA cohort of patients with breast cancer were accessed at https://www.cbioportal.org/. Out of all provisional cases (n = 1,105), breast cancer samples (n = 914) that had clinical information regarding receptor status of ER, PR (also known as PGR) and HER2 (also known as ERBB2) based on immunohistochemistry analysis as well as RNA-seq data for BACH1-related genes were analysed. The TNBC subgroup among these 914 samples were identified as samples that are negative for all three receptors (n = 115). If the immunohistochemistry results were positive, indeterminate or equivocal for any of the three receptors, those samples were grouped in non-TNBC (n = 799). BACH1 status of the samples were based on an arbitrary 0.5 cut-off for the z-score transformed RNA-seq expression values for the BACH1 gene (245 BACH1-high cases with z-score >0.5; 301 BACH1-low cases with z-score <0.5). b, Frequency (%) of patient tumours with overexpression of BACH1 compared to their matched normal tissues across multiple TCGA cancer types. Numbers of patients relative to healthy controls are indicated in the plot. c, Enriched BACH1 expression (RNA-seq) in TCGA provisional cancer datasets. Red bar indicates median BACH1 expression level in breast cancers. d, Extended plots from KEGG pathway analyses in Fig. 4f, carried out using DAVID using Benjamini-corrected P values (FDR), of genes that are negatively correlated with BACH1 expression. The top eight most-significantly enriched pathways with FDR values (−log(FDR)) are shown for each cancer type: colorectal, liver, lung, skin, ovary, pancreas, prostate and TNBC. e, Co-expression plots of UQCRC1 and BACH1 in TCGA cancers such as prostate (n = 497), skin (n = 472), liver (n = 371) and colon (n = 379). Pearson’s (<−0.3) and Spearman’s (<−0.3) correlation coefficients are shown.

Source data

Extended Data Fig. 9 OncoPrint analyses of multiple cancer types.

ad, Heat maps demonstrating upregulation (red) or downregulation (blue) of BACH1 and ETC genes across TCGA tumours (a, prostate carcinoma TCGA provisional, n = 497; b, patients with skin cutaneous cancer TCGA provisional, n = 472; c, patients with liver cancer TCGA provisional, n = 371; d, patients with colorectal cancer TCGA provisional, n = 379).

Source data

Extended Data Table 1 List of primers for gene expression analysis using qRT–PCR and ChIP assays

Supplementary information

Supplementary Information

This file contains the uncropped data scans and Supplementary Data Table 1.

Reporting Summary

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, J., Yesilkanal, A.E., Wynne, J.P. et al. Effective breast cancer combination therapy targeting BACH1 and mitochondrial metabolism. Nature 568, 254–258 (2019). https://doi.org/10.1038/s41586-019-1005-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1005-x

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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